Next Article in Journal
Enriched Z-Contractions and Fixed-Point Results with Applications to IFS
Previous Article in Journal
Differential Geometry and Its Application, 2nd Edition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Integrated Integrable Hierarchy Arising from a Broadened Ablowitz–Kaup–Newell–Segur Scenario

1
Department of Mathematics, Zhejiang Normal University, Jinhua 321004, China
2
Department of Mathematics, King Abdulaziz University, Jeddah 21589, Saudi Arabia
3
Department of Mathematics and Statistics, University of South Florida, Tampa, FL 33620-5700, USA
4
Material Science Innovation and Modelling, Department of Mathematical Sciences, North-West University, Mafikeng Campus, Mmabatho 2735, South Africa
Axioms 2024, 13(8), 563; https://doi.org/10.3390/axioms13080563
Submission received: 9 July 2024 / Revised: 31 July 2024 / Accepted: 16 August 2024 / Published: 19 August 2024
(This article belongs to the Special Issue New Perspectives in Lie Algebras)

Abstract

:
This study introduces a 4 × 4 matrix eigenvalue problem and develops an integrable hierarchy with a bi-Hamiltonian structure. Integrability is ensured by the zero-curvature condition, while the Hamiltonian structure is supported by the trace identity. Explicit derivations yield second-order and third-order integrable equations, illustrating the integrable hierarchy.

1. Introduction

In soliton theory, Lax pairs play a crucial role in exploring integrable models. The concept of a Lax pair [1] involves formulating a linear eigenvalue problem associated with a given nonlinear partial differential equation. By constructing an appropriate Lax pair, we can generate a compatible set of model equations that possess remarkable integrable properties, such as infinitely countless symmetries and conserved quantities. These equations exhibit soliton solutions, making them amenable to analytical techniques and providing deep insights into their dynamics [2,3].
To construct integrable models using Lax pairs, we typically start with a column potential vector u, and denote an eigenvalue parameter by k. The formulation entails defining a set of linear differential equations, referred to as the Lax pair, that are related through the compatibility condition, ensuring the integrability of the associated nonlinear equations. The Lax pair consists of two eigenvalue equations:
ϕ x = P ( u , k ) ϕ , ϕ t = Q ( u , k ) ϕ ,
where ϕ is the eigenfunction, P ( u , k ) is the spatial spectral matrix, and Q ( u , k ) is the temporal spectral matrix. These matrices depend on both the potential vector u and the eigenvalue parameter k. The zero-curvature condition, or the compatibility condition, is given by
P t Q x + [ P , Q ] = 0 ,
where [ P , Q ] = P Q Q P is the commutator of P and Q . This condition ensures that the eigenfunction ϕ evolves consistently in both the spatial and temporal directions, leading to an integrable system.
To illustrate, consider the AKNS (Ablowitz–Kaup–Newell–Segur) system, which is a well-known framework for generating integrable equations. The AKNS system [4] defines P and Q as follows:
P ( u , k ) = k p q k , Q ( u , k ) = A B C A ,
where p and q are components of the potential vector u, and A , B and C are functions of u and k. The specific forms of A , B and C depend on the particular integrable model under consideration. The zero-curvature condition, Equation (2), then leads to a set of nonlinear partial differential equations for p and q. For instance, in the case of the nonlinear Schrödinger (NLS) equation, the condition results in
p t = p x x + 2 p 2 q , q t = q x x 2 p q 2 .
Solving these equations reveals the integrable structure of the system, characterized by soliton solutions, infinite symmetries, and conserved quantities.
By appropriately choosing P and Q , one can derive various integrable models such as the sine-Gordon equation, the Korteweg–de Vries (KdV) equation, and others, all exhibiting remarkable integrable properties and amenable to powerful analytical techniques like the inverse scattering transform.
Hamiltonian structures are fundamental in the study of integrable systems, as they provide a framework for exploring the integrability of the resultant models. One method to generate Hamiltonian structures is by utilizing the trace identity or the variational identity. The trace identity, in particular, is a robust technique in this context.
The trace identity reads as follows (see [5] for details):
δ δ u tr R P k d x = k τ k k τ tr R P u ,
where δ δ u denotes the variational derivative with respect to u and tr stands for the trace of a matrix; τ remains invariant with respect to the eigenvalue parameter k. Here, R solves
R x = [ P , R ] ,
where P is the spectral matrix. The trace identity connects the variational derivative of an integral involving the eigenvalue parameter k to the trace of a matrix expression, linking the eigenvalue problem with the system’s Hamiltonian structure.
A plethora of Liouville integrable hierarchies of soliton Hamiltonian equations can be derived using the aforementioned Lax pair formulation, utilizing loop algebras derived from both special linear algebras (see, for instance, [4,5,6,7,8,9,10,11,12,13,14]), special orthogonal algebras (see, for example, [15,16]), and non-semisimple Lie algebras (see, e.g., [17,18,19,20,21,22,23,24]). These hierarchies are pivotal in the study of integrable models, providing a structured framework to explore the solutions and properties of soliton equations.
This paper proposes a novel spectral matrix and constructs a Liouville integrable hierarchy comprising four-component bi-Hamiltonian equations using the Lax pair formulation. The resulting soliton equations exhibit established bi-Hamiltonian structures, demonstrated through the application of the trace identity. Several demonstrative examples are provided, including four-component coupled integrable NLS equations and modified Korteweg–de Vries (mKdV) equations. The conclusion in the final section summarizes the findings and offers summary remarks.

2. Commuting Integrable Hamiltonian Models

Motivated by a study on non-perturbation-type integrable couplings within the AKNS hierarchy via the Lax pair formulation [25], we consider a newly proposed matrix eigenvalue problem of the following form:
ϕ x = P ϕ = P ( u , k ) ϕ , P = ξ 1 k u 1 η 1 k u 3 u 2 ξ 2 k u 4 η 2 k η 1 k u 3 ξ 1 k u 1 u 4 η 2 k u 2 ξ 2 k ,
where k is again the eigenvalue parameter and u stands for the dependent variable consisting of four components:
u = u ( x , t ) = ( u 1 , u 2 , u 3 , u 4 ) T .
If the bottom-left 2 × 2 block is taken to be zero, then this spectral problem with ξ 1 = ξ 2 = 1 and η 1 = η 2 = 1 becomes the one discussed in the above reference. To guarantee that an integrable hierarchy can be generated via the Lax pair formation from this new spectral problem, we need to impose a necessary and sufficient condition:
ξ 2 η 2 0 , ξ = ξ 1 ξ 2 , η = η 1 η 2 .
When η 1 = η 2 = 0 and u 3 = u 4 = 0 , the spectral problem reduces to two identical copies of the standard AKNS eigenvalue problem [4], and thus, it provides a broadened version of the AKNS eigenvalue problem.
To establish a corresponding four-component Liouville integrable hierarchy, we initially solve the associated stationary zero-curvature Equation (4) by seeking a specific Laurent series solution:
R = a b e f c a g e e f a b g e c a = n 0 k n R { n } ,
with six fundamental components assumed to be expanded in Laurent series of the eigenvalue parameter k:
a = n 0 k n a { n } , b = n 0 k n b { n } , c = n 0 k n c { n } , e = n 0 k n e { n } , f = n 0 k n f { n } , g = n 0 k n g { n } .
It is evident that the corresponding associated stationary zero-curvature Equation (4) leads to the following relations:
a x = u 1 c u 2 b + u 3 g u 4 f , b x = ξ k b + η k f 2 u 1 a 2 u 3 e , c x = ξ k c η k g + 2 u 2 a + 2 u 4 e , e x = u 1 g u 2 f + u 3 c u 4 b , f x = η k b + ξ k f 2 u 1 e 2 u 3 a , g x = η k c ξ k g + 2 u 2 e + 2 u 4 a .
This gives
k ξ η η ξ b f = b x + 2 u 1 a + 2 u 3 e f x + 2 u 1 e + 2 u 3 a
and
k ξ η η ξ c g = c x + 2 u 2 a + 2 u 4 e g x + 2 u 2 e + 2 u 4 a .
Therefore, the condition, Equation (7), which guarantees the invertibility of the coefficient matrix in the above two systems, is both necessary and sufficient to ensure that we can recursively determine a Laurent series solution R . Furthermore, we observe that the system, Equation (10), yields the initial requirements
a x { 0 } = e x { 0 } = 0 , b { 0 } = c { 0 } = f { 0 } = g { 0 } = 0 ,
and the recursion relations used to define the Laurent series solution:
b { n + 1 } = ξ ξ 2 η 2 ( b x { n } + 2 u 1 a { n } + 2 u 3 e { n } ) η ξ 2 η 2 ( f x { n } + 2 u 1 e { n } + 2 u 3 a { n } ) , f { n + 1 } = η ξ 2 η 2 ( b x { n } + 2 u 1 a { n } + 2 u 3 e { n } ) + ξ ξ 2 η 2 ( f x { n } + 2 u 1 e { n } + 2 u 3 a { n } ) ,
c { n + 1 } = ξ ξ 2 η 2 ( c x { n } + 2 u 2 a { n } + 2 u 4 e { n } ) η ξ 2 η 2 ( g x { n } + 2 u 2 e { n } + 2 u 4 a { n } ) , g { n + 1 } = η ξ 2 η 2 ( c x { n } + 2 u 2 a { n } + 2 u 4 e { n } ) + ξ ξ 2 η 2 ( g x { n } + 2 u 2 e { n } + 2 u 4 a { n } ) ,
a x { n + 1 } = u 1 c { n + 1 } u 2 b { n + 1 } + u 3 g { n + 1 } u 4 f { n + 1 } , e x { n + 1 } = u 1 g { n + 1 } u 2 f { n + 1 } + u 3 c { n + 1 } u 4 b { n + 1 } ,
where n 0 . As usual, to determine a specific Laurent series solution, we introduce arbitrary constant initial data
a { 0 } = 1 2 μ , e { 0 } = 1 2 ν ,
and assume the integration constants to be zero:
a { n } | u = 0 = 0 , e { n } | u = 0 = 0 , n 1 .
Through these conditions, one can derive all sequences of { a { n } , b { n } , c { n } , e { n } , f { n } , g { n } } for n 1 . The first sequence reads
b { 1 } = 1 ξ 2 η 2 [ ξ ( μ u 1 + ν u 3 ) η ( ν u 1 + μ u 3 ) ] , f { 1 } = 1 ξ 2 η 2 [ ξ ( ν u 1 + μ u 3 ) η ( μ u 1 + ν u 3 ) ] , c { 1 } = 1 ξ 2 η 2 [ ξ ( μ u 2 + ν u 4 ) η ( ν u 2 + μ u 4 ) ] , g { 1 } = 1 ξ 2 η 2 [ ξ ( ν u 2 + μ u 4 ) η ( μ u 2 + ν u 4 ) ] , a { 1 } = e { 1 } = 0 .
The second sequence reads
b { 2 } = 1 ( ξ 2 η 2 ) 2 ( p 2 , 1 u 1 , x p 2 , 2 u 3 , x ) , f { 2 } = 1 ( ξ 2 η 2 ) 2 ( p 2 , 2 u 1 , x + p 2 , 1 u 3 , x ) , c { 2 } = 1 ( ξ 2 η 2 ) 2 ( p 2 , 1 u 2 , x p 2 , 2 u 4 , x ) , g { 2 } = 1 ( ξ 2 η 2 ) 2 ( p 2 , 2 u 2 , x p 2 , 1 u 4 , x ) ,
a { 2 } = 1 ( ξ 2 η 2 ) 2 [ ( p 2 , 1 u 2 + p 2 , 2 u 4 ) u 1 + ( p 2 , 2 u 2 + p 2 , 1 u 4 ) u 3 ] , e { 2 } = 1 ( ξ 2 η 2 ) 2 [ ( p 2 , 2 u 2 + p 2 , 1 u 4 ) u 1 + ( p 2 , 1 u 2 + p 2 , 2 u 4 ) u 3 ] ,
where p 2 , 1 and p 2 , 2 are two special polynomials of second order in terms of ξ and η :
p 2 , 1 = ξ 2 μ 2 ξ η ν + η 2 μ , p 2 , 2 = ξ 2 ν 2 ξ η μ + η 2 ν .
The third sequence reads
b { 3 } = 1 ( ξ 2 η 2 ) 3 [ p 3 , 1 u 1 , x x + p 3 , 2 u 3 , x x 2 ( p 3 , 1 u 2 + p 3 , 2 u 4 ) u 1 2 4 ( p 3 , 2 u 2 + p 3 , 1 u 4 ) u 1 u 3 2 ( p 3 , 1 u 2 + p 3 , 2 u 4 ) u 3 2 ] , f { 3 } = 1 ( ξ 2 η 2 ) 3 [ p 3 , 2 u 1 , x x + p 3 , 1 u 3 , x x 2 ( p 3 , 2 u 2 + p 3 , 1 u 4 ) u 1 2 4 ( p 3 , 1 u 2 + p 3 , 2 u 4 ) u 1 u 3 2 ( p 3 , 2 u 2 + p 3 , 1 u 4 ) u 3 2 ] ,
c { 3 } = 1 ( ξ 2 η 2 ) 3 [ p 3 , 1 u 2 , x x + p 3 , 2 u 4 , x x 2 ( p 3 , 1 u 1 + p 3 , 2 u 3 ) u 2 2 4 ( p 3 , 2 u 1 + p 3 , 1 u 3 ) u 2 u 4 2 ( p 3 , 1 u 1 + p 3 , 2 u 3 ) u 4 2 ] , g { 3 } = 1 ( ξ 2 η 2 ) 3 [ p 3 , 2 u 2 , x x + p 3 , 1 u 4 , x x 2 ( p 3 , 2 u 1 + p 3 , 1 u 3 ) u 2 2 4 ( p 3 , 1 u 1 + p 3 , 2 u 3 ) u 2 u 4 2 ( p 3 , 2 u 1 + p 3 , 1 u 3 ) u 4 2 ] ,
a { 3 } = 1 ( ξ 2 η 2 ) 3 [ ( p 3 , 1 u 2 + p 3 , 2 u 4 ) u 1 , x + ( p 3 , 1 u 1 + p 3 , 2 u 3 ) u 2 , x ( p 3 , 2 u 2 + p 3 , 1 u 4 ) u 3 , x + ( p 3 , 2 u 1 + p 3 , 1 u 3 ) u 4 , x ] , e { 3 } = 1 ( ξ 2 η 2 ) 3 [ ( p 3 , 2 u 2 p 3 , 1 u 4 ) u 1 , x + ( p 3 , 2 u 1 + p 3 , 1 u 3 ) u 2 , x ( p 3 , 1 u 2 + p 3 , 2 u 4 ) u 3 , x + ( p 3 , 1 u 1 + p 3 , 2 u 3 ) u 4 , x ] ,
where p 3 , 1 and p 3 , 2 are two special polynomials of third order in terms of ξ and η :
p 3 , 1 = ξ 3 μ 3 ξ 2 η ν + 3 ξ η 2 μ η 3 ν , p 3 , 2 = ξ 3 ν 3 ξ 2 η μ + 3 ξ η 2 ν η 3 μ .
The fourth sequence reads
b { 4 } = 1 ( ξ 2 η 2 ) 4 [ p 4 , 1 u 1 , x x x + p 4 , 2 u 3 , x x x 6 ( p 4 , 1 u 1 u 2 + p 4 , 2 u 1 u 4 + p 4 , 2 u 2 u 3 + p 4 , 1 u 3 u 4 ) u 1 , x 6 ( p 4 , 2 u 1 u 2 + p 4 , 1 u 1 u 4 + p 4 , 1 u 2 u 3 + p 4 , 2 u 3 u 4 ) u 3 , x ] , f { 4 } = 1 ( ξ 2 η 2 ) 4 [ p 4 , 2 u 1 , x x x + p 4 , 1 u 3 , x x x 6 ( p 4 , 2 u 1 u 2 + p 4 , 1 u 1 u 4 + p 4 , 1 u 2 u 3 + p 4 , 2 u 3 u 4 ) u 1 , x 6 ( p 4 , 1 u 1 u 2 + p 4 , 2 u 1 u 4 + p 4 , 2 u 2 u 3 + p 4 , 1 u 3 u 4 ) u 3 , x ] ,
c { 4 } = 1 ( ξ 2 η 2 ) 4 [ p 4 , 1 u 2 , x x x p 4 , 2 u 4 , x x x + 6 ( p 4 , 1 u 1 u 2 + p 4 , 2 u 1 u 4 + p 4 , 2 u 2 u 3 + p 4 , 1 u 3 u 4 ) u 2 , x + 6 ( p 4 , 2 u 1 u 2 + p 4 , 1 u 1 u 4 + p 4 , 1 u 2 u 3 + p 4 , 2 u 3 u 4 ) u 4 , x ] , g { 4 } = 1 ( ξ 2 η 2 ) 4 [ p 4 , 2 u 2 , x x x p 4 , 1 u 4 , x x x + 6 ( p 4 , 2 u 1 u 2 + p 4 , 1 u 1 u 4 + p 4 , 1 u 2 u 3 + p 4 , 2 u 3 u 4 ) u 2 , x + 6 ( p 4 , 1 u 1 u 2 + p 4 , 2 u 1 u 4 + p 4 , 2 u 2 u 3 + p 4 , 1 u 3 u 4 ) u 4 , x ] ,
a { 4 } = 1 ( ξ 2 η 2 ) 4 [ ( p 4 , 1 u 2 + p 4 , 2 u 4 ) u 1 , x x ( p 4 , 1 u 1 + p 4 , 2 u 3 ) u 2 , x x ( p 4 , 2 u 2 + p 4 , 1 u 4 ) u 3 , x x ( p 4 , 2 u 1 + p 4 , 1 u 3 ) u 4 , x x + p 4 , 1 u 1 , x u 2 , x + p 4 , 2 u 1 , x u 4 , x + p 4 , 2 u 2 , x u 3 , x + p 4 , 1 u 3 , x u 4 , x + 3 ( p 4 , 1 u 2 2 + 2 p 4 , 2 u 2 u 4 + p 4 , 1 u 4 2 ) u 1 2 + 6 ( p 4 , 2 u 2 2 + 2 p 4 , 1 u 2 u 4 + p 4 , 2 u 4 2 ) u 1 u 3 3 ( p 4 , 1 u 2 2 2 p 4 , 2 u 2 u 4 p 4 , 1 u 4 2 ) u 3 2 ] ,
e { 4 } = 1 ( ξ 2 η 2 ) 4 [ ( p 4 , 2 u 2 + p 4 , 1 u 4 ) u 1 , x x ( p 4 , 2 u 1 + p 4 , 1 u 3 ) u 2 , x x ( p 4 , 1 u 2 + p 4 , 2 u 4 ) u 3 , x x ( p 4 , 1 u 1 + p 4 , 2 u 3 ) u 4 , x x + p 4 , 2 u 1 , x u 2 , x + p 4 , 1 u 1 , x u 4 , x + p 4 , 1 u 2 , x u 3 , x + p 4 , 2 u 3 , x u 4 , x + 3 ( p 4 , 2 u 2 2 + 2 p 4 , 1 u 2 u 4 + p 4 , 2 u 4 2 ) u 1 2 + 6 ( p 4 , 1 u 2 2 + 2 p 4 , 2 u 2 u 4 + p 4 , 1 u 4 2 ) u 1 u 3 + 3 ( p 4 , 2 u 2 2 + 2 p 4 , 1 u 2 u 4 + p 4 , 2 u 4 2 ) u 3 2 ] ,
where p 4 , 1 and p 4 , 2 are two special polynomials of fourth order in terms of ξ and η :
p 4 , 1 = ξ 4 μ 4 ξ 3 η ν + 6 ξ 2 η 2 μ 4 ξ η 3 ν + η 4 μ , p 4 , 2 = ξ 4 ν 4 ξ 3 η μ + 6 ξ 2 η 2 ν 4 ξ η 3 μ + η 4 ν .
On the basis of these computations, we can set Δ m = 0 , m 0 , to formulate
ϕ t m = Q { m } ϕ = Q { m } ( u , k ) ϕ , Q { m } = ( k m R ) + = n = 0 m k n R { m n } , m 0 .
These are the temporal matrix eigenvalue problems within the Lax pair formulation. The conditions ensuring solvability for the spatial and temporal matrix eigenvalue problems in Equations (5) and (20) are given by the following zero-curvature equations:
P t m Q x { m } + [ P , Q { m } ] = 0 , m 0 .
These compatibility equations engender a hierarchy of integrable models with four dependent variables:
u t m = Ƶ { m } = ( ξ b { m + 1 } + η f { m + 1 } , ξ c { m + 1 } η g { m + 1 } , η b { m + 1 } + ξ f { m + 1 } , η c { m + 1 } ξ g { m + 1 } ) T ,
or more concretely,
u 1 , t m = ξ b { m + 1 } + η f { m + 1 } , u 2 , t m = ξ c { m + 1 } η g { m + 1 } , u 3 , t m = η b { m + 1 } + ξ f { m + 1 } , u 4 , t m = η c { m + 1 } ξ g { m + 1 } ,
in which m 0 .
As particular examples, this soliton hierarchy contains various coupled systems of integrable NLS equations and coupled systems of integrable mKdV equations. If taking
ξ = 1 , η = 0 , μ = 1 , ν = 0 ,
one obtains a coupled system of the following integrable NLS equations:
u 1 , t 2 = u 1 , x x 2 u 1 2 u 2 4 u 1 u 3 u 4 2 u 2 u 3 2 , u 2 , t 2 = u 2 , x x + 2 u 1 u 2 2 + 2 u 1 u 4 2 + 4 u 2 u 3 u 4 , u 3 , t 2 = u 3 , x x 2 u 1 2 u 4 4 u 1 u 2 u 3 2 u 3 2 u 4 , u 4 , t 2 = u 4 , x x + 4 u 1 u 2 u 4 + 2 u 2 2 u 3 + 2 u 3 u 4 2 ,
and a coupled system of integrable mKdV equations:
u 1 , t 3 = u 1 , x x x 6 ( u 1 u 2 + u 3 u 4 ) u 1 , x 6 ( u 1 u 4 + u 2 u 3 ) u 3 , x , u 2 , t 3 = u 2 , x x x 6 ( u 1 u 2 + u 3 u 4 ) u 2 , x 6 ( u 1 u 4 + u 2 u 3 ) u 4 , x , u 3 , t 3 = u 3 , x x x 6 ( u 1 u 4 + u 2 u 3 ) u 1 , x 6 ( u 1 u 2 + u 3 u 4 ) u 3 , x , u 4 , t 3 = u 4 , x x x 6 ( u 1 u 4 + u 2 u 3 ) u 2 , x 6 ( u 1 u 2 + u 3 u 4 ) u 4 , x .
If taking
ξ = 1 , η = 0 , μ = 1 , ν = 1 ,
one obtains a coupled system of the following combined integrable NLS equations:
u 1 , t 2 = u 1 , x x + u 3 , x x 2 ( u 2 + u 4 ) u 1 2 4 ( u 2 + u 4 ) u 1 u 3 2 ( u 2 + u 4 ) u 3 2 , u 2 , t 2 = u 2 , x x u 4 , x x + 2 ( u 1 + u 3 ) u 2 2 + 4 ( u 1 + u 3 ) u 2 u 4 + 2 ( u 1 + u 3 ) u 4 2 , u 3 , t 2 = u 1 , x x + u 3 , x x 2 ( u 2 + u 4 ) u 1 2 4 ( u 2 u 4 ) u 1 u 3 2 ( u 2 + u 4 ) u 3 2 , u 4 , t 2 = u 2 , x x u 4 , x x + 2 ( u 1 + u 3 ) u 2 2 + 4 ( u 1 + u 3 ) u 2 u 4 + 2 ( u 1 + u 3 ) u 4 2 ,
and a coupled system of the following combined integrable mKdV equations:
u 1 , t 3 = u 1 , x x x + u 3 , x x x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 1 , x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 3 , x , u 2 , t 3 = u 2 , x x x + u 4 , x x x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 2 , x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 4 , x , u 3 , t 3 = u 1 , x x x + u 3 , x x x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 1 , x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 3 , x , u 4 , t 3 = u 2 , x x x + u 4 , x x x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 2 , x 6 [ u 1 ( u 2 + u 4 ) + u 3 ( u 2 + u 4 ) ] u 4 , x .
These four systems represent typical coupled integrable models, expanding the class of coupled integrable NLS equations and mKdV equations (see, e.g., [26,27]).

3. Bi-Hamiltonian Structures

The introduction of bi-Hamiltonian structures into the soliton hierarchy, Equation (23), can be achieved by employing the classical trace identity, Equation (3), on the spatial matrix eigenvalue problem, Equation (5).
The trace identity uses the solution R defined by Equation (8). One can then readily work out the Hamiltonian structures for the resultant hierarchy of soliton models. Through applying the classical trace identity to the spatial matrix eigenvalue problem, the Hamiltonian densities and the associated flows can be systematically derived. Concretely, we have
tr R P k = 2 ξ a + 2 η e , tr R P u = ( 2 c , 2 b , 2 g , 2 f ) T ,
and consequently, the classical trace identity gives
δ δ u k ( n + 1 ) ( ξ a { n + 1 } + η e { n + 1 } ) d x = k τ k k τ n ( c { n } , b { n } , g { n } , f { n } ) T , n 0 .
When checked with n = 2 , it results in τ = 0 , and as a consequence, one obtains
δ δ u H { n } = ( c { n + 1 } , b { n + 1 } , g { n + 1 } , f { n + 1 } ) T , n 0 ,
in which the required Hamiltonian quantities are computed as follows:
H { n } = ξ a { n + 2 } + η e { n + 2 } n + 1 d x , n 0 .
This allows us to establish the Hamiltonian structures for the soliton hierarchy, Equation (23):
u t m = Ƶ { m } = J 1 δ H { m } δ u , J 1 = 0 ξ 0 η ξ 0 η 0 0 η 0 ξ η 0 ξ 0 , m 0 ,
where J 1 is, obviously, Hamiltonian, and H [ m ] are the functionals defined by Equation (33). It is important to note that Hamiltonian structures exhibit a significant property, namely, the interrelation S = J 1 δ H δ u between a conserved quantity H and a symmetry S within the same nonlinear model.
The standard soliton theory expresses that those vector fields Ƶ { n } commute
[ [ Ƶ { n 1 } , Ƶ { n 2 } ] ] = Ƶ { n 1 } ( u ) [ Ƶ { n 2 } ] Ƶ { n 2 } ( u ) [ Ƶ { n 1 } ] = 0 , n 1 , n 2 0 ,
which can bee seen from an algebra of temporal spectral matrices:
[ [ Q { n 1 } , Q { n 2 } ] ] = Q { n 1 } ( u ) [ Ƶ { n 2 } ] Q { n 2 } ( u ) [ Ƶ { n 1 } ] + [ Q { n 1 } , Q { n 2 } ] = 0 , n 1 , n 2 0 .
One can also verify this property directly by analyzing the relationship between the isospectral zero-curvature equations.
Moreover, utilizing the recursion relation Ƶ m + 1 = Φ Ƶ m , a straightforward yet lengthy computation results in a recursion operator Φ = ( Φ j k ) 4 × 4 , which is established as hereditary [28], for the soliton hierarchy, Equation (23). This hereditary recursion operator Φ reads
Φ 11 = 1 ξ 2 η 2 ( ξ 2 ξ u 1 1 u 2 2 ξ u 3 1 u 4 + 2 η u 1 1 u 4 + 2 η u 3 1 u 2 ) , Φ 12 = 1 ξ 2 η 2 ( 2 ξ u 1 1 u 1 2 ξ u 3 1 u 3 + 2 η u 1 1 u 3 + 2 η u 3 1 u 1 ) , Φ 13 = 1 ξ 2 η 2 ( η + 2 ξ u 1 1 u 4 2 ξ u 3 1 u 2 + 2 η u 1 1 u 2 + 2 η u 3 1 u 4 ) , Φ 14 = 1 ξ 2 η 2 ( 2 ξ u 1 1 u 3 2 ξ u 3 1 u 1 + 2 η u 1 1 u 1 + 2 η u 3 1 u 3 ) ;
Φ 21 = 1 ξ 2 η 2 ( 2 ξ u 2 1 u 2 + 2 ξ u 4 1 u 4 2 η u 2 1 u 4 2 η u 4 1 u 2 ) , Φ 22 = 1 ξ 2 η 2 ( ξ + 2 ξ u 2 1 u 1 + 2 ξ u 4 1 u 3 2 η u 2 1 u 3 2 η u 4 1 u 1 ) , Φ 23 = 1 ξ 2 η 2 ( 2 ξ u 2 1 u 4 + 2 ξ u 4 1 u 2 2 η u 2 1 u 2 2 η u 4 1 u 4 ) , Φ 24 = 1 ξ 2 η 2 ( η + 2 ξ u 2 1 u 3 + 2 ξ u 4 1 u 1 2 η u 2 1 u 1 2 η u 4 1 u 3 ) ;
Φ 31 = 1 ξ 2 η 2 ( η 2 ξ u 1 1 u 4 2 ξ u 3 1 u 2 + 2 η u 1 1 u 2 + 2 η u 3 1 u 4 ) , Φ 32 = 1 ξ 2 η 2 ( 2 ξ u 1 1 u 3 2 ξ u 3 1 u 1 + 2 η u 1 1 u 1 + 2 η u 3 1 u 3 ) , Φ 33 = 1 ξ 2 η 2 ( ξ 2 ξ u 1 1 u 2 2 ξ u 3 1 u 4 + 2 η u 1 1 u 4 + 2 η u 3 1 u 2 ) , Φ 34 = 1 ξ 2 η 2 ( 2 ξ u 1 1 u 1 2 ξ u 3 1 u 3 + 2 η u 1 1 u 3 + 2 η u 3 1 u 1 ) ;
and
Φ 41 = 1 ξ 2 η 2 ( 2 ξ u 2 1 u 4 + 2 ξ u 4 1 u 2 2 η u 2 1 u 2 2 η u 4 1 u 4 ) , Φ 42 = 1 ξ 2 η 2 ( η + 2 ξ u 2 1 u 3 + 2 ξ u 4 1 u 1 2 η u 2 1 u 1 2 η u 4 1 u 3 ) , Φ 43 = 1 ξ 2 η 2 ( 2 ξ u 2 1 u 2 + 2 ξ u 4 1 u 4 2 η u 2 1 u 4 2 η u 4 1 u 2 ) , Φ 44 = 1 ξ 2 η 2 ( ξ + 2 ξ u 2 1 u 1 + 2 ξ u 4 1 u 3 2 η u 2 1 u 3 2 η u 4 1 u 1 ) .
Let us show the idea for computing this recursion operator using the recursion relations in Equations (12)–(14). Assume that Ƶ { m } = ( Ƶ 1 { m } , Ƶ 2 { m } , Ƶ 3 { m } , Ƶ 4 { m } ) T , m 0 . Here, we focus solely on the third component of Ƶ { m + 1 } and perform the following computation:
Ƶ 3 { m + 1 } = η b { m + 2 } + ξ f { m + 2 } = 1 ξ 2 η 2 [ ξ η ( b x { m + 1 } + 2 u 1 a { m + 1 } + 2 u 3 e { m + 1 } ) η 2 ( f x { m + 1 } + 2 u 1 e { m + 1 } + 2 u 3 a { m + 1 } ) ξ η ( b x { m + 1 } + 2 u 1 a { m + 1 } + 2 u 3 e { m + 1 } ) + ξ 2 ( f x { m + 1 } + 2 u 1 e { m + 1 } + 2 u 3 a { m + 1 } ) ] = 1 ξ 2 η 2 [ ξ Ƶ 3 , x { m } η Ƶ 1 , x { m } + 2 ( ξ u 1 η u 3 ) ( η a { m + 1 } + ξ e { m + 1 } ) + 2 ( ξ u 3 η u 1 ) ( η e { m + 1 } + ξ a { m + 1 } ) ] .
On the other hand, we have
η a { m + 1 } + ξ e { m + 1 } = 1 ( u 1 Ƶ 4 { m } + u 2 Ƶ 3 { m } + u 3 Ƶ 2 { m } + u 4 Ƶ 1 { m } ) , η e { m + 1 } + ξ a { m + 1 } = 1 ( u 1 Ƶ 2 { m } + u 2 Ƶ 1 { m } + u 3 Ƶ 4 { m } + u 4 Ƶ 3 { m } ) .
All this yields the third row of the recursion operator Φ , defined by Equation (39). The remaining rows can be derived in a completely similar manner.
The recursion operator, defined by Equations (37)–(40), involves two constant parameters, ξ and η , which are not simultaneously zero, exhibiting the diversity of the recursion structure in the integrable hierarchy. Despite the nonlocality of the recursion operator, the locality of the isospectral ( k t m = 0 ) flows is maintained. This implies that each flow in the hierarchy preserves the integrable structure, ensuring that the derived soliton equations remain solvable by inverse scattering techniques and other methods applicable to local equations.
Further, with some detailed analysis, we can see that J 1 and J 2 = Φ J 1 form a Hamiltonian pair. Therefore, the soliton hierarchy Equation (23) exhibits the following bi-Hamiltonian structures [29]:
u t m = Ƶ { m } = J 1 δ H { m } δ u = J 2 δ H { m 1 } δ u , m 1 .
It can then be observed that the resulting Hamiltonian quantities commute under their respective Poisson brackets:
{ H { n 1 } , H { n 2 } } J i = 0 , n 1 , n 2 0 , i = 1 , 2 ,
where
{ H , K } J i = δ H δ u T J i δ K δ u d x , i = 1 , 2 .
The two equalities established above, Equations (35) and (42), imply that all isospectral flows possess infinitely many conserved quantities and symmetries inherent to the integrable system. Additionally, based on the recursion and bi-Hamiltonian structures, the resulting conserved quantities and symmetries can be effectively computed and utilized. This property is crucial for the practical application and analysis of these integrable models, as it ensures that the solutions exhibit well-defined physical behavior and can be systematically studied.
In summary, the soliton hierarchy, Equation (23), exhibits specific bi-Hamiltonian structures, demonstrating Liouville integrability. Each model features infinitely many commuting conserved quantities H { n } n = 0 and symmetries Ƶ { n } n = 0 . The concrete examples provided in Equations (25), (26), (28), and (29) highlight special nonlinear coupled Liouville integrable models with bi-Hamiltonian structures, contributing to the ongoing discourse in the literature (see, for instance, [30,31,32,33,34,35]).

4. Concluding Remarks

This research explores integrable hierarchies and their relationship to specific matrix eigenvalue problems formulated under zero-curvature. Generating integrable models with bi-Hamiltonian structures is essential for comprehending the dynamics inherent in these systems.
Employing Laurent series solutions for solving the stationary zero-curvature equation proves to be a robust method, enabling researchers to reveal the integrability characteristics of the models under investigation. Furthermore, applying the trace identity to the matrix isospectral eigenvalue problem provides deeper insights into the bi-Hamiltonian structures embedded within these systems.
The concrete examples presented provide specific coupled systems of nonlinear uncombined and combined integrable models. These examples, which belong to M 2 -extensions [36], demonstrate the practical application of the theoretical framework discussed earlier and highlight the integrability and rich structure of the resulting equations.
Exploring the structures of explicit soliton solutions in the resulting integrable models is of interest, employing advanced methods in soliton theory such as the Zakharov–Shabat dressing method [37], the Riemann–Hilbert technique [38], the determinant approach [39], and the Darboux transformation (see, e.g., [40,41,42,43,44]). Additionally, other significant solutions including breather, kink, anti-kink, lump and rogue wave solutions, as well as the corresponding mixed solutions (see, e.g., [45,46,47,48,49,50,51,52]), can be derived from specific wave number reductions of solitons. Novel reduced integrable equations involving reflection points can also be obtained through nonlocal reduced matrix eigenvalue problems under similarity transformations (see, e.g., [53]).
Most certainly, increasing the number of dependent variables in the spatial spectral matrix can indeed lead to the generation of larger integrable models (see, e.g., [54,55,56]). However, it is worth noting that as the number of dependent variables increases, the complexity of the resulting equations also grows. This can make the analysis and understanding of the system more challenging. Nevertheless, the study of larger integrable models remains a fruitful area of research [36,53], offering insights into the fundamental principles governing nonlinear dynamics and integrability in mathematical physics.

Funding

This work was supported in part by the Ministry of Science and Technology of China (G2021016032L and G2023016011L) and the National Natural Science Foundation of China (12271488 and 11975145).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Lax, P.D. Integrals of nonlinear equations of evolution and solitary waves. Comm. Pure Appl. Math. 1968, 21, 467–490. [Google Scholar] [CrossRef]
  2. Ablowitz, M.J.; Segur, H. Solitons and the Inverse Scattering Transform; SIAM: Philadelphia, PA, USA, 1981. [Google Scholar]
  3. Das, A. Integrable Models; World Scientific: Teaneck, NJ, USA, 1989. [Google Scholar]
  4. Ablowitz, M.J.; Kaup, D.J.; Newell, A.C.; Segur, H. The inverse scattering transform-Fourier Analysis for nonlinear problems. Stud. Appl. Math. 1974, 53, 249–315. [Google Scholar] [CrossRef]
  5. Tu, G.Z. On Liouville integrability of zero-curvature equations and the Yang hierarchy. J. Phys. A Math. Gen. 1989, 22, 2375–2392. [Google Scholar]
  6. Drinfel′d, V.; Sokolov, V.V. Lie algebras and equations of Korteweg—de Vries type. Sov. J. Math. 1985, 30, 1975–2036. [Google Scholar] [CrossRef]
  7. Antonowicz, M.; Fordy, A.P. Coupled KdV equations with multi-Hamiltonian structures. Physica D 1987, 28, 345–357. [Google Scholar] [CrossRef]
  8. Geng, X.G. A hierarchy of non-linear evolution equations, its Hamiltonian structure and classical integrable system. Physica A 1992, 180, 241–251. [Google Scholar] [CrossRef]
  9. Guo, F.K. A variant of Lax representations and Lax representations of hierarchies of Hamilton’s equations. Acta Math. Sinica 1994, 37, 515–522. Available online: https://actamath.cjoe.ac.cn/Jwk_sxxb_cn/EN/10.12386/A1994sxxb0069 (accessed on 6 July 2024).
  10. Zhao, Q.L.; Li, Y.X.; Li, X.Y.; Sun, Y.P. The finite-dimensional super integrable system of a super NLS-mKdV equation. Commun. Nonlinear Sci. Numer. Simulat. 2012, 17, 4044–4052. [Google Scholar] [CrossRef]
  11. Zhaqilao. A generalized AKNS hierarchy, bi-Hamiltonian structure, and Darboux transformation. Commun. Nonlinear Sci. Numer. Simul. 2012, 17, 2319–2332. [Google Scholar] [CrossRef]
  12. Manukure, S. Finite-dimensional Liouville integrable Hamiltonian systems generated from Lax pairs of a bi-Hamiltonian soliton hierarchy by symmetry constraints. Commun. Nonlinear Sci. Numer. Simul. 2018, 57, 125–135. [Google Scholar] [CrossRef]
  13. Liu, T.S.; Xia, T.C. Multi-component generalized Gerdjikov-Ivanov integrable hierarchy and its Riemann-Hilbert problem. Nonlinear Anal. Real World Appl. 2022, 68, 103667. [Google Scholar] [CrossRef]
  14. Wang, H.F.; Zhang, Y.F. Application of Riemann-Hilbert method to an extended coupled nonlinear Schrödinger equations. J. Comput. Appl. Math. 2023, 420, 114812. [Google Scholar] [CrossRef]
  15. Ma, W.X. Novel Liouville integrable Hamiltonian models with six components and three signs. Chin. J. Phys. 2023, 86, 292–299. [Google Scholar] [CrossRef]
  16. Ma, W.X. A four-component hierarchy of combined integrable equations with bi-Hamiltonian formulations. Appl. Math. Lett. 2024, 153, 109025. [Google Scholar] [CrossRef]
  17. Xia, T.C.; Yu, F.J.; Zhang, Y. The multi-component coupled Burgers hierarchy of soliton equations and its multi-component integrable couplings system with two arbitrary functions. Physica A 2004, 343, 238–246. [Google Scholar] [CrossRef]
  18. Li, Z.; Dong, H.H. Two integrable couplings of the Tu hierarchy and their Hamiltonian structures. Comput. Math. Appl. 2008, 55, 2643–2652. [Google Scholar] [CrossRef]
  19. Xu, X.X. Integrable couplings of relativistic Toda lattice systems in polynomial form and rational form, their hierarchies and bi-Hamiltonian structures. J. Phys. A Math. Theoret. 2009, 42, 395201. [Google Scholar] [CrossRef]
  20. Xu, X.X. An integrable coupling hierarchy of the Mkdv_integrable systems, its Hamiltonian structure and corresponding nonisospectral integrable hierarchy. Appl. Math. Comput. 2010, 216, 344–353. [Google Scholar] [CrossRef]
  21. You, F.C. Nonlinear super integrable Hamiltonian couplings. J. Math. Phys. 2011, 52, 123510. [Google Scholar] [CrossRef]
  22. Wu, J.Z.; Xing, X.Z.; Geng, X.G. Integrable couplings of fractional L-hierarchy and its Hamiltonian structures. Math. Methods Appl. Sci. 2016, 39, 3925–3931. [Google Scholar] [CrossRef]
  23. Yao, Y.Q.; Li, C.X.; Shen, S.F. Completion of the integrable coupling systems. arXiv 2017, arXiv:1711.04073. [Google Scholar]
  24. Wang, H.F.; Zhang, Y.F. A new multi-component integrable coupling and its application to isospectral and nonisospectral problems. Commun. Nonlinear Sci. Numer. Simul. 2022, 105, 106075. [Google Scholar] [CrossRef]
  25. Zhang, Y.J.; Ma, W.X.; Ünsal, Ö. A novel kind of AKNS integrable couplings and their Hamiltonian structures. Turk. J. Math. 2017, 41, 1467–1476. [Google Scholar] [CrossRef]
  26. Ma, W.X. A generalized hierarchy of combined integrable bi-Hamiltonian equations from a specific fourth-order matrix spectral problem. Mathematics 2024, 12, 927. [Google Scholar] [CrossRef]
  27. Ma, W.X. A combined Kaup-Newell type integrable Hamiltonian hierarchy with four potentials and a hereditary recursion operator. Discrete Cont. Dyn. Syst. S 2024, 17, 108775. [Google Scholar] [CrossRef]
  28. Fuchssteiner, B.; Fokas, A.S. Symplectic structures, their Bäcklund transformations and hereditary symmetries. Physica D 1981, 4, 47–66. [Google Scholar] [CrossRef]
  29. Magri, F. A simple model of the integrable Hamiltonian equation. J. Math. Phys. 1978, 19, 1156–1162. [Google Scholar] [CrossRef]
  30. Yang, J.Y.; Ma, W.X. Four-component Liouville integrable models and their bi-Hamiltonian formulations. Rom. J. Phys. 2024, 69, 101. [Google Scholar] [CrossRef]
  31. Ma, W.X. A combined Liouville integrable hierarchy associated with a fourth-order matrix spectral problem. Commun. Theor. Phys. 2024, 76, 075001. [Google Scholar] [CrossRef]
  32. Ma, W.X. Four-component combined integrable equations possessing bi-Hamiltonian formulations. Mod. Phys. Lett. B 2024, 38, 2450319. [Google Scholar] [CrossRef]
  33. Li, C.X. A hierarchy of coupled Korteweg—de Vries equations and the corresponding finite-dimensional integrable system. J. Phys. Soc. Jpn. 2004, 73, 327–331. [Google Scholar] [CrossRef]
  34. Zhao, Q.L.; Cheng, H.B.; Li, X.Y.; Li, C.Z. Integrable nonlinear perturbed hierarchies of NLS-mKDV equation and soliton solutions. Electr. J. Differ. Equ. 2022, 2022, 71. [Google Scholar] [CrossRef]
  35. Zhou, R.G.; Zhu, H.Y. An integrable matrix NLS equation on star graph and symmetry-dependent connection conditions of vertex. Comput. Appl. Math. 2023, 42, 69. [Google Scholar] [CrossRef]
  36. Ma, W.X. Integrable couplings and two-dimensional unital algebras. Axioms 2024, 13, 481. [Google Scholar] [CrossRef]
  37. Doktorov, E.V.; Leble, S.B. A Dressing Method in Mathematical Physics; Springer: Dordrecht, The Netherlands, 2007. [Google Scholar]
  38. Novikov, S.P.; Manakov, S.V.; Pitaevskii, L.P.; Zakharov, V.E. Theory of Solitons: The Inverse Scattering Method; Consultants Bureau: New York, NY, USA, 1984. [Google Scholar]
  39. Aktosun, T.; Busse, T.; Demontin, F.; van der Mee, C. Symmetries for exact solutions to the nonlinear Schrödinger equation. J. Phys. A Math. Theoret. 2010, 43, 025202. [Google Scholar] [CrossRef]
  40. Matveev, V.B.; Salle, M.A. Darboux Transformations and Solitons; Springer-Verlag: Berlin/Heidelberg, Germany, 1991. [Google Scholar]
  41. Geng, X.G.; Li, R.M.; Xue, B. A vector general nonlinear Schrödinger equation with (m+n) components. J. Nonlinear Sci. 2020, 30, 991–1013. [Google Scholar] [CrossRef]
  42. Ma, W.X. Binary Darboux transformation of vector nonlocal reverse-time integrable NLS equations. Chaos Solitos Fractals 2024, 180, 114539. [Google Scholar] [CrossRef]
  43. Ye, R.S.; Zhang, Y. A vectorial Darboux transformation for the Fokas-Lenells system. Chaos Solitons Fractals 2023, 169, 113233. [Google Scholar] [CrossRef]
  44. Ma, W.X.; Huang, Y.H.; Wang, F.D.; Zhang, Y.; Ding, L.Y. Binary Darboux transformation of vector nonlocal reverse-space nonlinear Schrödinger equations. Int. J. Geom. Methods Mod. Phys. 2024, 21, 2450182. [Google Scholar] [CrossRef]
  45. Cheng, L.; Zhang, Y.; Lin, M.J. Lax pair and lump solutions for the (2+1)-dimensional DJKM equation associated with bilinear Bäcklund transformations. Anal. Math. Phys. 2019, 9, 1741–1752. [Google Scholar] [CrossRef]
  46. Sulaiman, T.A.; Yusuf, A.; Abdeljabbar, A.; Alquran, M. Dynamics of lump collision phenomena to the (3+1)-dimensional nonlinear evolution equation. J. Geom. Phys. 2021, 169, 104347. [Google Scholar] [CrossRef]
  47. Yusuf, A.; Sulaiman, T.A.; Abdeljabbar, A.; Alquran, M. Breathem waves, analytical solutions and conservation lawn using Lie–Bäcklund symmetries to the (2+1)-dimensional Chaffee-Infante equation. J. Ocean Eng. Sci. 2023, 8, 145–151. [Google Scholar] [CrossRef]
  48. Manukure, S.; Chowdhury, A.; Zhou, Y. Complexiton solutions to the asymmetric Nizhnik-Novikov-Veselov equation. Int. J. Mod. Phys. B 2019, 33, 1950098. [Google Scholar] [CrossRef]
  49. Zhou, Y.; Manukure, S.; McAnally, M. Lump and rogue wave solutions to a (2+1)-dimensional Boussinesq type equation. J. Geom. Phys. 2021, 167, 104275. [Google Scholar] [CrossRef]
  50. Manukure, S.; Zhou, Y. A study of lump and line rogue wave solutions to a (2+1)-dimensional nonlinear equation. J. Geom. Phys. 2021, 167, 104274. [Google Scholar] [CrossRef]
  51. Yang, S.X.; Wang, Y.F.; Zhang, X. Conservation laws, Darboux transformation and localized waves for the N-coupled nonautonomous Gross-Pitaevskii equations in the Bose-Einstein condensates. Chaos Solitons Fractals 2023, 169, 113272. [Google Scholar] [CrossRef]
  52. Cheng, L.; Zhang, Y. Grammian-type determinant solutions to generalized KP and BKP equations. Comput. Math. Appl. 2017, 74, 727–735. [Google Scholar] [CrossRef]
  53. Ma, W.X. Type (λ*,λ) reduced nonlocal integrable AKNS equations and their soliton solutions. Appl. Numer. Math. 2024, 199, 105–113. [Google Scholar] [CrossRef]
  54. Zhang, Y.F.; Han, Z.; Zhao, Z.L. Applications of a few Lie algebras. Acta Math. Appl. Sin. Engl. Ser. 2016, 32, 289–304. [Google Scholar] [CrossRef]
  55. Gerdjikov, V.S. Nonlinear evolution equations related to Kac-Moody algebras Ar(1): Spectral aspects. Turk. J. Math. 2022, 46, 1828–1844. [Google Scholar] [CrossRef]
  56. Geng, X.G.; Jia, M.X.; Xue, B.; Zhai, Y.Y. Application of tetragonal curves to coupled Boussinesq equations. Lett. Math. Phys. 2024, 114, 30. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ma, W.-X. An Integrated Integrable Hierarchy Arising from a Broadened Ablowitz–Kaup–Newell–Segur Scenario. Axioms 2024, 13, 563. https://doi.org/10.3390/axioms13080563

AMA Style

Ma W-X. An Integrated Integrable Hierarchy Arising from a Broadened Ablowitz–Kaup–Newell–Segur Scenario. Axioms. 2024; 13(8):563. https://doi.org/10.3390/axioms13080563

Chicago/Turabian Style

Ma, Wen-Xiu. 2024. "An Integrated Integrable Hierarchy Arising from a Broadened Ablowitz–Kaup–Newell–Segur Scenario" Axioms 13, no. 8: 563. https://doi.org/10.3390/axioms13080563

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop