删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

Solutions of the nonlocal (2【-逻*辑*与-】plus;1)-D breaking solitons hierarchy and the negative order AK

本站小编 Free考研考试/2022-01-02

Jing Wang, Hua Wu, Da-jun Zhang,Department of Mathematics, Shanghai University, Shanghai 200444, China

Received:2019-12-5Revised:2020-02-14Accepted:2020-02-15Online:2020-03-27


Abstract
The (2+1)-dimensional nonlocal breaking solitons AKNS hierarchy and the nonlocal negative order AKNS hierarchy are presented. Solutions in double Wronskian form of these two hierarchies are derived by means of a reduction technique from those of the unreduced hierarchies. The advantage of our method is that we start from the known solutions of the unreduced bilinear equations, and obtain solitons and multiple-pole solutions for the variety of classical and nonlocal reductions. Dynamical behaviors of some obtained solutions are illustrated. It is remarkable that for some real nonlocal equations, amplitudes of solutions are related to the independent variables that are reversed in the real nonlocal reductions.
Keywords: nonlocal;(2+1)-dimensional breaking solitons AKNS hierarchy;negative order AKNS hierarchy;double Wronskian solutions;reduction


PDF (1592KB)MetadataMetricsRelated articlesExportEndNote|Ris|BibtexFavorite
Cite this article
Jing Wang, Hua Wu, Da-jun Zhang. Solutions of the nonlocal (2+1)-D breaking solitons hierarchy and the negative order AKNS hierarchy. Communications in Theoretical Physics, 2020, 72(4): 045002- doi:10.1088/1572-9494/ab7705

1. Introduction

Recently, ${ \mathcal P }{ \mathcal T }$-symmetric integrable nonlocal nonlinear Schrödinger equation$\begin{eqnarray}{\rm{i}}{q}_{t}(x,t)+{q}_{{xx}}(x,t)\pm {q}^{2}(x,t){q}^{* }(-x,t)=0,\end{eqnarray}$was proposed by Ablowitz and Musslimani [1], which is reduced from the second-order Ablowitz–Kaup–Newell–Segur (AKNS) system with reduction $r(x,t)=\mp {q}^{* }(-x,t)$. It is ${ \mathcal P }{ \mathcal T }$-symmetric because of the potential $V(x,t)\,=\pm q(x,t){q}^{* }(-x,t)={V}^{* }(-x,t)$ (see [2, 3]). Since then, more and more integrable nonlocal systems, such as nonlocal Davey–Stewartson model, nonlocal Korteweg–de Vries model and so forth, were proposed (e.g. [49]). Classical solving methods, such as the inverse scattering transform and Darboux transformation (e.g. [1, 5, 1014]), have been successfully used to find solutions of nonlocal integrable systems. However, the bilinear method cannot be directly applied to the nonlocal case as it is difficult to define a ‘nonlocal’ Hirota’s bilinear operator. Recently, Chen et al [1518] developed a reduction technique of double Wronskians on unreduced bilinear equations to obtain solutions to those reduced equations, including nonlocal ones. One advantage of this method is that one only needs to make use of the solutions (most of them are known) of the unreduced bilinear equations. The other advantage is one can obtain N-soliton solutions for the whole hierarchy, rather than implementing reductions solution by solution and equation by equation (see [19]).

The (2+1)-dimensional breaking solitons equations ware first systematically investigated by Bogoyavlenskii [2026]. Such models can be used to describe two-dimensional interaction of a Riemann wave with transverse long waves (see [20, 21]). Mathematically, these equations can be constructed based on (1+1)-D Lax pairs by imposing an evolution in y-direction in both spectral parameters and eigenfunctions (see [21, 24]). The (2+1)-D breaking AKNS(BAKNS in brief) equations were first proposed in [24], and has been investigated from many aspects (e.g. [2731]). And recently, bilinear form and double Wronskian solutions of (2+1)-D BAKNS hierarchy were worked out [32]. In this paper, we will take into account of bilinearization-reduction technique and derive double Wronskian solutions for the classical and nonlocal (2+1)-D BAKNS hierarchy and the negative order AKNS hierarchy. Note that the negative order AKNS equations also exhibits interesting dynamical behaviors (see [33]).

This paper is organized as follows. In section 2 we recall the unreduced (2+1)-D BAKNS hierarchy, its bilinear form and double Wronskian solutions. In section 3 we present classical and nonlocal reductions of the (2+1)-D BAKNS hierarchy. In section 4, solutions in double Wronskian form of the reduced (2+1)-D BAKNS hierarchies are derived by means of the reduction technique from those of the unreduced hierarchy. Section 5 is contributed to the negative order AKNS hierarchy. Finally, section 6 consists of concluding remarks.

2. Unreduced (2+1)-D BAKNS hierarchy and solutions

Let us recall the unreduced (2+1)-D BAKNS hierarchy, its bilinear form and double Wronskian solutions.

The unreduced (2+1)-D BAKNS hierarchy reads [27]$\begin{eqnarray}{u}_{{t}_{n}}={K}_{n}=\left(\begin{array}{c}{K}_{1,n}\\ {K}_{2,n}\end{array}\right)={L}^{n}{u}_{y},\,\,n=1,2,3,\cdot \cdot \cdot ,\end{eqnarray}$where $u={(q,r)}^{{\rm{T}}}$, L is a recursion operator$\begin{eqnarray}L=\left(\begin{array}{cc}-{\partial }_{x}+2q{\partial }_{x}^{-1}r & 2q{\partial }_{x}^{-1}q\\ -2r{\partial }_{x}^{-1}r & {\partial }_{x}-2r{\partial }_{x}^{-1}q\end{array}\right),\end{eqnarray}$in which ${\partial }_{x}=\tfrac{\partial }{{\partial }_{x}}$ and ${\partial }_{x}{\partial }_{x}^{-1}={\partial }_{x}^{-1}{\partial }_{x}=1$. This hierarchy is related to the AKNS spectral problem [24]$\begin{eqnarray}{\phi }_{x}=\left(\begin{array}{cc}\lambda & q\\ r & -\lambda \end{array}\right)\phi ,\end{eqnarray}$and$\begin{eqnarray*}{\phi }_{{t}_{n}}={\left(2\lambda \right)}^{n}{\phi }_{y}+\left(\begin{array}{cc}{A}_{n} & {B}_{n}\\ {C}_{n} & -{A}_{n}\end{array}\right)\phi ,\end{eqnarray*}$where ${\lambda }_{{t}_{n}}={\left(2\lambda \right)}^{n}{\lambda }_{y}$ and$\begin{eqnarray*}\begin{array}{rcl}{A}_{n} & = & {\partial }^{-1}(r,q){\left(-{B}_{n},{C}_{n}\right)}^{{\rm{T}}},\,\,{\left(-{B}_{n},{C}_{n}\right)}^{{\rm{T}}}\\ & = & \sum _{j=1}^{n}{2}^{n-j}{L}^{j-1}{\left({q}_{y},{r}_{y}\right)}^{{\rm{T}}}{\lambda }^{n-j}.\end{array}\end{eqnarray*}$

The hierarchy (2) can be rewritten as$\begin{eqnarray}{q}_{{t}_{n}}=-{q}_{x,{t}_{n-1}}+2q{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{n-1}},\end{eqnarray}$$\begin{eqnarray}{r}_{{t}_{n}}={r}_{x,{t}_{n-1}}-2r{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{n-1}}\end{eqnarray}$with n=1, 2, 3, ⋯ and setting t0=y. Introducing transformation$\begin{eqnarray}q=\displaystyle \frac{g}{f},\,\,r=-\displaystyle \frac{h}{f},\end{eqnarray}$the hierarchy is written as the following bilinear form [32]$\begin{eqnarray}({D}_{{t}_{n}}+{D}_{x}{D}_{{t}_{n}-1})g\cdot f=0,\end{eqnarray}$$\begin{eqnarray}({D}_{{t}_{n}}-{D}_{x}{D}_{{t}_{n}-1})h\cdot f=0,\end{eqnarray}$$\begin{eqnarray}{D}_{x}^{2}f\cdot f=2{gh},\end{eqnarray}$for n≥1, where Hirota’s bilinear operator D is defined as [34]$\begin{eqnarray*}\begin{array}{l}{D}_{x}^{m}{D}_{y}^{n}f(x,y)\cdot g(x,y)\\ \quad =\,{\left({\partial }_{x}-{\partial }_{x^{\prime} }\right)}^{m}{\left({\partial }_{y}-{\partial }_{y^{\prime} }\right)}^{n}f(x,y)g(x^{\prime} ,y^{\prime} ){| }_{x^{\prime} =x,y^{\prime} =y}.\end{array}\end{eqnarray*}$

System (7) admits double Wronskian solutions [32]$\begin{eqnarray}f={W}^{n+1,m+1}(\varphi ,\psi )=| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(m)}| ,\end{eqnarray}$$\begin{eqnarray}g=2{W}^{n+2,m}(\varphi ,\psi )=2| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(m-1)}| ,\end{eqnarray}$$\begin{eqnarray}h=2{W}^{n,m+2}(\varphi ,\psi )=2| {\widehat{\varphi }}^{(n-1)};{\widehat{\psi }}^{(m+1)}| .\end{eqnarray}$Here, $\varphi $ and $\psi $ are respectively $(n+m+2)$th order column vectors$\begin{eqnarray}\varphi ={\left({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{n+m+2}\right)}^{{\rm{T}}},\,\,\psi ={\left({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{n+m+2}\right)}^{{\rm{T}}}\end{eqnarray}$defined by$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \exp \left(-{Ax}+{A}^{2}y+\sum _{j=1}^{\infty }{2}^{j}{A}^{j+2}{t}_{j}\right){C}^{+},\\ \psi & = & \exp \left({Ax}-{A}^{2}y-\sum _{j=1}^{\infty }{2}^{j}{A}^{j+2}{t}_{j}\right){C}^{-},\end{array}\end{eqnarray}$where $A$ is a $(n+m+2)\times (n+m+2)$ constant matrix in ${{\mathbb{C}}}_{(m+n+2)\times (m+n+2)}$ and$\begin{eqnarray}{C}^{\pm }={\left({c}_{1}^{\pm },{c}_{2}^{\pm },\cdots ,{c}_{n+m+2}^{\pm }\right)}^{{\rm{T}}},\,\,{c}_{i}^{\pm }\in {\mathbb{C}},\end{eqnarray}$${\widehat{\varphi }}^{(n)}$ and ${\widehat{\psi }}^{(m)}$ respectively denote $(n+m+2)\times (n+1)$ and $(n+m+2)\times (m+1)$ Wronski matrices$\begin{eqnarray}\begin{array}{rcl}{\widehat{\varphi }}^{(n)} & = & \left(\varphi ,{\partial }_{x}\varphi ,{\partial }_{x}^{2}\varphi ,\cdots ,{\partial }_{x}^{n}\varphi \right),\\ {\widehat{\psi }}^{(m)} & = & \left(\psi ,{\partial }_{x}\psi ,{\partial }_{x}^{2}\psi ,\cdots ,{\partial }_{x}^{m}\psi \right).\end{array}\end{eqnarray}$

It is easy to find that for the odd members in the unreduced (2+1)-D BAKNS hierarchy (2)$\begin{eqnarray}{u}_{{t}_{2l+1}}={K}_{2l+1},\,\,l=0,1,2,\cdots ,\end{eqnarray}$their solutions are given through (6) and (8) where$\begin{eqnarray}\displaystyle \begin{array}{rcl}\varphi & = & \exp \left(-{Ax}+{A}^{2}y+\sum _{j=0}^{\infty }{2}^{2j+1}{A}^{2j+3}{t}_{2j+1}\right){C}^{+},\\ \psi & = & \exp \left({Ax}-{A}^{2}y-\sum _{j=0}^{\infty }{2}^{2j+1}{A}^{2j+3}{t}_{2j+1}\right){C}^{-},\end{array}\end{eqnarray}$and for the even members$\begin{eqnarray}{u}_{{t}_{2l}}={K}_{2l},\,\,l=1,2,\cdots ,\end{eqnarray}$their solutions are given through (6) and (8) where$\begin{eqnarray}\displaystyle \begin{array}{rcl}\varphi & = & \exp \left(-{Ax}+{A}^{2}y+\sum _{j=1}^{\infty }{2}^{2j}{A}^{2j+2}{t}_{2j}\right){C}^{+},\\ \psi & = & \exp \left({Ax}-{A}^{2}y-\sum _{j=1}^{\infty }{2}^{2j}{A}^{2j+2}{t}_{2j}\right){C}^{-}.\end{array}\end{eqnarray}$

3. Reductions of the (2+1)-D BAKNS hierarchy

3.1. Classical reductions

The odd hierarchy (13) allows a classical complex reduction$\begin{eqnarray}r(x,y,t)=\delta {q}^{* }(x,y,t),\,\,\delta =\pm 1,\,y\to {\rm{i}}y,\end{eqnarray}$where i is the imaginary unit and ∗ stands for complex conjugate. And the representative one is$\begin{eqnarray}{q}_{{t}_{1}}={\rm{i}}{q}_{{xy}}-2{\rm{i}}\delta q{\partial }_{x}^{-1}{\left({{qq}}^{* }\right)}_{y},\,\,\delta =\pm 1.\end{eqnarray}$

The even hierarchy (15) allows classical complex reduction$\begin{eqnarray}r(x,y,t)=\delta {q}^{* }(x,y,t)\,\,\delta =\pm 1,\,\,y\to {\rm{i}}y,{t}_{2l}\to {\rm{i}}{t}_{2l},\end{eqnarray}$when l=1, the corresponding equation is$\begin{eqnarray}\begin{array}{rcl}{q}_{{t}_{2}} & = & {q}_{{xxy}}-4\delta {q}^{* }{{qq}}_{y}-2\delta {q}_{x}{\partial }_{x}^{-1}{\left({{qq}}^{* }\right)}_{y}\\ & & -2\delta q{\partial }_{x}^{-1}({q}_{x}{q}_{y}^{* }-{q}_{x}^{* }{q}_{y}),\,\,\delta =\pm 1,\end{array}\end{eqnarray}$

3.2. Nonlocal reductions

For the odd hierarchy (13), first, it allows a real nonlocal reduction$\begin{eqnarray}r(x,y,t)=\delta q(\sigma x,-y,\sigma t),\,\,\delta ,\sigma =\pm 1,\,\,x,y,t\in {\mathbb{R}},\end{eqnarray}$where the simplest one-component equation reads$\begin{eqnarray}{q}_{{t}_{1}}=-{q}_{{xy}}+2\delta q{\partial }_{x}^{-1}{\left({qq}(\sigma x,-y,\sigma t\right)}_{y},\,\,\delta ,\sigma =\pm 1.\end{eqnarray}$Equation (13) also allows a complex nonlocal reduction$\begin{eqnarray}r(x,y,t)=\delta {q}^{* }(\sigma x,-y,\sigma t),\,\,\delta ,\sigma =\pm 1,\,\,x,y,t\in {\mathbb{R}},\end{eqnarray}$and the resulted simplest one-component equation is$\begin{eqnarray}{q}_{{t}_{1}}=-{q}_{{xy}}+2\delta q{\partial }_{x}^{-1}{\left({{qq}}^{* }(\sigma x,-y,\sigma t\right)}_{y},\,\,\delta ,\sigma =\pm 1.\end{eqnarray}$

The even hierarchy (15) also allows two types of reductions. One is real$\begin{eqnarray}r(x,y,t)=\delta q(\sigma x,-y,-t),\,\,\delta ,\sigma =\pm 1,\end{eqnarray}$and the resulted simplest one-component equation is$\begin{eqnarray}\begin{array}{rcl}{q}_{{t}_{2}} & = & {q}_{{xxy}}-4\delta \widetilde{q}{{qq}}_{y}-2\delta {q}_{x}{\partial }_{x}^{-1}{\left(q\widetilde{q}\right)}_{y}\\ & & -2\delta q{\partial }_{x}^{-1}({q}_{x}{\widetilde{q}}_{y}-{\widetilde{q}}_{x}{q}_{y}),\,\,\delta ,\sigma =\pm 1,\end{array}\end{eqnarray}$where$\begin{eqnarray}\widetilde{q}=q(\sigma x,-y,-t).\end{eqnarray}$The other reduction is complex$\begin{eqnarray}r(x,y,t)=\delta {q}^{* }(\sigma x,-y,-t),\,\,\delta ,\sigma =\pm 1,\end{eqnarray}$and the resulted simplest one-component equation is$\begin{eqnarray}\begin{array}{rcl}{q}_{{t}_{2}} & = & {q}_{{xxy}}-4\delta {\widetilde{q}}^{* }{{qq}}_{y}-2\delta {q}_{x}{\partial }_{x}^{-1}{\left(q{\widetilde{q}}^{* }\right)}_{y}\\ & & -2\delta q{\partial }_{x}^{-1}({q}_{x}{\widetilde{q}}_{y}^{* }-{\widetilde{q}}_{x}^{* }{q}_{y}),\,\,\delta ,\sigma =\pm 1,\end{array}\end{eqnarray}$where $\widetilde{q}$ is given as (27).

It is remarkable that the integration operator ${\partial }_{x}^{-1}$ should specially take the form$\begin{eqnarray}{\partial }_{x}^{-1}=\displaystyle \frac{1}{2}\left({\int }_{-\infty }^{x}-{\int }_{x}^{+\infty }\right)\cdot {\rm{d}}x,\end{eqnarray}$which has played an important role in nonlocal reduction involved with x (see [16, 18]). For example, for the integration term ${\partial }_{x}^{-1}{\left({qr}\right)}_{y}$ in (5) with n=1, one can handle it as the following. First, introduce$\begin{eqnarray*}z(x,y,t)={\partial }_{x}^{-1}{\left({qr}\right)}_{y}.\end{eqnarray*}$Then we have$\begin{eqnarray*}\begin{array}{l}z(\sigma x,-y,\sigma t)\\ =\,\displaystyle \frac{-1}{2}\left({\displaystyle \int }_{-\infty }^{\sigma x}-{\displaystyle \int }_{\sigma x}^{+\infty }\right){\left[q(x,-y,\sigma t)r(x,-y,\sigma t)\right]}_{y}{\rm{d}}x\\ =\,\displaystyle \frac{-\sigma }{2}\left({\displaystyle \int }_{-\infty }^{x}-{\displaystyle \int }_{x}^{+\infty }\right){\left[q(\sigma x^{\prime} ,-y,\sigma t)r(\sigma x^{\prime} ,-y,\sigma t)\right]}_{y}{\rm{d}}x^{\prime} \\ =\,\displaystyle \frac{-\sigma }{2}\left({\displaystyle \int }_{-\infty }^{x}-{\displaystyle \int }_{x}^{+\infty }\right){\left[r(x^{\prime} ,y,t)q(x^{\prime} ,y,t)\right]}_{y}{\rm{d}}x^{\prime} \\ =\,-\sigma z(x,y,t),\end{array}\end{eqnarray*}$which provides a clear expression for the integration term with nonlocal reduction.

4. Reduction of solutions

4.1. Nonlocal cases

In the following we implement the reduction procedure on the double Wronskian solutions, by which solutions for the reduced nonlocal hierarchies can be obtained from those of the unreduced (2+1)-D BAKNS hierarchy (2) presented in theorem 1.

4.1.1. Reduction (21)

The nonlocal hierarchy$\begin{eqnarray}{q}_{{t}_{2l+1}}={K}_{\mathrm{1,2}l+1}{| }_{(21)},\,\,l=0,1,2,\cdots \end{eqnarray}$allows the following solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$(i.e. m=n in (9)), defined by (14) and satisfy$\begin{eqnarray}\psi (x,y,t)=T\varphi (\sigma x,-y,\sigma t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+},\end{eqnarray}$in which $T$ is a constant matrix determined through$\begin{eqnarray}{AT}+\sigma {TA}=0,\,\,\end{eqnarray}$$\begin{eqnarray}{T}^{2}=\sigma \delta I,\,\,\sigma ,\delta =\pm 1.\end{eqnarray}$

It follows from (14), (33b) and (34a) that$\begin{eqnarray*}\begin{array}{rcl}\psi (x,y,t) & = & \exp \left({Ax}-{A}^{2}y-\sum _{j=0}^{\infty }{2}^{2j+1}{A}^{2j+3}{t}_{2j+1}\right){C}^{-}\\ & = & \exp \left(-\sigma ({{TAT}}^{-1})x-{A}^{2}y\right.\\ & & \left.-\sum _{j=0}^{\infty }{2}^{2j+1}(-\sigma {{TA}}^{2j+3}{T}^{-1}){t}_{2j+1}\right){{TC}}^{+}\\ & = & T\exp \left(-A(\sigma x)+{A}^{2}(-y)\right.\\ & & \left.+\sum _{j=0}^{\infty }{2}^{2j+1}{A}^{2j+3}(\sigma {t}_{2j+1}\right){C}^{+}\\ & = & T\varphi (\sigma x,-y,\sigma t)^{\prime} ,\end{array}\end{eqnarray*}$which gives rise to the the constraint (33a). Thus, f, g, h in (8) (with m=n) are expressed as$\begin{eqnarray}\begin{array}{rcl}f(x,y,t) & = & | {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| \\ & = & | {\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[x]};T{\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| ,\end{array}\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}g(x,y,t) & = & 2| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| \\ & = & 2| {\widehat{\varphi }}^{(n+1)}{\left(x,y,t\right)}_{[x]};T{\widehat{\varphi }}^{(n-1)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| ,\end{array}\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}h(x,y,t) & = & 2| {\widehat{\varphi }}^{(n-1)};{\widehat{\psi }}^{(n+1)}| \\ & = & 2| {\widehat{\varphi }}^{(n-1)}{\left(x,y,t\right)}_{[x]};T{\widehat{\varphi }}^{(n+1)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| ,\end{array}\end{eqnarray}$where we employed a notation (see [15])$\begin{eqnarray*}{\widehat{\varphi }}^{(n)}{\left({ax}\right)}_{[{bx}]}=\left(\varphi ({ax}),{\partial }_{{bx}}\varphi ({ax}),{\partial }_{{bx}}^{2}\varphi ({ax}),\cdots ,{\partial }_{{bx}}^{n}\varphi ({ax}\right).\end{eqnarray*}$Next, with assumption (34b), we have$\begin{eqnarray*}\begin{array}{l}f(\sigma x,-y,\sigma t)\\ =\,| {\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[\sigma x]};T{\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[\sigma x]}| \\ =\,| T| | {T}^{-1}{\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[\sigma x]};{\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[\sigma x]}| \\ =\,{\left(-1\right)}^{{\left(n+1\right)}^{2}}| T| | {\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[x]};{T}^{-1}{\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| \\ =\,{\left(-1\right)}^{{\left(n+1\right)}^{2}}| T| | {\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[x]};\sigma \delta T{\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| \\ =\,{\left(-1\right)}^{{\left(n+1\right)}^{2}}{\left(\sigma \delta \right)}^{n+1}| T| | {\widehat{\varphi }}^{(n)}{\left(x,y,t\right)}_{[x]};T{\widehat{\varphi }}^{(n)}{\left(\sigma x,-y,\sigma t\right)}_{[x]}| \\ =\,{\left(-1\right)}^{{\left(n+1\right)}^{2}}{\left(\sigma \delta \right)}^{n+1}| T| f(x,y,t),\end{array}\end{eqnarray*}$and in a similar way$\begin{eqnarray*}g(\sigma x,-y,\sigma t)={\left(-1\right)}^{(n+2)n}{\sigma }^{n+1}{\delta }^{n}| T| h(x,y,t),\end{eqnarray*}$which then give rise to$\begin{eqnarray*}\begin{array}{rcl}\delta q(\sigma x,-y,\sigma t) & = & \delta \displaystyle \frac{g(\sigma x,-y,\sigma t)}{f(\sigma x,-y,\sigma t)}\\ & = & -\displaystyle \frac{h(x,y,t)}{f(x,y,t)}=r(x,y,t).\end{array}\end{eqnarray*}$This is nothing but the reduction (21). Thus we complete the proof.

Solutions of (34) can be written out by assuming T and A to be block matrices$\begin{eqnarray}T=\left(\begin{array}{cc}{T}_{1} & {T}_{2}\\ {T}_{3} & {T}_{4}\end{array}\right),\,\,A=\left(\begin{array}{cc}{K}_{1} & 0\\ 0 & {K}_{4}\end{array}\right),\end{eqnarray}$where Ti and Ki are (n+1)×(n+1) matrices and Ki is a complex matrix. We list them in table 1.

One can prove that A and any of its similar forms lead to same solution (32). We only need to consider canonical form of A. When$\begin{eqnarray}{{\bf{K}}}_{n+1}=\mathrm{Diag}({k}_{1},{k}_{2},\cdots ,{k}_{n+1}),\end{eqnarray}$we get$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\zeta ({k}_{1})},{c}_{2}{{\rm{e}}}^{\zeta ({k}_{2})},\cdots ,{c}_{n+1}{{\rm{e}}}^{\zeta ({k}_{n+1})},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\zeta (-{k}_{1})},{d}_{2}{{\rm{e}}}^{\zeta (-{k}_{2})},\cdots ,{d}_{n+1}{{\rm{e}}}^{\zeta (-{k}_{n+1})}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$When ${{\bf{K}}}_{n+1}$ is a $(n+1)\times (n+1)$ Jordan matrix ${J}_{n+1}(k)$$\begin{eqnarray}{J}_{n+1}(k)={\left(\begin{array}{cccc}k & 0 & \cdots & 0\\ 1 & k & \cdots & 0\\ \vdots & \ddots & \ddots & \vdots \\ 0 & \cdots & 1 & k\end{array}\right)}_{(n+1)\times (n+1)},\end{eqnarray}$we get$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\zeta (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\zeta (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\zeta (k)}),\right.\\ & & {\left.d{{\rm{e}}}^{\zeta (-k)},\displaystyle \frac{{\partial }_{k}}{1!}(d{{\rm{e}}}^{\zeta (-k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(d{{\rm{e}}}^{\zeta (-k)}\right)}^{{\rm{T}}},\end{array}\end{eqnarray}$where$\begin{eqnarray}\zeta ({k}_{i})=-{k}_{i}x+{k}_{i}^{2}y+\sum _{j=0}^{\infty }{2}^{2j+1}{k}_{i}^{2j+3}{t}_{2j+1}.\end{eqnarray}$

As examples, for equation (22) with different (σ, δ), its one-soliton solution (1SS) are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{4{cdk}{{\rm{e}}}^{2{k}^{2}y}}{{c}^{2}{{\rm{e}}}^{-2{kx}+4{k}^{3}t}+{d}^{2}{{\rm{e}}}^{2{kx}-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{4{cdk}{{\rm{e}}}^{2{k}^{2}y}}{{c}^{2}{{\rm{e}}}^{-2{kx}+4{k}^{3}t}-{d}^{2}{{\rm{e}}}^{2{kx}-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{4k{{\rm{e}}}^{2{k}^{2}y}}{-{{\rm{e}}}^{-2{kx}+4{k}^{3}t}-{{\rm{e}}}^{2{kx}-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{4k{{\rm{e}}}^{2{k}^{2}y}}{-{\mathrm{ie}}^{-2{kx}+4{k}^{3}t}-{\mathrm{ie}}^{2{kx}-4{k}^{3}t}},\end{eqnarray}$where we have taken $k={k}_{1},\,t={t}_{1},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

Next, let us quickly investigate dynamics of ${q}_{(\sigma ,\delta )}={q}_{(1,-1)}$ which is governed by equation$\begin{eqnarray}{q}_{{t}_{1}}=-{q}_{{xy}}+2\delta q{\partial }_{x}^{-1}{\left({qq}(x,-y,t\right)}_{y}.\end{eqnarray}$Its 1SS (42a) is rewritten as$\begin{eqnarray}q(x,y,t)=\displaystyle \frac{2{kcd}\,{{\rm{e}}}^{2{k}^{2}y}}{| {cd}| }{\rm{sech}} \left(-2{kx}+4{k}^{3}t+\mathrm{ln}\displaystyle \frac{| c| }{| d| }\right),\end{eqnarray}$which is a moving wave with an initial phase $\mathrm{ln}\tfrac{| c| }{| d| }$ and a y-dependent amplitude $\tfrac{2{cdk}\,{{\rm{e}}}^{2{k}^{2}y}}{| {cd}| }$. The top trajectory is $x(t)=2{k}^{2}t+\tfrac{\mathrm{ln}\left|\tfrac{c}{d}\right|}{2k}$, and travel velocity is 2k2. Fixing y, the trajectory of one-solton in coordinate frame {x, t} is depicted as figure 1(a).

Figure 1.

New window|Download| PPT slide
Figure 1.(a). Shape and motion of 1SS (42a) for equation (43), in which k=1, c=1, d=1, y=0. (b). Shape and motion of 2SS (45) for equation (43), in which ${k}_{1}=0.8,{k}_{2}=-0.6,{c}_{1}=1$, ${c}_{2}=1,{d}_{1}=1,{d}_{2}=2,y=0$. (c). Shape and motion of 2SS (45) for equation (43), in which ${k}_{1}=1.2,{k}_{2}=-0.8,{c}_{1}=1,{c}_{2}=1,{d}_{1}=-1,{d}_{2}=2,y=0$. 1(d). Breather provided by (45) for equation (43), in which ${k}_{1}=0.8+0.6i,{k}_{2}=0.8-0.6i,{c}_{1}=1$, ${c}_{2}=1,{d}_{1}=1,{d}_{2}=1,y=0$.


The 2SS of equation (43) can be written as$\begin{eqnarray}q(x,y,t)=\displaystyle \frac{A}{B},\end{eqnarray}$where$\begin{eqnarray*}\begin{array}{rcl}A & = & 4\left({k}_{1}^{2}-{k}_{2}^{2}\right){{\rm{e}}}^{2\left({k}_{1}^{2}+{k}_{2}^{2}\right)y}\left({c}_{2}{c}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}\left(2{k}_{2}^{2}t+{k}_{2}y+x\right)+8{k}_{1}^{3}t}\right.\\ & & \left.+{c}_{2}{d}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{4{k}_{2}^{3}t+2{k}_{2}x+4{k}_{1}x+2{k}_{2}^{2}y}\right)\\ & & -4{c}_{1}{d}_{1}{k}_{1}\left({k}_{1}^{2}-{k}_{2}^{2}\right)\left({c}_{2}^{2}{{\rm{e}}}^{8{k}_{2}^{3}t}\right.\\ & & \left.+{d}_{2}^{2}{{\rm{e}}}^{4{k}_{2}x}\right){{\rm{e}}}^{\left(2{k}_{1}\left(2{k}_{1}^{2}t+{k}_{1}y+x\right)+2\left({k}_{1}^{2}+{k}_{2}^{2}\right)y\right)},\end{array}\end{eqnarray*}$$\begin{eqnarray*}\begin{array}{rcl}B & = & -{c}_{2}^{2}\left({{\rm{e}}}^{2{k}_{2}^{2}\left(4{k}_{2}t+y\right)+2{k}_{1}^{2}y}\right)\\ & & \times \left({c}_{1}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}{{\rm{e}}}^{8{k}_{1}^{3}t}+{d}_{1}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}{{\rm{e}}}^{4{k}_{1}x}\right)\\ & & +4{c}_{1}{c}_{2}{d}_{1}{d}_{2}{k}_{1}{k}_{2}\left({{\rm{e}}}^{4{k}_{1}^{2}y}+{{\rm{e}}}^{4{k}_{2}^{2}y}\right){{\rm{e}}}^{4{k}_{1}^{3}t+4{k}_{2}^{3}t+2{k}_{1}x+2{k}_{2}x}\\ & & -{d}_{2}^{2}{{\rm{e}}}^{4{k}_{2}x+2{k}_{1}^{2}y+2{k}_{2}^{2}y}\left({c}_{1}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}{{\rm{e}}}^{8{k}_{1}^{3}t}\right.\\ & & \left.+{d}_{1}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}{{\rm{e}}}^{4{k}_{1}x}\right),\end{array}\end{eqnarray*}$and we note that 2SS is non-singular if ${c}_{1}{c}_{2}{d}_{1}{d}_{2}{k}_{1}{k}_{2}\lt 0$. As shown in figures 1(b)–(d), there are three types of two-soliton interactions, respectively are soliton-soliton interaction, soliton-anti-soliton interaction and breather soliton. To investigate asymptotic behavior of the first two types solution interactions, we make asymptotic analysis for t going to infinity in the case of ${k}_{1}\gt -{k}_{2}\gt 0$. To do that, we first rewrite the 2SS (45) in the following coordinate$\begin{eqnarray}({X}_{1}=x-2{k}_{1}^{2}t,t),\end{eqnarray}$fix X1, let $t\to \pm \infty $, and we get$\begin{eqnarray}\begin{array}{l}q(x,y,t)\\ \sim \,\left\{\begin{array}{ll}\tfrac{2{c}_{1}{d}_{1}{k}_{1}{{\rm{e}}}^{2{k}_{1}^{2}y}}{| {c}_{1}| | {d}_{1}| }{\rm{sech}} (2{k}_{1}{X}_{1}+\mathrm{ln}\tfrac{| {d}_{1}| ({k}_{1}+{k}_{2})}{| {c}_{1}| ({k}_{1}-{k}_{2})}), & t\to +\infty ,\\ \tfrac{2{c}_{1}{d}_{1}{k}_{1}{{\rm{e}}}^{2{k}_{1}^{2}y}}{| {c}_{1}| | {d}_{1}| }{\rm{sech}} (2{k}_{1}{X}_{1}+\mathrm{ln}\tfrac{| {d}_{1}| ({k}_{1}-{k}_{2})}{| {c}_{1}| ({k}_{1}+{k}_{2})}), & t\to -\infty .\end{array}\right.\end{array}\end{eqnarray}$It shows that along the line X1=constant, there is a soliton; when $t\to \pm \infty $, the soliton asymptotically follows$\begin{eqnarray}{\rm{top}}\,{\rm{trajectory}}:\,\,\,x(t)=2{k}_{1}^{2}t-\displaystyle \frac{\mathrm{ln}| \tfrac{{d}_{1}}{{c}_{1}}| }{2{k}_{1}}\pm \displaystyle \frac{\mathrm{ln}\tfrac{{k}_{1}+{k}_{2}}{{k}_{1}-{k}_{2}}}{2{k}_{1}},\end{eqnarray}$$\begin{eqnarray}{\rm{amplitude}}:\,\,\,\displaystyle \frac{2{c}_{1}{d}_{1}{k}_{1}{{\rm{e}}}^{2{k}_{1}^{2}y}}{| {c}_{1}| | {d}_{1}| },\end{eqnarray}$and phase shift after interaction is $\tfrac{{\rm{ln}}\tfrac{{k}_{1}+{k}_{2}}{{k}_{1}-{k}_{2}}}{{k}_{1}}$.

We can also rewrite the 2SS (45) in the coordinate $({X}_{2}=x-2{k}_{2}^{2}t,t)$. By fixing X2, we can obtain$\begin{eqnarray}\begin{array}{l}q(x,y,t)\\ \sim \,\left\{\begin{array}{ll}\tfrac{-2{c}_{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}^{2}y}}{| {c}_{2}| | {d}_{2}| }{\rm{sech}} (2{k}_{2}{X}_{2}+\mathrm{ln}\tfrac{| {d}_{2}| ({k}_{1}+{k}_{2})}{| {c}_{2}| ({k}_{1}-{k}_{2})}), & t\to +\infty ,\\ \tfrac{-2{c}_{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}^{2}y}}{| {c}_{2}| | {d}_{2}| }{\rm{sech}} (2{k}_{2}{X}_{2}+\mathrm{ln}\tfrac{| {d}_{2}| ({k}_{1}-{k}_{2})}{| {c}_{2}| ({k}_{1}+{k}_{2})}), & t\to -\infty .\end{array}\right.\end{array}\end{eqnarray}$It indicates that along the line X2= constant, there is a soliton; when $t\to \pm \infty $, the soliton asymptotically follows$\begin{eqnarray}{\rm{top}}\,{\rm{trajectory}}:\,\,\,x(t)=2{k}_{2}^{2}t-\displaystyle \frac{\mathrm{ln}| \tfrac{{d}_{2}}{{c}_{2}}| }{2{k}_{2}}\pm \displaystyle \frac{\mathrm{ln}\tfrac{{k}_{1}+{k}_{2}}{{k}_{1}-{k}_{2}}}{2{k}_{2}},\end{eqnarray}$$\begin{eqnarray}{\rm{amplitude}}:\,\,\,\displaystyle \frac{-2{c}_{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}^{2}y}}{| {c}_{2}| | {d}_{2}| },\end{eqnarray}$and the phase shift after interaction is $\tfrac{{\rm{ln}}\tfrac{{k}_{1}+{k}_{2}}{{k}_{1}-{k}_{2}}}{{k}_{2}}$.

Supposing ${k}_{1}\gt -{k}_{2}\gt 0,{c}_{1}{c}_{2}{d}_{1}{d}_{2}{k}_{1}{k}_{2}\lt 0$, from the asymptotic analysis, we can obtain that if $\mathrm{sgn}[{c}_{1}{d}_{1}]\,=\mathrm{sgn}[{c}_{2}{d}_{2}]$, it will appear soliton-soliton interaction, such as figure 1(b); if $\mathrm{sgn}[{c}_{1}{d}_{1}]=-\mathrm{sgn}[{c}_{2}{d}_{2}]$, it will appear soliton-anti-soliton interaction, such as figure 1(c).

We list out solutions for other cases of nonlocal reductions without giving proof.

4.1.2. Reduction (23)

For the reduction (23), the reduced nonlocal hierarchy$\begin{eqnarray}{q}_{{t}_{2l+1}}={K}_{\mathrm{1,2}l+1}{| }_{(23)},\,\,l=0,1,2,\cdots \end{eqnarray}$allow a solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi $ and $\psi $ are $(2n+2)$th order column vectors (i.e. $m=n$ in (9)), defined by (14) and satisfy$\begin{eqnarray}\psi (x,y,t)=T{\varphi }^{* }(\sigma x,-y,\sigma t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+* },\end{eqnarray}$in which $T$ is a constant matrix determined through$\begin{eqnarray}{AT}+\sigma {{TA}}^{* }=0,\,\,\end{eqnarray}$$\begin{eqnarray}{{TT}}^{* }=\sigma \delta I,\,\,\sigma ,\delta =\pm 1.\end{eqnarray}$

When T and A are block matrices (36), solutions to equations (54) are given in table 2.


Table 2.
Table 2.T and A for (54).
σ,δTA
(1, −1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}=-{T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}=-{K}_{4}^{* }={{\bf{K}}}_{n+1}$
(1, 1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}={T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}=-{K}_{4}^{* }={{\bf{K}}}_{n+1}$
(−1, −1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}={T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}={K}_{4}^{* }={{\bf{K}}}_{n+1}$
(−1, 1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}=-{T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}={K}_{4}^{* }={{\bf{K}}}_{n+1}$

New window|CSV

When ${{\bf{K}}}_{n+1}$ is a diagonal matrix (37), the vector φ in (52) can be given as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\left.\theta ({k}_{1}\right)},{c}_{2}{{\rm{e}}}^{\left.\theta ({k}_{2}\right)},\cdots ,{c}_{n+1}{{\rm{e}}}^{\left.\theta ({k}_{n+1}\right)},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\theta (-\sigma {k}_{1}^{* })},{d}_{2}{{\rm{e}}}^{\theta (-\sigma {k}_{2}^{* })},\cdots ,{d}_{n+1}{{\rm{e}}}^{\theta (-\sigma {k}_{n+1}^{* })}\right)}^{{\rm{T}}},\end{array}\end{eqnarray}$where$\begin{eqnarray}\theta ({k}_{i})=-{k}_{i}x+{k}_{i}^{2}y+\sum _{j=0}^{\infty }{2}^{2j+1}{k}_{i}^{2j+3}{t}_{2j+1}.\end{eqnarray}$When ${{\bf{K}}}_{n+1}={J}_{n+1}(k)$ (39), we have$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\theta (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\theta (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\theta (k)}),d{{\rm{e}}}^{\theta (-\sigma {k}^{* })},\right.\\ & & {\left.\displaystyle \frac{{\partial }_{{k}^{* }}}{1!}(d{{\rm{e}}}^{\theta (-\sigma {k}^{* })}),\cdots ,\displaystyle \frac{{\partial }_{{k}^{* }}^{n}}{n!}(d{{\rm{e}}}^{\theta (-\sigma {k}^{* })}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$

For the equation (24) with different (σ, δ), its 1SS are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{k}^{* 2}y+4{k}^{* 3}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{k}^{* 2}y+4{k}^{* 3}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{2{cd}(k-{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x-2{k}^{* 2}y+4{k}^{* 3}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{2{cd}(k-{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x-2{k}^{* 2}y+4{k}^{* 3}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{1},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

4.1.3. Reduction (25)

For the reduction (25), the reduced nonlocal hierarchy$\begin{eqnarray}{q}_{{t}_{2l}}={K}_{\mathrm{1,2}l}{| }_{(25)},\,\,l=1,2,\cdots \end{eqnarray}$allow a solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\hat{\varphi }}^{(n+1)};{\hat{\psi }}^{(n-1)}| }{| {\hat{\varphi }}^{(n)};{\hat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$, defined by (16) and satisfy$\begin{eqnarray}\psi (x,y,t)=T\varphi (\sigma x,-y,-t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+},\end{eqnarray}$in which $T$ is a constant matrix satisfying (34).

If T and A are block matrices (36), solutions to (34) haven been given by table 1. In the case that ${{\bf{K}}}_{n+1}$ is a diagonal matrix (37), the vector φ in (60) can be given as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\xi ({k}_{1})},{c}_{2}{{\rm{e}}}^{\xi ({k}_{2})},\cdots ,{c}_{n+1}{{\rm{e}}}^{\xi ({k}_{n+1})},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\xi (-{k}_{1})},{d}_{2}{{\rm{e}}}^{\xi (-{k}_{2})},\cdots ,{d}_{n+1}{{\rm{e}}}^{\xi (-{k}_{n+1})}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$where$\begin{eqnarray}\xi ({k}_{i})=-{k}_{i}x+{k}_{i}^{2}y+\sum _{j=1}^{\infty }{2}^{2j}{k}_{i}^{2j+2}{t}_{2j}.\end{eqnarray}$When ${{\bf{K}}}_{n+1}={J}_{n+1}(k)$ (39), we have$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\xi (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\xi (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\xi (k)}),\right.\\ & & {\left.d{{\rm{e}}}^{\xi (-k)},\displaystyle \frac{{\partial }_{k}}{1!}(d{{\rm{e}}}^{\xi (-k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(d{{\rm{e}}}^{(-k)}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$


Table 1.
Table 1.T and A for (34).
(σ, δ)TA
(1, −1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}=-{T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}=-{K}_{4}={{\bf{K}}}_{n+1}$
(1, 1)${T}_{1}={T}_{4}={{\bf{0}}}_{n+1},{T}_{3}={T}_{2}={{\bf{I}}}_{n+1}$${K}_{1}=-{K}_{4}={{\bf{K}}}_{n+1}$
(−1, −1)${T}_{1}=-{T}_{4}={{\bf{I}}}_{n+1},{T}_{3}={T}_{2}={{\bf{0}}}_{n+1}$${K}_{1}=-{K}_{4}={{\bf{K}}}_{n+1}$
(−1, 1)${T}_{1}=-{T}_{4}=i{{\bf{I}}}_{n+1},{T}_{3}={T}_{2}={{\bf{0}}}_{n+1}$${K}_{1}=-{K}_{4}={{\bf{K}}}_{n+1}$

New window|CSV

For the equation (26) with different (σ, δ), its 1SS are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{4{cdk}{{\rm{e}}}^{2{k}^{2}y+8{k}^{4}t}}{{c}^{2}{{\rm{e}}}^{-2{kx}}+{d}^{2}{{\rm{e}}}^{2{kx}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{4{cdk}{{\rm{e}}}^{2{k}^{2}y+8{k}^{4}t}}{{c}^{2}{{\rm{e}}}^{-2{kx}}-{d}^{2}{{\rm{e}}}^{2{kx}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{4k{{\rm{e}}}^{2{k}^{2}y+8{k}^{4}t}}{-{{\rm{e}}}^{-2{kx}}-{{\rm{e}}}^{2{kx}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{4{\rm{i}}k{{\rm{e}}}^{2{k}^{2}y+8{k}^{4}t}}{-{{\rm{e}}}^{-2{kx}}-{{\rm{e}}}^{2{kx}}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{2},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

As an example we consider the dynamics of ${q}_{(\sigma ,\delta )}={q}_{(1,-1)}$, which is governed by equation$\begin{eqnarray}{q}_{{t}_{2}}={q}_{{xxy}}+4\widetilde{q}{{qq}}_{y}+2{q}_{x}{\partial }_{x}^{-1}{\left(q\widetilde{q}\right)}_{y}+2q{\partial }_{x}^{-1}({q}_{x}{\widetilde{q}}_{y}-{\widetilde{q}}_{x}{q}_{y}),\end{eqnarray}$where $\widetilde{q}=q(x,-y,-t)$. Its 1SS (65a) can be rewritten as$\begin{eqnarray}q(x,y,t)=\displaystyle \frac{2{cdk}\,{{\rm{e}}}^{2{k}^{2}y+8{k}^{4}t}}{| {cd}| }{\rm{{\rm{sech}} }}\left(-2{kx}+\mathrm{ln}| \displaystyle \frac{c}{d}| \right).\end{eqnarray}$

Depicted as figure 2(a), this is a stationary wave with an initial phase $\mathrm{ln}\left|\tfrac{c}{d}\right|$ and an amplitude that exponentially increases with time t and y, and the top trajectory is $x(t)=\tfrac{\mathrm{ln}\left|\tfrac{c}{d}\right|}{2k}$. It is interesting that the amplitude is y and t-dependent, which looks like in nonisospectral case (see [35]) where amplitudes are changed due to time-dependent eigenvalues (spectral parameters). However, here the amplitude changes might be caused by the reversed y and t in the nonlocal reduction, rather than nonspectral parameters as the eigenvalues (see A in table 1) are still constant.

Figure 2.

New window|Download| PPT slide
Figure 2.(a). Shape and motion of 1SS (65a) for equation (66), in which $k=0.67,c=1,d=1,y=0$. (b). Shape and motion of 2SS (68) for equation (66), in which ${k}_{1}=0.8,{k}_{2}=-0.6,{c}_{1}=2,{c}_{2}=1$, ${d}_{1}=1,{d}_{2}=1,y=0$. (c). Shape and motion of $-{{\rm{e}}}^{-8t{({k}_{1}{k}_{2})}^{2}}{q}_{1,-1}(x,t)$ where ${q}_{1,-1}(x,t)$ is depicted in (b).


The 2SS of equation (66) can be written as$\begin{eqnarray}q(x,y,t)=\displaystyle \frac{G}{F},\end{eqnarray}$where$\begin{eqnarray*}\begin{array}{rcl}G & = & -4{c}_{1}{d}_{1}{k}_{1}\left({k}_{1}^{2}-{k}_{2}^{2}\right)\left({c}_{2}^{2}\right.\\ & & \left.+{d}_{2}^{2}{{\rm{e}}}^{4{k}_{2}x}\right){{\rm{e}}}^{8{k}_{1}^{4}t+8{k}_{2}^{4}t+2{k}_{1}^{2}y+2{k}_{2}^{2}y+2{k}_{1}\left(4{k}_{1}^{3}t+{k}_{1}y+x\right)}\\ & & +4\left({k}_{1}^{2}-{k}_{2}^{2}\right){{\rm{e}}}^{8{k}_{1}^{4}t+8{k}_{2}^{4}t+2{k}_{1}^{2}y+2{k}_{2}^{2}y}\left({c}_{2}{c}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}\left(4{k}_{2}^{3}t+{k}_{2}y+x\right)}\right.\\ & & \left.+{c}_{2}{d}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{8{k}_{2}^{4}t+2{k}_{2}x+4{k}_{1}x+2{k}_{2}^{2}y}\right),\end{array}\end{eqnarray*}$$\begin{eqnarray*}\begin{array}{rcl}F & = & -{d}_{1}^{2}\left({c}_{2}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}\right.\\ & & \left.+{d}_{2}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}{{\rm{e}}}^{4{k}_{2}x}\right){{\rm{e}}}^{\left(8{k}_{1}^{4}t+8{k}_{2}^{4}t+4{k}_{1}x+2{k}_{1}^{2}y+2{k}_{2}^{2}y\right)}\\ & & +4{c}_{1}{c}_{2}{d}_{1}{d}_{2}{k}_{1}{k}_{2}{{\rm{e}}}^{2\left({k}_{1}+{k}_{2}\right)x}\left({{\rm{e}}}^{4{k}_{1}^{2}\left(4{k}_{1}^{2}t+y\right)}+{{\rm{e}}}^{4{k}_{2}^{2}\left(4{k}_{2}^{2}t+y\right)}\right)\\ & & -{c}_{1}^{2}\left({{\rm{e}}}^{8{k}_{1}^{4}t+8{k}_{2}^{4}t+2{k}_{1}^{2}y+2{k}_{2}^{2}y}\right)\\ & & \times \left({c}_{2}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}+{d}_{2}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}{{\rm{e}}}^{4{k}_{2}x}\right).\end{array}\end{eqnarray*}$

4.1.4. Reduction (28)

For the reduction (28), the reduced nonlocal hierarchy$\begin{eqnarray}{q}_{{t}_{2l}}={K}_{\mathrm{1,2}l}{| }_{(28)},\,\,l=1,2,\cdots \end{eqnarray}$allow a solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi $ and $\psi $ are defined by (9) with $m=n$ and (16), satisfying$\begin{eqnarray}\psi (x,y,t)=T{\varphi }^{* }(\sigma x,-y,-t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+* },\end{eqnarray}$in which $T$ is a constant matrix satisfied (54).

When T and A are block matrices (36), solutions to (54) are already given by table 2. In the case that ${{\bf{K}}}_{n+1}$ is a diagonal matrix (37), the vector φ in (5) can be written as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\eta ({k}_{1})},{c}_{2}{{\rm{e}}}^{\eta ({k}_{2})},\cdots ,{c}_{n+1}{{\rm{e}}}^{\eta ({k}_{n+1})},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\eta (-\sigma {k}_{1}^{* })},{d}_{2}{{\rm{e}}}^{\eta (-\sigma {k}_{2}^{* })},\cdots ,{d}_{n+1}{{\rm{e}}}^{\eta (-\sigma {k}_{n+1}^{* })}\right)}^{{\rm{T}}},\end{array}\end{eqnarray}$where$\begin{eqnarray}\eta ({k}_{i})=-{k}_{i}x+{k}_{i}^{2}y+\sum _{j=1}^{\infty }{2}^{2j}{k}_{i}^{2j+2}{t}_{2j}.\end{eqnarray}$When ${{\bf{K}}}_{n+1}={J}_{n+1}(k)$ (39), we get$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\eta (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\eta (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\eta (k)}),\right.\\ & & {\left.d{{\rm{e}}}^{\eta (-\sigma {k}^{* })},\displaystyle \frac{{\partial }_{{k}^{* }}}{1!}(d{{\rm{e}}}^{\eta (-\sigma {k}^{* })}),\cdots ,\displaystyle \frac{{\partial }_{{k}^{* }}^{n}}{n!}(d{{\rm{e}}}^{\eta (-\sigma {k}^{* })}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$

For equation (29) with different (σ, δ), its 1SS are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{k}^{* 2}y-8{k}^{* 4}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-8{k}^{4}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{k}^{* 2}y-8{k}^{* 4}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-8{k}^{4}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{2{cd}(k-{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x-2{k}^{* 2}y-8{k}^{* 4}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-8{k}^{4}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{2{cd}(k-{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x-2{k}^{* 2}y-8{k}^{* 4}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{k}^{2}y-8{k}^{4}t}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{2},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

4.2. Classical cases

For the classical reduction (17), the reduced local hierarchy$\begin{eqnarray}{q}_{{t}_{2l+1}}={K}_{\mathrm{1,2}l+1}{| }_{(17)},\,\,l=0,1,2,\cdots \end{eqnarray}$allow a solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\hat{\varphi }}^{(n+1)};{\hat{\psi }}^{(n-1)}| }{| {\hat{\varphi }}^{(n)};{\hat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi $ and $\psi $ are defined by (9) with $m=n$ and (14), satisfying$\begin{eqnarray}\psi (x,y,t)=T{\varphi }^{* }(x,y,t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+* },\end{eqnarray}$in which $T$ is a constant matrix determined through$\begin{eqnarray}{AT}+{{TA}}^{* }=0,\,\,\end{eqnarray}$$\begin{eqnarray}{{TT}}^{* }=\delta I,\,\,\delta =\pm 1.\end{eqnarray}$

When T and A are block matrices (36), solutions to equations (79) are given in table 2 with σ=1. When ${{\bf{K}}}_{n+1}$ is a diagonal matrix (37), the vector φ in (77) can be given as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\theta ({k}_{1})},{c}_{2}{{\rm{e}}}^{\theta ({k}_{2})},\cdots ,{c}_{n+1}{{\rm{e}}}^{\theta ({k}_{n+1})},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\theta (-{k}_{1}^{* })},{d}_{2}{{\rm{e}}}^{\theta (-{k}_{2}^{* })},\cdots ,{d}_{n+1}{{\rm{e}}}^{\theta (-{k}_{n+1}^{* })}\right)}^{{\rm{T}}},\end{array}\end{eqnarray}$where$\begin{eqnarray}\theta ({k}_{j})=-{k}_{j}x+{\rm{i}}{k}_{j}^{2}y+\sum _{l=0}^{\infty }{2}^{2l+1}{k}_{j}^{2l+3}{t}_{2l+1}.\end{eqnarray}$When ${{\bf{K}}}_{n+1}={J}_{n+1}(k)$ (39), we have$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\theta (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\theta (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\theta (k)}),\right.\\ & & {\left.d{{\rm{e}}}^{\theta (-{k}^{* })},\displaystyle \frac{{\partial }_{{k}^{* }}}{1!}(d{{\rm{e}}}^{\theta (-{k}^{* })}),\cdots ,\displaystyle \frac{{\partial }_{{k}^{* }}^{n}}{n!}(d{{\rm{e}}}^{\theta (-{k}^{* })}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$In addition, we present the 1SS of the equation (18)$\begin{eqnarray}{q}_{\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{\rm{i}}{k}^{* 2}y+4{k}^{* 3}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{\rm{i}}{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\delta =-1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{\rm{i}}{k}^{* 2}y+4{k}^{* 3}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{\rm{i}}{k}^{2}y-4{k}^{3}t}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{1},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

For the reduction (19), the reduced local hierarchy$\begin{eqnarray}{q}_{{t}_{2l}}={\rm{i}}{K}_{\mathrm{1,2}l}{| }_{(19)},\,\,l=1,2,\cdots \end{eqnarray}$allow a solution$\begin{eqnarray}q(x,y,t)=2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$ expressed as (16) and satisfy the condition (78), in which $T$ is a constant matrix satisfied (79).

When T and A are block matrices (36), solutions to (79) are already given by table 2. In the case that ${{\bf{K}}}_{n+1}$ is a diagonal matrix (37), the vector φ in (85) can be written as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left({c}_{1}{{\rm{e}}}^{\eta ({k}_{1})},{c}_{2}{{\rm{e}}}^{\eta ({k}_{2})},\cdots ,{c}_{n+1}{{\rm{e}}}^{\eta ({k}_{n+1})},\right.\\ & & {\left.{d}_{1}{{\rm{e}}}^{\eta (-{k}_{1}^{* })},{d}_{2}{{\rm{e}}}^{\eta (-{k}_{2}^{* })},\cdots ,{d}_{n+1}{{\rm{e}}}^{\eta (-{k}_{n+1}^{* })}\right)}^{{\rm{T}}},\end{array}\end{eqnarray}$where$\begin{eqnarray}\eta ({k}_{j})=-{k}_{j}x+{\rm{i}}{k}_{j}^{2}y+\sum _{l=1}^{\infty }{2}^{2l}{\rm{i}}{k}_{j}^{2l+2}{t}_{2l}.\end{eqnarray}$When ${{\bf{K}}}_{n+1}={J}_{n+1}(k)$ (39), we get$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \left(c{{\rm{e}}}^{\eta (k)},\displaystyle \frac{{\partial }_{k}}{1!}(c{{\rm{e}}}^{\eta (k)}),\cdots ,\displaystyle \frac{{\partial }_{k}^{n}}{n!}(c{{\rm{e}}}^{\eta (k)}),\right.\\ & & {\left.d{{\rm{e}}}^{\eta (-{k}^{* })},\displaystyle \frac{{\partial }_{{k}^{* }}}{1!}(d{{\rm{e}}}^{\eta (-{k}^{* })}),\cdots ,\displaystyle \frac{{\partial }_{{k}^{* }}^{n}}{n!}(d{{\rm{e}}}^{\eta (-{k}^{* })}\right)}^{{\rm{T}}}.\end{array}\end{eqnarray}$For equation (20), its 1SS with different δ are given as$\begin{eqnarray}{q}_{\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{\rm{i}}{k}^{* 2}y-8{\rm{i}}{k}^{* 4}t}-| d{| }^{2}{{\rm{e}}}^{2{kx}-2{\rm{i}}{k}^{2}y-8{\rm{i}}{k}^{4}t}},\end{eqnarray}$$\begin{eqnarray}{q}_{\delta =-1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-2{\rm{i}}{k}^{* 2}y-8{\rm{i}}{k}^{* 4}t}+| d{| }^{2}{{\rm{e}}}^{2{kx}-2{\rm{i}}{k}^{2}y-8{\rm{i}}{k}^{4}t}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{2},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

5. Negative order AKNS hierarchy

In this section, we list double Wronskian solutions for the negative order AKNS hierarchy and consider several cases of reductions.

5.1. Solutions

For the negative order AKNS hierarchy [33]$\begin{eqnarray}\left(\begin{array}{c}{q}_{{t}_{n}}\\ {r}_{{t}_{n}}\end{array}\right)={K}_{n}=\left(\begin{array}{c}{K}_{1,n}\\ {K}_{2,n}\end{array}\right)={L}^{-n}\left(\begin{array}{c}-q\\ r\end{array}\right),\,\,n=1,2,3,\cdots ,\end{eqnarray}$where L is defined by (3). The hierarchy can be rewritten as [36]$\begin{eqnarray}q={q}_{x,{t}_{1}}-2q{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{1}},\end{eqnarray}$$\begin{eqnarray}r={r}_{x,{t}_{1}}-2r{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{1}},\end{eqnarray}$$\begin{eqnarray}{q}_{{t}_{n-1}}=-{q}_{x,{t}_{n}}+2q{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{n}},\end{eqnarray}$$\begin{eqnarray}{r}_{{t}_{n-1}}={r}_{x,{t}_{n}}-2r{\partial }_{x}^{-1}{\left({qr}\right)}_{{t}_{n}},\,\,(n=2,3,\cdots ).\end{eqnarray}$By the transformation$\begin{eqnarray}q=-\displaystyle \frac{g}{f},\,\,r=\displaystyle \frac{h}{f},\end{eqnarray}$its bilinear form turns out to be [36]$\begin{eqnarray}{D}_{x}^{2}f\cdot f=2{gh},\end{eqnarray}$$\begin{eqnarray}{D}_{x}{D}_{{t}_{1}}g\cdot f={gf},\end{eqnarray}$$\begin{eqnarray}{D}_{x}{D}_{{t}_{1}}h\cdot f={hf},\end{eqnarray}$$\begin{eqnarray}({D}_{{t}_{n-1}}+{D}_{x}{D}_{{t}_{n}})g\cdot f=0,\end{eqnarray}$$\begin{eqnarray}({D}_{{t}_{n-1}}-{D}_{x}{D}_{{t}_{n}})h\cdot f=0,\,\,(n=2,3\cdots ),\end{eqnarray}$which allow double Wronskian solution [32]$\begin{eqnarray}\begin{array}{rcl}f & = & | {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(m)}| ,\,\,g=2| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(m-1)}| ,\\ h & = & 2| {\widehat{\varphi }}^{(n-1)};{\widehat{\psi }}^{(m+1)}| .\end{array}\end{eqnarray}$where φ and ψ are defined as$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \exp \left(-{Ax}-\sum _{n=1}^{\infty }{2}^{-(n+1)}{A}^{-n}{t}_{n}\right){C}^{+},\\ \psi & = & \exp \left({Ax}+\sum _{n=1}^{\infty }{2}^{-(n+1)}{A}^{-n}{t}_{n}\right){C}^{-},\end{array}\end{eqnarray}$$A\in {{\mathbb{C}}}_{(n+m+2)\times (n+m+2)}$ and ${C}^{\pm }$ are $(n+m+2)$th-order constant column vector.

5.2. Reductions and solutions

The odd members in the hierarchy (91) admit a real reduction$\begin{eqnarray}r(x,t)=\delta q(\sigma x,\sigma t),\,\,\sigma ,\delta =\pm 1,\end{eqnarray}$under which we have$\begin{eqnarray}{q}_{{t}_{2l+1}}={K}_{\mathrm{1,2}l+1}{| }_{(96)},\,\,l=0,1,2,\cdot \cdot \cdot \end{eqnarray}$and the representative one$\begin{eqnarray}q={q}_{x,{t}_{1}}-2\delta q{\partial }_{x}^{-1}{\left({qq}(\sigma x,\sigma t\right)}_{{t}_{1}}.\end{eqnarray}$In addition, the odd members in the hierarchy (91) also admit a complex reduction$\begin{eqnarray}r(x,t)=\delta {q}^{* }(\sigma x,\sigma t),\,\,\sigma ,\delta =\pm 1,\end{eqnarray}$under which we have$\begin{eqnarray}{q}_{{t}_{2l+1}}={K}_{\mathrm{1,2}l+1}{| }_{(99)},\,\,l=0,1,2,\cdot \cdot \cdot \end{eqnarray}$and$\begin{eqnarray}q={q}_{x,{t}_{1}}-2\delta q{\partial }_{x}^{-1}{\left({{qq}}^{* }(\sigma x,\sigma t\right)}_{{t}_{1}}\end{eqnarray}$as the first member.

For those even members in the hierarchy (91), first, they allow a real reduction$\begin{eqnarray}r(x,t)=\delta q(\sigma x,-t),\,\,\sigma ,\delta =\pm 1,\end{eqnarray}$which generate$\begin{eqnarray}{q}_{{t}_{2l}}={K}_{\mathrm{1,2}l}{| }_{(102)},\,\,l=1,2,\cdot \cdot \cdot ,\end{eqnarray}$the simplest one is$\begin{eqnarray}\begin{array}{rcl}q & = & -{q}_{{{xxt}}_{2}}+4\delta \widehat{q}{{qq}}_{{t}_{2}}+2\delta {q}_{x}{\partial }_{x}^{-1}{\left(q\widehat{q}\right)}_{{t}_{2}}\\ & & +2\delta q{\partial }_{x}^{-1}({q}_{x}{\widehat{q}}_{{t}_{2}}-{\widehat{q}}_{x}{q}_{{t}_{2}}),\,\,\delta ,\sigma =\pm 1,\end{array}\end{eqnarray}$where $\widehat{q}=q(\sigma x,-t)$. And second, a complex reduction$\begin{eqnarray}r(x,t)=\delta {q}^{* }(\sigma x,t),\,\,\sigma ,\delta =\pm 1,\,\,{t}_{2j}\to {\rm{i}}{t}_{2j},\end{eqnarray}$leads to$\begin{eqnarray}\ {\rm{i}}{q}_{{t}_{2l}}=-{K}_{\mathrm{1,2}l}{| }_{(105)},\,\,l=1,2,\cdot \cdot \cdot ,\end{eqnarray}$in which the representative one is$\begin{eqnarray}\begin{array}{rcl}\ {\rm{i}}q & = & -{q}_{{{xxt}}_{2}}+4\delta {\widehat{q}}^{* }{{qq}}_{{t}_{2}}+2\delta {q}_{x}{\partial }_{x}^{-1}{\left(q{\widehat{q}}^{* }\right)}_{{t}_{2}}\\ & & +2\delta q{\partial }_{x}^{-1}({q}_{x}{\widehat{q}}_{{t}_{2}}^{* }-{\widehat{q}}_{x}^{* }{q}_{{t}_{2}}),\,\,\delta ,\sigma =\pm 1,\end{array}\end{eqnarray}$where $\widehat{q}=q(\sigma x,t)$.

In the following we list out double Wronskian solutions and we skip proof.

The system (97) admits the following solutions$\begin{eqnarray}q(x,t)=-2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$, are defined by$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \exp \left(-{Ax}-\sum _{j=0}^{\infty }{2}^{-(2j+2)}{A}^{-(2j+1)}{t}_{2j+1}\right){C}^{+},\\ \psi & = & \exp \left({Ax}+\sum _{j=0}^{\infty }{2}^{-(2j+2)}{A}^{-(2j+1)}{t}_{2j+1}\right){C}^{-}\end{array}\end{eqnarray}$and satisfy$\begin{eqnarray}\psi (x,t)=T\varphi (\sigma x,\sigma t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+},\end{eqnarray}$in which $T$ is a constant matrix determined by (34).

For equation (98) with different (σ, δ), its 1SS are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=-\displaystyle \frac{4{cdk}}{{c}^{2}{{\rm{e}}}^{-2{kx}-\tfrac{t}{2k}}+{d}^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{4{cdk}}{{d}^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}-{c}^{2}{{\rm{e}}}^{-2{kx}-\tfrac{t}{2k}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{4k}{{{\rm{e}}}^{-2{kx}-\tfrac{t}{2k}}+{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{4k{\rm{i}}}{{{\rm{e}}}^{-2{kx}-\tfrac{t}{2k}}+{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{1},\,{C}^{+}=(c,d)$.

The system (100) admits solutions$\begin{eqnarray}q(x,t)=-2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$, are defined by (109) and satisfy$\begin{eqnarray}\psi (x,t)=T{\varphi }^{* }(\sigma x,\sigma t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+* },\end{eqnarray}$in which $T$ is a constant matrix determined by (54).

1SS of the equation (101) with different (σ, δ) are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{-2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-\tfrac{t}{2{k}^{* }}}+| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}-| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x-\tfrac{t}{2{k}^{* }}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{2{cd}({k}^{* }-k)}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x+\tfrac{t}{2{k}^{* }}}-| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{2{cd}({k}^{* }-k)}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x+\tfrac{t}{2{k}^{* }}}+| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{2k}}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{1},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

The hierarchy (103) admit$\begin{eqnarray}q(x,t)=-2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where $\varphi ={({\varphi }_{1},{\varphi }_{2},\cdots ,{\varphi }_{2n+2})}^{{\rm{T}}}$, $\psi ={({\psi }_{1},{\psi }_{2},\cdots ,{\psi }_{2n+2})}^{{\rm{T}}}$ are defined by$\begin{eqnarray}\begin{array}{rcl}\varphi & = & \exp \left(-{Ax}-\sum _{j=1}^{\infty }{2}^{-(2j+1)}{A}^{-2j}{t}_{2j}\right){C}^{+},\\ \psi & = & \exp \left({Ax}+\sum _{j=1}^{\infty }{2}^{-(2j+1)}{A}^{-2j}{t}_{2j}\right){C}^{-}\end{array}\end{eqnarray}$and satisfy$\begin{eqnarray}\psi (x,t)=T\varphi (\sigma x,-t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+},\end{eqnarray}$in which $T$ is a constant matrix determined by (34).

For equation (104), its 1SS with different (σ, δ) are$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{-4{cdk}}{{c}^{2}{{\rm{e}}}^{-2{kx}+\tfrac{t}{4{k}^{2}}}+{d}^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{4{k}^{2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{4{cdk}}{{d}^{2}{{\rm{e}}}^{2{kx}+\tfrac{t}{4{k}^{2}}}-{c}^{2}{{\rm{e}}}^{-2{kx}+\tfrac{t}{4{k}^{2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{4k}{{{\rm{e}}}^{-2{kx}+\tfrac{t}{4{k}^{2}}}+{{\rm{e}}}^{2{kx}+\tfrac{t}{4{k}^{2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{-4{\rm{i}}k}{{{\rm{e}}}^{-2{kx}+\tfrac{t}{4{k}^{2}}}+{{\rm{e}}}^{2{kx}+\tfrac{t}{4{k}^{2}}}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{2},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

As an example, let us illustrate dynamics of ${q}_{(\sigma ,\delta )}={q}_{(1,-1)}$ which is governed by the equation$\begin{eqnarray}\begin{array}{rcl}q & = & -{q}_{{{xxt}}_{2}}-4\widehat{q}{{qq}}_{{t}_{2}}-2{q}_{x}{\partial }_{x}^{-1}{\left(q\widehat{q}\right)}_{{t}_{2}}\\ & & -2q{\partial }_{x}^{-1}({q}_{x}{\widehat{q}}_{{t}_{2}}-{\widehat{q}}_{x}{q}_{{t}_{2}}),\end{array}\end{eqnarray}$where $\widehat{q}=q(x,-t)$, (118a) describes a single soliton$\begin{eqnarray}q(x,t)=\displaystyle \frac{2{cdk}}{| {cd}| }{{\rm{e}}}^{-\tfrac{t}{4{k}^{2}}}{\rm{{\rm{sech}} }}\left(-2{kx}+\mathrm{ln}\left|\displaystyle \frac{c}{d}\right|\right),\end{eqnarray}$with an initial phase $\mathrm{ln}\left|\tfrac{c}{d}\right|$, top trajectory $x(t)=\tfrac{\mathrm{ln}\left|\tfrac{c}{d}\right|}{2k}$, and an amplitude that decreases with time, see figure 3(a).

Figure 3.

New window|Download| PPT slide
Figure 3.(a). Shape and motion of 1SS (118a) for equation (119), in which $k=0.9,c=5,d=1$. (b). Shape and motion of 2SS (121) for equation (119), in which ${k}_{1}=3,{k}_{2}=1,{c}_{1}=0.5$, ${c}_{2}=-1,{d}_{1}=1,{d}_{2}=1$.


The 2SS of equation (119) is$\begin{eqnarray}q(x,t)=\displaystyle \frac{{G}^{{\prime} }}{{F}^{{\prime} }},\end{eqnarray}$where$\begin{eqnarray*}\begin{array}{rcl}{G}^{{\prime} } & = & -4\left({k}_{1}^{2}-{k}_{2}^{2}\right)\left(-{c}_{1}{d}_{1}{k}_{1}{{\rm{e}}}^{2{k}_{1}x}\left({c}_{2}^{2}+{d}_{2}^{2}{{\rm{e}}}^{\tfrac{t}{4{k}_{2}^{2}}+4{k}_{2}x}\right)\right.\\ & & \left.+{c}_{2}{d}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{\tfrac{t}{4{k}_{1}^{2}}+4{k}_{1}x+2{k}_{2}x}\right)\\ & & -4\left({k}_{1}^{2}-{k}_{2}^{2}\right){c}_{2}{c}_{1}^{2}{d}_{2}{k}_{2}{{\rm{e}}}^{2{k}_{2}x},\end{array}\end{eqnarray*}$$\begin{eqnarray*}\begin{array}{rcl}{F}^{{\prime} } & = & {c}_{1}^{2}\left({c}_{2}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}+{d}_{2}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}{{\rm{e}}}^{\tfrac{t}{4{k}_{2}^{2}}+4{k}_{2}x}\right)\\ & & -4{c}_{2}{c}_{1}{d}_{1}{d}_{2}{k}_{1}{k}_{2}\left({{\rm{e}}}^{\tfrac{t}{4{k}_{1}^{2}}}+{{\rm{e}}}^{\tfrac{t}{4{k}_{2}^{2}}}\right){{\rm{e}}}^{2\left({k}_{1}+{k}_{2}\right)x}\\ & & +{d}_{1}^{2}{{\rm{e}}}^{\tfrac{t}{4{k}_{1}^{2}}+4{k}_{1}x}\left({c}_{2}^{2}\left({k}_{1}+{k}_{2}\right){}^{2}+{d}_{2}^{2}\left({k}_{1}-{k}_{2}\right){}^{2}{{\rm{e}}}^{\tfrac{t}{4{k}_{2}^{2}}+4{k}_{2}x}\right),\end{array}\end{eqnarray*}$which we depict in figure 3(b).

The hierarchy (106) admit solutions$\begin{eqnarray}q(x,t)=-2\displaystyle \frac{| {\widehat{\varphi }}^{(n+1)};{\widehat{\psi }}^{(n-1)}| }{| {\widehat{\varphi }}^{(n)};{\widehat{\psi }}^{(n)}| },\end{eqnarray}$where φ and ψ are $(2n+2)$th order column vectors (i.e. $m=n$ in (9)), are defined by (116) and satisfy$\begin{eqnarray}\psi (x,t)=T{\varphi }^{* }(\sigma x,t),\end{eqnarray}$$\begin{eqnarray}{C}^{-}={{TC}}^{+* },\end{eqnarray}$in which $T$ is a constant matrix determined by (54).

The equation (107) with different (σ, δ) has the following 1SS:$\begin{eqnarray}{q}_{\sigma =1,\delta =-1}=\displaystyle \frac{-2{cd}(k+{k}^{* })}{| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x+\tfrac{{\rm{i}}t}{4{k}^{* 2}}}+| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{{\rm{i}}t}{4{k}^{2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =1,\delta =1}=\displaystyle \frac{2{cd}(k+{k}^{* })}{| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{{\rm{i}}t}{4{k}^{2}}}-| c{| }^{2}{{\rm{e}}}^{-2{k}^{* }x+\tfrac{{\rm{i}}t}{4{k}^{* 2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =-1}=\displaystyle \frac{2{cd}({k}^{* }-k)}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x+\tfrac{{\rm{i}}t}{4{k}^{* 2}}}-| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{{\rm{i}}t}{4{k}^{2}}}},\end{eqnarray}$$\begin{eqnarray}{q}_{\sigma =-1,\delta =1}=\displaystyle \frac{2{cd}({k}^{* }-k)}{| c{| }^{2}{{\rm{e}}}^{2{k}^{* }x+\tfrac{{\rm{i}}t}{4{k}^{* 2}}}+| d{| }^{2}{{\rm{e}}}^{2{kx}+\tfrac{{\rm{i}}t}{4{k}^{2}}}},\end{eqnarray}$where $k={k}_{1},\,t={t}_{2},\,{C}^{+}={(c,d)}^{{\rm{T}}}$.

6. Conclusions

We have discussed possible nonlocal reductions of the (2+1)-D BAKNS hierarchy and the negative order AKNS hierarchy. By means of the reduction technique (see [1518]), we derived N-soliton solutions and multiple-pole solutions in double Wronskian form of the reduced hierarchies from those of the unreduced hierarchies. Note that these hierarchies contain integration w.r.t. x, the integration operator with form of (30) is necessary and helpful in considering nonlocal reductions. In addition, this reduction method is based on bilinear forms and double Wronskians, As we can see, it allows us to obtain N-soliton solutions and multiple-pole solutions for the whole reduced hierarchy. Compared with [14, 16, 19] where only some single equations in the (2+1)-D BAKNS hierarchy and the negative order AKNS hierarchy were considered, here we solved the two hierarchies and obtained more solutions. Finally, it is remarkable that for some real nonlocal equations, amplitudes of solutions are related to the independent variables that are reversed in the real nonlocal reductions, e.g. (44), (67) and (120).

Acknowledgments

This work was supported by the NSF of China [grant numbers 11 875 040, 11 631 007, 11 571 225].


Reference By original order
By published year
By cited within times
By Impact factor

Ablowitz M J Musslimani Z H 2013 Integrable nonlocal nonlinear Schrödinger equation
Phys. Rev. Lett. 110 064105

DOI:10.1103/PhysRevLett.110.064105 [Cited within: 2]

Bender C M Boettcher S 1998 Real spectra in non-Hermitian Hamiltonians having PT symmetry
Phys. Rev. Lett. 80 5243 5246

DOI:10.1103/PhysRevLett.80.5243 [Cited within: 1]

Konotop V V Yang J K Zezyulin D A 2016 Nonlinear waves in PT-symmetric systems
Rev. Mod. Phys. 88 035002

DOI:10.1103/RevModPhys.88.035002 [Cited within: 1]

Fokas A S 2016 Integrable multidimensional versions of the nonlocal nonlinear Schrödinger equation
Nonlinearity 29 319 324

DOI:10.1088/0951-7715/29/2/319 [Cited within: 1]

Ablowitz M J Musslimani Z H 2016 Inverse scattering transform for the integrable nonlocal nonlinear Schrödinger equation
Nonlinearity 29 915

DOI:10.1088/0951-7715/29/3/915 [Cited within: 1]

Ablowitz M J Musslimani Z H 2017 Integrable nonlocal nonlinear equations
Stud. Appl. Math. 139 7 59

DOI:10.1111/sapm.12153

Yang J K 2018 Physically significant nonlocal nonlinear Schrödinger equation and its soliton solutions
Phys. Rev. 98 042202

DOI:10.1103/PhysRevE.98.042202

Lou S Y 2018 $\widehat{P}-\widehat{T}-\widehat{C}$ symmetry invariant and symmetry breaking soliton solutions
J. Math. Phys. 59 083507

DOI:10.1063/1.5051989

Ablowitz M J Musslimani Z H 2019 Integrable nonlocal asymptotic reductions of physically significant nonlinear equations
J. Phys. A: Math. Theor. 52 15LT02

DOI:10.1088/1751-8121/ab0e95 [Cited within: 1]

Zhou Z X 2018 Darboux transformations and global explicit solutions for nonlocal Davey–Stewartson: I. Equation
Stud. Appl. Math. 141 186 204

DOI:10.1111/sapm.12219 [Cited within: 1]

Yang B Chen Y 2018 Reductions of Darboux transformations for the PT-symmetricnonlocal Davey–Stewartson equations
Appl. Math. Lett. 82 43 49

DOI:10.1016/j.aml.2017.12.025

Yang B Chen Y 2018 Several reverse-time integrable nonlocal nonlinear equations: Rogue-wave solutions
Chaos 28 053104

DOI:10.1063/1.5019754

Yang B Chen Y 2018 Dynamics of high-order solitons in the nonlocal nonlinear Schrödinger equations
Nonlinear Dyn 94 489 502

DOI:10.1007/s11071-018-4373-0

Zhu X M Zuo D F 2019 Some (2+1)-dimensional nonlocal ‘breaking soliton’-type systems
Appl. Math. Lett. 91 181 187

DOI:10.1016/j.aml.2018.12.011 [Cited within: 2]

Chen K Zhang D J 2018 Solutions of the nonlocal nonlinear Schrödinger hierarchy via reduction
Appl. Math. Lett. 75 82 88

DOI:10.1016/j.aml.2017.05.017 [Cited within: 3]

Chen K Deng X Lou S Y Zhang D J 2018 Solutions of nonlocal equations reduced from the AKNS hierarchy
Stud. Appl. Math. 141 113 141

DOI:10.1111/sapm.12215 [Cited within: 2]

Deng X Lou S Y Zhang D J 2018 Bilinearisation-reduction approach to the nonlocal discrete nonlinear Schrödinger equations
Appl. Math. Comput. 332 477 483

DOI:10.1016/j.amc.2018.03.061

Chen K Liu S M Zhang D J 2019 Covariant hodograph transformations between nonlocal short pulse models and AKNS(−1) system
Appl. Math. Lett. 88 230 236

DOI:10.1016/j.aml.2018.09.005 [Cited within: 3]

Gürses M Pekcan A 2019 (2 + 1)-dimensional AKNS(−N) Systems: N=3, 4
(arXiv:1910.11298 [nlin.SI])

[Cited within: 2]

Bogoyavlenskiĭ O I 1990 Overturning solitons in new two-dimensional integrable equations
Math. USSR-Izv. 34 245 259

DOI:10.1070/IM1990v034n02ABEH000628 [Cited within: 2]

Bogoyavlenskiĭ O I 1990 Breaking solitons in 2+1-dimensional integrable equations
Russ. Math. Surv. 45 1 86

DOI:10.1070/RM1990v045n04ABEH002377 [Cited within: 2]

Bogoyavlenskiĭ O I 1990 Breaking solitons. II
Math. USSR-Izv. 35 245 248

DOI:10.1070/IM1990v035n01ABEH000700

Bogoyavlenskiĭ O I 1991 Breaking solitons. III
Math. USSR-Izv. 36 129 137

DOI:10.1070/IM1991v036n01ABEH001953

Bogoyavlenskiĭ O I 1991 Breaking solitons. IV
Math. USSR-Izv. 37 475 487

DOI:10.1070/IM1991v037n03ABEH002154 [Cited within: 3]

Bogoyavlenskiĭ O I 1992 Breaking solitons: V. Systems of hydrodynamic type
Math. USSR-Izv. 38 451 465

DOI:10.1070/IM1992v038n03ABEH002209

Bogoyavlenskiĭ O I 1992 Breaking solitons: VI. Extension of systems of hydrodynamic type
Math. USSR-Izv. 39 959 973

DOI:10.1070/IM1992v039n02ABEH002233 [Cited within: 1]

Li Y S Zhang Y J 1993 Symmetries of a (2 + 1)-dimensional breaking soliton equation
J. Phys. A: Math. Gen. 26 7487 7494

DOI:10.1088/0305-4470/26/24/021 [Cited within: 2]

Gao Y T Tian B 1995 New family of overturning soliton solutions for a typical breaking soliton equation
Comput. Math. Appl. 30 97 100

DOI:10.1016/0898-1221(95)00176-Y

Li B Chen Y Xuan H N Zhang H Q 2003 Symbolic computation and construction of soliton-like solutions for a breaking soliton equation
Chaos Solitons Fractals 17 885 893

DOI:10.1016/S0960-0779(02)00570-2

Yao Y Q Chen D Y Zhang D J 2008 Multisoliton solutions to a nonisospectral (2 + 1)-dimensional breaking soliton equation
Phys. Lett. A 372 2017 2025

DOI:10.1016/j.physleta.2007.10.096

Hao H H Zhang D J Zhang J B Yao Y Q 2010 Rational and periodic solutions for a (2 + 1)-dimensional breaking soliton equation associated with ZS-AKNS hierarchy
Commun. Theor. Phys. 53 430 434

DOI:10.1088/0253-6102/53/3/05 [Cited within: 1]

Zhao S L Zhang D J Ji J 2010 Exact solutions for two equation hierarchies
Chin. Phys. Lett. 27 020201

[Cited within: 4]

Zhang D J Ji J Zhao S L 2009 Soliton scattering with amplitude changes of a negative order AKNS equation
Phys. D 238 2361 2367

DOI:10.1016/j.physd.2009.09.018 [Cited within: 2]

Hirota R 1974 A new form of Bäcklund transformations and its relation to the inverse scattering problem
Prog. Theor. Phys. 52 1498 1512

DOI:10.1143/PTP.52.1498 [Cited within: 1]

Silem A Zhang C Zhang D J 2019 Dynamics of three nonisospectral nonlinear Schrödinger equations
Chin. Phys. B 28 020202

DOI:10.1088/1674-1056/28/2/020202 [Cited within: 1]

Ji J Zhang B J Zhang D J 2009 Soliton solutions for a negative order AKNS equation hierarchy
Commun. Theor. Phys. 52 395 397

DOI:10.1088/0253-6102/52/3/03 [Cited within: 2]

相关话题/Solutions nonlocal breaking