Journal of Nonlinear Mathematical Physics

Volume 25, Issue 3, July 2018, Pages 399 - 432

The description of reflection coefficients of the scattering problems for finding solutions of the Korteweg–de Vries equations

Authors
Pham Loi Vu
Institute of Mechanics, Vietnam Academy of Science and Technology, 264 Doi Can Street, Hanoi, Vietnam,phamloivu77@gmail.com
Received 31 October 2017, Accepted 12 February 2018, Available Online 6 January 2021.
DOI
10.1080/14029251.2018.1494777How to use a DOI?
Keywords
Necessary and sufficient conditions; left- and right-reflection coefficients; time-evolution of scattering matrix; time-dependence of the reflection coefficients; soliton-solutions
Abstract

The results of inverse scattering problem associated with the initial-boundary value problem (IBVP) for the Korteweg–de Vries (KdV) equation with dominant surface tension are formulated. The necessary and sufficient conditions for given functions to be the left- and right-reflection coefficients of the scattering problem are established. The time-dependence t, t > 0 of each element of the scattering matrix s(k, t) is found in respective sector of the k-spectral plane by expansion formulas which are constructed from the known initial and boundary conditions of the IBVP. Knowing the right-reflection coefficient calculated from the elements of s(k, t), we solve the Gelfand–Levitan–Marchenko (GLM) equation in the inverse problem. Then the solution of the IBVP is expressible through the solution of the GLM equation. The asymptotic behavior at infinity of time of the solution of the IBVP is shown

Copyright
© 2018 The Authors. Published by Atlantis and Taylor & Francis
Open Access
This is an open access article distributed under the CC BY-NC 4.0 license (http://creativecommons.org/licenses/by-nc/4.0/).

1. Introduction

After a series of papers by Fokas and Its [2], it became clear that under arbitrary boundary conditions solving the initial-boundary value problem (IBVP) for the nonlinear equations like the Korteweg-de Vries (KdV) or nonlinear Schrödinger (NLS) equations had not met the same success as solving the Cauchy problem for the KdV equation on the whole line. But there is a specific class of boundary conditions that are completely consistent with the integrability property. Under these conditions, the IBVPs are effectively embedded in the ISM schema. A number of examples of such boundary conditions were discussed in [1,3,4,11,13]. In the present paper we study inverse scattering problem (ISP) associated with the IBVP for the KdV equation:

ptpxxx+6ppx=0,x>0,t>0,(1.1)
p|x=0=0,pxx|x=0=0,(1.2)
p|t=0=p(x),p(x)|x=0,(1.3)
where the boundary conditions (1.2) are consistent at the corner point, i.e., p(0, 0) = pxx(0, 0) = 0, and the function p(x) that determines the initial condition (1.3) is required to satisfy the following conditions, which will be referred to as Conditions I:

Conditions I: The function p(x) is real-valued infinitely smooth and tends to zero at infinity in the Schwartz sense, [10], i.e., p(x) and all its derivatives decrease faster than any positive power of x−1. At x = 0 p(x) vanishes together with all derivatives, and 0p(x)dx0.

In [11] the problem of solving the considered IBVP is reduced to that of solving two ISPs. The first scattering problem (SP) is associated with the KdV equation (1.1), the second SP is self-conjugate.

In the present paper we prove the theorem of the necessary and sufficient conditions for given functions to be the left- and right-reflection coefficients of the first SP. The proof of this principal theorem is absent in [1,3,4,11 ]. Further, in Sec. 3 the scattering function of the self-conjugate SP is expressible through elements of the given scattering matrix s(k) = s(k, 0) of the first SP. Knowing the scattering function, we solve the inverse SP for finding the unknown potential self-conjugate matrix. In Sec. 4 the time-dependence of elements of the matrix s(k, t) is found in every respective sector of the k-spectral plane. The solution of the IBVP is expressible through the solution of the Galfand–Levitan–Marchenko (GLM) equation. The behaviour at infinity of t-time of the solution of the IBVP is shown. Exact soliton-solutions of the Cauchy problem for the KdV equation are presented in Sec. 5.

2. The direct and inverse SP

The IBVP (1.1)(1.2)(1.3) is associated with the SP on a half-line for a system of equations:

ddx(yyx)=(01p(x)λ0)(yyx),0x<(2.1)
with the boundary conditions as x → 0:
(y(k,x),yx(k,x))=(1,ik)+o(1),λ=k2.(2.2)

The SP for the Schrödinger equation on the whole line is well studied, therefore it is convenient to reduce the SP (2.1)(2.2) on the half-line to a problem on the whole line by continuing the potential p(x) from the positive half-line to the whole line. The potential is continued trivially by setting p(x) ≡ 0 for all x < 0. According to this way, we write system (2.1) in the form:

yx=Uy,U=(01pλ0),y=(y1(k,x),y2(k,x)),λ=k2,<x<,(2.3)
where the function p(x) satisfies Condition I, p(x) ≡ 0 for x < 0, y2 = y1x and y1 satisfies the Schrödinger equation:
yxx+p(x)y=k2y,<x<.(2.4)

The formulas (2.5)(2.23) presented below are deduced from the known facts of the scattering theory for equation (2.4) (see [5, 7]). We construct the matrix solutions of system (2.1):

E=(E+,E)=(e(k,x)e(k,x)ex(k,x)ex(k,x)),W=(W,W+)=(ω(k,x)ω(k,x)ωx(k,x)ωx(k,x)),
with the conditions for all real k=λ:
E=T(k)(eikx00eikx)+o(1)(x),(2.5)
W=T(k)+o(1)(x0),T(k)=(11ikik).(2.6)

The function e(k, x) satisfies the integral equation:

e(k,x)=eikx+xsink(sx)kp(s)e(k,s)ds,(2.7)
and can be represented as:
e(k,x)=eikx+xK(x,ξ)eikξdξ,Imk0,(2.8)
where the function K(x, ξ) is the kernel of the transformation operator and is determined from the integral equation of the Volterra type:
K(x,ξ)=12x+ξ2p(s)ds+x+ξ2dα0ξx2p(αβ)K(αβ,α+β)dβ,(2.9)
with K(x, ξ) = 0 for ξ < x. Consequently, K(x, ξ) satisfies the estimate:
|K(x,ξ)|Cx+ξ2|p(s)|ds,(2.10)
and the conditions:
2ddxK(x,x)=p(x),(ddxK(x,x))x=0=12p(0)=0.(2.11)

The function ω(k, x) satisfies the integral equation:

ω(k,x)=eikx+0sink(xs)kp(s)ω(k,s)ds,Imk0,
and can be represented as:
ω(k,x)={eikx+xK(x,ξ)eikξdξforx>0,eikxforx0,(2.12)
where K(x, ξ) ≡ 0 for ξx ≤ 0, and 2ddxK(x,x)={p(x)forx>0,0forx0.

In view of the reality of p(x), the functions K(x, ξ) and K(x, ξ) are real-valued and therefore

e(k,x)=e(k,x)¯,ex(k,x)=ex(k,x)¯,ω(k,x)=ω(k,x)¯,ωx(k,x)=ωx(k,x)¯.(2.13)

The solutions e(k, x) and ω(k, x) of system (2.3) admit from the real line an analytical continuation into the upper half-plane Imk > 0. Since the Wronskians of the solutions do not depend on x, then

W{ω(k,x),ω(k,x)}=W{e(k,x),e(k,x)}=e(k,x)ex(k,x)ex(k,x)e(k,x)=2ikforrealk0.

Hence, (e(k, x),e(−k, x)) and (ω(−k, x), ω(k, x)) for real k ≠ 0 are bases of solutions of system (2.3), therefore

(e(k,x)e(k,x)ex(k,x)ex(k,x))=(ω(k,x)ω(k,x)ωx(k,x)ωx(k,x))(s11(k)s12(k)s21(k)s22(k)),k0.(2.14)
(ω(k,x)ω(k,x)ωx(k,x)ωx(k,x))=(e(k,x)e(k,x)ex(k,x)ex(k,x))(s11(k)s21(k)s21(k)s11(k)),k0.(2.15)

The matrix s(k)=(s11(k)s12(k)s21(k)s22(k)) thus defined is called the scattering matrix of the SP (2.3), (2.6). Using (2.6), (2.14), we determine the entries sij(k) of s(k) in terms of boundary values (BVs) (e(k), ex(k)) = (e(k, 0), ex(k,0)) for k ≠ 0:

s11(k)=12e(k)+12ikex(k),s12(k)=12e(k)+12ikex(k),s21(k)=12e(k)12ikex(k),s22(k)=12e(k)12ikex(k),(2.16)
where due to (2.6) and (2.13) the BVs (e(k), ex(k)) are different from zero for all real k ≠ 0.

Remark 2.1 ([5,7]).

For real values of the parameter k the entries sij(k) of s(k) possess the properties:

  1. 1.

    The involutions: s11(k)=s11(k)¯, s21(k)=s21(k)¯, s22(k) = s11(−k), s12(k) = s21(−k).

  2. 2.

    The constraint: dets(k) = 1 = |s11(k)|2|s12(k)|2, |t(k)|2+|r˜(k)|2=1, t(k)=s111(k),

    r˜(k)=s21(k)s11(k),R˜(k)=s12(k)s11(k)=s11(k)s11(k)r˜(k),s11(k)0.

The functions s11(k) and s21(k) are called the refraction and reflection coefficients, respectively. The functions R˜(k) and r˜(k) are called the right- and left-reflection coefficients, respectively for the waves incident on the potential p(x) from the right.

Substituting (2.7) for x = 0 into (2.16) with due regard for (2.8), we obtain:

2ik(s11(k)1)=2ik(s22(k)1)=0eiksp(s)e(k,s)ds=0p(ξ)dξ0eikξdξ0p(s)K(s,ξ+s)ds,Imk0,(2.17)
2iks21(k)=2iks12(k)=0eiksp(s)e(k,s)ds=120eikξp(ξ2)dξ+0eikξdξ0ξ2p(s)K(s,ξs)ds,Imk0.(2.18)

Owing to Conditions I and (2.10) the functions sij(k) have the asymptotic behavior:

s11(k)1=s22(k)1=O(1|k|)as|k|,(2.19)
s21(k)=s12(k)=o(1|k|)as|k|.(2.20)

The integral representations for sij(k) are obtained from (2.17) and (2.18):

2iks12(k)=2iks21(k)=0eikξA(ξ)dξImk0,(2.21)
2ik(s11(k)1)=2ik(s22(k)=1)=0p(ξ)dξ0eikξB(ξ)dξImk0,(2.22)
where
A(ξ)=12p(ξ2)+0ξ2p(s)K(s,ξs)ds,B(ξ)=0p(s)K(s,ξ+s)ds.(2.23)

Lemma 2.1.

The coefficients s21(k) and s11(k) of the scattering matrix s(k) of the SP (2.3), (2.6) are infinitely differentiable for functions k ≠ 0, Imk ≥ 0. Their derivatives satisfy the estimates as k → ∞:

|s21(m)(k)|Cm,j|k|j,m,j=0,1,2,(2.24)
|s11(k)1|C|k|1,|s11(m)(k)|Cm|k|1,m=1,2,(2.25)

The functions 2iks11(k) and 2iks21(k) are continuous in the closed half-plane Imk ≥ 0.

Proof.

Owing to Conditions I and (2.10) of the potential p, the function K(x, ξ) from Eq. (2.9) is real-valued and infinitely differentiable with respect to each variable of x and ξ. Furthermore, K(x, ξ) and all its derivatives decrease faster than any positive power of x−1 and ξ−1. Therefore, the functions A(ξ) and B(ξ) defined by (2.23) are infinitely differentiable and decrease faster than any positive power of ξ−1:

|A(ξ)|CA|ξ|j,|B(ξ)|CB|ξ|j,j=0,1,2,(2.26)

Using Condition I of p and the smoothness and estimates of K, from (2.23) we get:

|A(m)(ξ)|CAm|ξ|j,m,j=0,1,2,(2.27)

Since p(0) = 0, then A(0) = 0. It can be proved by induction that

|A(m)(ξ)|ξ=0=0,m=1,2,(2.28)

Due to (2.26) the functions s21(k) and s11(k) defined by (2.21) and (2.22), respectively are infinitely differentiable for k ≠ 0. Then, from (2.21)(2.23) it follows that the functions 2iks11(k) and 2iks21(k) are continuous in the closed half-plane Imk ≥ 0. The estimate (2.24) can be proved by induction with respect to m. Indeed, by using equality (2.26)(2.28) and integrating the Fourier integral (2.21) by parts j times, we get:

2ik(ik)js21(k)=0eikξA(j)(ξ)dξ,
whence it follows that |s21(k)|C0,j|k|j, j = 0, 1, 2,... and m = 0. Thus, the estimate (2.24) is true for j = 0, 1, 2,... and m = 0. Supposing that the estimate (2.24) is proved for m − 1:
|s21(m1)(k)|Cm1,j|k|j.

Differentiating (2.21) m times and using the Leibniz’s Rule, we obtain:

(2iks21(k))(m)=2iks21(m)(k)+2ims21(m1)(k)=0eikξ(iξ)mA(ξ)dξ.(2.29)

Integrating the integral in the right-hand side of (2.29) by parts j − 1 times, using (2.20), (2.26)(2.28), we calculate:

0eikξ(iξ)mA(ξ)dξ=1ikeikξ[(iξ)mA(ξ)]ξ=0,1ik0eikξ[(iξ)mA(ξ)]dξ==1(ik)j1eikξ[(iξ)mA(ξ)]ξ=0,(j2)1(ik)j10eikξ[(iξ)mA(ξ)](j1)dξ=1(ik)j10eikξ[(iξ)mA(ξ)]j1dξ.

Using the right-hand side of the above equality and the induction hypothesis for s21(m1)(k), from (2.29) we obtain the estimate (2.24) for s21(m)(k).

The first inequality of (2.25) is deduced from estimate (2.19). We prove the second inequality of (2.25). Differentiating (2.22) m times, yield:

(2iks11(k))(m)=2iks11(m)(k)+2ims11(m1)(k)=0eikξ(iξ)mB(ξ)dξ.

Using (2.26) and the induction hypothesis for s11(m1)(k), from the latest equality we obtain the second estimate of (2.25). The number B(0) is nonzero. Indeed, using (2.9) and Conditions I of p, we calculate:

B(0)=0p(s)K(s,s)ds=120p(s)dssp(ξ)dξ=12{0p(s)ds}2120p(s)ds0sp(ξ)dξ=12{0p(s)ds}2B(0).

Hence, B(0)=14{0p(s)ds}20. On account of this fact, the function s11(m)(k) obeys the estimate (2.25) as the first power of k−1. Lemma 2.1 is proved.

The following remark is deduced from Lemma 2.1, Remark 2.1 and properties of the solutions of problems (2.3), (2.5) and (2.3), (2.6).

Remark 2.2 ([5, 7]).

  1. 1.

    The analytic continuation of the function s11(k) with respect to k from the real axis into the upper half-plane Imk > 0 can have a finite number of simple zeros on the positive imaginary axis at kj = j, μj > 0, j = 1,...,N;

  2. 2.

    There is a one-to-one correspondence between the simple pole j, μj > 0 of r˜(k) Imk > 0 and the simple negative eigenvalues μj2, j = 1,...,N of the SP (2.1)(2.2).

Suppose that the potential function p(x) is to be subjected to the following restriction, which will be referred to as the condition II.

Condition II. The potential function p(x) is to be subjected to the condition that Eq. (2.4) must not have a discrete spectrum. Then by Remark 2.2, s11(k) ≠ 0 for all k, Imk > 0.

The scattering matrix s(k) gives complete information about the continuous spectrum of the Schrödinger operator. By Condition II, Remark 2.2 and the dispersion relation, we can show that essentially, all information about s(k) is contained in the right-reflection coefficient R˜(k). Denote by 𝒮 the class of all real-valued functions satisfying Conditions I and II. From (2.14) and (2.15) it follows that the functions:

[s11(k)]1e(k,x)=ω(k,x)+r˜(k)ω(k,x),(2.30)
[s11(k)]1ω(k,x)=e(k,x)+R˜(k)e(k,x),(2.31)
are bounded solutions of system (2.1) with the potential p(x) belonging to the class 𝒮. The behaviour of the function ks11(k) as k → 0 is closely related to that of functions R˜(k) and r˜(k) as k → 0. In fact, upon rewriting (2.30) and (2.31) and letting k → 0, gives:
0=e(0,x)limk0ks11(k){R˜(k)+1}=ω(0,x)limk0ks11(k){r˜(k)+1}.Consequently,limk0ks11(k){R˜(k)+1}=limk0ks11(k){r˜(k)+1}=0.(2.32)

From condition (2.32) we have the estimate: k{s11(k) + s21(k)} = o(1) as k → 0.

To fulfill this estimate, the following condition must be satisfied:

s11(k)+s21(k)=O(1)ask0,(2.33)

Conversely, if the condition (2.33) is fulfilled, then the condition (2.32) is satisfied.

Lemma 2.2.

The left-reflection coefficient r˜(k) of the SP for system (2.1) with the potential p(x) belonging to the class 𝒮 and boundary condition (2.2) obeys the following conditions:

  1. 1.

    For all k, Imk ≥ 0, the function r˜(k) is completely continuous and infinitely differentiable. r˜(k) and all its derivatives decrease faster than any positive power of k−1, and

    r˜(k)¯=r˜(k),|r˜(k)|<1forrealk0.(2.34)

    If the residues of the functions s11(k) and s21(k) at k = 0 are different from zero, then r˜(0)=1.

  2. 2.

    The function r˜(k) admits the Fourier integral representation:

    r˜(k)=0eikxr(x)dxforallk,Imk0,(2.35)
    where r(x) is a completely continuous and rapidly decreasing function, which is defined by the inverse Fourier transform:
    r(x)=12πeikxr˜(k)dk,forx>0.(2.36)

    The function r(x) is infinitely differentiable:

    r(m)(x)=12π(ik)meikxr˜(k)dkforx>0,m=1,2,

    In addition r(x) and r(m)(x) are real-valued functions vanishing at x = 0.

Proof.

The validity of assertion 1 for k ≠ 0, Imk ≥ 0 follows from Lemma 2.1 and Remarks 2.1, 2.2. We prove the smoothness of r˜(k) and its derivatives at k = 0. From (2.16) we have the relations:

2iks21(k)=ike(k)ex(k),2iks11(k)=ike(k)+ex(k).

These relations make clear that the entries sij(k), i, j = 1, 2 of s(k) have generally a simple pole at k = 0. The residues of functions 2iks11(k) and 2iks21(k) at k = 0 are defined by:

Res[2iks11(k),0]=Res[2iks21(k),0]=ex(0).(2.37)

If ex(0) ≠ 0, then the function r˜(k) and its derivatives are continuous and smooth at k = 0:

limk0r˜(k)=limk02iks21(k)2iks11(k)=limk0ex(k)ex(k)=ex(0)ex(0)=1.

If ex(0) = 0, i.e., the residues (2.37) are zero, then the functions s11(k) and s21(k) are continuous and analytic at k = 0. The constraint condition |s11(k)| 2 − |s21(k)| 2 = 1 implies that |s11(k)| > 1 for k ∈ ℝ. Therefore, the function r˜(k) and its derivatives are continuous and smooth at k = 0.

By Lemma 2.1 the functions r˜(k) and (ik)mr˜(k) are analytic in the upper half-plane Imk ≥ 0 and rapidly decrease at infinity for any nonnegative integer m. On account of this fact, the function r˜(k) admits the Fourier integral representation (2.35) of a function r(x) for x > 0. Since the Fourier transform (2.35) and its inverse Fourier transform (2.36) maps 𝒮 onto 𝒮 mutually continuously one-to-one [10], then r(x) and its derivatives r(m)(x) for x > 0 defined by (2.36) are rapidly decreasing functions. Due to (2.34) these functions are real-valued. We have from (2.36):

r(m)(0)=12π(ik)mr˜(k)dk;m=0,1,2,
where the integrand function G(k)=(ik)mr˜(k) and its derivatives G(j)(k) are analytic in the upper half-plane Imk ≥ 0 and rapidly decrease at infinity for nonnegative integer m, and G(∞) = G(j)(∞) = 0, j = 1, 2,... Hence, the function G(k) is analytic in the closed upper half-plane Imk ≥ 0, therefore, by the Cauchy’s Theorem the right-hand side of the last formula vanishes, i.e., r(m)(0) = 0, m = 0, 1, 2,... Lemma 2.2 is proved.

The further conditions of R˜(k) are deduced from Lemma 2.1, Remarks 2.1 and 2.2.

Lemma 2.3.

The right reflection coefficient R˜(k) of the SP for system (2.1) with the potential p(x) belonging to the class 𝒮 and boundary condition (2.2) obeys the following conditions:

  1. 1.

    The function R˜(k) is completely continuous and infinitely differentiable for all real k ∈ (−∞, ∞). R˜(k) and all its derivatives decrease faster than any positive power of k−1, and

    R˜(k)¯=R˜(k),|R˜(k)|=|r˜(k)|<1forrealk0;(2.38)

    If the residues of the functions s11(k) and s12(k) at k = 0 are different from zero, then R˜(0)=1;

  2. 2.

    For real k the function R˜(k) admits the Fourier transform:

    R˜(k)=eikxR(x)dx,(2.39)
    where R(x) is a real completely continuous and rapidly decreasing function, which is defined by the inverse Fourier transform:
    R(x)=12πeikxR˜(k)dk.(2.40)

    The function R(x) is infinitely differentiable:

    R(m)(x)=12π(ik)meikxR˜(k)dk,m=1,2,
    where the Fourier transform (2.39) and its inverse Fourier transform (2.40) maps 𝒮 onto 𝒮 mutually continuously one-to-one, [10]. Due to this fact and (2.38) R(x) and R(m)(x), m = 1, 2,... are rapidly decreasing and real-valued functions.

To recover the SP (2.1)(2.2) from the right-reflection coefficient R˜(k), we derive the fundamental integral equation connecting the given R˜(k) with the kernels of the transformation operator. The following equation is derived from (2.31) (see [5, 7]):

R(x+y)+K(x,y)+xR(y+ξ)K(x,ξ)dξ=0fory>x,(2.41)
which is the Gelfand–Levitan–Machenko equation in the case of a purely continuous spectrum. In Eq. (2.41) x is a parameter, R(x + y) is a known function satisfying conditions enumerated in Lemma 2.3, and K(x, y) is an unknown function of y for every x ∈ [0, ∞). Owing to conditions (2.38)(2.40) of the function R(x + y), Eq. (2.41) has a unique solution K(x, y) either in L1[x, ∞) or L2[x, ∞).

Analogously, the following fundamental integral equation is derived from (2.30):

r(x+y)+K(x,y)+xr(y+ξ)K(x,ξ)dξ=0fory<x(2.42)

Owing to conditions (2.34)(2.36) of the known function r(x + y) Eq. (2.42) has a unique solution in either L1(−∞, x] or L2(−∞, x].

We use Eq. (2.41) to extract information on the solution K(x, y), y > x from the conditions of the known function R(x) in this equation. In fact, the function K(x, y), y > x satisfies conditions, which are analogous to the conditions of the function R(x). As has been proved in [5, 7] that the solution K(x, y) of Eq. (2.41) is the kernel of the transformation operator and the function constructed from K(x, y):

e(k,x)=eikx+xK(x,ξ)eikξdξ,Imk0(2.43)
satisfies the Schrödinger equation (2.4):
e(k,x)+p(x)e(k,x)=k2e(k,x)(2.44)
with the potential p(x) constructed from the solution of Eq. (2.41) by the formula:
p(x)=2ddxK(x,x)forx>0.(2.45)

Substituting (2.43) into (2.7), we obtain integral Eq. (2.9) with the constructed potential (2.45). Since the solution of (2.9) is unique, then K(x, y) satisfies conditions (2.11).

By an argument analogous to the previous one, we can prove that the function constructed from the solution K(x, y) of Eq. (2.42):

ω(k,x)=eikx+xK(x,ξ)eikξdξ,Imk0,(2.46)
satisfies the Schrödinger equation (2.4):
ω(k,x)+p(x)ω(k,x)=k2ω(k,x)(2.47)
with the potential p(x) constructed by p(x)=2ddxK(x,x) for x > 0.

The principal mathematical problem in inverse problem consists in the proof of the fact that under conditions (properties) of functions s11(k), s21(k) enumerated in Remark 2.1 and Lemma 2.1, the procedure actually leads to the same differential equation, i.e., p(x) ≡ p(x). This leads to describing the scattering data, i.e., to establishing the necessary and sufficient conditions of functions r˜(k) and R˜(k) to be the left- and right-reflection coefficients of the considered problem.

Theorem 2.1.

Suppose that the functions s11(k) and s21(k), −∞ < k < ∞ satisfy the conditions enumerated in Remark 2.1, Lemma 2.1, condition (2.33) and the function s11(k) admits an analytical continuation into the upper half-plane Imk > 0 and has no zeros there. Then

  1. 1.

    The functions e(k, x) and ω(k, x) constructed from the solutions K(x, y) ∈ Lj[x, ∞) and K(x, y) ∈ Lj(−∞, x], j = 1, 2 of Eqs. (2.41) and (2.42), respectively satisfy the same Shrödinger equation (2.4) with the constructed potential:

    p(x)p(x)=2ddxK(x,x),(2.48)
    p(0)=2ddx(K(x,x))x=0=0.(2.49)

  2. 2.

    The conditions of functions s11(k) and s21(k) are both necessary and sufficient for the ratios of the type:

    r˜(k)=s21(k)s11(k).Imk0andR˜(k)=s12(k)s11(k),<k<
    to be the left-reflection and right-reflection coefficients of the SP for one and the same system (2.1) with boundary condition (2.2) and constructed potential (2.48) belonging to the class 𝒮. The Schrödinger equation (2.4) is restored precisely from R˜(k).

Proof.

The functions given by (2.43) and (2.46) admit analytic continuations into the upper half-plane Imk > 0. Extend the domain of the function K(x, y) by setting: K(x, y) = 0 for y > x.

Further, we put:

Φ˜(x,y)=K(x,y)+r(x+y)+xr(y+ξ)K(x,ξ)dξforallrealy.(2.50)

Due to Eq. (2.42), Φ˜(x,y)=0 for y < x. For y > x:

Φ˜(x,y)=r(x+y)+xr(y+ξ)K(x,ξ)dξ,|Φ˜(x,y)|dy=x|Φ˜(x,y)|dy<.

Multiply both sides of equality (2.50) by eiky, then integrate with respect to y and apply the inverse Fourier formula to r(x + y):

xΦ˜(x,y)eikydy=xK(x,y)eikydy+s21(k)s11(k)eikx+s21(k)s11(k)xK(x,ξ)eikξdξ.

Adding eikx to the right- and left- hand sides of the last equality and using (2.46), gives:

eikx+xΦ˜(x,y)eikydy=ω(k,x)+s21(k)s11(k)ω(k,x).(2.51)

Multiply (2.51) by s11(k), then

s11(k)ω(k,x)+s21(k)ω(k,x)=e*(k,x),(2.52)
e*(k,x)=s11(k){eikx+xΦ˜(x,y)eikydy}.(2.53)

Replacing k by −k in (2.52), yields:

s11(k)ω(k,x)+s21(k)ω(k,x)=e*(k,x).(2.54)

Solving the system (2.52), (2.54) for ω(k, x), using conditions of s11(k) and s21(k), gives:

ω(k,x)=s11(k)e*(k,x)s21(k)e*(k,x)(2.55)

In order to prove the identity (2.48), we need to prove that

e*(k,x)e(k,x)(2.56)

Indeed, if identity (2.56) will be proved, then due to (2.15) and (2.44), it follows from (2.55):

ω(k,x)=s11(k)e(k,x)+s21(k)e(k,x)=[k2p(x)]ω(k,x),
because of (2.47), we obtain identity (2.48).

To prove identity (2.56), certain properties of the function e*(k, x) should be established.

  1. a./

    The function e*(k, x) defined by formula (2.53) admits an analytic continuation into the upper half-plane Imk > 0. Using estimate (2.25) for large k, from (2.53) we obtain the following estimate:

    |e*(k,x)eikx|=O(exImk|k|)ask.(2.57)

  2. b./

    The function ke*(k, x) is continuous in the closed upper half-plane Imk ≥ 0 and in the neigh-bourhood of the point k = 0, this function satisfies uniformly the estimate:

    ke*(k,x)=o(1)ask0.(2.58)

    By the Lemma 2.1, the continuity of function ke*(k, x) follows from that of function ks11(k). In proving estimate (2.58) two case may arise.

    1. (1)

      The function s11(k) is bounded in neighborhood of point k = 0, in which case the function e*(k, x) is also bounded in a neighborhood of this point, therefore, ke*(k, x) → 0 as k → 0. Hence, the estimate (2.58) follows from the continuity of ke*(k, x).

    2. (2)

      The function s11(k) is not bounded in a neighborhood of the point k = 0, in which case there exists a sequence kn → 0 such that s11(kn) → ∞. It follows from condition (2.32) that limn→∞ kns11(kn) = O(1), limns21(kn)s11(kn)=1.

      Putting k = kn in (2.51), passing to the limit as kn → 0, gives: 1+xΦ˜(x,y)dy=ω(0,x)ω(0,x)=0. Therefore, limn→∞ kne*(kn, x) = 0, which proves estimate (2.58) in this case too, due to the continuity of the function ke*(k, x).

  3. c./

    e*(k,x)eikxL2(,).(2.59)

    Since the function s11(k) satisfies the first estimate of (2.25) for large k, therefore it suffices to show that e*(k, x) is square integrable in a neighborhood of the point k = 0. We need to show that e*(k, x) is a bounded function in a neighborhood of the point k = 0. We write equality (2.52) in the form:

    e*(k,x)=[s11(k)+s21(k)]ω(k,x)+s21(k)[ω(k,x)ω(k,x)].

    Taking into account that: ks21(k) = o(1) as k → 0, we get the estimate:

    s21(k)[ω(k,x)ω(k,x)]=2is21(k)sinkx2ixK(x,ξ)s21(k)sinkξdξ=2isinkxk[ks21(k)]2ixK(x,ξ)sinkξk[ks21(k)]dξ=O(x)+O(x|ξ||K(x,ξ)||dξ|)=O(1)ask0.

From this estimate and assumption (2.33), it follows that the function e*(k, x) is bounded in a neighborhood of the point k = 0. Thus, the assertion c./ is proved.

Now we can prove identity (2.56). Consider a function [e*(k, x) − eikx]eiky for y < x, which is analytic in the upper half-plane Imk > 0. Integrating this function along the contour represented in Figure 1. Due to properties a./ and b./ of e*(k, x), the integrals along the small and large semi-circles tend to zero as ρ → 0 and R → ∞. Hence, by also the property c./ and Cauchy’s Theorem: limRRR[e*(k,x)eikx]eikydk=0 for y < x. Due to the property c./, there exists a function K*(x,y):12πx[e*(k,x)eikx]eikydk=K*(x,y) for y > x. Hence, K*(x, y), y > x is the inverse Fourier transformation of the function e*(k · x) − eikx, which belongs to the space L2(−∞, ∞) and

Fig. 1.

The path of integration.

e*(k,x)eikx=xK*(x,y)eikydy.(2.60)

The Fourier transform (2.60) and its inverse Fourier transform map L2[x, ∞) onto L2[x, ∞) mutually continuously one-to-one, therefore K*(x, y) ∈ L2(x, ∞). Dividing equality (2.55) by s11(k):

1s11(k)ω(k,x)eikx=xK*(x,ξ)eikξdξs21(k)s11(k)[eikx+xK*(x,ξ)eikξdξ].(2.61)

The function s11(k) is analytical in the upper half-planeImk > 0 and has no zero there, then the function:

h(k,x)=[1s11(k)ω(k,x)eikx]eiky=[e*(k,x)s21(k)s11(k)e*(k,x)eikx]eiky=O(e(yx)Imk|k|)as|k|,y>x,
is also analytical in this upper half-plane.

Multiply both sides of identity (2.61) by 12πeiky, and integrate with respect to k, then due to the analytical property of the function h(k, x) and its estimate, the left-hand side of the obtained equality vanishes when y > x, while the right-hand side of this equality gives:

R(x+y)+K*(x,y)+xR(y+ξ)K*(x,ξ)dξ.

Therefore, the function K*(x, y) satisfies the integral fundamental equation (2.41). From the uniqueness of a solution of Eq. (2.41), it follows that K*(x, y) = K(x, y). Then by (2.43) and (2.60) e*(k, x) ≡ e(k, x). Since the solution of equation (2.41) is the kernel of the transformation operator, then it satisfies Eq. (2.9). Due to the uniqueness of solution of Eq. (2.9), the solution K(x, y) of Eq. (2.41) is related to the potential by formula (2.48) and satisfies condition (2.49). Therefore, the restored potential (2.48) belongs to the class 𝒮. Thus, the Schrödinger equation (2.4) is restored with the potential (2.48) satisfying condition (2.49). In addition, the negative spectrum of the restored equation is absent. Thus, the first assertion is proved.

We proceed to prove the second assertion. Let s11(k) and s21(k), −∞ < k < ∞ be any given functions satisfying the sufficient conditions enumerated in Theorem 2.1. We prove that these given functions are the scattering data of the SP (2.1)(2.2) with the restored potential belonging to the class 𝒮. In fact, consider the SP with the restored potential (2.48) satisfying condition (2.49). Let s˜11(k) and s˜21(k) be scattering data, and let K˜(x,y), y > x be the kernel of the transformation operator of this considered SP. Then the function K˜(x,y) satisfies the equation for y > x:

R^(x+y)+K˜(x,y)+xR^(y+ξ)K˜(x,ξ)dξ=0,R^(x+y)=12πs˜21(k)s˜11(k)eik(x+y)dk.(2.62)

Since the solutions K(x, y) and K˜(x,y) of Eqs. (2.41) and (2.62), respectively, satisfy the same integral equation (2.9) with the potential (2.48), then owing to the uniqueness of solution of Eq. (2.9), we obtain the identity: K(x,y)K˜(x,y) for y > x. Taking this identity into account and subtracting Eq. (2.62) from Eq. (2.41), we have:

R(x+y)R^(x+y)+x[R(y+ξ)R^(y+ξ)]K(x,ξ)dξ=0,fory>x,(2.63)

For a sufficiently large positive x = x0, the integral operator in homogeneous equation (2.63) is a contracting operator in the space of functions bounded on the interval (x0, ∞). Hence, for xx0 and yx:R^(x+y)R(x+y), from which it follows that, R^(x)=R(x) for x ≥ 2x0. Therefore:

R(x+y)R^(x+y)+x2x0y[R(y+ξ)R^(y+ξ)]K(x,ξ)dξ=0fory>x.(2.64)

For fixed x0 and y, Eq. (2.64) is a Volterra homogeneous, therefore R(x+y)R^(x+y)0 for any x and yx, i.e., R^(x) is identical to R(x), or, in expanded form:

12πs21(k)s11(k)eikxdk=12πs˜21(k)s˜11(k)eikxdk,(2.65)

Then owing to uniqueness of expansion of a function in the Fourier integral: s˜21(k)s˜11(k)s21(k)s11(k). Since the functions s11(k) and s˜11(k) have no zeros in the upper half-plane Imk > 0, then s11(k) and s˜11(k) are uniquely restorable from |R˜(k)|=|s21(k)s11(k)|, whence it follows that:

s11(k)s˜11(k),s21(k)s˜21(k),s12(k)=R˜(k)s11(k),r˜(k)=s11(k)s11(k)R˜(k).

Thus, the ratios: R˜(k)=s12(k)s11(k) and r˜(k)=s21(k)s11(k) satisfy conditions enumerated in Lemmas 2.2 and 2.3, respectively. Therefore, they are right- and left-reflection coefficients of the considered SP, and Eq. (2.4) is restored precisely from R˜(k). Thus, the sufficiency of conditions of right-reflection and left-reflection coefficients R˜(k) and r˜(k) is proved. The necessity of conditions of these coefficients has been established by Lemmas 2.1, 2.2 and 2.3. The proof of Theorem 2.1 is completed.

The KdV equation (1.1) is derived from the Lax condition for compatibility of two systems:

Yx=UY,Y=(y1,y2),(2.66)
Yt=VY,(2.67)
where U and V are the given matrices having the form:
U=(01pλ0),V=(px4λ2ppxx(4λ+2p)(pλ)px),k2=λ.

The potential p(x, t) for every t > 0 belongs to the class 𝒮, therefore the systems (2.66) and (2.67) are compatible, i.e.,

y2xt=y2txory1xxt=y1xtx.

The above equality will be referred to as the compatibility condition for systems (2.66) and (2.67). It is easy to verify that this compatibility condition is equivalent to the KdV equation (1.1). The boundary conditions (1.2) are satisfied if and only if the system (2.67) along the line x = 0 takes the form:

Yt=V0Y,V0=(px(0,t)4λ4λ2px(0,t)),λ=k2.(2.68)

Upon differentiating equality (2.14) with respect to t, we have the equality:

Et(k,x,t)=Wt(k,x,t)s(k,t)+W(k,x,t)st(k,t),(2.69)

Substituting matrix expressions Et and Wt into equality (2.69), using conditions (1.2), we derive the system of linear differential equations governing the time-dependence of s(k, t):

st(k,t)=4ik3[s,σ3]+px(0,t)σ1s(k,t),σ1=(0110),σ3=(1001).(2.70)

The time-dependence of the scattering matrix s(k, t) defined by system (2.70) implicitly depends on the time t. This is the main difference between the IBVP (1.1)(1.2)(1.3) and the Cauchy problem, therein lies the difficulty in passing from the Cauchy problem to this IBVP. The system (2.70) is undetermined, because the function px(0, t) entering coefficients of this system is unknown. In Sec. 3 we shall prove that the unknown object px(0, t) can be expressed through entries of the given s(k).

3. The self-conjugate problem

The system (2.68) describes the time-evolution of the eigenfunction for the boundary point x = 0. Using the linear change of dependent variables: Y (k, t) = J(k)(μ, t), we reduce (2.68) to the form:

y˜t(k,t)=(J1VJ)y˜(k,t),(3.1)
where J is a matrix, which is to be taken so that J−1σ3J = σ1. From the equality: 1 = σ3J, it follows that the matrix J takes the form: J=(ααββ), α and β are arbitrary constants and α ≠ 0, β ≠ 0. Taking α = 1 and β = ±ik, which are the roots of the equation: β +k2 = 0, we reduce system (3.1) to the form:
y˜t(μ,t)=(±4iμσ3+px(0,t)σ1)y˜(μ,t),μ=k3.(3.2)

Let β = −ik, then the matrix J in (3.1) coincides with the matrix T(k) defined by:

J(k)=(11ikik)=T(k),Y(k,t)=T(k)y˜(μ,t),μ=k3,0t<.(3.3)

Thus, by virtue of the linear change of dependent variables given by (3.3), we lead system (3.2) into the system of first-order ordinary differential equations on the half-line:

iσ3y˜t(μ,t)+C(t)y˜(μ,t)=4μy˜(μ,t),μ=k3,0t<,(3.4)
where the potential matrix C(t) is self-conjugate:
C(t)=(0c1(t)c2(t)0)=C(t)¯,(3.5)
c1(t) = −ipx(0, t), c2(t) = ipx(0, t), the potential px(0, t) is a real-valued function.

Denote (μ, t) by y(μ, t) and consider the self-conjugate SP generated by system (3.4):

iσ3yt(μ,t)+C(t)y(μ,t)=4μy(μ,t),μ=k3,0t<(3.6)
with the boundary condition at the origin t = 0:
y1(μ,0)=y2(μ,0).(3.7)

We also consider the problem for system (3.6) for real μ with boundary conditions at infinity:

y1(μ,t)=A(μ)e4iμt+o(1),y2(μ,t)=B(μ)e4iμt+o(1),(t).(3.8)

Assume that the potential function px(0, t) in (3.5) satisfies the estimate:

|px(0,t)|Const.1+t1+ε,t>0,ε>0.(3.9)

The problems (3.6), (3.8) and (3.6)(3.7) with the potential satisfying estimate (3.9) have been solved in [14]. Since the potential matrix (3.5) is a particular case of the potential self-conjugate matrix of the problem investigated in [14], then the following Propositions are deduced from corresponding assertions proved in [14] without proving.

Proposition 3.1.

The problem (3.6)(3.7) has a unique bounded solution (y1(μ, t), y2(μ, t)) for real μ such that for any given number A(μ) there exists a unique number B(μ) defined from A(μ) so that the asymptotics (3.8) are satisfied. This solution has the representation:

y1(μ,t)=A(μ)e4iμt(1+0H11(t,t+ξ)e4iμξdξ)+B(μ)e4iμt0H21(t,t+ξ)e4iμξdξ,y2(μ,t)=A(μ)e4iμt0H21(t,t+ξ)e4iμξdξ+B(μ)e4iμt(1+0H11(t,t+ξ)e4iμξ),(3.10)
where the matrix H(t,s)=(H11(t,s)H21(t,s)H21(t,s)H11(t,s)), 0 ≤ ts is an analog of kernel of the transformation operator. The functions H11(t, s) and H21(t, s) satisfy the estimate:
|Hj1(t,s)|Const.1+(t+s)1+ε,0ts,j=1,2.(3.11)

Using condition (3.5), wherein the functions c1(t), c2(t) are pure imaginary functions, from the integral equations for kernels we obtain that the solutions Hjk(t, s), j, k = 1, 2 of these equations are real-valued functions, and

H11(t,s)=H22(t,s),H21(t,s)=H12(t,s),0ts.

The kernel function H21(t, t + ξ) is related to the potential px(0, t) by the formula:

2H21(t,t)=px(0,t)fort>0.(3.12)

Proposition 3.2.

There exist the bounded Jost solutions e(1)(μ,t)=(e1(1)(μ,t),e2(1)(μ,t)) and e(2)(μ,t)=(e1(2)(μ,t),e2(2)(μ,t)) of system problem (3.6) with the boundary conditions at infinity:

e(1)(μ,t)e4iμt=(1+o(1),o(1)),Imμ0(t),e(2)(μ,t)e4iμt=(o(1),1+o(1)),Imμ0(t).

By an argument analogous to problem (3.6)(3.7) on the half-line 0 ≤ t < ∞, for any t ≥ 0 we consider the problem generated by system (3.6) on a half-line tτ < ∞ with the boundary condition at τ = t:

z1(μ,t,t)=z2(μ,t,t),(3.13)
and the problem for this system with the boundary condition at infinity:
z1(μ,τ,t)=A(μ,t)e4iμ(τt)+o(1),z2(μ,τ,t)=B(μ,t)e4iμ(τt)+o(1),(τ),tτ<.

Definition 3.1.

The one-to-one correspondence between numbers A(μ, t) and B(μ, t) determines the scattering function S(μ, t): S(μ, t)A(μ, t) = B(μ, t), −∞ < μ < ∞ for the SP generated by system (3.6) on the half-line: tτ < ∞ with condition (3.13).

By Definition 3.1 and using (3.10), (3.13), we derive the factorization of S(μ, t):

S(μ,t)=(1+H(μ,t))(1+H+(μ,t))1,<μ<,(3.14)
where H(μ,t)=0H(t,ξ)e4iμξdξ, H+(μ,t)=0H(t,ξ)e4iμξdξ, H(t, ξ) = H11(t, t + ξ) − H21(t, t + ξ) and Hj1(t, t + ξ), j = 1, 2 satisfies estimate (3.11).

By virtue of estimate (3.11) and the self-conjugate property of matrix (3.5), the function H(t, ξ) is absolutely integrable with respect to ξ, the numerator 1 + H(μ, t) and denominator 1 + H+(μ, t) of ratio (3.14) are different from zero and analytic in the half-planes Im μ ≤ 0 andIm μ ≥ 0, respectively. There exists an absolutely integrable with respect to ξ function K(t, ξ) such that

(1+0H(t,ξ)e4iμξdξ)1=1+0K(t,ξ)e4iμξdξ.

Furthermore, for any t ≥ 0 the scattering function S(μ, t) possesses the property:

S(μ,t)¯=S1(μ,t)=S(μ,t),<μ<.(3.15)

Proposition 3.3.

For any t ≥ 0 the scattering functions S(μ, t) − 1 and S−1(μ, t) − 1 for the SP (3.6), (3.13) are the Fourier transformations:

S(μ,t)=1+F(μ,t)=1+f(t,ξ)e4iμξdξ,(3.16)
S1(μ,t)=1+G(μ,t)=1+g(t,ξ)e4iμξdξ,(3.17)
where f(t, ξ) and g(t, ξ) are real-valued functions defined by the formulas:
f(t,ξ)=H(t,ξ)+K(t,ξ)+0K(t,ζ)H(t,ζζ)dζ,(3.18)
g(t,ξ)=H(t,ξ)+K(t,ξ)+0K(t,ζ)H(t,ξζ)dζ.(3.19)

Proposition 3.4.

There exists uniquely a solution φ(μ, t) = (φ1(μ, t), φ2(μ, t)) of (3.6) with the initial condition: φ1(μ, 0) = φ2(μ, 0) = 1. The function φ(μ, t) is entire analytic of μ, and:

D1(μ)ϕ(μ,t)=e(1)(μ,t)+S(μ)e(2)(μ,t),Imμ0,N1(μ)ϕ(μ,t)=S1(μ,t)+e(1)(μ,t)+e(2)(μ,t),Imμ0,(3.20)
where N(μ)=e1(1)(μ,0)e2(1)(μ,0)=1+H(μ,0)0 for Im μ ≤ 0,
D(μ)=e2(2)(μ,0)e1(2)(μ,0)=N(μ)¯0forImμ0,N(μ)=1+o(1)(|μ|),D(μ)=1+o(1)(|μ|),S(μ)=N(μ)D(μ),S1(μ)=S(μ)¯=S(μ),<μ<,(3.21)

Proposition 3.5.

For any t ≥ 0 the functions f(t, ξ) and g(t, ξ) defined by (3.18) and (3.19) are closely related to f(ξ) and g(ξ), respectively by the formula:

f(t,ξ)=f(ξ2t)forξ<0,g(t,ξ)=g(ξ+2t)forξ>0,f(ξ2t)=g(ξ+2t)forξ>0.(3.22)

The functions f(−ξ) and g(ξ), ξ > 0 satisfy the estimate of the type (3.9):

|f(ξ)|const1+ξ1+ε,|g(ξ)|const1+ξ1+ε,ξ>0,ε>0.

Proposition 3.6.

For every fixed t ≥ 0 the following Fredholm system:

{H22(t,t+ξ)+0H21(t,t+ζ)g(2t+ξ+ζ)dζ=0,g(2t+ξ)+H21(t,t+ξ)+0H22(t,t+ζ)g(2t+ξ+ζ)dζ=0(3.23)
has a unique solution H22(t, t + ξ), H21(t, t + ξ) in the space L1[0, ∞).

Proposition 3.7.

For the given function S(μ) to be the scattering function for the self-conjugate problem (3.6)(3.7), it is necessary and sufficient that there exists a function S(μ, t) such that S(μ) = S(μ, 0) and

  1. 1.

    the function S(μ, t) admits the factorization of the form (3.14);

  2. 2.

    for any t ≥ 0 the functions S(μ, t) − 1 and S−1(μ, t) − 1 are the Fourier transformations (3.16) and (3.17) of the functions (3.18) and (3.19), respectively;

  3. 3.

    for any t ≥ 0 the functions f(t, ξ) and g(t, ξ) are closely related to f(−ξ) and g(ξ) by formula (3.22), respectively. f(−ξ) and g(ξ) are real-valued functions satisfying the estimate of type (3.9).

To solve the inverse SP (3.6)(3.7) for finding the unknown object px(0, t), we need to find the unknown scattering function S(μ) of this problem first. To find S(μ), we express the function S(μ) through known elements of the scattering matrix s(k) of the SP (2.1)(2.2). This is the key step to recover the potential matrix (3.5), i.e., the system (3.6).

The matter is that the first SP (2.1), (2.2) and the second SP (3.6)(3.7) are formulated on two different spectral planes. The first SP is considered on the k-plane, whereas the second SP is considered on the μ-plane, (μ = k3). To compare functions on the k-plane with those on the μ-plane, we use the conjugation contour. The contour Im μ = Imk3 = 0 splits into a system of rays {nj}j=16 coming from the origin of the k-plane with the slope angles γj = (j − 1)π/3 with respect to the positive direction of the line Imk = 0. We denote by Nj the interior sector confined between the rays nj+1 and nj (see the Figure 2). Taking into account that arg μ = argk3 = 3argk, we have:

Fig. 2.

The conjugation contour.

IfkN1N3N5,thenImμ>0.(3.24)
IfkN2N4N6,thenImμ<0.(3.25)

We compose the matrix functions ψ+(μ, t) and ψ(μ, t) according to the rule:

ψ+(μ,t)=(ϕ(μ,t)D(μ),e(2)(μ,t))e4iμtσ3forkN1N3N5, (3.26)
ψ(μ,t)=(e(1)(μ,t),ϕ(μ,t)N(μ))e4iμtσ3forkN2N4N6.(3.27)

Since the matrix functions (ϕ(μ,t)D(μ),e(2)(μ,t)) and (e(1)(μ,t),ϕ(μ,t)N(μ)) satisfy system (3.6), then the matrix functions ψ+(μ, t) and ψ(μ, t) satisfy system (2.70). Hence, all the matrix functions s(k, t), ψ+(μ, t) and ψ(μ, t) satisfy the same system (2.70). It is easily persuaded that the functions ψ+(μ, t) and ψ(μ, t) are the fundamental solutions of system (2.70). Denote the columns of s(k, t) by s1(k, t) = (s11(k, t), s21(k, t)), s2(k, t) = (s12(k, t), s22(k, t)). Both columns are solutions of system (2.70), but they are defined in different half-planes, therefore we consider them separately for convenience. Every column solution sj(k, t), j = 1, 2 can be represented in a form:

s1(k,t)=ψ+(μ,t)(α(k,t)β(k,t))forkN1N3,(3.28)
s1(k,t)=ψ(μ,t)(α(k,t)β(k,t))forkN2,(3.29)
s2(k,t)=ψ(μ,t)(α(k,t)β(k,t))forkN4N6,(3.30)
s2(k,t)=ψ+(μ,t)(α(k,t)β(k,t))forkN5,(3.31)
where the coefficients α(k, t) and β(k, t) are unknown, which will be defined below.

The functions s1(k, t) and s2(k, t) are defined in the half-planes Imk ≥ 0 and Imk ≤ 0, respectively. We compose the matrix functions Φ + and Φ by columns of matrices (2.5) and (2.6) at x = 0 according to the rule:

Φ+(k,0,t)=(s111(k,t)E+(k,0,t),W+(k,0,t))Imk>0,Φ(k,0,t)=(W(k,0,t),s221(k,t)E(k,0,t),)Imk<0.

Using the consistency condition for systems (2.66) and (2.67) in the quarter of the plane x ≥ 0, t ≥ 0, we calculate α(k, 0) and β(k, 0). In view of system (2.68) for the eigenfunction along the line x = 0 and the change of variables (3.3), this consistency condition means that the matrix functions T(k)ψ+(μ, t), T(k)ψ(μ, t) defined on μ-plane and Φ +(k, 0, t), Φ (k, 0, t) defined on k-plane must be consistent at the corner point (x, t) = (0, 0) for all values of k. Making use of this fact, we calculate α(k, 0) and β(k, 0) for kNj, j = 1, 2, ...,6, [11]:

α(k,0)=s11(k,0)r12(ω2k)s21(k,0),β(k,0)=s21(k,0)forkN1,(3.32)
α(k,0)=s11(k,0),β(k,0)=0forkN2,(3.33)
α(k,0)=s11(k,0)r12(ωk)s21(k,0),β(k,0)=s21(k,0)forkN3,(3.34)
α(k,0)=s12(k,0),β(k,0)=s22(k,0)r21(ω2k)s12(k,0)forkN4,(3.35)
α(k,0)=0,β(k,0)=s22(k,0)forkN5,(3.36)
α(k,0)=s12(k,0),β(k,0)=s22(k,0)r21(ωk)s12(k,0)forkN6.(3.37)

Using formula (3.21), (3.26) and (3.27), from (3.28)(3.31) for t = 0, we get:

D(μ)β(k,0)={s21(k,0)s11(k,0)forkN1N3s22(k,0)s12(k,0)forkN5(3.38)
N(μ)α(k,0)={s12(k,0)s22(k,0)forkN4N6s11(k,0)s21(k,0)forkN2.(3.39)

Here the coefficients α(k, 0), β(k, 0) are calculated by formulas (3.32)(3.37):

β(k,0)=s21(k,0)forkN1N3,α(k,0)=s12(k,0)forkN4N6,(3.40)
β(k,0)=s22(k,0)forkN5,α(k,0)=s11(k,0)forkN2.(3.41)

Due to (2.20), from (3.40) we have the asymptotics as |k| → ∞:

β(k,0)=s21(k,0)0forkN1N3,α(k,0)=s12(k,0)0forkN4N6,
but the equalities (3.38) for kN1N3 and (3.39) for kN4N6 do not tend to zero when |k| → ∞. Thus, the sought coefficients β(k, 0)and α(k, 0) in (3.38) and (3.39) are determined by formulas (3.41). The equalities (3.38) for all kN5 and (3.39) for all kN2 are derived for establishing relationship between the sought quantities N(μ), D(μ) and entries sij(k) of the known scattering matrix s(k). From these derived equalities, using Lemma 2.1, we obtain the formulas for calculating the sought quantities expressed through given entries of s(k):
D(μ)=1s12(k,0)s22(k,0)forkN5,Imμ>0,(3.42)
N(μ)=D(μ)¯=1s21(k,0)s11(k,0)forkN2,Imμ<0.(3.43)

Using the self-conjugate property of the SP (3.6)(3.7) and (2.16), (2.19), (2.20) we get:

D(μ)=1ike(k)+ex(k)ike(k)ex(k)=2+o(1)ask0,D(μ)=1+o(1)as|μ|,N(μ)=1ike(k)ex(k)ike(k)+ex(k)=2+o(1)ask0,N(μ)=1+o(1)as|μ|,S(μ)=N(μ)D(μ),S(μ)¯=S1(μ)=S(μ)<μ<.(3.44)

Hence, the found quantities D(μ), N(μ) and S(μ) satisfy all properties assembled in Proposition 3.4.

4. The time-evolution of s(k, t) and solution of the IBVP

The differential equations for functions s1(k, t) and s2(k, t) are derived from system (2.70) for columns of the matrix s(k, t):

ddt(s11(k,t)s21(k,t))=8ik3(0s21(k,t))+px(0,t)(s21(k,t)s11(k,t)),Imk0,(4.1)
ddt(s12(k,t)s22(k,t))=8ik3(s12(k,t)0)+px(0,t)(s22(k,t)s12(k,t)),Imk0.(4.2)

Further, differentiating equality (3.28) with respect to t and taking into account that the matrix functions s(k, t) and ψ+(μ,t)=[ψij+(μ,t)] satisfy the same system (2.70), gives:

ddt(s11(k,t)s21(k,t))={8ik3(0ψ12+(μ,t)ψ21+(μ,t)0)+px(0,t)σ1ψ+(μ,t)}×(α(k,t)β(k,t))+ψ+(μ,t)ddt(α(k,t)β(k,t)),kN1.

By comparison of the last equality with equality (4.1), using (3.28) for kN1 we derive the evolution equation:

ddt(α(k,t)β(k,t))=8ik3β(k,t)(ψ+(μ,t))1(ψ12+(μ,t)ψ22+(μ,t))=8ik3β(k,t)(ψ+(μ,t)1ψ+(μ,t))(01)=8ik3β(k,t)(01),kN1.(4.3)

Using (3.32), from (4.3) we obtain the explicit formulas for the coefficient (α(k, t), β(k, t)) :

α(k,t)=α(k,0)=s11(k,0)r12(ω2k)s21(k,0),β(k,t)=s21(k,0)e8ik3tforkN1.

Thus, the evolution in time t, t > 0 of s1(k, t) for kN1 is derived from (3.28) with this coefficient:

s1(k,t)=ψ+(μ,t)(s11(k,0)r12(ω2k)s21(k,0)s21(k,0)e8ik3t)forkN1.(4.4)

In the same way as in the previous case, by using (4.1), (4.2), (3.28)(3.31) and (3.33)(3.37), analogously derive:

s1(k,t)=ψ(μ,t)(s11(k,0)0)forkN2,(4.5)
s1(k,t)=ψ+(μ,t)(s11(k,0)r12(ωk)s21(k,0)s21(k,0)e8ik3t)forkN3.(4.6)
s2(k,t)=ψ(μ,t)(s12(k,0)e8ik3ts22(k,0)r21(ω2k)s12(k,0))forkN4,(4.7)
s2(k,t)=ψ+(μ,t)(0s22(k,0))forkN5,(4.8)
s2(k,t)=ψ(μ,t)(s12(k,0)e8ik3ts22(k,0)r21(ωk)s12(k,0))forkN6(4.9)

Hence, the obtained columns s1(k, t) and s2(k, t) are expressible by expansion formulas (4.4)(4.6) and (4.7)(4.9), respectively in their sectors in terms of entries of s(k, 0) and fundamental solutions ψ±(μ, t) of system (2.70). The solutions Ψ ±(μ, t) are calculated from known conditions (1.2) and (1.3). The condition t ≥ 0 is important precisely here. Indeed, for t < 0 the functions s1(k, t) and s2(k, t) are therefore, no longer bounded at infinity of t.

We are now to solve the IBVP (1.1)(1.3). By Theorems 2.1 and results presented in Secs. 3 and 4, this problem is reduced to that of solving the GLM time-dependent equation (2.41):

R(x+y,t)+K(x,y,t)+xR(y+ξ,t)K(x,ξ,t)dξ=0fory>x,t0,(4.10)
where x and t enter Eq. (4.10) as parameters, K(x, y, t), y > x is an unknown function of y for every (x, t) ∈ [0, ∞) × [0, ∞), and R(x + y, t) is the function defined by (2.40) for y > x, t > 0:
R(x+y,t)=12πeik(x+y)R˜(k,t)dk,R˜(k,t)=s12(k,t)s11(k,t),s12(k,t)=s21(k,t),(4.11)
and for every t > 0 the functions s11(k, t) and s21(k, t) satisfy the sufficient conditions of Theorem 2.1.

From the unique solvability of Eq. (2.41), it follows that for every t ≥ 0 the Eq. (4.10) has a unique solution in either L2[x, ∞) or L1[x, ∞). By condition (2.38) and Parseval’s relation, we find that ‖RL2 < 1 in L2[x, ∞). Consequently, Eq. (4.10) can be solved by the method of successive approximations. The solution of Eq. (4.10) may be represented as a convergent Neumann series:

K(x,y;t)=R(x+y,t)+xR(y+ξ1;t)R(x+ξ1,t)dξ1xR(ξ2+y,t)dξ2xR(ξ1+ξ2,t)R(x+ξ1,t)dξ1+(4.12)

The solution p(x, t) of the IBVP (1.1)(1.3) is constructed by formula (2.11) expressed through the solution (4.12) of Eq. (4.10):

p(x,t)=2ddxK(x,x;t).(4.13)

Hence, the solution (4.13) of Eq. (1.1) corresponding to solution (4.12) of Eq. (4.10) is determined by:

p(x,t)=2ddxR(2x,t)+2R2(2x,t)4{xR(x+ξ,t)Rx(x+ξ,t)dξ+R(2x,t)xR2(x+ξ,t)dξxRx(x+ξ2,t)dξ2xR(ξ1+ξ2,t)R(x+ξ1,t)dξ1}+(4.14)

One can verify directly that the solution (4.14) satisfies Eq. (1.1) to any desired order in powers of R. The presentation (4.14) in formality is similar to that of the solution of the KdV equation with the positive coefficient of the dispersive term on the whole-line, that evolves from a purely continuous spectrum, [9].

By Theorem 2.1 the solution of Eq. (4.10) coincides with the kernel of the transformation operator of the SP (2.1)(2.2) with potential (4.14). Hence, p(x, t) determined by (4.14) belongs to the class 𝒮, and therefore it satisfies condition (2.49).

Consider the asymptotic behaviour of s1(k, t) and s2(k, t) at infinity of time (t). Substituting (3.26) into (4.4), (4.6) and (4.8), using Propositions 3.2, and 3.4 gives:

(s11(k,t)s21(k,t))=(s11(k,0)r12s21(k,0)+o(1)o(1))astforkN1N3,Imk3>0,(s12(k,t)s22(k,t))=(o(1)s22(k,0)+o(1))astforkN5,Imk3<0,
where r12 = r12(ω2k) for kN1 and r12 = r12(ωk) for kN3.

Substituting (3.27) into (4.5), (4.7) and (4.9), using Propositions 3.2, and 3.4, gives:

(s11(k,t)s21(k,t))=(s11(k,0)+o(1)o(1))astforkN2,Imk3>0,(s12(k,t)s22(k,t))=(o(1)s22(k,0)r21s12(k,0)+o(1))astforkN4N6,Imk3<0,
where r21 = r21(ω2k) for kN4 and r21 = r21(ωk) for kN6.

From the obtained asymptotics it follows that R˜(k,t)=o(1), r˜(k,t)=o(1) as t → ∞. Hence, R(x + y, t) = o(1) as t → ∞ and for any x > 0:

p(x,t)=o(1)ast.(4.15)

It is known that if the spectrum of Eq. (2.4) is purely continuous, then the asymptotic solution p as t → ∞ of the KdV equation on the whole line is still slowly varying wave train, oscillating about p = 0, (see [9]). Hence, the asymptotic solution (4.15) at infinity of t of the IBVP (1.1)(1.3) is different from that of the KdV equation on the whole line.

5. Exact soliton-solutions of the Cauchy problem for the KdV equation

We consider the Cauchy problem for the KdV equation:

pt6ppx+pxxx=0,(x,t)[0,)×(,),(5.1)
with the known initial condition:
p(x,0)=p(x),(5.2)
where p(x, t) is a real-valued function satisfying the condition for any t ∈ (−∞, ∞) and some ε > 0:
0eεx|p(x,t)|dx<.(5.3)

The Cauchy problem (5.1)(5.2) is associated with the SP for the Schrödinger equation:

yxx+p(x,t)y=ρ2y,(x,t)[0,)×(,)(5.4)
with the boundary condition:
y(ρ,0;t)=0,(5.5)
where the potential p(x, t) is a real-valued function satisfying the condition (5.3).

5.1. The direct and inverse SP (5.4)(5.5)

The SP (5.4)(5.5) with the potential p(x) satisfying condition (5.3) has been investigated in the works [6,7]. In the subsection 5.1 we recall the known results of this SP from these works and omit the proof.

Eq. (5.4) with the potential p(x) satisfying condition (5.3) has a solution e(ρ, x), which for each x ≥ 0 is a holomorphic function of ρ when Imρ>ε2 and satisfies the asymptotic condition as x → ∞:

e(ρ,x)=eiρx[1+o(1)],ex(ρ,x)=eiρx[iρ+o(1)].(5.6)

For each ρ0 > 0 Eq. (5.4) has a solution e1(ρ, x), which for each x ≥ 0 is holomorphic function of ρ in the domain |ρ| > ρ0,Imρ > 0, and satisfies the asymptotic condition as x → ∞:

e1(ρ,x)=eiρx[1+o(1)],e1x(ρ,x)=eiρx[iρ+o(1)].(5.7)
uniformly in ρ in the domain |ρ| > ρ0 > 0.

The functions e(ρ, x), e(−ρ, x) and e(ρ, x), e1(ρ, x) form the fundamental systems of solutions of Eq. (5.4) and their Wronskians are equal to:

W[e(ρ,x),e(ρ,x)]=2iρfor|Imρ|<ε2,(5.8)
W[e(ρ,x),e1(ρ,x)]=2iρfor|ρ|>ρ0,Imρ>0,ρ0>0.(5.9)

The solutions of Eq. (5.4) can be represented in the form:

e(ρ,x)=eiρx+xK(x,ξ)eiρξdξ,(5.10)
where the kernel K(x, ξ) has first-order continuous partial derivatives with respect to x and ξ.

Denote by ω(ρ, x) the solution of the eigenvalue problem generated by Eq. (5.4):

ωxx(ρ,x)+p(x)ω(ρ,x)=ρ2ω(ρ,x),x[0,)(5.11)
with the initial conditions:
ω(ρ,0)=0,ωx(ρ,0)=1.(5.12)

By virtue of (5.8) and (5.9), the solution of the problem (5.11)(5.12) is represented in the form:

ω(ρ,x)=e(ρ)e(ρ,x)e(ρ)e(ρ,x)2iρfor|Imρ|<ε2,(5.13)
ω(ρ,x)=e1(ρ)e(ρ,x)e(ρ)e1(ρ,x)2iρfor|ρ|>ρ0,ρ>0,(5.14)
where e(ρ) = e(ρ, 0) and e1(ρ) = e1(ρ, 0).

Differentiating the equality (5.14) with respect to x, using the initial condition (5.12), we have:

e1(ρ)=2iρ+e(ρ)e1x(ρ)ex(ρ)for|ρ|>ρ0,ρ>0.(5.15)

By L we mean the operator generated in the space L2[0, ∞) by Eq. (5.4) and boundary condition (5.5). The potential p(x) in the operator L is a real-valued function. Consider an eigenfunction Ω(ρ, x) of the operator L normalized in the following way:

Ω(ρ,x)=2iρω(ρ,x)e(ρ),Imρ>ε2.(5.16)

By (5.13) and (5.14), the normalized eigenfunction Ω(ρ, x) is represented in the form:

Ω(ρ,x)=S(ρ)e(ρ,x)e(ρ,x)for|Imρ|<ε2,
Ω(ρ,x)=S1(ρ)e(ρ,x)e1(ρ,x)for|ρ|>ρ0,
where
S(ρ)=e(ρ)e(ρ),for|Imρ|<ε2,S1(ρ)=e1(ρ)e(ρ),for|ρ|>ρ0,Imρ>0.(5.17)

The functions S(ρ) and S1(ρ) are called the scattering function and the reflection coefficient of the operator L, respectively.

Since the potential in Eq. (5.11) is a real-valued function satisfying estimate (5.3), then all the zeros ρj of the function e(ρ) are simple and lie on the imaginary axis, i.e., ρj = j, μj > ε0 > 0, j = 1,...,N. By virtue of this fact, using the expression (5.15), we calculate:

fj(x)=iRes|ρ=iμj{e1(ρ)e(ρ)eiρx}=iRes|ρ=iμj{2iρ+e(ρ)e1x(ρ)e(ρ)ex(ρ)eiρx}=iRes|iμj{2iρeiρxe(ρ)ex(ρ)}=i2i(iμj)eμjxeρ(iμj)ex(iμj),(5.18)
where eμjx fj(x) is expressed through the square of norm mj62 of the Jost solution e(j, x) in L2[0, ∞), [7]:
eμjxfj(x)=2iμjeρ(iμj)ex(iμj)=(0|e(iμj,x)|2dx)1=mj2>0,j=1,,N.(5.19)

We introduce the function:

FS(x)=12π+iη++iη[S(ρ)1]eiρxdρ,(5.20)
where η is a number satisfying the equality: 0 < η < ε0.

The integral (5.20) is applied to analytic function S(ρ) − 1 in the strip 0 < |Imρ| < ε0, therefore its value will not depend on η.

The function FS(x) like the scattering function S(ρ), is a spectral characteristic of the operator L on its continuous spectrum. While the functions fj(x) defined by (5.18) characterize the operator L on its point spectrum.

The scattering function S(ρ), the nonsingular numbers j,...,N and the normalization multipliers m12,,mN2 are called the scattering data of the operator L. The scattering data are not independent of each other. The scattering data uniquely determine the self-adjoint operator L. To reconstruct this operator from the scattering data, we construct the function [7]:

F(x)=FS(x)+j=1Nfj(x)=FS(x)+j=1Nmj2eμjx.(5.21)

The kernel K(x, y) from (5.10) satisfies the GLM equation:

F(x+y)+K(x,y)+xK(x,ξ)F(y+ξ)dξ=0,0x<y<.(5.22)

Eq. (5.22) has a unique solution K(x, y), and the potential p(x) is recovered through the found solution by the equality [7]:

p(x)=2ddxK(x,x),x0.(5.23)
where the reconstructed function (5.23) is real-valued and satisfies the same estimate (5.3), as the estimate for the potential in the Schrödinger equation (5.4).

5.2. Non-scattering potentials

There exists a remarkable class of potentials, for which Eq. (5.22) can be solved exactly. These are non-scattering potentials on the half-line, for which the inverse Fourier transform FS(x) defined by (5.20) in the sense of generalized functions is equal to zero, [12]. Hence, in the class of non-scattering potentials the functions FS(x) and F(x) defined by (5.20) and (5.21), respectively, are

FS(x)=0,F(x)=j=1Nmj2eμjx.(5.24)

Our definition of non-scattering potential is similar to the definition of reflectionless of potentials, for which the reflection coefficient is identically zero [8].

Under the condition (5.24) Eq. (5.22) can be solved exactly. Indeed, the solution K(x, y) of this equation is to be sought in the form:

K(x,y)=j=1NKj(x)eμjy,0xy,μj>ε0,j=1,,N.(5.25)

Substituting (5.25) into Eq. (5.22), after some simple transformations, we obtain a system of linear algebraic equations for Kj(x):

Kj(x)+fj(x)n=1Neμnxμn+μjKn(x)=fj(x),j=1,,N(5.26)

Let D(x) denote an N × N square matrix consisting of the elements:

Djn(x)=δjn+fj(x)eμnxμn+μj,j,n=1,,N.(5.27)

From linear algebra, we know that the solution of the system (5.26) is

Kj(x)=detD(j)(x)detD(x),j=1,,N.
where D(j)(x) stands for the matrix obtained from the matrix D(x) on substituting the elements in its j-th column by the elements − fn(x):
Dnj(j)(x)=fn(x)=mn2eμnx,n=1,,N.

Since the potential p(x) is determined by K(x, x), then we calculate it with the help of (5.25):

K(x,x)=(detD(x))1j=1NdetD(j)(x)eμjx,x0.

Using the rule of differentiation of determinants, we find that the numerator in this expression is equal to the derivative of detD(x), because K(x,x)=ddxlndetD(x). Hence, the formula (5.23) for the potential p(x) in the class of non-scattering potentials is written in a compact form:

p(x)=2d2dx2lndetD(x),x0.(5.28)

The expression (5.28) completely describes the whole family of non-scattering potentials.

5.3. The time-dependence of the reflection coefficient

It is known that the KdV equation (5.1) is identical to the equation defined by the Lax representation [8]:

L˙=[L,A]=LAAL,(5.29)
where L=d2dx2+p, A=4d3dx33px6pddxγ, L˙ is derivative of L(t) with respect to t, and γ is some constant, which will be determined below.

The potential p(x, t) in the operator L(t) is called isospectral if the spectrum of L(t) is invariant with t, i.e., λ˙=0. The Lax pair for the KdV equation (5.1) consists of the operator L(t) for the spectral problem and the operator A governing the time-dependence of eigenfunctions. Namely,

LΩ=λΩ,λ=ρ2,Imρ>ε2,(5.30)
Ω˙=AΩ,(x,t)[0,)×(,),(5.31)
where Ω is the normalized eigenfunction defined by (5.16).

Differentiating Eq. (5.30) with respect to t and using (5.31), we have:

L˙Ω+LΩ˙=L˙ΩLAΩ=λ˙Ω+λΩ˙=λ˙ΩλAΩ,{L˙(LAAL)}Ω=λ˙Ω.(5.32)

It follows from (5.32) that the Lax representation (5.29) for the nontrivial eigenfunction Ω holds if and only if λ˙=0.

Lemma 5.1.

If the potential p(x, t) in the operator L(t) satisfies the KdV equation (5.1), then the time-dependence of the normalization eigenfunction (5.16) is defined by the evolution equation:

Ω˙=(4iρ34d3dx3+6pddx+3px)Ω,Imρ>ε2,(x,t)[0,)×(,)(5.33)
and the reflection coefficient S1(ρ; t) evolves according to the equation:
S1(ρ;t)=e1(ρ;t)e(ρ;t)=e1(ρ)e(ρ)e8iρ3tfor|ρ|>ρ0,Imρ>0,t(,).(5.34)

Proof.

Let the potential p(x, t) in L(t) satisfy the KdV equation (5.1), then the time-dependence of the normalized eigenfunction (5.16) is given by the evolution equation (5.31). Using (5.29), we write Eq. (5.31) in the form:

Ω˙=(γpx)Ω+(4λ+2p)Ωx,λ=ρ2.(5.35)

Due to (5.6) and (5.7) the normalization eigenfunction (5.16) obeys the asymptotic condition as x → ∞:

Ω(ρ,x;t)=e1(ρ;t)e(ρ;t)eiρxeiρx+o(1)for|ρ|>ρ0>0,Imρ>0.(5.36)

Since p(x, t) is a solution of the KdV equation (5.1), then the potential p(x, t) is a isospectral potential. Using this fact and (5.3), (5.36), in (5.35) letting x tend to ∞, we find

ddt(e1(ρ;t)e(ρ;t))eiρx=γ(e1(ρ;t)e(ρ;t)eiρxeiρx)+4iρ3(e1(ρ;t)e(ρ;t)eiρx+eiρx),
whence, it follows that for |ρ| > ρ0 > 0, Imρ > 0:
γeiρx+4iρ3eiρx=0,ddt(e1(ρ;t)e(ρ;t))eiρx=(γ+4iρ3)e1(ρ;t)e(ρ;t)eiρx.(5.37)

Hence, γ = 43, and the time-dependence of the functions Ω and S1 defined by evolution equations (5.33) and (5.34) are deduced from (5.31) and (5.37), respectively. The lemma is proved.

The Lemma 5.1 enables us to find the time-dependent potential p(x, t) in the class of non-scattering potentials. In fact, the time-dependent matrix D(x;t) is obtained from the matrix D(x) given by (5.27) with the help of the following substitution:

Djn(x)=δjn+fj(x)eμnxμn+μjDjn(x;t)=δjn+fj(x;t)eμnxμn+μj,
where fj(x, t) is calculated by the formulas (5.34), (5.18) and (5.19):
fj(x;t)=iRes|ρ=iμj{S1(ρ;t)eiρx}=iRes|ρ=iμj{S1(ρ)e8iρ3t+iρx}=2μjieρ(iμj)ex(iμj)eμjx+8μj3t=mj2eμjx+8μj3t=fj(x)e8μj3t,(5.38)
in addition fj(x, 0) = fj(x), j = 1,...,N, S1(ρ; 0) = S1(ρ), (x, t) ∈ [0, ∞) × (−∞, ∞).

The formulas (5.28) and (5.38) give exact soliton-solutions of the KdV equation (5.1) in the class of non-scattering potentials:

p(x)=2d2dx2lndetD(x;t),(5.39)
where
Djn(x)=δjn+2μji(μj+μn)eρ(iμj)ex(iμj)e(μj+μn)x+8μj3t,j,n=1,,N.(5.40)

The soliton-solution (5.39) of the KdV equation (5.1) is constructed from the non-scattering data s of the associated scattering problem (5.4)(5.5):

s={S(ρ;t)1for|Imρ|<ε02,ρj=iμj,μj>ε0>0,mj2>0,j=1,,N}.(5.41)
The non-scattering data (5.41) are formulated from the known initial condition p(x) = p(x, 0) of the Cauchy problem for the KdV equation (5.1) considered in the class of non-scattering potentials.

Theorem 5.1.

Let the function p(x) in the operator L be an isospectral non-scattering potential which is a real-valued continuous function satisfying the estimate (5.3). Then the normalization multipliers mj2 are defined by formula (5.19), and the time-dependence of the normalization eigenfunction and the reflection coefficient is defined by formulas (5.33) and (5.34), respectively. By these formulas, the non-scattering potential (5.39) constructed from the given non-scattering data (5.41) describes the whole family of non-scattering potentials which are soliton-solutions of the Cauchy problem for the KdV equation (5.1) with the known initial condition p(x).

5.4. An example

Example. Let the non-scattering data (5.41) consist of two simple poles ρ1 = 1 and ρ2 = 2, μ1 > μ2 > 0. In this case the elements Djn(x, t) of the matrix D(x, t) are calculated by (5.39) and (5.40):

D11(x,t)=1+m122μ1e2μ1x+8μ13t,D12(x,t)=m12μ1+μ2e(μ1+μ2)x+8μ13t,D21(x,t)=m22μ1+μ2e(μ1+μ2)x+8μ23t,D22(x,t)=1+m222μ2e2μ2x+8μ23t(5.42)
where m12 and m22 are defined by the formula (5.19).

Putting

ξ=μ1x4μ13t+ξ0,η=μ2x4μ23t+η0,(5.43)
ξ0=12ln(2μ1m12μ1+μ2μ1μ2),η0=12ln(2μ2m22μ1+μ2μ1μ2).(5.44)
and using (5.42), we calculate the determinant D of the matrix D(x, t):
D=detD(x,t)=1+μ1+μ2μ1μ2e1+μ1+μ2μ1μ2e2+e1e2,(5.45)
where
e1=e2ξ=m122μ1μ1μ2μ1+μ2e2μ1x+8μ13t,e2=e2η=m222μ2μ1μ2μ1+μ2e2μ2x+8μ23t(5.46)

The non-scattering real-valued potential p(x, t) is calculated by formulas (5.39)(5.40)

p(x)=2d2dx2detD(x,t)=2DDxx(Dx)2D2,(x,t)[0,)×(,)(5.47)
where the first and the second partial derivatives Dx and Dxx of the determinant D are found from (5.45), using (5.46):
Dx=2μ1A12e12μ2A12e22(μ1+μ2)e1e2,Dxx=4μ12A12e1+4μ22A12e2+4(μ1+μ2)2e1e2,A12=μ1+μ2μ1μ2.(5.48)

It is easy to verify that

D=1+A12e2ξ+A12e2η+e2ξ2η=eξημ1μ2{(μ1μ2)eξ+η+(μ1+μ2)eξ+η+(μ1+μ2)eξη+(μ1μ2)eξη}=eξημ1μ2{μ1(eξ+eξ)(eη+eη)μ2(eξeξ)(eηeη)}=eη(eη+eη)eξ(eξeξ)μ1μ2{μ1(eξ+eξ)(eξeξ)μ2(eηeη)(eη+eη)}=(1+e2η)(1e2ξ)μ1μ2(μ1cthξμ2thη).

Hence,

D2=(1+e2)2(1e1)2(μ1μ2)2(μ1cthξμ2thη)2.(5.49)

Using (5.45), (5.46) and (5.48), we calculate:

DDxxDx2=(1+A12e1+A12e2+e1e2)×(4μ12A12e1+4μ22A12e2+4(μ1+μ2)2e1e2)4μ12A122e124μ22A122e224(μ1+μ2)2e12e228μ1(μ1+μ2)A12e12e28μ2(μ1+μ2)A12e1e228μ1μ2A122e1e2=4μ12A12e1+4μ22A12e2+[4(μ1+μ2)2+4(μ12+μ222μ1μ2)A122]e1e2+[4(μ1+μ2)2+4μ128μ1(μ1+μ2)]A12e12e2+[4(μ1+μ2)2+4μ228μ2(μ1+μ2)]A12e1e22=4μ12A12e1+4μ22A12e2+8(μ1+μ2)2e1e2+4μ22A12e12e2+4μ12A12e1e22=4μ12A12e1+4μ22A12e2+8(μ12μ22)A12e1e2+4μ22A12e12e2+4μ12A12e1e22=4μ12A12e1(1+2e2+e22)+4μ22A12e2(12e1+e12)=4μ12A12e1(1+e2)2+4μ22A12e2(1e1)2=A12(1+e2)2(1e1)2{μ124e1(1e1)2+μ224e2(1+e2)2}=μ1+μ2μ1μ2(1+e2)2(1e1)2{μ124(eξeξ)2+μ224(eηeη)2}.(5.50)

The explicit soliton-solution p(x, t) of the KdV equation (5.1) with two bound states is obtained from (5.47), using (5.48), (5.49) and (5.50):

p(x,t)=2(μ12μ22)μ12cosech2ξ+μ22sech2η(μ1cthξμ2thη)2,(x,t)[0,)×(,),(5.51)
where ξ and η are determined by (5.43) and (5.44), μ1 > μ2 > 0.

Thus, p(x, t) represents the nonlinear superposition of two forms, one traveling with speed 4μ12, the other traveling with speed 4μ22. If t is very large negative, then using (5.45), (5.46) and (5.48) from (5.47) we have:

p(x,t)=o(1)ast.

We suppose that μ1 > μ2, then from (5.43)(5.44) it follows that for every x:

ξη=(μ1μ2)[x4(μ12+μ1μ2+μ22)t]+ξ0η0ast,
i.e., for t much less than zero and for those valued of x, where ξ is about one and η is very negative, we have
sechη0,andthη1,thereforep(x,t)2(μ12μ22)μ12cosech2ξ(μ1cthξμ2)2=2μ12sech2(ξ+Δ)
where Δ=12ln(μ1+μ2μ1μ2).

While in the region of x, where η is about one and ξ is very positive, we obtain:

ξηast,andcosechξ0,cthξ1,p(x,t)2(μ12μ22)μ22sech2η(μ1μ2thη)2=2μ22sech2(ηΔ).

That is, for very large negative t the solution looks like two solitary pulses, the large one to the left of the small one.

After a long time, when t is large positive, it follows from (5.43)(5.44) that for every x:

ηξ=(μ2μ1)[x4(μ12+μ1μ2+μ22)t]+η0ξ0ast,
i.e., for large positive t and for those values of x, where η is order one and ξ very large negative, then
cosechξ0,andcthξ1asξ,therefore,p(x,t)2(μ12μ22)μ22sech2η(μ1+μ2thη)2=2μ22sech2(η+Δ).

While if ξ is order one, and η very large positive, then

sechη0,andthη1,asη,therefore,p(x,t)2(μ12μ22)μ12cosech2ξ(μ1cthξμ2)2=2μ12sech2(ξΔ)

That is after a long time the large solitary pulse is to the right of the small solitary pulse. They have coalesced and reemerged with their shaped unscathed. The only remnant of the interaction is the phase shift Δ=12ln(μ1+μ2μ1μ2). That is the large pulse is moved forward by an amount 2Δ1 relative to where it would have been in the absence of an interaction, and the small pulse is retarded by an amount 2Δ2 relative to where it would have been in an unperturbed situation. In Figure 3 we show a sketch of this soliton-solution of the KdV equation (5.1) with two bound states at four successive times.

Fig. 3.

Two soliton-solution of the KdV equation (5.1) with two bound states at the four successive moments of time t = t0, t1, t2 and t3.

In general, the non-scattering solution with N bound states has a similar behavior. In this case the non-scattering data (5.41) consist of N simple poles: ρj = j, μj > 0, j = 1,...,N. With N bound states the solution resembles the superposition of N solitary pulses whose speeds and amplitudes are determined by the positive values μj, j = 1,...,N. The solitary pulses emerge unscathed from interaction except for a phase shift given by the sum of phase shifts from all possible pairwise interactions.

6. Conclusions

By Propositions 3.6, 3.7 and formulas (3.42)(3.44), the self-conjugate matrix (3.5) is found uniquely from the known conditions (1.2) and (1.3). Then the time-dependence of s(k, t) is derived by (4.4)(4.9). The known function R˜(k,t) in Eq. (4.10) is defined by ratio (4.11), in which s11(k; t) and s12(k; t) for every t ≥ 0 are any given complex-valued functions satisfying the sufficient conditions of Theorem 2.1. Thus, the application of obtained results to solving the IBVP (1.1)(1.2)(1.3) is consistent and is effectively embedded in the ISM schema.

By Theorem 5.1 the non-scattering potential (5.39) constructed from the non-scattering data (5.41) of the scattering problem (5.4)(5.5) describes the whole family of non-scattering potentials which are soliton-solutions of the Cauchy problem for the KdV equation (5.1).

Acknowledgments

This work has been supported by the Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant number 107.03-2015.31.

References

[3]I.T. Habibullin, Initial boundary value problem for the KdV equation on a semiaxis with homogeneous boundary conditions, Theoretical and Mathematical Physics, Vol. 130, 2002, pp. 25-44.
[5]B.M. Levitan, Inverse Sturm–Liouville problems, VNU Science Press BC, Utrecht, The Netherlands, 1984.
[8]S. Novikov, S.V. Manakov, L.P. Pitaevskii, and V.E. Zakharov, Theory of solitons: The inverse scattering method, Springer Science & Business Media, 1984.
[14]P.L. Vu. The inverse scattering problem for a system of Dirac equations on a half-line, Selected works”Linear boundary value problems of Mathematical Physics”, Institute of Mathematics, National Academy of Sciences of the Ukraine (1973) 174–207 (in Russian)
Journal
Journal of Nonlinear Mathematical Physics
Volume-Issue
25 - 3
Pages
399 - 432
Publication Date
2021/01/06
ISSN (Online)
1776-0852
ISSN (Print)
1402-9251
DOI
10.1080/14029251.2018.1494777How to use a DOI?
Copyright
© 2018 The Authors. Published by Atlantis and Taylor & Francis
Open Access
This is an open access article distributed under the CC BY-NC 4.0 license (http://creativecommons.org/licenses/by-nc/4.0/).

Cite this article

TY  - JOUR
AU  - Pham Loi Vu
PY  - 2021
DA  - 2021/01/06
TI  - The description of reflection coefficients of the scattering problems for finding solutions of the Korteweg–de Vries equations
JO  - Journal of Nonlinear Mathematical Physics
SP  - 399
EP  - 432
VL  - 25
IS  - 3
SN  - 1776-0852
UR  - https://doi.org/10.1080/14029251.2018.1494777
DO  - 10.1080/14029251.2018.1494777
ID  - Vu2021
ER  -