This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dbar dressing method to nonlinear Schrödinger equation with nonzero boundary conditions

Junyi Zhua, Xueling Jiangb and Xueru Wanga
a School of Mathematics and Statistics, Zhengzhou University, Zhengzhou, Henan 450001, China
b Zhengzhou E-Commerce Vocational College, Zhengzhou, Henan 450048, China
[email protected]@qq.com
Abstract

The Dbar dressing method is extended to study the focusing/defocusing nonlinear Schrödinger (NLS) equation with nonzero boundary condition. A special type of complex function is considered. The function is meromorphic outside an annulus with center 0 and satisfies a local Dbar problem inside the annulus. The theory of such function is extended to construct the Lax pair of the NLS equation with nonzero boundary condition. In this procedure, the relation between the NLS potential and the solution of the Dbar problem is established. A certain distribution for the Dbar problem is introduced to obtain the focusing/defocusing NLS equation and the conservation laws. The explicit solutions of the focusing/defocusing NLS equation with nonzero boundary condition are given from special distributions.

Keywords: Dbar dressing method, nonlinear Schrödinger equation, nonzero boundary condition

1 Introduction

Nonlinear integrable equations with nonzero boundary condition (NZBC) have been well studied. Among the methods, the inverse scattering transform or the Riemann-Hilbert problem play an important role [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 29, 30, 28]. It is worth noting that Jaulent, Manna, et al. introduced the spatial transform method based on certain Dbar equation to study the integrable systems, such as KdV, Toda and AKNS hierarchy [31, 32, 33, 34].

The ¯\bar{\partial} (Dbar) problem is a powerful tools to study the nonlinear integrable equations, such as multidimensional equations [35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47], differential-difference equations [48, 49, 50, 47], (1+1) dimensional equations [38, 51, 52, 53, 54, 55, 56]. The Dbar-steepest descent method is developed to study the the asymptotic behavior [57, 58, 59]. The Dbar problem can also be used to consider the well-posedness of integrable equations [60, 61]. To our knowledge, very few of the nonlinear integrable equations with NZBCs are considered by the Dbar problem. we note that multi-lump solutions of KP equation with integrable boundary uy|y=0=0u_{y}|_{y=0}=0 via ¯\bar{\partial}-dressing method were given in [62].

In this paper, we give a different view to know about the nonlinear integrable equations with NZBCs. As an example, we extend the Dbar approach to discuss the focusing and defocusing nonlinear Schrödinger (NLS) equation with NZBC. The associated theory is developed, and can also be used to discuss other nonlinear integrable equations with NZBCs. For convenience, we consider the NLS equation with nonzero boundary condition in the following form [63]

iqt+qxx2ν(|q|2q02)q=0,ν=±1,iq_{t}+q_{xx}-2\nu(|q|^{2}-q_{0}^{2})q=0,\quad\nu=\pm 1, (1.1)

and

q(x,t)ρ,|x|,q(x,t)\rightarrow\rho,\quad|x|\to\infty, (1.2)

where ρ\rho is a constant and |ρ|=q00|\rho|=q_{0}\neq 0. Equation (1.1) is the compatibility condition of the linear system

φx=Uφ,φt=Vφ,\varphi_{x}=U\varphi,\quad\varphi_{t}=V\varphi, (1.3)

where

U=(ikqνq¯ik),V=(2ik2iν(|q|2q02)2kq+iqx2kνq¯iνq¯x2ik2+iν(|q|2q02)).U=\left(\begin{matrix}ik&q\\ \nu\bar{q}&-ik\end{matrix}\right),\quad V=\left(\begin{matrix}-2ik^{2}-i\nu(|q|^{2}-q_{0}^{2})&-2kq+iq_{x}\\ -2k\nu\bar{q}-i\nu\bar{q}_{x}&2ik^{2}+i\nu(|q|^{2}-q_{0}^{2})\end{matrix}\right). (1.4)

It is noted that the eigenvalues of the matrix U0=U(q=ρ)U_{0}=U(q=\rho) have double branches, and the associated spectral space for the nonlinear Schrödinger (NLS) equation with NZBC is multi-sheeted Riemann surface [9, 8]. To use the Dbar approach solving the NLS equation with nonzero boundary condition, one needs to transform the multi-sheeted Riemann surface into a Riemann sphere. This can be done by introducing the the uniformization variable zz defined by z=k+λz=k+\lambda and

λ(z)=12(zνq02z),k(z)=12(z+νq02z).\lambda(z)=\frac{1}{2}(z-\nu\frac{q_{0}^{2}}{z}),\quad k(z)=\frac{1}{2}(z+\nu\frac{q_{0}^{2}}{z}). (1.5)

Hence, the eigenfunction of the spectral problem (1.3) as q=ρq=\rho can be given as

(I+izσ3Q0)eiθ(z;x,t)σ3,\left(I+\frac{i}{z}\sigma_{3}Q_{0}\right)e^{i\theta(z;x,t)\sigma_{3}}, (1.6)

where

Q0=(0ρνρ¯0),θ(z;x,t)=λ(z)(x2k(z)t).\displaystyle Q_{0}=\left(\begin{matrix}0&\rho\\ \nu\bar{\rho}&0\end{matrix}\right),\quad\theta(z;x,t)=\lambda(z)(x-2k(z)t). (1.7)

In the following, we consider the Dbar problem in the extended complex zz plane. To do this, we construct an annulus with center at 0, that is, 0 and \infty are outside the annulus. In Section 2, we introduce a special complex function which satisfies a Dbar problem in the annulus, and is meromorphic outside the annulus. Thus the Laurent series near the points 0 and \infty play the role of non-canonical normalization conditions to the Dbar problem. The Dbar problem with normalization conditions is equivalent to an inhomogeneous integral equation, and the inhomogeneous terms are given by the normalization conditions. We present the following theorem proved in Appendix to fullfill the Dbar dressing.

Theorem    Suppose that f(z)f(z) admits ¯f(z)0\bar{\partial}f(z)\neq 0 in 0={0}\mathbb{C}^{0}=\mathbb{C}\setminus\{0\}. If f(z)f(z) satisfies the following asymptotic behaviors in ={}\mathbb{C}^{*}=\mathbb{C}\cup\{\infty\}

f(z)=j=1mamjzj+O(1),z0,\displaystyle f(z)=\sum\limits_{j=1}^{m}\frac{a_{m-j}}{z^{j}}+O(1),\quad z\to 0, (1.8)
f(z)=j=0nbjzj+O(1/z),z,\displaystyle f(z)=\sum\limits_{j=0}^{n}b_{j}z^{j}+O(1/z),\quad z\to\infty,

then for the circles ΓR={z:|z|=R>0}\Gamma_{R}=\{z:|z|=R>0\} and Γε={z:|z|=ε>0}\Gamma_{\varepsilon}=\{z:|z|=\varepsilon>0\}

12πiΓR+Γεf(k)kzdk=j=0nbjzj+j=1mamjzj,ε<|z|<R,\frac{1}{2\pi i}\oint_{\Gamma_{R}+\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k=\sum\limits_{j=0}^{n}b_{j}z^{j}+\sum\limits_{j=1}^{m}\frac{a_{m-j}}{z^{j}},\quad\varepsilon<|z|<R, (1.9)

where RR\to\infty and ε0\varepsilon\to 0.

The Dbar dressing method is based on the hypothesis that the homogeneous integral equation has only zero solution. To establish the relation between the NLS potential and the solution of the Dbar problem, we construct, in Section 3, the Lax pair of the NLS equation with NZBC. To this end, it is important to find two sets of operator which have same normalization conditions.

A special distribution (or spectral transform matrix) for the Dbar problem is introduced in Section 4 to construct the NLS equation under the nonzero boundary conditions and the conservation laws. The determinant of the associated eigenfunction (or the solution of the Dbar problem) is shown to be analytic in the annulus. In this procedure, we introduce a symmetry matrix function about the eigenfunction, and give its evolution equation in terms of the Lax pair. By substituting the expansion of the symmetry matrix function into the evolution equation and taking the O(zl)O(z^{l}) terms, we find the NLS equation with nonzero boundary condition from the off-diagonal parts, and the conservation laws from the diagonal parts. We note that the AKNS hierarchy and infinite conservation laws are shown in [31].

In Section 5, the explicit solutions of the focusing and defocusing NLS equation with NZBC are obtained from two special distributions, which make sure the small norm of the operator in the integral equation associated with the Dbar problem. NN-soliton solutions of the NLS equation with NZBC are given, and for the defocusing NLS equation, dark one-soliton and dark two-soliton are presented. We show that that the collision angle of dark two-soliton in xx-tt plane is determined not only by the eigenvalues but also by the boundary condition.

The conclusions are given in Section 6. At last, the theory of the normalization part of the associated eigenfunctions are presented in the Appendixes.

2 Dbar-problem with non-canonical normalization conditions

We consider the following ¯\bar{\partial}(Dbar)-problem

¯χ(z;x,t):=χ(z;x,t)z¯=χ(z;x,t)r(z),z0,\bar{\partial}\chi(z;x,t):=\frac{\partial\chi(z;x,t)}{\partial\bar{z}}=\chi(z;x,t)r(z),\quad z\in\mathbb{C}^{0}, (2.1)

where χ(z;x,t),r(z)\chi(z;x,t),r(z) are 2×22\times 2 matrices, the distribution r(z)r(z) is independent of xx and tt. To study the NLS equation with NZBC, we introduce the following normalization condition

χ(z;x,t)\displaystyle\chi(z;x,t) eiθ(z;x,t)σ3,z,\displaystyle\sim e^{i\theta(z;x,t)\sigma_{3}},\quad z\rightarrow\infty, (2.2)
χ(z;x,t)\displaystyle\chi(z;x,t) izσ3Q0eiθ(z;x,t)σ3,z0.\displaystyle\sim\frac{i}{z}\sigma_{3}Q_{0}e^{i\theta(z;x,t)\sigma_{3}},\quad z\rightarrow 0.

We note that in the procedure of the inverse scattering transform, one can introduce the Jost functions which tend to (1.6) as |x||x|\to\infty. It is remarked that the condition (1.6) implies the normalization condition (2.2) for the Dbar problem.

For simplicity, we introduce a new function

χ^(z;x,t)=χ(z;x,t)eiθ(z;x,t)σ3,\hat{\chi}(z;x,t)=\chi(z;x,t)e^{-i\theta(z;x,t)\sigma_{3}}, (2.3)

then it satisfies the asymptotic behavior

χ^(z;x,t)I,z,\hat{\chi}(z;x,t)\sim I,\quad z\rightarrow\infty, (2.4)
χ^(z;x,t)izσ3Q0,z0.\hat{\chi}(z;x,t)\sim\frac{i}{z}\sigma_{3}Q_{0},\quad z\rightarrow 0. (2.5)

and the generalised Cauchy integral formula

χ^(z)=12πiΓR+Γεχ^(k)kzdk+12πiε<|k|<R¯χ^(k)kzdkdk¯,\hat{\chi}(z)=\frac{1}{2\pi i}\int_{\Gamma_{R}+\Gamma_{\varepsilon}^{-}}\frac{\hat{\chi}(k)}{k-z}\mathrm{d}k+\frac{1}{2\pi i}\iint_{\varepsilon<|k|<R}\frac{\bar{\partial}\hat{\chi}(k)}{k-z}\mathrm{d}k\wedge\mathrm{d}{\bar{k}}, (2.6)

where ΓR\Gamma_{R} and Γε\Gamma_{\varepsilon} are oriented circle with center at origin of zz plane and radius RR and ε\varepsilon, respectively. Here, RR\to\infty and ε0\varepsilon\to 0. For simplicity, we define the first Cauchy integral on the right hand side of (2.6) as the normalization part of χ\chi, and denote it as 𝒩(χ)\mathcal{N}(\chi), that is

𝒩(χ)=12πiΓR+Γεχ^(k)kzdk.\mathcal{N}(\chi)=\frac{1}{2\pi i}\int_{\Gamma_{R}+\Gamma_{\varepsilon}^{-}}\frac{\hat{\chi}(k)}{k-z}\mathrm{d}k. (2.7)

Then for χ\chi in (2.2), we find from the Theorem that

𝒩(χ)=I+izσ3Q0.\mathcal{N}(\chi)=I+\frac{i}{z}\sigma_{3}Q_{0}. (2.8)

As RR\rightarrow\infty and ε0\varepsilon\rightarrow 0, (2.6) reduces to

χ^(z)=I+izσ3Q0+Jχ^(z),\hat{\chi}(z)=I+\frac{i}{z}\sigma_{3}Q_{0}+J\hat{\chi}(z), (2.9)

where

Jχ^(z)=12πi0χ^(k)eiθ(k)σ3r(k)eiθ(k)σ3kzdkdk¯.J\hat{\chi}(z)=\frac{1}{2\pi i}\iint_{\mathbb{C}^{0}}\frac{\hat{\chi}(k)e^{i\theta(k)\sigma_{3}}r(k)e^{-i\theta(k)\sigma_{3}}}{k-z}\mathrm{d}{k}\wedge\mathrm{d}{\bar{k}}. (2.10)

It is important to assume that the homogeneous equation of (2.9) only has zero solution, that is,

(IJ)f=0f=0.(I-J)f=0\Rightarrow f=0. (2.11)

It is valid for small norm of the operator JJ.

Now we introduce the following solution space of the Dbar-problem (2.1) as

={χ(z;x,t)|¯χ(z;x,t)=χ(z;x,t)r(z),z0}.\mathcal{F}=\{\chi(z;x,t)|\bar{\partial}\chi(z;x,t)=\chi(z;x,t)r(z),\quad z\in\mathbb{C}^{0}\}. (2.12)

In particular, let ψ(x,;z)\psi(x,;z)\in\mathcal{F} and 𝒩(ψ)=I+izσ3Q0\mathcal{N}(\psi)=I+\frac{i}{z}\sigma_{3}Q_{0}. Note that the distribution r(z)r(z) is independent of the variables xx and tt. To study the NLS equation with NZBC via the Dbar-problem (2.1), one needs to introduce certain constraint in the physic space. Here, we suppose

ψ(z;x,t)(I+izσ3Q0)eiθ(z;x,t)σ3,|x|,\psi(z;x,t)\sim\left(I+\frac{i}{z}\sigma_{3}Q_{0}\right)e^{i\theta(z;x,t)\sigma_{3}},\quad|x|\to\infty, (2.13)

and

ψ(z;x,t)=izψ(νq02z;x,t)σ3Q0.\psi(z;x,t)=\frac{i}{z}\psi(\nu\frac{q_{0}^{2}}{z};x,t)\sigma_{3}Q_{0}. (2.14)

Then from (2.9), we know that ψ(z;x,t)\psi(z;x,t) has the following asymptotic behaviors

ψ(z;x,t)=(I+l=1al(x,t)zl)eiθ(z;x,t)σ3,z,\psi(z;x,t)=\left(I+\sum_{l=1}^{\infty}a_{l}(x,t)z^{-l}\right)e^{i\theta(z;x,t)\sigma_{3}},\quad z\to\infty, (2.15)
ψ(z;x,t)=(m=1bm(x,t)zm)eiθ(z;x,t)σ3,z0,\psi(z;x,t)=\left(\sum_{m=-1}^{\infty}b_{m}(x,t)z^{m}\right)e^{i\theta(z;x,t)\sigma_{3}},\quad z\to 0, (2.16)

where

al(x,t)=δl,1iσ3Q012πiDψ(z;x,t)r(z)eiθ(z)σ3zl1dzdz¯,a_{l}(x,t)=\delta_{l,1}\cdot i\sigma_{3}Q_{0}-\frac{1}{2\pi i}\iint_{D}\psi(z;x,t)r(z)e^{-i\theta(z)\sigma_{3}}z^{l-1}\mathrm{d}z\wedge\mathrm{d}{\bar{z}}, (2.17)

and

bm(x,t)={δm,0+12πiDψ(z;x,t)r(z)eiθ(z)σ3zm1dzdz¯,m0,iσ3Q0,m=1.b_{m}(x,t)=\left\{\begin{array}[]{ll}\delta_{m,0}+\frac{1}{2\pi i}\iint_{D}\psi(z;x,t)r(z)e^{-i\theta(z)\sigma_{3}}z^{-m-1}\mathrm{d}z\wedge\mathrm{d}{\bar{z}},&m\geq 0,\\ i\sigma_{3}Q_{0},&m=-1.\end{array}\right. (2.18)

It is remarked that the coefficients al(x,t)a_{l}(x,t) and bm(x,t)b_{m}(x,t) are not independent in terms of the symmetry condition (2.14). In fact,

b1=iσ3Q0,bm1(x,t)=iνmq02mam(x,t)σ3Q0.b_{-1}=i\sigma_{3}Q_{0},\quad b_{m-1}(x,t)=\frac{i}{\nu^{m}q_{0}^{2m}}a_{m}(x,t)\sigma_{3}Q_{0}. (2.19)

3 Dbar dressing for NLS equation with NZBC

We note that equations (2.11) and (2.8) imply that, for χ1,χ2\chi_{1},\chi_{2}\in\mathcal{F},

𝒩(χ1)=𝒩(χ2)χ1=χ2.\mathcal{N}(\chi_{1})=\mathcal{N}(\chi_{2})\Leftrightarrow\chi_{1}=\chi_{2}. (3.1)

This result can be used to construct the Lax pair of NLS equation (1.1). In fact, Since the distribution r(z)r(z) is independent of the variables xx and tt, we know that α(z;x,t)xψ+β(z;x,t)tψ+A(z;x,t)ψ\alpha(z;x,t)\partial_{x}\psi+\beta(z;x,t)\partial_{t}\psi+A(z;x,t)\psi\in\mathcal{F}, if ψ\psi\in\mathcal{F}. Here α(z;x,t),β(z;x,t)\alpha(z;x,t),\beta(z;x,t) and A(z;x,t)A(z;x,t) are some 2×22\times 2 matrices.

To obtain the spatial linear spectral problem, a little manipulation is needed. Let ψ\psi\in\mathcal{F}. Here and after, ψ=ψ(z;x,t)\psi=\psi(z;x,t). From (2.15), we find, at zz\to\infty, that

ψx=\displaystyle\psi_{x}= (i2σ3z+i2a1σ3+i2l=1al+1σ3zl\displaystyle\left(\frac{i}{2}\sigma_{3}z+\frac{i}{2}a_{1}\sigma_{3}+\frac{i}{2}\sum_{l=1}^{\infty}a_{l+1}\sigma_{3}z^{-l}\right. (3.2)
+iq022(σ3z+l=2al1σ3zl)+l=1al,xzl)eiθ(z)σ3,\displaystyle\quad\left.+\frac{iq_{0}^{2}}{2}(\frac{\sigma_{3}}{z}+\sum_{l=2}^{\infty}a_{l-1}\sigma_{3}z^{-l})+\sum_{l=1}^{\infty}a_{l,x}z^{-l}\right)e^{i\theta(z)\sigma_{3}},
ikσ3ψ=i2(σ3z+σ3a1+l=1σ3al+1zlq02(σ3z+l=2σ3al1zl))eiθ(z)σ3,ik\sigma_{3}\psi=\frac{i}{2}\left(\sigma_{3}z+\sigma_{3}a_{1}+\sum_{l=1}^{\infty}\sigma_{3}a_{l+1}z^{-l}-q_{0}^{2}(\frac{\sigma_{3}}{z}+\sum_{l=2}^{\infty}\sigma_{3}a_{l-1}z^{-l})\right)e^{i\theta(z)\sigma_{3}}, (3.3)
i2[σ3,a1]ψ=i2([σ3,a1]l=1[σ3,a1]alzl)eiθ(z)σ3,-\frac{i}{2}[\sigma_{3},a_{1}]\psi=\frac{i}{2}\left(-[\sigma_{3},a_{1}]-\sum_{l=1}^{\infty}[\sigma_{3},a_{1}]a_{l}z^{-l}\right)e^{i\theta(z)\sigma_{3}}, (3.4)

where k=k(z)k=k(z) is defined in (1.5), then

ψx=(i2σ3z+i2a1σ3+O(1z))eiθ(z)σ3,z,\psi_{x}=\left(\frac{i}{2}\sigma_{3}z+\frac{i}{2}a_{1}\sigma_{3}+O(\frac{1}{z})\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow\infty, (3.5)
ikσ3ψi2[σ3,a1]ψ\displaystyle ik\sigma_{3}\psi-\frac{i}{2}[\sigma_{3},a_{1}]\psi =(i2σ3z+i2a1σ3+O(1z))eiθ(z)σ3,z,\displaystyle=\left(\frac{i}{2}\sigma_{3}z+\frac{i}{2}a_{1}\sigma_{3}+O(\frac{1}{z})\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow\infty, (3.6)

with [σ3,a1]=σ3a1a1σ3[\sigma_{3},a_{1}]=\sigma_{3}a_{1}-a_{1}\sigma_{3}. Equations (3.5) and (3.6) imply that the Laurent series of xψ\partial_{x}\psi and ikσ3ψi2[σ3,a1]ψik\sigma_{3}\psi-\frac{i}{2}[\sigma_{3},a_{1}]\psi at zz\to\infty share the same principal part.

Similarly, from (2.16), we know that, as z0z\to 0

ψx=(b1,xz1i2νq02(b0σ3z1+b1σ3z2)+m=0bm,xzm\displaystyle\psi_{x}=\left({b_{-1,x}z^{-1}-\frac{i}{2}\nu q_{0}^{2}(b_{0}\sigma_{3}z^{-1}+b_{-1}\sigma_{3}z^{-2})+\sum_{m=0}^{\infty}b_{m,x}z^{m}}\right. (3.7)
+i2m=0bm1σ3zmi2νq02m=0bm+1σ3zm)eiθ(z)σ3,\displaystyle\left.{+\frac{i}{2}\sum_{m=0}^{\infty}b_{m-1}\sigma_{3}z^{m}-\frac{i}{2}\nu q_{0}^{2}\sum_{m=0}^{\infty}b_{m+1}\sigma_{3}z^{m}}\right)e^{i\theta(z)\sigma_{3}},
ikσ3ψ=\displaystyle ik\sigma_{3}\psi= (i2νq02(σ3b1z2+σ3b0z1)+i2m=0σ3bm1zm\displaystyle\left(\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{-1}z^{-2}+\sigma_{3}b_{0}z^{-1})+\frac{i}{2}\sum_{m=0}^{\infty}\sigma_{3}b_{m-1}z^{m}\right. (3.8)
+i2νq02m=0σ3bm+1zm)eiθ(z)σ3,\displaystyle\quad\left.+\frac{i}{2}\nu q_{0}^{2}\sum_{m=0}^{\infty}\sigma_{3}b_{m+1}z^{m}\right)e^{i\theta(z)\sigma_{3}},
i2νq02(σ3b0+b0σ3)b11ψ=(i2νq02(σ3b0+b0σ3)z1\displaystyle-\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})b_{-1}^{-1}\psi=\left(-\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})z^{-1}\right. (3.9)
i2νq02(σ3b0+b0σ3)b11m=0bmzm)eiθ(z)σ3,\displaystyle\qquad\left.-\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})b_{-1}^{-1}\sum_{m=0}^{\infty}b_{m}z^{m}\right)e^{i\theta(z)\sigma_{3}},

and further

ψx=(i2νq02(b1σ3z2+b0σ3z1)+O(1))eiθ(z)σ3,z0,\psi_{x}=\left(-\frac{i}{2}\nu q_{0}^{2}(b_{-1}\sigma_{3}z^{-2}+b_{0}\sigma_{3}z^{-1})+O(1)\right)e^{i\theta(z)\sigma_{3}},\quad z\to 0, (3.10)
ikσ3ψi2νq02(σ3b0+b0σ3)b11ψ\displaystyle ik\sigma_{3}\psi-\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})b_{-1}^{-1}\psi (3.11)
=(i2νq02(b1σ3z2+b0σ3z1)+O(1))eiθ(z)σ3,z0,\displaystyle\quad=\left(-\frac{i}{2}\nu q_{0}^{2}(b_{-1}\sigma_{3}z^{-2}+b_{0}\sigma_{3}z^{-1})+O(1)\right)e^{i\theta(z)\sigma_{3}},\quad z\to 0,

which implies that the Laurent series of xψ\partial_{x}\psi and ikσ3ψ+i2q02(σ3b0+b0σ3)b11ψik\sigma_{3}\psi+\frac{i}{2}q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})b_{-1}^{-1}\psi at z=0z=0 share the same principal part.

Now, using the relation (2.19), we find that the coefficient of second item on the left hand side of (3.11) is equivalent to that of (3.6),

i2νq02(σ3b0+b0σ3)b11=i2[σ3,a1].\frac{i}{2}\nu q_{0}^{2}(\sigma_{3}b_{0}+b_{0}\sigma_{3})b_{-1}^{-1}=\frac{i}{2}[\sigma_{3},a_{1}]. (3.12)

Since xψ\partial_{x}\psi defined by (3.5) and (3.10) belongs to the space \mathcal{F}, then, from (2.7) and the Theorem, we find

𝒩(ψx)=i2σ3z+i2a1σ3+i2q02(b0σ3z1+b1σ3z2).\mathcal{N}(\psi_{x})=\frac{i}{2}\sigma_{3}z+\frac{i}{2}a_{1}\sigma_{3}+\frac{i}{2}q_{0}^{2}(b_{0}\sigma_{3}z^{-1}+b_{-1}\sigma_{3}z^{-2}). (3.13)

Similarly, for ikσ3ψi2[σ3,a1]ψik\sigma_{3}\psi-\frac{i}{2}[\sigma_{3},a_{1}]\psi\in\mathcal{F} in terms of (3.6) and (3.11), we get

𝒩(ψx)=𝒩(ikσ3ψi2[σ3,a1]ψ),\mathcal{N}(\psi_{x})=\mathcal{N}(ik\sigma_{3}\psi-\frac{i}{2}[\sigma_{3},a_{1}]\psi), (3.14)

which gives the spatial linear spectral problem

ψx=ikσ3ψ+Qψ,Q=i2[σ3,a1],\psi_{x}=ik\sigma_{3}\psi+Q\psi,\quad Q=-\frac{i}{2}[\sigma_{3},a_{1}], (3.15)

in view of the identity (3.1).

Substituting the expansion (2.15) into (3.15) and taking the O(z1)O(z^{-1}) terms, we get the following equation

i2a2σ3=iνq02σ3i2σ3a2+a1,xQa1,-\frac{i}{2}a_{2}\sigma_{3}=-i\nu q_{0}^{2}\sigma_{3}-\frac{i}{2}\sigma_{3}a_{2}+a_{1,x}-Qa_{1}, (3.16)

which can also be derived from equations (3.2)-(3.4). Equations (3.15) and (3.16) imply that

a1,x=iσ3(QxQ2+νq02).a_{1,x}=i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2}). (3.17)

Similarly, substituting the expansion (2.16) into (3.15), and taking the O(z0)O(z^{0}) terms, we get

i2νq02b1σ3=b0,x+Q0i2νq02σ3b1Qb0.\frac{i}{2}\nu q_{0}^{2}b_{1}\sigma_{3}=b_{0,x}+Q_{0}-\frac{i}{2}\nu q_{0}^{2}\sigma_{3}b_{1}-Qb_{0}. (3.18)

Here b1=iσ3Q0b_{-1}=i\sigma_{3}Q_{0} has been used.

Next, we will derive the temporal linear spectral problem. As zz\to\infty, from (3.14), we have

ψt=\displaystyle\psi_{t}= (i2σ3z2i2a1σ3zi2a2σ3+l=1al,tzl\displaystyle\left({-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}a_{1}\sigma_{3}z-\frac{i}{2}a_{2}\sigma_{3}+\sum_{l=1}^{\infty}a_{l,t}z^{-l}}\right. (3.19)
i2l=1al+2σ3zl+i2q04(σ3z2+l=3al2σ3zl))eiθ(z)σ3,\displaystyle\left.{-\frac{i}{2}\sum_{l=1}^{\infty}a_{l+2}\sigma_{3}z^{-l}+\frac{i}{2}q_{0}^{4}(\frac{\sigma_{3}}{z^{2}}+\sum_{l=3}^{\infty}a_{l-2}\sigma_{3}z^{-l})}\right)e^{i\theta(z)\sigma_{3}},
2ik2σ3ψ=(i2σ3z2i2σ3a1ziνq02σ3i2σ3a2i2σ3l=1al+2zl\displaystyle-2ik^{2}\sigma_{3}\psi=\left({-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}\sigma_{3}a_{1}z-i\nu q_{0}^{2}\sigma_{3}-\frac{i}{2}\sigma_{3}a_{2}-\frac{i}{2}\sigma_{3}\sum_{l=1}^{\infty}a_{l+2}z^{-l}}\right. (3.20)
iνq02σ3l=1alzli2q04σ3(1z2+l=3al2zl))eiθ(z)σ3,\displaystyle\left.{-i\nu q_{0}^{2}\sigma_{3}\sum_{l=1}^{\infty}a_{l}z^{-l}-\frac{i}{2}q_{0}^{4}\sigma_{3}(\frac{1}{z^{2}}+\sum_{l=3}^{\infty}a_{l-2}z^{-l})}\right)e^{i\theta(z)\sigma_{3}},
2kQψ=(QzQa1l=1Qal+1zlνq02(Qz+l=2Qal1zl))eiθ(z)σ3,-2kQ\psi=\left(-Qz-Qa_{1}-\sum_{l=1}^{\infty}Qa_{l+1}z^{-l}-\nu q_{0}^{2}(\frac{Q}{z}+\sum_{l=2}^{\infty}Qa_{l-1}z^{-l})\right)e^{i\theta(z)\sigma_{3}}, (3.21)
iσ3(QxQ2+νq02)ψ=(iσ3(QxQ2q02)+l=1iσ3(QxQ2q02)alzl)eiθ(z)σ3,i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi=\left(i\sigma_{3}(Q_{x}-Q^{2}-q_{0}^{2})+\sum_{l=1}^{\infty}i\sigma_{3}(Q_{x}-Q^{2}-q_{0}^{2})a_{l}z^{-l}\right)e^{i\theta(z)\sigma_{3}}, (3.22)

which imply that

ψt=(i2σ3z2i2a1σ3zi2a2σ3+O(1z))eiθ(z)σ3,z,\psi_{t}=\left(-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}a_{1}\sigma_{3}z-\frac{i}{2}a_{2}\sigma_{3}+O(\frac{1}{z})\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow\infty, (3.23)

and

2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ\displaystyle-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi (3.24)
=(i2σ3z2i2σ3a1zQzQa1iνq02σ3\displaystyle\quad=\left(-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}\sigma_{3}a_{1}z-Qz-Qa_{1}-i\nu q_{0}^{2}\sigma_{3}\right.
i2σ3a2+iσ3(QxQ2q02)+O(1z))eiθ(z)σ3,z.\displaystyle\quad\qquad\left.-\frac{i}{2}\sigma_{3}a_{2}+i\sigma_{3}(Q_{x}-Q^{2}-q_{0}^{2})+O(\frac{1}{z})\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow\infty.

Using (3.15)-(3.17), equation (3.24) can be further reduced to

2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ\displaystyle-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi (3.25)
=(i2σ3z2i2a1σ3zi2a2σ3+O(1z))eiθ(z)σ3,z.\displaystyle\quad=\left(-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}a_{1}\sigma_{3}z-\frac{i}{2}a_{2}\sigma_{3}+O(\frac{1}{z})\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow\infty.

Equations (3.23) and (3.25) imply that the Laurent series of tψ\partial_{t}\psi and 2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi at zz\to\infty share the same principal part.

Using (2.19), (3.14) and (3.15), we find from (2.16) that

ψt=(i2q04(b1σ3z1+b0σ3z2+b1σ3z3)+O(1))eiθ(z)σ3,z0,\psi_{t}=\left(\frac{i}{2}q_{0}^{4}(b_{1}\sigma_{3}z^{-1}+b_{0}\sigma_{3}z^{-2}+b_{-1}\sigma_{3}z^{-3})+O(1)\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow 0, (3.26)

and

2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ\displaystyle-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi (3.27)
=(i2q04(b1σ3z1+b0σ3z2+b1σ3z3)+O(1))eiθ(z)σ3,z0.\displaystyle=\left(\frac{i}{2}q_{0}^{4}(b_{1}\sigma_{3}z^{-1}+b_{0}\sigma_{3}z^{-2}+b_{-1}\sigma_{3}z^{-3})+O(1)\right)e^{i\theta(z)\sigma_{3}},\quad z\rightarrow 0.

Since tψ\partial_{t}\psi\in\mathcal{F} and has the asymptotic behaviors given by (3.23) and (3.26), then

𝒩(ψt)=\displaystyle\mathcal{N}(\psi_{t})= i2σ3z2i2a1σ3zi2a2σ3\displaystyle-\frac{i}{2}\sigma_{3}z^{2}-\frac{i}{2}a_{1}\sigma_{3}z-\frac{i}{2}a_{2}\sigma_{3} (3.28)
+i2q04(b1σ3z1+b0σ3z2+b1σ3z3).\displaystyle+\frac{i}{2}q_{0}^{4}(b_{1}\sigma_{3}z^{-1}+b_{0}\sigma_{3}z^{-2}+b_{-1}\sigma_{3}z^{-3}).

From (3.25) and (3.27), we find that

𝒩(ψt)=𝒩(2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ).\mathcal{N}(\psi_{t})=\mathcal{N}\left(-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi\right). (3.29)

From this equation and the fact that 2ik2σ3ψ2kQψ+iσ3(QxQ2q02)ψ-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}-q_{0}^{2})\psi\in\mathcal{F}, we obtain the temporal linear spectral problem

ψt=2ik2σ3ψ2kQψ+iσ3(QxQ2+νq02)ψ,\psi_{t}=-2ik^{2}\sigma_{3}\psi-2kQ\psi+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2})\psi, (3.30)

in view of identity (3.1).

To derive the focusing/defocusing NLS equation (1.1) with NZBC (1.2), we need to introduce the symmetry condition on the off-diagonal matrix QQ in (3.15) as

σνQ¯σν=Q,σν={σ2,ν=1,σ1,ν=1,,\sigma_{\nu}\bar{Q}\sigma_{\nu}=Q,\quad\sigma_{\nu}=\left\{\begin{matrix}\sigma_{2},&\nu=-1,\\ \sigma_{1},&\nu=1,\end{matrix}\right., (3.31)

which implies that the potential QQ takes the following form

Q=(0qνq¯0),Q=\left(\begin{matrix}0&q\\ \nu\bar{q}&0\end{matrix}\right), (3.32)

and σνU(x,t;z¯)¯σν=U(z;x,t)\sigma_{\nu}\overline{U(x,t;\bar{z})}\sigma_{\nu}=U(z;x,t), where U(z;x,t)=ik(z)σ3+QU(z;x,t)=ik(z)\sigma_{3}+Q. For the linear system (3.15) and (3.30) with the boundary condition (2.13), in addition to the first symmetry condition (2.14), the matrix eigenfunction ψ(z;x,t)\psi(z;x,t) admits another symmetry condition

ψ(z;x,t)=σνψ(x,t;z¯)¯σν.\psi(z;x,t)=\sigma_{\nu}\overline{\psi(x,t;\bar{z})}\sigma_{\nu}. (3.33)

It is remarked that the first symmetry condition (2.14) plays the role of construction of the linear spectral problems (3.15) and (3.30), and the second symmetry condition (3.33) is about to curb the potential matrix QQ. As a result, the focusing NLS equation (1.1), is equivalent to the compatibility condition of the linear system (3.15) and (3.30) with QQ given by (3.32).

4 NLS equation and conservation laws

In this section, we consider that the 2×22\times 2 distribution R(z;x,t)R(z;x,t) admits the following properties

(i) The matrix R(z;x,t)R(z;x,t) has zero diagonal part;

(ii) The time evolution of R(z;x,t)R(z;x,t) is tR(z;x,t)=p(z;x,t)σ3R(z;x,t)\partial_{t}R(z;x,t)=p(z;x,t)\sigma_{3}R(z;x,t), where pp is a scaler function

p(z;x,t)=2itθ(z)=4iλ(z)k(z).p(z;x,t)=2i\partial_{t}\theta(z)=-4i\lambda(z)k(z).

Here λ(z),k(z)\lambda(z),k(z) and θ(z)=θ(z;x,t)\theta(z)=\theta(z;x,t) are defined in (1.7).

Given the distribution, we can define a new Dbar problem

¯χ^(z;x,t)=χ^(z;x,t)R(z;x,t),z0.\bar{\partial}\hat{\chi}(z;x,t)=\hat{\chi}(z;x,t)R(z;x,t),\quad z\in\mathbb{C}^{0}. (4.1)

and denote the associated solution space by ^\mathcal{\hat{F}}. It is verified that if χ^^\hat{\chi}\in\mathcal{\hat{F}}, then

V(z;x,t)χ^^,χ^t+12p(z;x,t)χ^σ3^,V(z;x,t)\hat{\chi}\in\mathcal{\hat{F}},\quad\hat{\chi}_{t}+\frac{1}{2}p(z;x,t)\hat{\chi}\sigma_{3}\in\mathcal{\hat{F}}, (4.2)

for some matrix V(z;x,t)=n=MNVn(x,t)znV(z;x,t)=\sum_{n=-M}^{N}V_{n}(x,t)z^{n}, where N,MN,M\in\mathbb{N}. In fact, the first property is obviously. To prove the second property, we let X^(z;x,t)=tχ^+12p(z;x,t)χ^σ3\hat{X}(z;x,t)=\partial_{t}\hat{\chi}+\frac{1}{2}p(z;x,t)\hat{\chi}\sigma_{3}, then, using the properties (i) and (ii), we find for z0z\in\mathbb{C}^{0}

¯X^=\displaystyle\bar{\partial}\hat{X}= (χ^t)R+pχ^σ3R+12pχ^Rσ3\displaystyle(\hat{\chi}_{t})R+p\hat{\chi}\sigma_{3}R+\frac{1}{2}p\hat{\chi}R\sigma_{3}
=\displaystyle= (χ^t)R+pχ^σ3R12pχ^σ3R\displaystyle(\hat{\chi}_{t})R+p\hat{\chi}\sigma_{3}R-\frac{1}{2}p\hat{\chi}\sigma_{3}R
=\displaystyle= (χ^t+12pχ^σ3)R=X^R.\displaystyle(\hat{\chi}_{t}+\frac{1}{2}p\hat{\chi}\sigma_{3})R=\hat{X}R.

In particular, if

R(z;x,t)=eiθ(z)σ3r(z)eiθ(z)σ3,R(z;x,t)=e^{i\theta(z)\sigma_{3}}r(z)e^{-i\theta(z)\sigma_{3}},

and the distribution r(z)r(z) in (2.1) admit the property (i), then the matrix function ψ^(z;x,t)\hat{\psi}(z;x,t) is a solution of the Dbar problem (4.1) and takes the following asymptotic behavior

ψ^(z;x,t)=l=0al(x,t)zl,z,\displaystyle\hat{\psi}(z;x,t)=\sum\limits_{l=0}^{\infty}a_{l}(x,t)z^{-l},\quad z\to\infty, (4.3)
ψ^(z;x,t)=m=1bm(x,t)zm,z0,\displaystyle\hat{\psi}(z;x,t)=\sum\limits_{m=-1}^{\infty}b_{m}(x,t)z^{m},\quad z\to 0,

where al(x,t)a_{l}(x,t) and bm(x,t)b_{m}(x,t) are given in (2.17) and (2.18), respectively. By the properties (4.2), we know that there is a certain matrix T(z;x,t)T(z;x,t) admitting the equation

ψ^t(z;x,t)+12p(z;x,t)ψ^(z,x,t)σ3=T(z;x,t)ψ^(z;x,t).\hat{\psi}_{t}(z;x,t)+\frac{1}{2}p(z;x,t)\hat{\psi}(z,x,t)\sigma_{3}=T(z;x,t)\hat{\psi}(z;x,t). (4.4)

In fact, the matrix T(z;x,t)T(z;x,t) can be found in the following form

T(z;x,t)=2ik2σ32kQ+iσ3(QxQ2+νq02),T(z;x,t)=-2ik^{2}\sigma_{3}-2kQ+i\sigma_{3}(Q_{x}-Q^{2}+\nu q_{0}^{2}), (4.5)

in view of the temporal linear spectral problem (3.30).

We note that the trace of distribution R(z;x,t)R(z;x,t) is zero, then from the Dbar problem (4.1)

¯detψ^(z;x,t)=0,z0,\bar{\partial}\det\hat{\psi}(z;x,t)=0,\quad z\in\mathbb{C}^{0}, (4.6)

which implies that detψ^(z;x,t)\det\hat{\psi}(z;x,t) is analytic in 0\mathbb{C}^{0}. Then using the Cauchy integral formula and the asymptotic behaviors (4.3), we find

detψ^(z;x,t)=1νq02z2:=γ.\det\hat{\psi}(z;x,t)=1-\nu\frac{q_{0}^{2}}{z^{2}}:=\gamma. (4.7)

In fact, by the Cauchy integral formula the asymptotic behaviors (4.3), we know that

detψ^(z;x,t)=\displaystyle\det\hat{\psi}(z;x,t)= 12πiΓR+Γεdetψ^(μ;x,t)μz𝑑μ\displaystyle\frac{1}{2\pi i}\int_{\Gamma_{R}+\Gamma_{\varepsilon}^{-}}\frac{\det\hat{\psi}(\mu;x,t)}{\mu-z}d\mu
=\displaystyle= limR12πiΓR1μz𝑑μ+νlimε012πiΓεq02/(μz)μ2𝑑μ\displaystyle\lim\limits_{R\to\infty}\frac{1}{2\pi i}\int_{\Gamma_{R}}\frac{1}{\mu-z}d\mu+\nu\lim\limits_{\varepsilon\to 0}\frac{1}{2\pi i}\int_{\Gamma_{\varepsilon}}\frac{q_{0}^{2}/(\mu-z)}{\mu^{2}}d\mu
=\displaystyle= 1νq02z2.\displaystyle 1-\nu\frac{q_{0}^{2}}{z^{2}}.

For z±iq0,(ν=1)z\neq\pm iq_{0},(\nu=-1) or z±q0,(ν=1)z\neq\pm q_{0},(\nu=1), we have γψ^1=σ2ψ^Tσ2\gamma\hat{\psi}^{-1}=\sigma_{2}\hat{\psi}^{T}\sigma_{2}. Thus, equation (4.4) can be rewritten as

γT(z;x,t)σ2=(ψ^t)(z;x,t)σ2ψ^T(z;x,t)i2p(z;x,t)Ψ(z;x,t),\gamma T(z;x,t)\sigma_{2}=(\hat{\psi}_{t})(z;x,t)\sigma_{2}\hat{\psi}^{T}(z;x,t)-\frac{i}{2}p(z;x,t)\Psi(z;x,t), (4.8)

where Ψ(z;x,t)=ψ^(z;x,t)σ1ψ^T(z;x,t)=ΨT(z;x,t)\Psi(z;x,t)=\hat{\psi}(z;x,t)\sigma_{1}\hat{\psi}^{T}(z;x,t)=\Psi^{T}(z;x,t). Note that Ψ(z;x,t)\Psi(z;x,t) admits the following asymptotic behavior as zz\to\infty

Ψ(z;x,t)=n=0Ψn(x,t)zn,z,\Psi(z;x,t)=\sum\limits_{n=0}^{\infty}\frac{\Psi_{n}(x,t)}{z^{n}},\quad z\to\infty, (4.9)

where

Ψ0=σ1,Ψn(x,t)=m=0nam(x,t)σ1anmT(x,t),(n1).\displaystyle\Psi_{0}=\sigma_{1},\quad\Psi_{n}(x,t)=\sum\limits_{m=0}^{n}a_{m}(x,t)\sigma_{1}a_{n-m}^{T}(x,t),\quad(n\geq 1). (4.10)

Similarly, at z=0z=0, we have

Ψ(z;x,t)=n=2Ψ~n(x,t)zn,z0,\Psi(z;x,t)=\sum\limits_{n=-2}^{\infty}\tilde{\Psi}_{n}(x,t)z^{n},\quad z\to 0, (4.11)

where

Ψ~2=νq02σ1,Ψ~n=m=1n+1bmσ1bnmT,(n1).\tilde{\Psi}_{-2}=\nu q_{0}^{2}\sigma_{1},\quad\tilde{\Psi}_{n}=\sum\limits_{m=-1}^{n+1}b_{m}\sigma_{1}b_{n-m}^{T},\quad(n\geq-1). (4.12)

Substituting the expansions (4.4) and (4.9) at zz\to\infty into (4.8), and considering the O(z1)O(z^{-1}) items, we have

a1,t(x,t)=12Ψ3(x,t)σ2.a_{1,t}(x,t)=\frac{1}{2}\Psi_{3}(x,t)\sigma_{2}. (4.13)

Since a1(x,t)a_{1}(x,t) can be regarded as the function of the potential matrix QQ by (3.15) and (3.17), so Ψ3(x,t)\Psi_{3}(x,t) should also be the function of QQ. Next, we will find the expression of Ψ3(x,t)\Psi_{3}(x,t) in terms of QQ.

Note that ψ^(z;x,t)\hat{\psi}(z;x,t) admits the linear problem

ψ^x(z;x,t)=iλ(z)ψ^(z;x,t)σ3+ik(z)σ3ψ^(z;x,t)+Qψ^(z;x,,t),\hat{\psi}_{x}(z;x,t)=-i\lambda(z)\hat{\psi}(z;x,t)\sigma_{3}+ik(z)\sigma_{3}\hat{\psi}(z;x,t)+Q\hat{\psi}(z;x,,t), (4.14)

in view of the spatial linear spectral problem (3.15). Hence, Ψ(z;x,t)\Psi(z;x,t) satisfies the following linear problem

Ψx(z;x,t)=ik(z)(σ3Ψ(z;x,t)+Ψ(z;x,t)σ3)+QΨ(z;x,t)+Ψ(z;x,t)QT,\Psi_{x}(z;x,t)=ik(z)(\sigma_{3}\Psi(z;x,t)+\Psi(z;x,t)\sigma_{3})+Q\Psi(z;x,t)+\Psi(z;x,t)Q^{T}, (4.15)

which can be rewritten as

Ψx[d](z;x,t)=2ik(z)σ3Ψ[d](z;x,t)+QΨ[o](z;x,t)+Ψ[o](z;x,t)QT,\Psi_{x}^{[d]}(z;x,t)=2ik(z)\sigma_{3}\Psi^{[d]}(z;x,t)+Q\Psi^{[o]}(z;x,t)+\Psi^{[o]}(z;x,t)Q^{T}, (4.16)

and

Ψx[o](z;x,t)=QΨ[d](z;x,t)+Ψ[d](z;x,t)QT,\Psi_{x}^{[o]}(z;x,t)=Q\Psi^{[d]}(z;x,t)+\Psi^{[d]}(z;x,t)Q^{T}, (4.17)

where Ψ[d]\Psi^{[d]} and Ψ[o]\Psi^{[o]} denote the diagonal part and off-diagonal part of the matrix Ψ\Psi. Substituting the expansion (4.9) at zz\to\infty into (4.16) and (4.17), we find

iσ3Ψ1[d]+2Qσ1=0,\displaystyle i\sigma_{3}\Psi_{1}^{[d]}+2Q\sigma_{1}=0, (4.18)
iσ3Ψ2[d]=Ψ1,x[d](QΨ1[o]+Ψ1[o]QT),\displaystyle i\sigma_{3}\Psi_{2}^{[d]}=\Psi_{1,x}^{[d]}-(Q\Psi_{1}^{[o]}+\Psi_{1}^{[o]}Q^{T}),
iσ3Ψn+1[d]=Ψn,x[d]+iq02σ3Ψn1[d](QΨn[o]+Ψn[o]QT),n=2,3,,\displaystyle i\sigma_{3}\Psi_{n+1}^{[d]}=\Psi_{n,x}^{[d]}+iq_{0}^{2}\sigma_{3}\Psi_{n-1}^{[d]}-(Q\Psi_{n}^{[o]}+\Psi_{n}^{[o]}Q^{T}),\quad n=2,3,\cdots,

and

Ψn,x[o]=QΨn[d]+Ψn[d]QT,n=1,2,3,.\Psi_{n,x}^{[o]}=Q\Psi_{n}^{[d]}+\Psi_{n}^{[d]}Q^{T},\quad n=1,2,3,\cdots. (4.19)

Here and after the variables (x,t)(x,t) are omitted for simplicity. Equations (4.18) and (4.19) are recurrent formula, from which we can find all Ψn\Psi_{n}, for example

Ψ1[d]=2Qσ2,Ψ1[o]=0;\displaystyle\Psi_{1}^{[d]}=2Q\sigma_{2},\quad\Psi_{1}^{[o]}=0; (4.20)
Ψ2[d]=2Qxσ1,Ψ2[o]=(2Q2νq02)σ1;\displaystyle\Psi_{2}^{[d]}=2Q_{x}\sigma_{1},\quad\Psi_{2}^{[o]}=(2Q^{2}-\nu q_{0}^{2})\sigma_{1};
Ψ3[d]=2Qxxσ2+4(Q2νq02)Qσ2,Ψ3[o]=2(QxQQQx)σ2;\displaystyle\Psi_{3}^{[d]}=-2Q_{xx}\sigma_{2}+4(Q^{2}-\nu q_{0}^{2})Q\sigma_{2},\quad\Psi_{3}^{[o]}=2(Q_{x}Q-QQ_{x})\sigma_{2};
Ψ4[d]=(2Qxxx+12Q2Qx6νq02Qx)σ2,\displaystyle\Psi_{4}^{[d]}=(-2Q_{xxx}+12Q^{2}Q_{x}-6\nu q_{0}^{2}Q_{x})\sigma_{2},
Ψ4[o]=2(QQxx+QxxQQx2)σ1+6(Q2νq02)Q2σ1,\displaystyle\Psi_{4}^{[o]}=-2(QQ_{xx}+Q_{xx}Q-Q_{x}^{2})\sigma_{1}+6(Q^{2}-\nu q_{0}^{2})Q^{2}\sigma_{1},
\displaystyle\qquad\cdots\qquad\cdots\qquad\cdots

With the expressions of Ψn\Psi_{n} in hand, from the off-diagonal part of equation (4.13), we find the nonlinear equation

iσ3Qt+Qxx2(Q2νq02)Q=0,i\sigma_{3}Q_{t}+Q_{xx}-2(Q^{2}-\nu q_{0}^{2})Q=0, (4.21)

which implies the nonlinear NLS equation (1.1). In addition, from the diagonal part of (4.13), we get the first conservation law

i(|q|2q02)t=(qq¯xqxq¯)x.i(|q|^{2}-q_{0}^{2})_{t}=(q\bar{q}_{x}-q_{x}\bar{q})_{x}. (4.22)

Similarly, the O(z2)O(z^{-2}) terms in the expansion of (4.8) by substituting (4.9) take the following form

a2,tσ2+a1,tσ2a1T=νq02Qxσ1+νq02(Q2νq02)σ1+12Ψ4.a_{2,t}\sigma_{2}+a_{1,t}\sigma_{2}a_{1}^{T}=\nu q_{0}^{2}Q_{x}\sigma_{1}+\nu q_{0}^{2}(Q^{2}-\nu q_{0}^{2})\sigma_{1}+\frac{1}{2}\Psi_{4}. (4.23)

It is noted that

a1[0]=iQσ3,a1,x[d]=i(Q2νq02)σ3,a2[0]=QxQx1(Q2νq02),\displaystyle a_{1}^{[0]}=-iQ\sigma_{3},\quad a_{1,x}^{[d]}=-i(Q^{2}-\nu q_{0}^{2})\sigma_{3},\quad a_{2}^{[0]}=Q_{x}-Q\partial_{x}^{-1}(Q^{2}-\nu q_{0}^{2}), (4.24)
a2,x[d]=QQx(Q2νq02)x1(Q2νq02),\displaystyle\quad a_{2,x}^{[d]}=QQ_{x}-(Q^{2}-\nu q_{0}^{2})\partial_{x}^{-1}(Q^{2}-\nu q_{0}^{2}),

then the diagonal part of equation (4.23) also reduces to the nonlinear equation (4.21), and the off-diagonal part gives the second conservation law

i(qq¯x)t=[qq¯xx|qx|2ν(|q|2q02)(|q|2+q02)]x.i(q\bar{q}_{x})_{t}=\big{[}q\bar{q}_{xx}-|q_{x}|^{2}-\nu(|q|^{2}-q_{0}^{2})(|q|^{2}+q_{0}^{2})\big{]}_{x}. (4.25)

It is remarked that the NLS equation (4.21) and conversation law (4.22) can also be derived from equation (4.8) by substituting the expansion (4.11) at z=0z=0 in terms of the symmetry condition (2.14). The other conservation laws can be found from the diagonal part of the items O(zn),(n=3,4,)O(z^{-n}),(n=3,4,\cdots) in the expansion of (4.8) at zz\to\infty [31]. It is also noted that the focusing and defocusing NLS equation share the same first conversation law (4.22), but have different second conversation law (4.25).

5 Explicit solutions

From (3.15) and (2.17), we know that

Q=Q0+14π[σ3,0ψ^(z)eiθ(z)σ3r(z)eiθ(z)σ3dzdz¯].Q=Q_{0}+\frac{1}{4\pi}\left[\sigma_{3},\iint_{\mathbb{C}^{0}}\hat{\psi}(z)e^{i\theta(z)\sigma_{3}}r(z)e^{-i\theta(z)\sigma_{3}}\mathrm{d}z\wedge\mathrm{d}{\bar{z}}\right]. (5.1)

where ψ^(z)\hat{\psi}(z) defined by (2.3) satisfies the integral equation (2.9). We note that the distribution r(z)r(z) in 0{\mathbb{C}^{0}} admits

r(z¯)¯=σνr(z)σν,\overline{r(\bar{z})}=\sigma_{\nu}r(z)\sigma_{\nu}, (5.2)

and

r(z)=1z¯2σ3Q0r(νq02z)σ3Q0,r(z)=\frac{1}{\bar{z}^{2}}\sigma_{3}Q_{0}r\big{(}\nu\frac{q_{0}^{2}}{z}\big{)}\sigma_{3}Q_{0}, (5.3)

in terms of the symmetry condition (3.33) and (2.14).

5.1 Solution of focusing NLS with NZBC

To obtain the soliton solutions of the focusing NLS equation (ν=1\nu=-1) with nonzero boundary condition (1.1), we choose, according to the symmetries (5.2) and (5.3), the distribution r(z)r(z) as

r(z)=πj=12N(0c¯jδ(zζ¯j)cjδ(zζj)0),\displaystyle r(z)=\pi\sum_{j=1}^{2N}\left(\begin{matrix}0&\bar{c}_{j}\delta(z-\bar{\zeta}_{j})\\ -c_{j}\delta(z-\zeta_{j})&0\end{matrix}\right), (5.4)

where cj,j=1,2,,Nc_{j}\in\mathbb{C},j=1,2,\cdots,N and

ζj=zj,ζj+N=q02z¯j,cN+j=(ρzj¯)2cj¯.\zeta_{j}=z_{j},\quad\zeta_{j+N}=-\frac{q_{0}^{2}}{\bar{z}_{j}},\quad c_{N+j}=-\left(\frac{\rho}{\bar{z_{j}}}\right)^{2}\bar{c_{j}}. (5.5)

Substituting (5.4) into (5.1) and (2.9), we find the solution of the focusing NLS equation with NZBC

q=ρ+idetMadetM,M=I+ΩΩ¯,q=\rho+i\frac{\det M^{a}}{\det M},\quad M=I+\Omega\bar{\Omega}, (5.6)

where

Ω=(Ωnj)(2N)×(2N),Ωjn=gjζjζn¯,Ma=(0g¯fTM),\displaystyle\Omega=(\Omega_{nj})_{(2N)\times(2N)},\quad\Omega_{jn}=\frac{g_{j}}{\zeta_{j}-\bar{\zeta_{n}}},\quad M^{a}=\left(\begin{matrix}0&\bar{g}\\ f^{T}&M\end{matrix}\right), (5.7)
g=(g1,g2,,g2N),f=(f1,f2,,f2N),\displaystyle g=(g_{1},g_{2},\cdots,g_{2N}),\quad f=(f_{1},f_{2},\cdots,f_{2N}),
gj=cje2iθ(ζj;x,t),fn=1+iρj=12NΩjnζj.\displaystyle g_{j}=c_{j}e^{-2i\theta(\zeta_{j};x,t)},\quad f_{n}=1+i\rho\sum\limits_{j=1}^{2N}\frac{\Omega_{jn}}{\zeta_{j}}.

It is remarked that if the eigenvalues zjz_{j} admit |zj|>q0|z_{j}|>q_{0} and zj>0\Im{z_{j}}>0, then the solution is in correspondence with that obtained by Riemann-Hilbert problem, that is to say, ρ\rho is the boundary condition at xx\to-\infty. Some special solutions can be obtained by choosing different parameters NN and ζj\zeta_{j}, see reference [9].

5.2 Solution of defocusing NLS with NZBC

Next, we consider the explicit solution of the defocusing NLS equation with NZBC. Since the self-adjointness of the linear spectral problem (3.15) and (3.30) for the defocusing NLS equation, the associated eigenvalues locate on the circle |z|=q02|z|=q_{0}^{2}. Hence, to obtain the soliton solution of defocusing NLS equation (ν=1\nu=1), we take the distribution in the following form

r(z)=πj=1N(0d¯jδ(zη¯j)djδ(zηj)0),\displaystyle r(z)=\pi\sum_{j=1}^{N}\left(\begin{matrix}0&\bar{d}_{j}\delta(z-\bar{\eta}_{j})\\ d_{j}\delta(z-\eta_{j})&0\end{matrix}\right), (5.8)

where |ηj|=q0|\eta_{j}|=q_{0}.

It is remarked that the focusing NLS equation with NZBC has two sets of eigenvalues in view of the symmetry conditions (5.2) and (5.3) with ν=1\nu=-1, while for defocusing NLS equation with NZBC, the symmetry conditions (5.2) and (5.3) with ν=1\nu=1 only reduce to one set of eigenvalues.

A similar procedure implies that the solution of the defocusing NLS equation with NZBC can be reconstructed as [63, 64, 65]

q(x)=ρ+idetDadetD,D=IiρG,q(x)=\rho+i\frac{\det D^{a}}{\det D},\quad D=I-i\rho{G}, (5.9)

where

G=(Gjl)N×N,Gjl=hl(η¯jηl)ηl,Da=(0h¯ETD),\displaystyle G=(G_{jl})_{N\times N},\quad G_{jl}=\frac{h_{l}}{(\bar{\eta}_{j}-\eta_{l})\eta_{l}},\quad D^{a}=\left(\begin{matrix}0&\bar{h}\\ E^{T}&D\end{matrix}\right), (5.10)
h=(h1,,hN),E=(1,,1)1×N,hj=dje2iθ(ηj).\displaystyle h=(h_{1},\cdots,h_{N}),\quad E=(1,\cdots,1)_{1\times N},\quad h_{j}=d_{j}e^{-2i\theta(\eta_{j})}.

Let ηj=q0eiαj,0<αj<π\eta_{j}=q_{0}e^{i\alpha_{j}},0<\alpha_{j}<\pi, and djd_{j} admit argdj=αjargρ\arg d_{j}=\alpha_{j}-\arg\rho and

iρdjηj(ηjη¯j)=|dj|2q0sinαj=e2ϵj,ϵj,\frac{i\rho d_{j}}{\eta_{j}(\eta_{j}-\bar{\eta}_{j})}=\frac{|d_{j}|}{2q_{0}\sin\alpha_{j}}=e^{2\epsilon_{j}},\quad\epsilon_{j}\in\mathbb{R}, (5.11)

then

iρhjηj(ηjη¯j)=e2ϑj,ϑj=q0sinαj(x2q0tcosαj)+ϵj.\frac{i\rho h_{j}}{\eta_{j}(\eta_{j}-\bar{\eta}_{j})}=e^{2\vartheta_{j}},\quad\vartheta_{j}=q_{0}\sin\alpha_{j}(x-2q_{0}t\cos\alpha_{j})+\epsilon_{j}. (5.12)

In this case, ρ\rho is correspondence with the boundary condition at x+x\to+\infty.

In particular, for N=1N=1, we obtain the dark one-soliton [8]

q(x,t)=ρeiα1[cosα1isinα1tanhϑ1].q(x,t)=\rho e^{-i\alpha_{1}}[\cos\alpha_{1}-i\sin\alpha_{1}\tanh\vartheta_{1}]. (5.13)

For N=2N=2, the two-soliton solution of the defocusing NLS equation (1.1) takes the following form

q(x,t)=ρ[1+e2ϑ^1+e2ϑ^2+ne2(ϑ^1+ϑ^2)1+e2ϑ1+e2ϑ2+ne2(ϑ1+ϑ2)],q(x,t)=\rho\left[\frac{1+e^{2\hat{\vartheta}_{1}}+e^{2\hat{\vartheta}_{2}}+ne^{2(\hat{\vartheta}_{1}+\hat{\vartheta}_{2})}}{1+e^{2\vartheta_{1}}+e^{2\vartheta_{2}}+ne^{2(\vartheta_{1}+\vartheta_{2})}}\right], (5.14)

where

ϑ^j=\displaystyle\hat{\vartheta}_{j}= ϑjiαj,j=1,2,\displaystyle\vartheta_{j}-i\alpha_{j},\quad j=1,2, (5.15)
n=\displaystyle n= 1cos(α1α2)1cos(α1+α2).\displaystyle\frac{1-\cos(\alpha_{1}-\alpha_{2})}{1-\cos(\alpha_{1}+\alpha_{2})}.

Since 0<αj<π,(j=1,2)0<\alpha_{j}<\pi,(j=1,2) and cos(α1α2)cos(α1+α2)=2sinα1sinα2>0\cos(\alpha_{1}-\alpha_{2})-\cos(\alpha_{1}+\alpha_{2})=2\sin\alpha_{1}\sin\alpha_{2}>0, then 0<n<10<n<1. We note from (5.14) that

q{ρ,x,ρe2i(α1+α2),x+.q\sim\left\{\begin{array}[]{cc}\rho,&x\to-\infty,\\ \rho e^{-2i(\alpha_{1}+\alpha_{2})},&x\to+\infty.\end{array}\right. (5.16)

Since α1\alpha_{1} and α2\alpha_{2} are symmetry in (5.14), for the convenience of discussion, we assume α1<α2\alpha_{1}<\alpha_{2} and let vj=2q0cosαj,(j=1,2)v_{j}=2q_{0}\cos\alpha_{j},(j=1,2), then v1>v2v_{1}>v_{2}. Let Xj=vjt+xj,ϵj=(q0sinαj)xjX_{j}=v_{j}t+x_{j},\epsilon_{j}=-(q_{0}\sin\alpha_{j})x_{j}, and define the compact domain Ωj\Omega_{j} containing the point x=Xjx=X_{j}.

As tt\to-\infty, the domains Ωj\Omega_{j} will separated, and Ω1\Omega_{1} is located on the left of Ω2\Omega_{2}. Now, in the domain Ω1\Omega_{1}, e2ϑ1e^{2\vartheta_{1}} and e2ϑ^1e^{2\hat{\vartheta}_{1}} are bounded, and e2ϑ2e^{2\vartheta_{2}} and e2ϑ^2e^{2\hat{\vartheta}_{2}} are O(e(v1v2)t)O(e^{(v_{1}-v_{2})t}). Thus, in Ω1\Omega_{1}, we find a dark one-soliton (denoted by S1S_{1}^{-})

ρeiα1[cosα1isinα1tanhϑ1].\rho e^{-i\alpha_{1}}\left[\cos\alpha_{1}-i\sin\alpha_{1}\tanh\vartheta_{1}\right]. (5.17)

In Ω2\Omega_{2}, we find another dark one-soliton (denoted by S2S_{2}^{-})

ρei(α2+2α1)[cosα2isinα2tanh(ϑ2τ)].\rho e^{-i(\alpha 2+2\alpha_{1})}\left[\cos\alpha_{2}-i\sin\alpha_{2}\tanh(\vartheta_{2}-\tau)\right]. (5.18)

where τ=ln(n)/2\tau=-\ln(n)/2.

In first case, α1<α2<π/2\alpha_{1}<\alpha_{2}<\pi/2, or v1>v2>0v_{1}>v_{2}>0, then S1S_{1}^{-} and S2S_{2}^{-} will move to right, and S1S_{1}^{-} will catch up with S2S_{2}^{-} as t0t\to 0. After collision, S1S_{1}^{-} will surpass the latter. In second case, α1<π/2<α2\alpha_{1}<\pi/2<\alpha_{2}, or v1>0>v2v_{1}>0>v_{2}, then S1S_{1}^{-} moves to right and S2S_{2}^{-} moves to left, and they will collide as t0t\to 0. In third case, π<α1<α2\pi<\alpha_{1}<\alpha_{2}, or v1<v2<0v_{1}<v_{2}<0, S2S_{2}^{-} and S1S_{1}^{-} will move to left, and S1S_{1}^{-} will also catch up with S2S_{2}^{-} as t0t\to 0.

Similarly, as t+t\to+\infty, Ω1\Omega_{1} is located on the right of Ω2\Omega_{2}. We also obtain two dark one-solitons

S1+:\displaystyle S_{1}^{+}: ρei(α1+2α2)[cosα1isinα1tanh(ϑ1τ)],xΩ1,\displaystyle\quad\rho e^{-i(\alpha_{1}+2\alpha_{2})}\left[\cos\alpha_{1}-i\sin\alpha_{1}\tanh(\vartheta_{1}-\tau)\right],\quad x\in\Omega_{1}, (5.19)
S2+:\displaystyle S_{2}^{+}: ρeiα2[cosα2isinα2tanhϑ2],xΩ2.\displaystyle\quad\rho e^{-i\alpha_{2}}\left[\cos\alpha_{2}-i\sin\alpha_{2}\tanh\vartheta_{2}\right],\qquad\qquad\quad x\in\Omega_{2}.

We note that wave heights of two solitons are q0|cosαj|,j=1,2q_{0}|\cos\alpha_{j}|,j=1,2.

As a consequence of the interaction, the phase shift of dark one-soliton in Ω1\Omega_{1} is τ-\tau and that of Ω2\Omega_{2} is τ\tau. For NN soliton, the interaction can be discussed similarly [64]. We note that there is an energy superposition as collision in second case, and no energy superposition for the first case and the third one. The plots of the interaction on above three cases are shown in Figure 1. In addition, the wave height at the collision point, that is, |q(0,0)||q(0,0)| (with q0=1,ϵj=0q_{0}=1,\epsilon_{j}=0) about α1\alpha_{1} and α2\alpha_{2} is shown in Figure 2. The value of |q(0,0)||q(0,0)| in Figure 2 and q0|cosαj|q_{0}|\cos\alpha_{j}| imply the energy superposition in different cases.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Plots of solution (5.14) with ρ=eiπ4,q0=1,ϵj=0\rho=e^{i\frac{\pi}{4}},q_{0}=1,\epsilon_{j}=0. In addtion, (left)α1=π4,α2=π3\alpha_{1}=\frac{\pi}{4},\alpha_{2}=\frac{\pi}{3}, (center)α1=π3,α2=3π4\alpha_{1}=\frac{\pi}{3},\alpha_{2}=\frac{3\pi}{4} and (right)α1=2π3,α2=3π4\alpha_{1}=\frac{2\pi}{3},\alpha_{2}=\frac{3\pi}{4} .
Refer to caption
Refer to caption
Figure 2: The wave height of |q(0,0)||q(0,0)| in solution (5.14) with q0=1,ϵj=0q_{0}=1,\epsilon_{j}=0.

Furthermore, We find that the two wave trains lie in lines in xx-tt plane with slopes κj=1/vj=(2q0cosαj)1\kappa_{j}=1/v_{j}=(2q_{0}\cos\alpha_{j})^{-1} which imply that the ”collision angle” denoted by Δα\Delta\alpha admits

tanΔα=2q0(cosα1cosα2)1+4q02cosα1cosα2:=F(q0).\tan\Delta\alpha=\frac{2q_{0}(\cos\alpha_{1}-\cos\alpha_{2})}{1+4q_{0}^{2}\cos\alpha_{1}\cos\alpha_{2}}:=F(q_{0}). (5.20)

If the two eigenvalues ηj=q0eiαj\eta_{j}=q_{0}e^{i\alpha_{j}} or αj,(0<α1<α2<π)\alpha_{j},(0<\alpha_{1}<\alpha_{2}<\pi) are fixed, then the angle is determined by the initial amplitude q0q_{0}. In fact,

F(q0)=2(cosα1cosα2)(14q02cosα1cosα2)(1+4q02cosα1cosα2)2.F^{\prime}(q_{0})=\frac{2(\cos\alpha_{1}-\cos\alpha_{2})(1-4q_{0}^{2}\cos\alpha_{1}\cos\alpha_{2})}{(1+4q_{0}^{2}\cos\alpha_{1}\cos\alpha_{2})^{2}}. (5.21)

For cosα1cosα2>0\cos\alpha_{1}\cos\alpha_{2}>0, the angle Δα\Delta\alpha in (5.20) takes the maximum

arctan(cosα1cosα22cosα1cosα2),\arctan\left(\frac{\cos\alpha_{1}-\cos\alpha_{2}}{2\sqrt{\cos\alpha_{1}\cos\alpha_{2}}}\right),

at q0=q~0q_{0}=\tilde{q}_{0}, where

q~0=12cos(α1)cos(α2).\tilde{q}_{0}=\frac{1}{2\sqrt{\cos(\alpha_{1})\cos(\alpha 2)}}.

We called collision in this case as the glancing-crossing collision. We note that in the glancing-crossing collision, there is no energy superposition. See Figure 3.

Refer to caption
Refer to caption
Refer to caption
Figure 3: Solution (5.14) with ρ=q0eiπ4,ϵj=0,α1=π4,α2=π3\rho=q_{0}e^{i\frac{\pi}{4}},\epsilon_{j}=0,\alpha_{1}=\frac{\pi}{4},\alpha_{2}=\frac{\pi}{3}. In addition, (left) with q0=0.1q_{0}=0.1; (center) with q0=12cos(α1)cos(α2)q_{0}=\frac{1}{2\sqrt{\cos(\alpha_{1})\cos(\alpha_{2})}}, and (right) with q0=2q_{0}=2.

For cosα1cosα2<0\cos\alpha_{1}\cos\alpha_{2}<0, the angle Δα\Delta\alpha in (5.20) monotonically increases from 0 to π\pi with respect to q0q_{0} changing from 0 to \infty, and takes π/2\pi/2 at the point

q0=12cos(α1)cos(α2).q_{0}^{*}=\frac{1}{2\sqrt{-\cos(\alpha_{1})\cos(\alpha 2)}}.

For example, for the eigenvalues in Figure 1 (center), we can change the initial amplitude q0q_{0} to modify the collision angle. The energy superposition still exists despite the glancing collision, see Figure 4 (q0=0.1,q0=1/(2cos(α1)cos(α2)q_{0}=0.1,q_{0}=1/(2\sqrt{-\cos(\alpha_{1})\cos(\alpha_{2})} and q0=2q_{0}=2).

Refer to caption
Refer to caption
Refer to caption
Figure 4: Solution (5.14) with ρ=q0eiπ4,ϵj=0,α1=π3,α2=3π4\rho=q_{0}e^{i\frac{\pi}{4}},\epsilon_{j}=0,\alpha_{1}=\frac{\pi}{3},\alpha_{2}=\frac{3\pi}{4}. In addition, (left) with q0=0.1q_{0}=0.1; (center) with q0=12cos(α1)cos(α2)q_{0}=\frac{1}{2\sqrt{-\cos(\alpha_{1})\cos(\alpha_{2})}}, and (right) with q0=2q_{0}=2.

In addition, the direction of the wave trains is also determined by q0q_{0}, that is, the two wave trains is vertical in xx-tt plane as q00q_{0}\to 0 and is horizontal as q0q_{0}\to\infty.

Next, for N=2N=2, we consider a particular case: η2=η¯1\eta_{2}=-\bar{\eta}_{1}, that is α1+α2=π,0<α1<π/2\alpha_{1}+\alpha_{2}=\pi,0<\alpha_{1}<\pi/2, then two-soliton (5.14) reduces to

q(x,t)=ρ[cos(α1)cosh(X)+cos(2α1)cosh(T)+isin(2α1)sinh(T)cos(α1)cosh(X)+cosh(T)],q(x,t)=\rho\left[\frac{\cos(\alpha_{1})\cosh(X)+\cos(2\alpha_{1})\cosh(T)+i\sin(2\alpha_{1})\sinh(T)}{\cos(\alpha_{1})\cosh(X)+\cosh(T)}\right], (5.22)

where

X=\displaystyle X= 2q0sin(α1)x+ln(cos(α1))+ϵ1+ϵ2,\displaystyle 2q_{0}\sin(\alpha_{1})\cdot x+\ln(\cos(\alpha_{1}))+\epsilon_{1}+\epsilon_{2}, (5.23)
T=\displaystyle T= 2q02sin(2α1)t+ϵ2ϵ1.\displaystyle 2q_{0}^{2}\sin(2\alpha_{1})\cdot t+\epsilon_{2}-\epsilon_{1}.

The solution is a special case of cosα1cosα2<0\cos\alpha_{1}\cos\alpha_{2}<0.

6 Conclusions

In this paper, we considered the Dbar-problem with non-canonical normalization condition, which was equivalent to an inhomogeneous integral equation. It is assumed that the associated homogeneous integral equation only has zero solution. Since the solution of the Dbar-problem was meromorphic outside an annulus with center 0 and satisfied a local Dbar problem inside the annulus. We developed the relevant theory to find new solutions of the Dbar problem with different inhomogeneous terms.

We extended the Dbar dressing method to study the focusing and defocusing NLS equation with nonzero boundary condition. To establish the relation between the NLS potential and the Dbar data, we constructed the linear system associated with the NLS equation. To obtain the focusing and defocusing NLS equation, we introduced different symmetry conditions, which implied different forms of the distribution or the spectral transform matrix. For focusing NLS equation with nonzero boundary condition, 2N2N discrete eigenvalues were introduced in the distribution, and NN-soliton solution was given.

For defocusing NLS equation, the associated linear spectral problem is self-adjoint, which implied that the eigenvalues located on the circle |z|=q0|z|=q_{0}. One-soliton, two-soliton and NN-soliton of the defocusing NLS equation with nonzero boundary condition were given. The collision angle of dark two-soltion was found to be related to the boundary condition.

A special distribution and a symmetry matrix function were introduced to construct the NLS equation( with nonzero bounary condition). The asymptotic expansion of the symmetry matrix was considered to obtain a series of equations. The off-diagonal part of these equations gave the NLS equation, and diagonal part provided the conservation laws.

It is seems that the analytical domains of the eigenfunctions are not important for the Dbar problem. Thus, unlike the Jost functions defined in the Riemann-Hilbert approach, the boundary condition ρ\rho at x+x\to+\infty or xx\to-\infty for the Dbar approach seems not crucial. However, if the special distribution with Dirac delta functions is chosen, the boundary condition can be confined to a certain direction (see (5.16)).

Acknowledgment

This work was supported by the National Natural Science Foundation of PR China [Grant number 11471295, 11971442].

Appendix

Lemma 1   Suppose f(z)f(z) is a meromorphic function in the domain DD and has only one mmth-order pole z0Dz_{0}\in D. Let the circle Γε={z:|zz0|=ε}\Gamma_{\varepsilon}=\{z:|z-z_{0}|=\varepsilon\} in DD, then

12πiΓεf(k)kzdk=j=1mamj(zz0)j,|zz0|>ε,\frac{1}{2\pi i}\int_{\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k=\sum\limits_{j=1}^{m}\frac{a_{m-j}}{(z-z_{0})^{j}},\quad|z-z_{0}|>\varepsilon, (A.1)

where Γε\Gamma_{\varepsilon}^{-} denotes a contour taking the negative direction of the circle Γε\Gamma_{\varepsilon} and

f(z)=j=1mamj(zz0)j+l=0a~l(zz0)l,0<|zz0|<ε.f(z)=\sum\limits_{j=1}^{m}\frac{a_{m-j}}{(z-z_{0})^{j}}+\sum\limits_{l=0}^{\infty}\tilde{a}_{l}(z-z_{0})^{l},\quad 0<|z-z_{0}|<\varepsilon. (A.2)

Proof. Suppose CC is a simple closed contour in DD and encloses the circle Γε\Gamma_{\varepsilon}. For all zz in the region with boundary CC, and admitting |zz0|>ε|z-z_{0}|>\varepsilon, we have

12πiCf(k)kzdk=f(z)+1(m1)!limkz0dm1dkm1f(k)(kz0)mkz.\frac{1}{2\pi i}\int_{C}\frac{f(k)}{k-z}{\rm d}k=f(z)+\frac{1}{(m-1)!}\lim\limits_{k\to z_{0}}\frac{{\rm d}^{m-1}}{{\rm d}k^{m-1}}\frac{f(k)(k-z_{0})^{m}}{k-z}. (A.3)

Since z0z_{0} is a mmth-order pole of ff, and 0<|kz0|<ε0<|k-z_{0}|<\varepsilon, then

(kz0)mf(k)=j=0m1aj(kz0)j+j=ma~jm(kz0)j.(k-z_{0})^{m}f(k)=\sum\limits_{j=0}^{m-1}a_{j}(k-z_{0})^{j}+\sum\limits_{j=m}^{\infty}\tilde{a}_{j-m}(k-z_{0})^{j}.

In addition, for |zz0|>ε|z-z_{0}|>\varepsilon, we have

1kz=1zz011kz0zz0=l=0(kz0)l(zz0)l+1,0<|kz0|<ε.\frac{1}{k-z}=\frac{-1}{z-z_{0}}\frac{1}{1-\frac{k-z_{0}}{z-z_{0}}}=-\sum\limits_{l=0}^{\infty}\frac{(k-z_{0})^{l}}{(z-z_{0})^{l+1}},\quad 0<|k-z_{0}|<\varepsilon.

Thus,

1(m1)!limkz0dm1dkm1f(k)(kz0)mkz=j=1mamj(zz0)j.\frac{1}{(m-1)!}\lim\limits_{k\to z_{0}}\frac{{\rm d}^{m-1}}{{\rm d}k^{m-1}}\frac{f(k)(k-z_{0})^{m}}{k-z}=-\sum\limits_{j=1}^{m}\frac{a_{m-j}}{(z-z_{0})^{j}}. (A.4)

Substituting (A.4) into (A.3) implies that

12πiCf(k)kzdk=f(z)j=1mamj(zz0)j.\frac{1}{2\pi i}\int_{C}\frac{f(k)}{k-z}{\rm d}k=f(z)-\sum\limits_{j=1}^{m}\frac{a_{m-j}}{(z-z_{0})^{j}}. (A.5)

On the other hand, for the region with boundary CC and Γε\Gamma_{\varepsilon}, we apply the Cauchy’s formula and find

f(z)=12πiC+Γεf(k)kzdk=12πiCf(k)kzdk+12πiΓεf(k)kzdk.f(z)=\frac{1}{2\pi i}\int_{C+\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k=\frac{1}{2\pi i}\int_{C}\frac{f(k)}{k-z}{\rm d}k+\frac{1}{2\pi i}\int_{\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k. (A.6)

The Lemma 1 is proved by comparing equations (A.5) and (A.6). It is noted that

Γεf(k)kzdk=limε0Γεf(k)kzdk.\int_{\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k=\lim\limits_{\varepsilon\to 0}\int_{\Gamma_{\varepsilon}^{-}}\frac{f(k)}{k-z}{\rm d}k.

Lemma 2  Suppose that f(z)f(z) is a meromorphic function in the region |z|>ϵ>0|z|>\epsilon>0 and satisfies the asymptotic behavior at the infinite point

f(z)=j=0nbjzj+O(1/z),z,f(z)=\sum\limits_{j=0}^{n}b_{j}z^{j}+O(1/z),\quad z\to\infty, (A.7)

then for the contour ΓR={z:|z|=R>ϵ}\Gamma_{R}=\{z:|z|=R>\epsilon\}

12πiΓRf(k)kzdk=j=0nbjzj,r<|z|<R,\frac{1}{2\pi i}\oint_{\Gamma_{R}}\frac{f(k)}{k-z}{\rm d}k=\sum\limits_{j=0}^{n}b_{j}z^{j},\quad r<|z|<R, (A.8)

for some r>ϵr>\epsilon.

Proof   Consider the contour integral on the circle Γr={k:|k|=r,ϵ<r<R}\Gamma_{r}=\{k:|k|=r,\epsilon<r<R\}

12πiΓrf(k)kzdk=12πiΓrf(k)kzdk,r<|z|<R.\frac{1}{2\pi i}\oint_{\Gamma_{r}^{-}}\frac{f(k)}{k-z}{\rm d}k=-\frac{1}{2\pi i}\oint_{\Gamma_{r}}\frac{f(k)}{k-z}{\rm d}k,\quad r<|z|<R.

Suppose z=ζ1z=\zeta^{-1} and make the transformation k=μ1,f(μ1)=g(μ)k=\mu^{-1},f(\mu^{-1})=g(\mu). Then the circle Γr\Gamma_{r} is transformed to the circle C={μ:|μ|=r1}C=\{\mu:|\mu|=r^{-1}\}. Noting the direction of the contours will change after the transformation, we have

12πiΓrf(k)kzdk=ζ2πiCg(μ)μ(μζ)dμ.\frac{1}{2\pi i}\oint_{\Gamma_{r}^{-}}\frac{f(k)}{k-z}{\rm d}k=\frac{\zeta}{2\pi i}\oint_{C}\frac{g(\mu)}{\mu(\mu-\zeta)}{\rm d}\mu. (A.9)

Since 0<|ζ|<r10<|\zeta|<r^{-1} and

g(μ)=j=0nbjμj+O(μ),μ0,g(\mu)=\sum\limits_{j=0}^{n}\frac{b_{j}}{\mu^{j}}+O(\mu),\quad\mu\to 0,

then

12πiCg(μ)μ(μζ)dμ=g(ζ)ζ+1n!limμ0dndμn[g(μ)μnμζ].\frac{1}{2\pi i}\oint_{C}\frac{g(\mu)}{\mu(\mu-\zeta)}{\rm d}\mu=\frac{g(\zeta)}{\zeta}+\frac{1}{n!}\lim\limits_{\mu\to 0}\frac{{\rm d}^{n}}{{\rm d}\mu^{n}}\big{[}\frac{g(\mu)\mu^{n}}{\mu-\zeta}\big{]}. (A.10)

For fixed ζ\zeta, we know that

g(μ)μnμζ=l=0n(j=0lbnl+jζj+1)μl+O(μn+1),μ0.\frac{g(\mu)\mu^{n}}{\mu-\zeta}=-\sum\limits_{l=0}^{n}\big{(}\sum\limits_{j=0}^{l}\frac{b_{n-l+j}}{\zeta^{j+1}}\big{)}\mu^{l}+O(\mu^{n+1}),\quad\mu\to 0. (A.11)

Substituting (A.11) into (A.10), we find

12πiCg(μ)μ(μζ)dμ=g(ζ)ζj=0nbjζj+1.\frac{1}{2\pi i}\oint_{C}\frac{g(\mu)}{\mu(\mu-\zeta)}{\rm d}\mu=\frac{g(\zeta)}{\zeta}-\sum\limits_{j=0}^{n}\frac{b_{j}}{\zeta^{j+1}}. (A.12)

Then, from (A.12) and (A.9), we have

12πiΓrf(k)kzdk=f(z)j=0nbjzj.\frac{1}{2\pi i}\oint_{\Gamma_{r}^{-}}\frac{f(k)}{k-z}{\rm d}k=f(z)-\sum\limits_{j=0}^{n}b_{j}z^{j}. (A.13)

On the other hand, in the annulus r<|z|<Rr<|z|<R, the Cauchy’s formula implies that

f(z)=12πiΓRf(k)kzdk+12πiΓrf(k)kzdk.f(z)=\frac{1}{2\pi i}\oint_{\Gamma_{R}}\frac{f(k)}{k-z}{\rm d}k+\frac{1}{2\pi i}\oint_{\Gamma_{r}^{-}}\frac{f(k)}{k-z}{\rm d}k. (A.14)

Equation (A.8) is obtained by comparing (A.13) and (A.14).

Lemma 1 and Lemma 2 give the proof the Theorem.

References

  • [1] V. E. Zakharov, A. B. Shabat, Interaction between solitons in a stable medium, Sov. Phys. -JETP 37 (1973) 823–828.
  • [2] T. Kawata, H. Inoue, Exact solutions of the derivative nonlinear Schrödinger equation under the nonvanishing conditions, J. Phys. Soc. Japan 44 (1978) 1968–1976.
  • [3] T. Kawata, H. Inoue, Inverse scattering method for nonlinear evolution equations under nonvanishing conditions, J. Phys. Soc. Japan 44 (1978) 1722–1729.
  • [4] X. J. Chen, W. K. Lam, Inverse scattering transform for the derivative nonlinear Schrödinger equation with nonvanishing boundary conditions, Phys. Rev. E 69 (2004) 066604.
  • [5] B. Prinari, M. J. Ablowitz, G. Biondini, Inverse scattering transform for vector nonlinear Schrödinger equation with non-vanishing boundary conditions, J. Math. Phys. 47 (2006) 063508.
  • [6] M. J. Ablowitz, G. Biondini, B. Prinari, Inverse scattering transform for the integrable discrete nonlinear Schrödinger equation with nonvanishing boundary conditions, Inverse Problem 23 (2007) 1711–1758.
  • [7] B. Prinari, G. Biondini, A. D. Trubatch, Inverse scattering transform for the multi-component nonlinear Schrödinger equation with nonzero boundary conditions, Stud. Appl. Math. 126 (2010) 245–302.
  • [8] F. Demontis, B. Prinari, C. van der Mee, F. Vitale, The inverse scattering transform for the defocusing nonlinear Schrödinger equations with nonzero boundary conditions, Stud. Appl. Math. 131 (2013) 1–40.
  • [9] G. Biondini, G. Kovačić, Inverse scattering transform for the focusing nonlinear Schrödinger equation with nonzero boundary conditions, J. Math. Phys. 55 (2014) 031506.
  • [10] F. Demontis, B. Prinari, C. van der Mee, F. Vitale, The inverse scattering transform for the focusing nonlinear Schrödinger equation with asymmetric boundary condition, J. Math. Phys. 55 (2014) 101505.
  • [11] G. Biondini, B. Prinari, On the spectrum of the Dirac operator and the existence of discrete eigenvalues for the defocusing nonlinear Schrödinger equation, Stud. Appl. Math. 132 (2014) 138–159.
  • [12] D. Kraus, G. K. G. Biondini, The focusing Manakov system with nonzero boundary conditions, Nonlinearity 28 (2015) 3101–3151.
  • [13] B. Prinari, F. Vitale, G. Biondini, Dark-bright soliton solutions with nontrivial polarization interactions for the three-component defocusing nonlinear Schrödinger equation with nonzero boundary conditions, J. Math. Phys. 56 (2015) 071505.
  • [14] B. Prinari, F. Vitale, Inverse scattering transform for the focusing nonlinear Schrödinger equation with one-sided nonzero boundary condition, Contemp. Math. 651 (2015) 157–194.
  • [15] C. van der Mee, Inverse scattering transform for the discrete focusing nonlinear Schrödinger equation with nonvanishing boundary conditions, J. Nonlinear Math. Phys. 22 (2015) 233–264.
  • [16] G. Biondini, D. Kraus, Inverse scattering transform for the defocusing Manakov system with nonzero boundary conditions, SIAM J. Math. Anal. 47 (2015) 706–757.
  • [17] G. Biondini, E. Fagerstrom, B. Prinari, Inverse scattering transform for the defocusing nonlinear Schrödinger equation with fully asymmetric non-zero boundary conditions, Physica D 333 (2016) 117–136.
  • [18] G. Biondini, D. K. Kraus, B. Prinari, The three-component defocusing nonlinear Schrödinger equation with nonzero boundary conditions, Commun. Math. Phys. 348 (2016) 475–533.
  • [19] M. Pichler, G. Biondini, On the focusing non-linear Schrödinger equation with non-zero boundary conditions and double poles, IMA J. Appl. Math. 82 (2017) 131–151.
  • [20] M. J. Ablowitz, X. D. Luo, Z. H. Musslimani, Inverse scattering transform for the nonlocal nonlinear Schrödinger equation with nonzero boundary conditions, J. Math. Phys. 59 (2018) 011501.
  • [21] B. Prinari, F. Demontis, S. T. Li, T. P. Horikis, Inverse scattering transform and soliton solutions for square matrix nonlinear Schrödinger equations with non-zero boundary conditions, Physica D 368 (2018) 22–49.
  • [22] B. F. Feng, X. D. Luo, M. J. Ablowitz, Z. H. Musslimani, General soliton solution to a nonlocal nonlinear Schrödinger equation with zero and nonzero boundary conditions, Nonlinearity 31 (2018) 5385–5409.
  • [23] J. Y. Zhu, L. L. Wang, Kuznetsov-Ma solution and Akhmediev breather for TD equation, Commun. Nonlinear Sci. Numer. Simulat. 67 (2019) 555–567.
  • [24] J. Y. Zhu, L. L. Wang, Z. J. Qiao, Inverse spectral transform for the Ragnisco-Tu equation with Heaviside initial condition, J. Math. Anal. Appl. 474 (2019) 452–466.
  • [25] G. Zhang, S. Y. Chen, Z. Y. Yan, Focusing and defocusing Hirota equations with non-zero boundary conditions: Inverse scattering transforms and soliton solutions, Commun. Nonlinear Sci. Numer. Simulat. 80 (2020) 104927.
  • [26] G. Q. Zhang, Z. Y. Yan, Inverse scattering transforms and soliton solutions of focusing and defocusing nonlocal mKdV equations with non-zero boundary conditions, Physica D 402 (2020) 132170.
  • [27] Z. C. Zhang, E. G. Fan, Inverse scattering transform for the Gerdjikov-Ivanov equation with nonzero boundary conditions, Z. Angew. Math. Phys. 71 (2020) 149.
  • [28] G. Biondini, J. Lottes, D. Mantzavinos, Inverse scatteringtransform for the focusing nonlinear Schrödinger equation with counterpropagating flows, Stud. Appl. Math. 146 (2021) 371-439.
  • [29] Y. Rybalko, D. Shepelsky, Long-time asymptotics for the integrable nonlocal focusing nonlinear Schrödinger equation for a family of step-kike initial data, Commun. Math. Phys. 382 (2021) 87–121.
  • [30] X. G. Geng, K. D. Wang, M. M. Chen, Long-time asymptotics for the spin-1 Gross-Pitaevskii equation, Commun. Math. Phys. 382 (2021) 585–611.
  • [31] M. Jaulent, M. Manna, The spatial transform method: partmacr derivation of the AKNS hierarchy, Phys. Lett. A 117 (1987) 62–66.
  • [32] M. Jaulent, M. Manna, Connection between the KDV and the AKNS spatial transforms, Inverse Problems 2 (1986) L35–L41.
  • [33] M. Jaulent, M. Manna, The spatial transform method for partmacr equations of nth linear order, Inverse Problems 3 (1987) L13–L18.
  • [34] M. Jaulent, M. Manna, L. M. Alonso, partmacr equations in the theory of integrable systems, Inverse Probl. 4 (1988) 123–150.
  • [35] V. Zakharov, S. V. Manakov, Construction of higher-dimensional nonlinear integrable systems and of their solutions, Func. Anal. Appl. 19 (1985) 89–101.
  • [36] R. Beals, R. R. Coifman, The D-bar approach to inverse scattering and nonlinear evolutions, Physica D 18 (1986) 242–249.
  • [37] L. V. Bogdanov, S. V. Manakov, The non-local partmacr problem and (2+1)-dimensional soliton equations, J. Phys. A: Math. Gen. 21 (1988) L537–L544.
  • [38] R. Beals, R. R. Coifman, Linear spectral problems, non-linear equations and the ¯\bar{\partial}-method, Inverse Problems 5 (1989) 87–130.
  • [39] B. G. Konopelchenko, Solitons in Multidimensions—Inverse Spectral transform Method, Word Scientific, Singapore, 1993.
  • [40] V. G. Dubrovsky, B. G. Konopelchenko, The (2+1)-dimensional integrable generalization of the sine-Gordon equation. II. localized solutions, Inverse Problem 9 (1993) 391–416.
  • [41] B. G. Konopelchenko, V. G. Dubrovsky, A (2+1)-dimensional integrable generalization of the sine-Gordon equation. i. ¯\bar{\partial}-\partial-dressing and the initial value problem, Stud. Appl. Math. 90 (1993) 189–223.
  • [42] V. G. Dubrovsky, B. G. Konopelchenko, ¯\bar{\partial}-dressing and exact solutions for the (2+1)-dimensional Harry Dym equation, J. Phys. A: Math. Gen. 27 (1994) 4619–4628.
  • [43] V. G. Dubrovsky, The application of the ¯\bar{\partial}-dressing method to some integrable (2+1)-dimensional nonlinear equations, J. Phys. A: Math. Gen. 29 (1996) 3617–3630.
  • [44] V. G. Dubrovsky, The ¯\bar{\partial}-dressing method and the solutions with constant asymptotic values at infinity of DS-II equation., J. Math. Phys. 38 (1997) 6382–6400.
  • [45] V. G. Dubrovsky, The construction of exact multiple pole solutions of some (2+1)-dimensional integrable nonlinear evolution equations via the ¯\bar{\partial}-dressing method, J. Phys. A: Math. Gen. 32 (1999) 369–390.
  • [46] V. G. Dubrovsky, I. B. Formusatik, The construction of exact rational solutions with constant asymptotic values at infinity of two-dimensional NVN integrable nonlinear evolution equations via the dbar-dressing method, J. Phys. A: Math. Gen. 34 (2001) 1837–1852.
  • [47] X. R. Wang, J. Y. Zhu, Z. J. Qiao, New solutions to the differential-difference KP equation, Appl. Math. Lett. 113 (2021) 106836.
  • [48] A. Doliwa, S. V. Manakov, P. M. Santini, ¯\bar{\partial}-reductions of the multidimensional quadrilateral lattice. The multidimensional circular lattice, Commun. Math. Phys. 196 (1998) 1–18.
  • [49] P. M. Santini, Transformations and reductions of integrable nonlinear equations and the dbar-problem, in: L. Mason, Y. Nutku (Eds.), Geometry and integrability, Cambridge University Press, Cambridge, 2003.
  • [50] Q. Li, S. Su, M. H. Li, J. B. Zhang, The inverse scattering transform for the differential-difference Kadomtsev-Petviashvili equation, J. Math. Phys. 56 (2015) 103507.
  • [51] E. V. Doktorov, S. B. Leble, Solitons, Nonlinear Evolution Equations and Inverse Scattering, Springer, Netherlands, 2007.
  • [52] J. Y. Zhu, X. G. Geng, A hierarchy of coupled evolution equations with self-consistent sources and the dressing method, J. Phys. A: Math. Gen. 46 (2013) 035204.
  • [53] J. Y. Zhu, X. Geng, The AB equations and the Dbar-dressing method in semi-characteristic coordinates, Math. Phys. Anal. Geo. 17 (2014) 49–65.
  • [54] Y. H. Kuang, J. Y. Zhu, A three-wave interaction model with self-consistent sources: the Dbar-dressing method and solutions, J. Math. Anal. Appl. 426 (2015) 783–793.
  • [55] Y. H. Kuang, J. Y. Zhu, The higher-order soliton solutions for the coupled Sasa-Satsuma system via the ¯\bar{\partial}-dressing method, Appl. Math. Lett. 69 (2017) 47–53.
  • [56] J. H. Luo, E. G. Fan, ¯\bar{\partial}-dressing method for the coupled Gerdjikov-Ivanov equation, Appl. Math. Lett. 110 (2020) 106589.
  • [57] K. T.-R. McLaughlin, P. D. Miller, The ¯\bar{\partial} steepest descent method and the asymptotic behavior of polynomials orthogonal on the unit circle with fixed and exponentially varying nonanalytic weights, Int. Math. Res. Pap. 2006 (2006) 1–77.
  • [58] K. T.-R. McLaughlin, P. D. Miller, The ¯\bar{\partial} steepest descent method for orthogonal polynomials on the real line with varying weights., Int. Math. Res. Notices (2008) rnn075.
  • [59] Y. L. Yang, E. G. Fan, Long-time asymptotic behavior of the modified schrödinger equation via ¯\bar{\partial}-steepest descent method, arXiv: 1912.10358.
  • [60] A. I. Nachman, Global uniqueness for a two-dimensional inverse boundary value problem, ann. Math. Lett. 143 (1996) 71–96.
  • [61] A. Nachman, I. Regev, D. Tataru, A nonlinear Plancherel theorem with applications to global well-posedness for the defocusing Davey-Stewartson equation and to the inverse boundary value problem of calderón, Int. Math. 220 (2020) 395–451.
  • [62] V. G. Dubrovsky, A. V. Topovsky, Multi-lump solutions of KP equation with integrable boundary via ¯\bar{\partial}-dressing method, Physica D 414 (2020) 132740.
  • [63] L. D. Faddeev, L. A. Takhtajan, Hamiltonian Methods in the Theory of Solitons, Springer, Berlin, 1987.
  • [64] N. Huang, Theory of Solitons and Method of Perturbations, SSTEPH, Shanghai, 1996, (In Chinese).
  • [65] J. W. Han, J. Yu, J. S. He, Determinant representation of n-times Darboux transformation for the defocusing nonlinear Schrödinger equation, Modern. Phys. Lett. B 27 (2013) 1350216.