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


Quasi-Gramian Solution of a Noncommutative Extension of the Higher-Order Nonlinear Schrödinger Equation


H. W. A. Riaz1 and J. Lin1
1 department of physics, zhejiang normal university, jinhua 321004, pr china
Abstract.

The nonlinear Schrödinger (NLS) equation, which incorporates higher-order dispersive terms, is widely employed in the theoretical analysis of various physical phenomena. In this study, we explore the non-commutative extension of the higher-order NLS equation (HNLS). We treat real or complex-valued functions, such as g1=g1(x,t)g_{1}=g_{1}(x,\;t) and g2=g2(x,t)g_{2}=g_{2}(x,\;t), as non-commutative, and employ the Lax pair associated with the evolution equation as in the commutation case. We derive the quasi-Gramian solution of the system by employing a binary Darboux transformation (DT). Moreover, the solution can be used to study the stability of plane waves and to understand the generation of periodic patterns in the context of modulational instability.

1. Introduction

The nonlinear Schrödinger (NLS) equation, which incorporates higher-order dispersive terms, is widely used in the theoretical analysis of various physical phenomena, including nonlinear optics, molecular systems, and fluid dynamics [1, 2, 3]. With the addition of fourth-order terms, known as the Lakshmanan-Porsezian-Daniel (LPD) equation, it describes higher-order molecular excitations with quadruple-quadruple coefficients and possesses integrability [4, 5, 6]. Lakshmanan et al. investigated its application to study nonlinear spin excitations involving bilinear and biquadratic interactions [5]. In recent years, Ankiewicz et al. introduced a further extension of the NLSE by incorporating third-order (odd) and fourth-order (even) dispersion terms [7]. The integrability of this extended NLSE, with certain parameter values, was confirmed in Ref. [8], where Lax operators were introduced. We now write this equation appears in the mentioned references with some modification as

iut+α2(uxx+2u|u|2)+iα1(uxxx+6ux|u|2)+γ(uxxxx+6u¯ux2+4u|ux|2\displaystyle\mathrm{i}u_{t}+\alpha_{2}\left(u_{xx}+2u|u|^{2}\right)+\mathrm{i}\alpha_{1}\left(u_{xxx}+6u_{x}|u|^{2}\right)+\gamma(u_{xxxx}+6\bar{u}u_{x}^{2}+4u|u_{x}|^{2} (1.1)
+8|u|2uxx+2u2u¯xx+6u|u|4)=0.\displaystyle+8|u|^{2}u_{xx}+2u^{2}\bar{u}_{xx}+6u|u|^{4})=0.

where u=u(x,t)u=u(x,\;t) is a complex-valued scalar function, and u¯\bar{u} represents its complex conjugate. This equation includes several particular cases, such as the standard nonlinear Schrödinger equation (NLSE) with α1=γ=0\alpha_{1}=\gamma=0 [11], the Hirota equation with γ=0\gamma=0 [12], and the Lakshmanan-Porsezian-Daniel equation with α1=0\alpha_{1}=0 [5].

In this study, we explore the non-commutative extension of the higher-order NLS (HNLS) equation (1.1). Noncommutative integrable systems have received considerable attention for their relevance in quantum field theories, DD-brane dynamics, and string theories [13, 14, 15]. Non-commutativity often arises from phase-space quantization, introducing non-commutativity among independent variables through a star product [16, 17]. Our approach to inducing non-commutativity in a given nonlinear evolution equation parallels the methods employed by Lechtenfeld et al. [18], Gilson and Nimmo [19], and Gilson and Macfarlane [20] for the non-commutative generalization of the sine-Gordon, Kadomtsev-Petviashvili (KP), and Davey-Stewartson (DS) equations, respectively.

We adopt a systematic method to extend the chosen equation to its non-commutative form, without explicitly specifying the nature of non-commutativity. We consider real or complex-valued functions, such as g1=g1(x,t)g_{1}=g_{1}(x,\;t) and g2=g2(x,t)g_{2}=g_{2}(x,\;t), as non-commutative and take advantage of the same Lax pair as in the commutative scenario to describe the equation of non-linear evolution.

In this paper, we investigate a non-commutative (nc) version of the HNLS equation. We define the Lax pair for the nc-HNLS equation in this context. To find solutions to the nc-HNLS equation, we construct the Darboux matrix and the binary Darboux matrix. We present explicit quasi-Gramian solutions for the non-commutative fields of the nc-HNLS equation, which, after reducing the non-commutativity limit, can be reduced to a ratio of Gramian solutions.

2. Modulation instability

For analyzing the modulation instability, we give a plane-wave solution of the system (1.1)

u(x,t)=cei(6c4γ+2α2c2)t,u(x,t)=ce^{i(6c^{4}\gamma+2\alpha_{2}c^{2})t}, (2.1)

The solution provided by equation (2.1) holds significant importance in the realm of optics, particularly within the context of the nonlinear Schrödinger equation. This solution represents a wave with constant amplitude that undergoes a non-linear evolution over time. The dynamics are determined by parameters such as the amplitude cc, the constant γ\gamma, and α2\alpha_{2}. This solution’s application extends to studying the stability of plane waves and comprehending the generation of periodic patterns through modulational instability. It serves as a prime example of how the nonlinear Schrödinger equation can lead to complex behavior in optical systems, thereby making it a crucial area of research in this field.

An approach to assess the stability of the plane wave solution involves introducing perturbations to the solution and examining the linearized evolution of these perturbations. To simplify the analysis, the common phase can be factored out of the equation. This leads to a first-order ordinary differential equation (ODE) that couples the complex field with its complex conjugate as a result of the perturbation. Substituting the perturbed function v(x,t)v(x,\;t) into equation (2.1), we obtain

u(x,t)=(c+v(x,t))ei(6c4γ+2α2c2)t,u(x,t)=(c+v(x,\;t))e^{i(6c^{4}\gamma+2\alpha_{2}c^{2})t}, (2.2)

Substituting (2.2) into (1.1) and after linearization, we have

ivt+α2vxx+iα1(vxxx+6c2vx)+γvxxxx+2γc2(4vxx+v¯xx)+c2(12γc2+2α2)(v+v¯)=0,iv_{t}+\alpha_{2}v_{xx}+i\alpha_{1}(v_{xxx}+6c^{2}v_{x})+\gamma v_{xxxx}+2\gamma c^{2}(4v_{xx}+\bar{v}_{xx})+c^{2}(12\gamma c^{2}+2\alpha_{2})(v+\bar{v})=0, (2.3)

In order to analyze the stability of the plane wave solution, the Fourier transform of the equation is taken. This results in a first-order ODE that governs the real and imaginary parts of evolution. The stability of the solution can then be determined by looking at the eigenvalues of this ODE. In particular, the eigenvalues represent the exponent in the time evolution of the solution. Thus, the Fourier transform of the evolution equation (2.3) is

iv^tα22k2v^α1k(k2+6c2)v^+γk4v^2γc2k2(4v^+v^¯)+c2(12γc2+2α2)(v^+v^¯)=0,i\hat{v}_{t}-\frac{\alpha_{2}}{2}k^{2}\hat{v}-\alpha_{1}k(-k^{2}+6c^{2})\hat{v}+\gamma k^{4}\hat{v}-2\gamma c^{2}k^{2}(4\hat{v}+\bar{\hat{v}})+c^{2}(12\gamma c^{2}+2\alpha_{2})(\hat{v}+\bar{\hat{v}})=0, (2.4)

The linear evolution equation for v^\hat{v} can be evaluated by separating it into its real and imaginary components. Thus, for v^=v1+iv2\hat{v}=v_{1}+iv_{2}, we have system of differential equation

yt=[0β6γc2k2β10γc2k2+2c2(12c2γ+2α2)0]y,y_{t}=\left[\begin{array}[]{cc}0&\beta-6\gamma\,c^{2}k^{2}\\ -\beta-10\gamma\,c^{2}k^{2}+2c^{2}\left(12c^{2}\gamma+2\alpha_{2}\right)&0\end{array}\right]y, (2.5)

where β=α2k22+α1k(6c2k2)γk4\beta=\frac{\alpha 2\,k^{2}}{2}+\alpha_{1}k\left(6c^{2}-k^{2}\right)-\gamma\,k^{4}, and y=[v1v2]Ty=[\begin{array}[]{cc}v_{1}&v_{2}\end{array}]^{T}. One can evaluate the stability of a system by analyzing exponential solutions in the form of y=νeωty=\nu e^{\omega t}, which leads to an eigenvalue problem where the eigenvalues are then given by

ω(k)=|k|2(α2(8c2+k2)2α1k(6c2+k2)2γ(24c410c2k2+k4))β1,\displaystyle\omega(k)={\frac{|k|}{2}\sqrt{\left(\alpha_{2}\left(-8c^{2}+k^{2}\right)-2\alpha_{1}k\left(-6c^{2}+k^{2}\right)-2\gamma\left(24c^{4}-10c^{2}k^{2}+k^{4}\right)\right)\beta_{1}}}, (2.6)
β1=kα22α1(6c2+k2)2γk(6c2+k2).\displaystyle\beta_{1}=k\alpha_{2}-2\alpha_{1}\left(-6c^{2}+k^{2}\right)-2\gamma k\left(6c^{2}+k^{2}\right).

The plot of the equation (2.6) is shown in Fig. 1

Refer to caption
Figure 1.

The stability of the solution becomes evident when examining a graphical representation of the real parts of eigenvalues plotted against different frequencies. If the real part is positive, the solution will exhibit growth; conversely, negative values indicate decay. The overall stability is established by observing whether the real eigenvalue remains positive or negative across various frequency ranges. This phenomenon is referred to as modulational instability.

3. Non-commutative HNLS equation

The spectral problem associated with the nc-HNLS equation is given by

Γ=xλ𝒥𝒰,\displaystyle\Gamma=\partial_{x}-\lambda\mathcal{J}-\mathcal{U}, (3.1)
Δ=t𝒱p,\displaystyle\Delta=\partial_{t}-\mathcal{B}-\mathcal{V}_{p}, (3.2)

where

𝒥=[I00I],𝒰=[0uu0],𝒱p=γ[𝒱1𝒱2𝒱3𝒱4],\displaystyle\mathcal{J}=\left[\begin{array}[]{cc}\mathrm{-I}&0\\ 0&\mathrm{I}\end{array}\right],\quad\mathcal{U}=\left[\begin{array}[]{cc}0&{u}\\ -{\mathit{u^{{\dagger}}}}&0\end{array}\right],\quad\mathcal{V}_{p}=\gamma\left[\begin{array}[]{cc}\mathcal{V}_{1}&\mathcal{V}_{2}\\ \mathcal{V}_{3}&\mathcal{V}_{4}\end{array}\right], (3.9)
=(ρ1+ρ2uuα1((xu)uu(xu))𝒜2𝒜1ρ1ρ2uuα1(u(xu)+(xu)u)),\displaystyle\mathcal{B}=\small{\begin{pmatrix}\rho_{1}+\rho_{2}{u}{\mathit{u^{{\dagger}}}}-\alpha_{1}\left(\left(\frac{\partial}{\partial x}{u}\right){\mathit{u^{{\dagger}}}}-{u}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)\right)&\mathcal{A}_{2}\\ \mathcal{A}_{1}&-\rho_{1}-\rho_{2}{\mathit{u^{{\dagger}}}}{u}-\alpha_{1}\left(-{\mathit{u^{{\dagger}}}}\left(\frac{\partial}{\partial x}{u}\right)+\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right){u}\right)\end{pmatrix}}, (3.10)
𝒜1=4λ2α1u+2λ(α2u+Iα1(xu))+Iα2(xu)α1(2uuu2x2u),\displaystyle\mathcal{A}_{1}=\small{-4\lambda^{2}\alpha_{1}{\mathit{u^{{\dagger}}}}+2\lambda\left(-\alpha_{2}{\mathit{u^{{\dagger}}}}+\mathrm{I}\alpha_{1}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)\right)+\mathrm{I}\alpha_{2}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)-\alpha_{1}\left(-2{\mathit{u^{{\dagger}}}}{u}{\mathit{u^{{\dagger}}}}-\frac{\partial^{2}}{\partial x^{2}}{\mathit{u^{{\dagger}}}}\right)},
𝒜2=4λ2α1u+2λ(α2u+Iα1(xu))+Iα2(xu)α1(2uuu+2x2u),\displaystyle\mathcal{A}_{2}=4\lambda^{2}\alpha_{1}{u}+2\lambda\left(\alpha_{2}{u}+\mathrm{I}\alpha_{1}\left(\frac{\partial}{\partial x}{u}\right)\right)+\mathrm{I}\alpha_{2}\left(\frac{\partial}{\partial x}{u}\right)-\alpha_{1}\left(2{u}{\mathit{u^{{\dagger}}}}{u}+\frac{\partial^{2}}{\partial x^{2}}{u}\right),
𝒱1=I((2x2u)u+u(2x2u)(xu)(xu))2λ(u(xu)(xu)u),\displaystyle\mathcal{V}_{1}=\mathrm{I}\left(\left(\frac{\partial^{2}}{\partial x^{2}}{u}\right){\mathit{u^{{\dagger}}}}+{u}\left(\frac{\partial^{2}}{\partial x^{2}}{\mathit{u^{{\dagger}}}}\right)-\left(\frac{\partial}{\partial x}{u}\right)\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)\right)-2\lambda\left({u}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)-\left(\frac{\partial}{\partial x}{u}\right){\mathit{u^{{\dagger}}}}\right),
𝒱2=4Iλ2(xu)+3I(uu(xu)+(xu)uu)+I(3x3u)+2λ(2x2u),\displaystyle\mathcal{V}_{2}=-4\,\mathrm{I}\lambda^{2}\left(\frac{\partial}{\partial x}{u}\right)+3\,\mathrm{I}\left({u}{\mathit{u^{{\dagger}}}}\left(\frac{\partial}{\partial x}{u}\right)+\left(\frac{\partial}{\partial x}{u}\right){\mathit{u^{{\dagger}}}}{u}\right)+\mathrm{I}\left(\frac{\partial^{3}}{\partial x^{3}}{u}\right)+2\lambda\left(\frac{\partial^{2}}{\partial x^{2}}{u}\right),
𝒱3=4Iλ2(xu)+3I(uu(xu)+(xu)uu)+I(3x3u)2λ(2x2u),\displaystyle\mathcal{V}_{3}=-4\,\mathrm{I}\lambda^{2}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)+3\,\mathrm{I}\left({\mathit{u^{{\dagger}}}}{u}\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)+\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right){u}{\mathit{u^{{\dagger}}}}\right)+\mathrm{I}\left(\frac{\partial^{3}}{\partial x^{3}}{\mathit{u^{{\dagger}}}}\right)-2\lambda\left(\frac{\partial^{2}}{\partial x^{2}}{\mathit{u^{{\dagger}}}}\right),
𝒱4=I((2x2u)u+u(2x2u)(xu)(xu))+2λ(u(xu)+(xu)u),\displaystyle\mathcal{V}_{4}=-\mathrm{I}\left(\left(\frac{\partial^{2}}{\partial x^{2}}{\mathit{u^{{\dagger}}}}\right){u}+{\mathit{u^{{\dagger}}}}\left(\frac{\partial^{2}}{\partial x^{2}}{u}\right)-\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right)\left(\frac{\partial}{\partial x}{u}\right)\right)+2\lambda\left(-{\mathit{u^{{\dagger}}}}\left(\frac{\partial}{\partial x}{u}\right)+\left(\frac{\partial}{\partial x}{\mathit{u^{{\dagger}}}}\right){u}\right),

where ρ1=4Iα1λ32Iλ2α2,ρ2=I(2λα1+α2)\rho_{1}=-4\,\mathrm{I}\alpha_{1}\lambda^{3}-2\,\mathrm{I}\lambda^{2}\alpha_{2},\;\rho_{2}=\,\mathrm{I}(2\lambda\alpha_{1}+\alpha_{2}), ρ3=3Iuuuu4Iλ2uu+8Iλ4,ρ4=3Iuuuu+4Iλ2uu8Iλ4\rho_{3}=3\,\mathrm{I}{u}{\mathit{u^{{\dagger}}}}{u}{\mathit{u^{{\dagger}}}}-4\,\mathrm{I}\lambda^{2}{u}{\mathit{u^{{\dagger}}}}+8\,\mathrm{I}\lambda^{4},\;\rho_{4}=-3\,\mathrm{I}{\mathit{u^{{\dagger}}}}{u}{\mathit{u^{{\dagger}}}}{u}+4\,\mathrm{I}\lambda^{2}{\mathit{u^{{\dagger}}}}{u}-8\,\mathrm{I}\lambda^{4}, ρ5=8λ3u+4λuuu,ρ6=8λ3u4λuuu\rho_{5}=-8\lambda^{3}{u}+4\lambda{u}{\mathit{u^{{\dagger}}}}{u},\;\rho_{6}=8\lambda^{3}{\mathit{u^{{\dagger}}}}-4\lambda{\mathit{u^{{\dagger}}}}{u}{\mathit{u^{{\dagger}}}}. The equation of motion for the system can be derived by setting the commutator of Γ\Gamma and Δ\Delta equal to zero and equating the coefficients at λ\lambda

iut+i(uxxx+3(uxuu+uuux))α1+(2uuu+uxx)α2+(uxxxx\displaystyle i{u}_{t}+i\left({u}_{xxx}+3({u}_{x}{\mathit{u^{\dagger}}}{u}+{u}{\mathit{u^{\dagger}}}{u}_{x})\right)\alpha_{1}+\left(2{u}{\mathit{u^{\dagger}}}{u}+{u}_{xx}\right)\alpha_{2}+({u}_{xxxx} (3.11)
+2(uxuxu+uuxux+uuxxu)+4(uxxuu+uuuxx)+6(uxuux+uuuuu))γ=0,\displaystyle+2({u}_{x}{\mathit{u^{\dagger}}}_{x}{u}+{u}{\mathit{u^{\dagger}}}_{x}{u}_{x}+{u}{\mathit{u^{\dagger}}}_{xx}{u})+4({u}_{xx}{\mathit{u^{\dagger}}}{u}+{u}{\mathit{u^{\dagger}}}{u}_{xx})+6({u}_{x}{\mathit{u^{\dagger}}}{u}_{x}+{u}{\mathit{u^{\dagger}}}{u}{\mathit{u^{\dagger}}}{u}))\gamma=0,

where u=u(x,t)u=u(x,\;t) is an nc object, denotes the adjoint (Hermitian conjugate), α1,α2\alpha_{1},\;\alpha_{2} and γ\gamma are real parameters, and λ\lambda is a spectral parameter (real or complex). The equation presented in (3.11) is a non-commutative generalization of the HNLS equation, as given in (1.1). This equation exhibits several interesting properties. For instance, when both α2\alpha_{2} and γ\gamma are zero, it reduces to an nc generalization of the complex modified Korteweg–de Vries (KdV) equation and to the standard modified KdV equation for real-valued uu. Moreover, setting γ\gamma to zero results in an nc generalization of the Hirota equation, while setting α1\alpha_{1} and α2\alpha_{2} to zero simultaneously yields the well-known Lakshmanan–Porsezian–Daniel (LPD) equation. Finally, when α1\alpha_{1} and γ\gamma are set to zero, the equation reduces to an nc generalization of the nonlinear Schrödinger (NLS) equation. After relaxing the noncommutativity condition, Eq. (3.11) corresponds to the commutative counterpart. The spectral problem linked with (1.1) remains the same as that of (3.1) and (3.2), with the exception that uu and u¯\bar{u} are now perceived as commutative functions.

3.1. Quasideterminants

In non-commutative algebra, quasideterminants serve as a replacement for ordinary determinants of matrices. They hold a similar significance in non-commutative algebra as ordinary determinants do in commutative algebra and have found vast applications in the domain of non-commutative integrable systems [19, 21, 22].

The quasideterminant |M|ij|M|_{ij} for i,j=1,,ni,\;j=1,\;...,\;n of an n×nn\times n matrix over a non-commutative ring R, expanded about the matrix mijm_{ij}, is defined as

|M|ij=|Mijcjirijmij|=mijrij(Mij)1cji,\left|M\right|_{ij}=\left|\begin{array}[]{cc}M^{ij}&c^{i}_{j}\\ r^{j}_{i}&\framebox{$m_{ij}$}\end{array}\right|=m_{ij}-r^{j}_{i}\left(M^{ij}\right)^{-1}c^{i}_{j}, (3.12)

where mijm_{ij} is referred to as the expansion point and represents the ijij-th entry of MM, rijr^{j}_{i} denotes the ii-th row of MM without the jj-th entry, cjic^{i}_{j} represents the jj-th column of MM without the ii-th row, and MijM^{ij} is the submatrix of MM obtained by removing the ii-th row and the jj-th column from MM.

Quasideterminants are not merely a generalization of usual commutative determinants but are also related to inverse matrices. The inverse of a matrix M=(m11m12m21m22)M=\left(\begin{smallmatrix}m_{11}&m_{12}\\ m_{21}&m_{22}\end{smallmatrix}\right) is defined as

M1=(|M|ji1)=(|m11m12m21m22|1|m11m12m21m22|1|m11m12m21m22|1|m11m12m21m22|1).\displaystyle M^{-1}=\left(|M|_{ji}^{-1}\right)=\left(\begin{array}[]{cc}\left|\begin{array}[]{cc}\framebox{$m_{11}$}&m_{12}\\ m_{21}&m_{22}\end{array}\right|^{-1}&\left|\begin{array}[]{cc}{m_{11}}&\framebox{$m_{12}$}\\ {m_{21}}&m_{22}\end{array}\right|^{-1}\\ \left|\begin{array}[]{cc}{m_{11}}&m_{12}\\ \framebox{$m_{21}$}&m_{22}\end{array}\right|^{-1}&\left|\begin{array}[]{cc}{m_{11}}&m_{12}\\ m_{21}&\framebox{$m_{22}$}\end{array}\right|^{-1}\end{array}\right). (3.23)

4. Darboux transformation

In this section, a Darboux transformation is introduced for the system of nc-HNLS equation (3.11) through the definition of the Darboux matrix

D(𝒴)=λI𝒴Λ𝒴1,\displaystyle D(\mathcal{Y})=\lambda I-\mathcal{Y}\Lambda\mathcal{Y}^{-1}, (4.1)

and the Lax operators Γ\Gamma and Δ\Delta

Γ=xλ𝒥𝒰,Δ=t𝒱p.\Gamma=\partial_{x}-\lambda\mathcal{J}-\mathcal{U},\quad\Delta=\partial_{t}-\mathcal{B}-\mathcal{V}_{p}. (4.2)

The spectral parameter λ\lambda, which can be real or complex, is incorporated along with the constant q×qq\times q matrix Λ\Lambda, and the nc objects 𝒰,𝒱p\mathcal{U},\;\mathcal{V}_{p}, and \mathcal{B} from (3.9) and (3.10) respectively, are utilized as entries in the Lax operators. Now, consider a function φ=φ(x,t)\varphi=\varphi(x,\;t) that is an eigenfunction of the Lax operators Γ\Gamma and Δ\Delta such that Γ(φ)=0\Gamma(\varphi)=0 and Δ(φ)=0\Delta(\varphi)=0. We can define a new function φ~\tilde{\varphi} using the Darboux matrix D(𝒴)D(\mathcal{Y}):

φ~\displaystyle\tilde{\varphi} :=\displaystyle:= D𝒴(φ),\displaystyle D_{\mathcal{Y}}(\varphi), (4.5)
=\displaystyle= λφ𝒴Λ𝒴1ϕ=|𝒴φ𝒴Λλφ|\displaystyle\lambda\varphi-\mathcal{Y}\Lambda\mathcal{Y}^{-1}\phi=\left|\begin{array}[]{cc}\mathcal{Y}&\varphi\\ \mathcal{Y}\Lambda&\framebox{$\lambda\varphi$}\end{array}\right|

The function φ~\tilde{\varphi} is a generic function of the new operators Γ~=D𝒴ΓD𝒴1\tilde{\Gamma}=D_{\mathcal{Y}}\Gamma D_{\mathcal{Y}}^{-1} and Δ~=D𝒴ΔD𝒴1\tilde{\Delta}=D_{\mathcal{Y}}\Delta D_{\mathcal{Y}}^{-1}. Next, we define 𝒴[1]=𝒴1\mathcal{Y}_{[1]}=\mathcal{Y}_{1} and φ[1]=φ\varphi_{[1]}=\varphi as generic eigenfunctions of Γ[1]=Γ\Gamma_{[1]}=\Gamma and Δ[1]=Δ\Delta_{[1]}=\Delta, where 𝒴1\mathcal{Y}_{1} and φ1\varphi_{1} are 2×22\times 2 matrices. We then define φ[2]=D𝒴[1](φ[1])\varphi_{[2]}=D_{\mathcal{Y}_{[1]}}\left(\varphi_{[1]}\right) and 𝒴[2]=φ[2]|φ𝒴2\mathcal{Y}_{[2]}=\varphi_{[2]}|_{\varphi\rightarrow\mathcal{Y}_{2}} to be eigenfunctions of the new operators Γ[2]=D𝒴[1]Γ[1]D𝒴[1]1{\Gamma}_{[2]}=D_{\mathcal{Y}_{[1]}}{\Gamma}_{[1]}D_{\mathcal{Y}_{[1]}}^{-1} and Δ[2]=D𝒴[1]Δ[1]D𝒴[1]1{\Delta}_{[2]}=D_{\mathcal{Y}_{[1]}}{\Delta}_{[1]}D_{\mathcal{Y}_{[1]}}^{-1}.

We define 𝒴i,i=1,,n\mathcal{Y}_{i},\;i=1,\;...,\;n (where each 𝒴i\mathcal{Y}_{i} is a matrix of size 2×22\times 2) be the set of particular eigenfunctions of Γ\Gamma and Δ\Delta. In the non-commutative case, we consider each entry of 𝒴i\mathcal{Y}_{i} as a matrix. We define the (n+1)(n+1)th eigenfunction as φ[n+1]=D𝒴[n](φ[n])\varphi_{[n+1]}=D_{\mathcal{Y}_{[n]}}\left(\varphi_{[n]}\right), which is a generic eigenfunction of the new operators Γ[n+1]=D𝒴[n]Γ[n]D𝒴[n]1{\Gamma}_{[n+1]}=D_{\mathcal{Y}_{[n]}}{\Gamma}_{[n]}D_{\mathcal{Y}_{[n]}}^{-1}, and Δ[n+1]=D𝒴[n]Δ[n]D𝒴[n]1{\Delta}_{[n+1]}=D_{\mathcal{Y}_{[n]}}{\Delta}_{[n]}D_{\mathcal{Y}_{[n]}}^{-1}. Here, 𝒴[n]\mathcal{Y}_{[n]} is a DT from Γ[n]\Gamma_{[n]} to Γ[n+1]\Gamma_{[n+1]} and Δ[n]\Delta_{[n]} to Δ[n+1]\Delta_{[n+1]}. Thus, the nnth-order DT is given by φ[n+1]=D𝒴[n](φ[n])=λφ[n]𝒴[n]Λn𝒴[n]1φ[n]\varphi_{[n+1]}=D_{\mathcal{Y}_{[n]}}\left(\varphi_{[n]}\right)=\lambda\varphi_{[n]}-\mathcal{Y}_{[n]}\Lambda_{n}\mathcal{Y}_{[n]}^{-1}\varphi_{[n]}, where 𝒴[j]=φ[j]|φ𝒴j\mathcal{Y}_{[j]}=\varphi_{[j]}|_{\varphi\rightarrow\mathcal{Y}_{j}}. Let us define a matrix Ξ=(𝒴1,,𝒴n)\Xi=(\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n}) comprising of the eigenfunctions 𝒴i,i=1,,n\mathcal{Y}_{i},\;i=1,\;...,\;n, and set φ[1]=φ\varphi_{[1]}=\varphi. Then, we can represent the nnth iteration of the DT in quasideterminant form as follows:

φ[n+1]=|ΞφΞ(n1)φ(n1)Ξ(n)φ(n)|,\varphi_{[n+1]}=\left|\begin{array}[]{cc}\Xi&\varphi\\ \vdots&\vdots\\ \Xi^{(n-1)}&\varphi^{(n-1)}\\ \Xi^{(n)}&\framebox{$\varphi^{(n)}$}\end{array}\right|, (4.6)

Here, φ(n)=λnφ\varphi^{(n)}=\lambda^{n}\varphi and Ξ(n)=ΞΛn\Xi^{(n)}=\Xi{\Lambda}^{n}., where each Λi,i=1,,n\Lambda^{i},\;i=1,\;...,\;n, is a constant matrix. Hence, we expressed a quasideterminant formula for φn+1\varphi_{n+1} in terms of the known eigenfunctions 𝒴i,i=1,,n\mathcal{Y}_{i},\;i=1,\;...,\;n and the eigenfunction φ\varphi of the “seed” Lax pair Γ=Γ1\Gamma=\Gamma_{1}, Δ=Δ1\Delta=\Delta_{1}.

5. Quasi-Wronskian solutions

In the upcoming analysis, we will examine how the DT D𝒴=λI𝒴Λ𝒴1D_{\mathcal{Y}}=\lambda I-\mathcal{Y}\Lambda\mathcal{Y}^{-1} affects the Lax operator Γ=Γ1\Gamma=\Gamma_{1}, where 𝒴\mathcal{Y} is an eigenfunction of Γ\Gamma (since Γ(𝒴)=0\Gamma(\mathcal{Y})=0 by definition) and Λ\Lambda is an eigenvalue matrix. It is important to note that the same results apply to the operator Δ=Δ1\Delta=\Delta_{1}. As a result of this transformation, the operator Γ\Gamma is converted to a new operator Γ~=Γ[2]\tilde{\Gamma}=\Gamma_{[2]}, which can be expressed as Γ~=D𝒴ΓD𝒴1\tilde{\Gamma}=D_{\mathcal{Y}}{\Gamma}D_{\mathcal{Y}}^{-1}. By substituting (3.1) and (4.1) into the latter equation and equating the coefficients at λj\lambda^{j}, we obtain two equations,

𝒰[2]𝒰[𝒥,𝒴Λ𝒴1]=0,\displaystyle\mathcal{U}_{[2]}-\mathcal{U}-[\mathcal{J},\;\mathcal{Y}\Lambda\mathcal{Y}^{-1}]=0, (5.1)

and

x(𝒴Λ𝒴1)+[𝒰,𝒴Λ𝒴1]+[𝒥,𝒴Λ𝒴1]𝒴Λ𝒴1=0.-\partial_{x}(\mathcal{Y}\Lambda\mathcal{Y}^{-1})+[\mathcal{U},\;\mathcal{Y}\Lambda\mathcal{Y}^{-1}]+[\mathcal{J},\;\mathcal{Y}\Lambda\mathcal{Y}^{-1}]\mathcal{Y}\Lambda\mathcal{Y}^{-1}=0. (5.2)

To confirm the validity of equation (5.2), we express equation (3.1) using a particular eigenfunction 𝒴\mathcal{Y} as 𝒴x=𝒥𝒴Λ+𝒰𝒴\mathcal{Y}_{x}=\mathcal{J}\mathcal{Y}\Lambda+\mathcal{U}\mathcal{Y}. By utilizing this equation, we can easily check that the condition expressed in (5.2) is satisfied. To simplify the notation, a matrix \mathcal{F} is introduced such that 𝒰=[,𝒥]\mathcal{U}=[\mathcal{F},\;\mathcal{J}]. This equation is satisfied if =12ı˙(0uu0)\mathcal{F}=\frac{1}{2\dot{\imath}}\left(\begin{smallmatrix}0&{u}\\ {u}^{{\dagger}}&0\end{smallmatrix}\right). Then, equation (5.1) with 𝒰=[,𝒥]\mathcal{U}=[\mathcal{F},\;\mathcal{J}] can be used to obtain [2]=𝒴(1)𝒴1\mathcal{F}_{[2]}=\mathcal{F}-\mathcal{Y}^{(1)}\mathcal{Y}^{-1}, where 𝒴(1)\mathcal{Y}^{(1)} is defined as 𝒴Λ\mathcal{Y}\Lambda. After nn repeated applications of the DT D𝒴D_{\mathcal{Y}}, we have

[n+1]=[n]𝒴[n](1)𝒴[n]1=k=1n𝒴[k](1)𝒴[k]1,\displaystyle\mathcal{F}_{[n+1]}=\mathcal{F}_{[n]}-\mathcal{Y}^{(1)}_{[n]}\mathcal{Y}^{-1}_{[n]}=\mathcal{F}-\sum_{k=1}^{n}\mathcal{Y}^{(1)}_{[k]}\mathcal{Y}^{-1}_{[k]}, (5.3)

where [1]=,𝒴[1]=𝒴1=𝒴\mathcal{F}_{[1]}=\mathcal{F},\;\mathcal{Y}_{[1]}=\mathcal{Y}_{1}=\mathcal{Y} and Λ1=Λ\Lambda_{1}=\Lambda.

Because our nc-HNLS equation (3.11) is expressed in terms of uu and uu^{{\dagger}}, it is more appropriate to express the quasi-Wronskian solution in terms of these objects. For this, we express each 𝒴i,i=1,,n\mathcal{Y}_{i},\;i=1,\;...,\;n as a 2×22\times 2 matrix as 𝒴i=(φ2i1φ2iχ2i1χ2i)\mathcal{Y}_{i}=\left(\begin{smallmatrix}\varphi_{2i-1}&\varphi_{2i}\\ \chi_{2i-1}&\chi_{2i}\end{smallmatrix}\right). For φ=φ(x,t)\varphi=\varphi(x,\;t) and χ=χ(x,t)\chi=\chi(x,\;t), we can express [n+1]\mathcal{F}_{[n+1]} as

[n+1]=+(|Ξ^f2n1φ(n)0||Ξ^f2nφ(n)0||Ξ^f2n1χ(n)0||Ξ^f2nχ(n)0|),\mathcal{F}_{[n+1]}=\mathcal{F}+\left(\begin{array}[]{cc}\left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n-1}\\ \varphi^{(n)}&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n}\\ \varphi^{(n)}&\framebox{$0$}\end{array}\right|\\ \\ \left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n-1}\\ \chi^{(n)}&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n}\\ \chi^{(n)}&\framebox{$0$}\end{array}\right|\end{array}\right), (5.4)

where Ξ^=(𝒴j(i1))i,j=1,,n\widehat{\Xi}=(\mathcal{Y}_{j}^{(i-1)})_{i,\;j=1,\;...,\;n} is the 2n×2n2n\times 2n matrix of 𝒴1,,𝒴n\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n}, and f2n1f_{2n-1} and f2nf_{2n} are the column vectors 2n×12n\times 1 with a 1 in the (2n1)(2n-1)th and (2n)(2n)th row, respectively, and zeros elsewhere, while φ(n)=(φ1(n),,φ2n(n)),χ(n)=(χ1(n),,χ2n(n))\varphi^{(n)}=\left(\varphi_{1}^{(n)},\;...,\;\varphi_{2n}^{(n)}\right),\;\chi^{(n)}=\left(\chi_{1}^{(n)},\;...,\;\chi_{2n}^{(n)}\right) denote the 1×2n1\times 2n row vectors. We can express the quasi-Wronskian solutions for uu and uu^{{\dagger}} by utilizing =12ı˙(0uu0)\mathcal{F}=\frac{1}{2\dot{\imath}}\left(\begin{smallmatrix}0&u\\ u^{\dagger}&0\\ \end{smallmatrix}\right) which leads to:

u[n+1]=u+2ı˙|Ξ^f2nφ(n)0|,u[n+1]=u+2ı˙|Ξ^f2n1χ(n)0|.u_{[n+1]}=u+2\dot{\imath}\left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n}\\ \varphi^{(n)}&\framebox{$0$}\end{array}\right|,\qquad u_{[n+1]}^{\dagger}=u^{\dagger}+2\dot{\imath}\left|\begin{array}[]{cc}\widehat{\Xi}&f_{2n-1}\\ \chi^{(n)}&\framebox{$0$}\end{array}\right|. (5.5)

We proceed in the next section to construct the binary DT for the nc-HNLS equation, using a strategy similar to that employed in [9].

6. Binary Darboux transformation

We introduce 𝒴1,,𝒴n\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n} as eigenfunctions of the Lax operators Γ\Gamma and Δ\Delta, and 𝒵1,,𝒵n\mathcal{Z}_{1},\;...,\;\mathcal{Z}_{n} as eigenfunctions of the adjoint Lax operators Γ\Gamma^{{\dagger}} and Δ\Delta^{\dagger}. Assuming φ[1]=φ\varphi_{[1]}=\varphi to be a generic eigenfunction of the Lax operators Γ\Gamma and Δ\Delta, and ψ[1]=ψ\psi_{[1]}=\psi to be a generic eigenfunction of the adjoint Lax operators Γ\Gamma^{{\dagger}} and Δ\Delta^{{\dagger}}, we define the binary DT and its adjoint as

D𝒴[1],𝒵[1]=I𝒴[1]Υ(𝒴[1],𝒵[1])1Ω𝒵[1],D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}=I-\mathcal{Y}_{[1]}\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})^{-1}\Omega^{-{\dagger}}\mathcal{Z}_{[1]}^{{\dagger}}, (6.1)
D𝒴[1],𝒵[1]=I𝒵[1]Υ(𝒴[1],𝒵[1])Λ𝒴[1].D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}^{{\dagger}}=I-\mathcal{Z}_{[1]}\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})^{-{\dagger}}\Lambda^{-{\dagger}}\mathcal{Y}_{[1]}^{{\dagger}}. (6.2)

In the context of the binary DT, we use 𝒴[1]=𝒴1\mathcal{Y}_{[1]}=\mathcal{Y}_{1} as the initial eigenfunction that characterizes the transformation from the Lax operators Γ\Gamma and Δ\Delta to the new operators Γ~\tilde{\Gamma} and Δ~\tilde{\Delta}. Similarly, we define 𝒵[1]=𝒵1\mathcal{Z}_{[1]}=\mathcal{Z}_{1} to represent the adjoint transformation, where the potential Υ\Upsilon

ΩΥ(𝒴[1],𝒵[1])+Υ(𝒴[1],𝒵[1])Λ=𝒵[1]𝒴[1],\displaystyle\Omega^{{\dagger}}\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})+\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})\Lambda=\mathcal{Z}_{[1]}^{{\dagger}}\mathcal{Y}_{[1]}, (6.3)
(λI+Ω)Υ(φ[1],𝒵[1])=𝒵[1]φ[1],\displaystyle\left(\lambda I+\Omega^{{\dagger}}\right)\Upsilon(\varphi_{[1]},\;\mathcal{Z}_{[1]})=\mathcal{Z}_{[1]}^{{\dagger}}\varphi_{[1]}, (6.4)
Υ(𝒴[1],ψ[1])(μI+Λ)=ψ[1]𝒴[1].\displaystyle\Upsilon(\mathcal{Y}_{[1]},\;\psi_{[1]})\left(\mu^{{\dagger}}I+\Lambda\right)=\psi_{[1]}^{{\dagger}}\mathcal{Y}_{[1]}. (6.5)

The transformed operators Γ~=Γ[2]\widetilde{\Gamma}=\Gamma_{[2]} and Δ~=Δ[2]\widetilde{\Delta}={\Delta}_{[2]} is defined as

Γ[2]\displaystyle{\Gamma}_{[2]} =\displaystyle= D𝒴[1],𝒵[1]Γ[1]D𝒴[1],𝒵[1]1,\displaystyle D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}{\Gamma}_{[1]}D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}^{-1}, (6.6)
Δ[2]\displaystyle{\Delta}_{[2]} =\displaystyle= D𝒴[1],𝒵[1]Δ[1]D𝒴[1],𝒵[1]1,\displaystyle D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}{\Delta}_{[1]}D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}^{-1}, (6.7)

with generic eigenfunction

φ[2]\displaystyle\varphi_{[2]} :=\displaystyle:= D𝒴[1],𝒵[1](φ[1]),\displaystyle D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}(\varphi_{[1]}), (6.8)
=\displaystyle= φ[1]𝒴[1]Υ(𝒴[1],𝒵[1])1(I+λIΩ)Υ(φ[1],𝒵[1]).\displaystyle\varphi_{[1]}-\mathcal{Y}_{[1]}\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})^{-1}(I+\lambda I\Omega^{-\dagger})\Upsilon(\varphi_{[1]},\;\mathcal{Z}_{[1]}).

and with generic adjoint eigenfunction,

ψ[2]\displaystyle\psi_{[2]} :=\displaystyle:= D𝒴[1],𝒵[1](ψ[1]),\displaystyle D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}^{-{\dagger}}(\psi_{[1]}), (6.9)
=\displaystyle= ψ[1]𝒵[1]Υ(𝒴[1],𝒵[1])(I+μIΛ)Υ(𝒴[1],ψ[1]).\displaystyle\psi_{[1]}-\mathcal{Z}_{[1]}\Upsilon(\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]})^{-{\dagger}}(I+\mu I\Lambda^{-\dagger})\Upsilon(\mathcal{Y}_{[1]},\;\psi_{[1]})^{{\dagger}}.

For the nth iteration of the binary DT, we choose the eigenfunction 𝒴[n]\mathcal{Y}_{[n]} that defines the transformation from Γ[n]\Gamma_{[n]}, Δ[n]\Delta_{[n]} to Γ[n+1]\Gamma_{[n+1]}, Δ[n+1]\Delta_{[n+1]}. Similarly, we choose the eigenfunction 𝒵[n]\mathcal{Z}_{[n]} for the adjoint transformation from Γ[n]\Gamma_{[n]}^{\dagger}, Δ[n]\Delta_{[n]}^{\dagger} to Γ[n+1]\Gamma_{[n+1]}^{\dagger}, Δ[n+1]\Delta_{[n+1]}^{\dagger}. The Lax operators Γ[n]\Gamma_{[n]} and Δ[n]\Delta_{[n]} exhibit covariance under the binary Darboux transformation.

D𝒴[n],𝒵[n]=I𝒴[n]Υ(𝒴[n],𝒵[n])1Ω𝒵[n],D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}=I-\mathcal{Y}_{[n]}\Upsilon(\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]})^{-1}\Omega^{-{\dagger}}\mathcal{Z}_{[n]}^{\dagger}, (6.10)

and adjoint binary DT

D𝒴[n],𝒵[n]=I𝒵[n]Υ(𝒴[n],𝒵[n])1Λ𝒴[n].D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}^{-{\dagger}}=I-\mathcal{Z}_{[n]}\Upsilon(\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]})^{-1}\Lambda^{-{\dagger}}\mathcal{Y}_{[n]}^{\dagger}. (6.11)

The transformed operators

Γ[n+1]\displaystyle{\Gamma}_{[n+1]} =\displaystyle= D𝒴[n],𝒵[n]Γ[n]D𝒴[n],𝒵[n]1,\displaystyle D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}{\Gamma}_{[n]}D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}^{-1}, (6.12)
Δ[n+1]\displaystyle{\Delta}_{[n+1]} =\displaystyle= D𝒴[n],𝒵[n]Δ[n]D𝒴[n],𝒵[n]1,\displaystyle D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}{\Delta}_{[n]}D_{\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]}}^{-1}, (6.13)

have generic eigenfunction

φ[n+1]=φ[n]𝒴[n]Υ(𝒴[n],𝒵[n])1(I+λIΩ)Υ(φ[n],𝒵[n]),\varphi_{[n+1]}=\varphi_{[n]}-\mathcal{Y}_{[n]}\Upsilon(\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]})^{-1}(I+\lambda I\Omega^{-{\dagger}})\Upsilon(\varphi_{[n]},\;\mathcal{Z}_{[n]}), (6.14)

and generic adjoint eigenfunction;

ψ[n+1]=ψ[n]𝒵[n]Υ(𝒴[n],𝒵[n])1(I+μIΛ)Υ(𝒴[n],ψ[n]).\psi_{[n+1]}=\psi_{[n]}-\mathcal{Z}_{[n]}\Upsilon(\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]})^{-1}(I+\mu I\Lambda^{-{\dagger}})\Upsilon(\mathcal{Y}_{[n]},\;\psi_{[n]})^{\dagger}. (6.15)

By introducing the matrices Ξ=(𝒴1,,𝒴n)\Xi=(\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n}) and Z=(𝒵1,,𝒵n)Z=(\mathcal{Z}_{1},\;...,\;\mathcal{Z}_{n}), we can represent these findings in the framework of quasi-Gramians, yielding the following expression:

φ[n+1]=|Υ(Ξ,Z)(I+λIΩ^)Υ(φ,Z)Ξφ|,ψ[n+1]=|Υ(Ξ,Z)(I+μIΛ^)Υ(Ξ,ψ)Zψ|,\varphi_{[n+1]}=\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&(I+\lambda I\hat{\Omega}^{-\dagger})\Upsilon(\varphi,\;Z)\\ \Xi&\framebox{$\varphi$}\end{array}\right|,\quad\psi_{[n+1]}=\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)^{{\dagger}}&(I+\mu I\hat{\Lambda}^{-\dagger})\Upsilon(\Xi,\;\psi)^{{\dagger}}\\ Z&\framebox{$\psi$}\end{array}\right|, (6.16)

where Ω^=diag(Ω,,Ω),Λ^=diag(Λ,,Λ)\hat{\Omega}=\text{diag}(\Omega,\;...,\;\Omega),\;\hat{\Lambda}=\text{diag}(\Lambda,\;...,\;\Lambda), with both Ω\Omega and Λ\Lambda being 2 × 2 matrices.

7. Quasi-Gramian solutions

In this section, we now determine the effect of binary DT D𝒴,𝒵=IξΥ(𝒴,𝒵)1Ω𝒵D_{\mathcal{Y},\;\mathcal{Z}}=I-\xi\Upsilon(\mathcal{Y},\;\mathcal{Z})^{-1}\Omega^{-{\dagger}}\mathcal{Z}^{{\dagger}} on the Lax operator Γ\Gamma, with 𝒴1,,𝒴n\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n} being eigenfunctions of Γ\Gamma. Similarly, let 𝒵1,,𝒵n\mathcal{Z}_{1},\;...,\;\mathcal{Z}_{n} denote the eigenfunctions of the adjoint Lax operator Γ{\Gamma}^{{\dagger}}. The same results apply to the operators Δ{\Delta} and Δ{\Delta}^{{\dagger}}.

As the binary DT D𝒴[1],𝒵[1]=D𝒴,𝒵D_{\mathcal{Y}_{[1]},\;\mathcal{Z}_{[1]}}=D_{\mathcal{Y},\;\mathcal{Z}} combines two ordinary DTs, D𝒴[1]=D𝒴D_{\mathcal{Y}_{[1]}}=D_{\mathcal{Y}} and D𝒴^[1]=D𝒴^D_{\hat{\mathcal{Y}}_{[1]}}=D_{\hat{\mathcal{Y}}}, it follows that the Lax operator Γ\Gamma is transformed into a new Lax operator Γ^\widehat{\Gamma} under the binary DT, given by:

Γ^=D𝒴^ΓD𝒴^1,\widehat{\Gamma}={D_{\widehat{\mathcal{Y}}}}{\Gamma}{D_{\widehat{\mathcal{Y}}}^{-1}}, (7.1)

we get

𝒰^=𝒰+[𝒥,𝒴Υ(𝒴,𝒵)1𝒵].\widehat{\mathcal{U}}=\mathcal{U}+[\mathcal{J},\;\mathcal{Y}\Upsilon(\mathcal{Y},\;\mathcal{Z})^{-1}\mathcal{Z}^{{\dagger}}]. (7.2)

Since 𝒰=[,𝒥]\mathcal{U}=[\mathcal{F},\;\mathcal{J}], so that

^=𝒴Υ(𝒴,𝒵)1𝒵.\widehat{\mathcal{F}}=\mathcal{F}-\mathcal{Y}\Upsilon(\mathcal{Y},\;\mathcal{Z})^{-1}\mathcal{Z}^{{\dagger}}. (7.3)

After nn iterations of applying the binary DT D𝒴,𝒵D_{\mathcal{Y},\;\mathcal{Z}}, the resulting expression is given by:

[n+1]=[n]𝒴[n]Υ(𝒴[n],𝒵[n])1𝒵[n]=i=1n𝒴[i]Υ(𝒴[i],𝒵[i])1𝒵[i].\displaystyle\mathcal{F}_{[n+1]}=\mathcal{F}_{[n]}-\mathcal{Y}_{[n]}\Upsilon(\mathcal{Y}_{[n]},\;\mathcal{Z}_{[n]})^{-1}\mathcal{Z}_{[n]}^{{\dagger}}=\mathcal{F}-\sum_{i=1}^{n}\mathcal{Y}_{[i]}\Upsilon(\mathcal{Y}_{[i]},\;\mathcal{Z}_{[i]})^{-1}\mathcal{Z}_{[i]}^{{\dagger}}. (7.4)

Using the notation [1]=,[2]=^,𝒴[1]=𝒴1\mathcal{F}_{[1]}=\mathcal{F},\;\mathcal{F}_{[2]}=\widehat{\mathcal{F}},\;\mathcal{Y}_{[1]}=\mathcal{Y}_{1} and 𝒵[1]=𝒵1\mathcal{Z}_{[1]}=\mathcal{Z}_{1}, and introducing the matrices Ξ=(𝒴1,,𝒴n)\Xi=(\mathcal{Y}_{1},\;...,\;\mathcal{Y}_{n}) and Z=(𝒵1,,𝒵n)Z=(\mathcal{Z}_{1},\;...,\;\mathcal{Z}_{n}), the result (7.4) can be expressed in the form of quasi-Gramian as follows:

[n+1]=+|Υ(Ξ,Z)ZΞ02|,\mathcal{F}_{[n+1]}=\mathcal{F}+\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Z^{{\dagger}}\\ \Xi&\framebox{$0_{2}$}\end{array}\right|, (7.5)

It is worth noting that each 𝒴i,𝒵i,i=1,,n\mathcal{Y}_{i},\;\mathcal{Z}_{i},\;i=1,\;...,\;n is a 2×22\times 2 matrix. Given that our system of the nc-HNLS equation is formulated in terms of non-commutative objects u,uu,\;u^{{\dagger}}, we find it more appropriate to express the quasi-Gramian solution (7.5) in terms of these objects. Thus, we introduce the matrices 𝒴i\mathcal{Y}_{i} following a similar approach as in the quasi-Wronskian case. We also define Z=ΞQZ=\Xi Q^{{\dagger}}, where QQ represents a constant matrix of size 2n×2n2n\times 2n and denotes the Hermitian conjugate. It is noted that Ξ\Xi and ZZ adhere to the same dispersion relation and remain unchanged when multiplied by a constant matrix. Consequently, the quasi-Gramian solution (7.5) can also be represented as follows:

[n+1]=+(|Υ(Ξ,Z)Qφφ0||Υ(Ξ,Z)Qχφ0||Υ(Ξ,Z)Qφχ0||Υ(Ξ,Z)Qχχ0|),\mathcal{F}_{[n+1]}=\mathcal{F}+\left(\begin{array}[]{cc}\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\varphi^{{\dagger}}\\ \varphi&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\chi^{{\dagger}}\\ \varphi&\framebox{$0$}\end{array}\right|\\ \\ \left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\varphi^{{\dagger}}\\ \chi&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\chi^{{\dagger}}\\ \chi&\framebox{$0$}\end{array}\right|\end{array}\right), (7.6)

where φ=(φ1,,φn)\varphi=(\varphi_{1},\;...,\;\varphi_{n}) and χ=(χ1,,χn)\chi=(\chi_{1},\;...,\;\chi_{n}) are row vectors. Thus, quasi-Gramian expression for uu and uu^{{\dagger}} are given by

u[n+1]=u+2ı˙|Υ(Ξ,Z)Qχφ0|,u[n+1]=u+2ı˙|Υ(Ξ,Z)Qφχ0|.\displaystyle u_{[n+1]}=u+2\dot{\imath}\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\chi^{{\dagger}}\\ \varphi&\framebox{$0$}\end{array}\right|,\quad u_{[n+1]}^{{\dagger}}=u^{{\dagger}}+2\dot{\imath}\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\varphi^{{\dagger}}\\ \chi&\framebox{$0$}\end{array}\right|. (7.11)

Equation (7.11) represents the quasi-Gramian solutions for the nc-HNLS equation (3.11). If we relax the non-commutativity condition, the equation can be simplified and expressed as a ratio of simple Gramians. In the limit of commutativity, we obtain the following expression

u[n+1]=u2ı˙|Υ(Ξ,Z)Qχϕ0||Δ(Ξ,Z)|,u¯[n+1]=u¯2ı˙|Υ(Ξ,Z)Qφχ0||Δ(Ξ,Z)|.\displaystyle u_{[n+1]}=u-2\dot{\imath}\frac{\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\chi^{{\dagger}}\\ \phi&0\end{array}\right|}{\left|\Delta(\Xi,\;Z)\right|},\quad\bar{u}_{[n+1]}=\bar{u}-2\dot{\imath}\frac{\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\varphi^{{\dagger}}\\ \chi&0\end{array}\right|}{\left|\Delta(\Xi,\;Z)\right|}. (7.16)

These expressions define the Gramian solutions of the higher-order NLS equation.

8. Explicit solutions

When u=0u=0, spectral problem (3.1) and (3.2) has the solution

φj=eı˙ζj,χj=eı˙ζj,ζj=λjx+2λj2(4γλj22λjα1α2)t,\varphi_{j}=e^{\dot{\imath}\zeta_{j}},\;\quad\chi_{j}=e^{-\dot{\imath}\zeta_{j}},\quad\zeta_{j}=-\lambda_{j}x+2\lambda_{j}^{2}(4\gamma\,\lambda_{j}^{2}-2\lambda_{j}\alpha_{1}-\alpha_{2})t, (8.1)

To simplify the notation and work with only φ1,,φn\varphi_{1},\ldots,\varphi_{n} and χ1,,χn\chi_{1},\ldots,\chi_{n}, we introduce the following relabeling. We redefine φi\varphi_{i} as φi+12\varphi_{\frac{i+1}{2}} for odd values of ii (i.e., i=1,,2n1i=1,\ldots,2n-1), and set φi=0\varphi_{i}=0 for even values of ii (i.e., i=2,4,,2ni=2,4,\ldots,2n). Similarly, we relabel χi\chi_{i} as χi2\chi_{\frac{i}{2}} for even values of ii, and χi=0\chi_{i}=0 for odd values of ii. We then have

φ=(φ1, 0,φ2, 0,,φn, 0),χ=(0,χ1, 0,χ2,, 0,χn),\varphi=(\varphi_{1},\;0,\;\varphi_{2},\;0,\;...,\;\varphi_{n},\;0),\quad\chi=(0,\;\chi_{1},\;0,\;\chi_{2},\;...,\;0,\;\chi_{n}), (8.2)

Using the notation 𝒴i=diag(φi,χi)\mathcal{Y}_{i}=\text{diag}(\varphi_{i},\chi_{i}) for i=1,,ni=1,\ldots,n, where φi\varphi_{i} and χi\chi_{i} are given in (8.1), we now focus on the commutative case. In this case, we can express the Gramian solution (7.16) as follows.

u[n+1]=2ı˙|Υ(Ξ,Z)Qχφ0||Υ(Ξ,Z)|=2ı˙𝒢,sayu_{[n+1]}=-2\dot{\imath}\frac{\left|\begin{array}[]{cc}\Upsilon(\Xi,\;Z)&Q\chi^{{\dagger}}\\ \varphi&0\end{array}\right|}{\left|\Upsilon(\Xi,\;Z)\right|}=-2\dot{\imath}\frac{\mathcal{G}}{\mathcal{R}},\;\text{say} (8.3)

where =det(𝒲)\mathcal{R}=\text{det}(\mathcal{W}) and

𝒲=Υ(Ξ,Z)=Z𝒥Ξ𝑑x+I2n,\mathcal{W}=\Upsilon(\Xi,\;Z)=\int Z^{{\dagger}}\mathcal{J}\Xi dx+I_{2n}, (8.4)

Here, I2nI_{2n} is the identity matrix with dimensions 2n×2n2n\times 2n. Constructing the matrix Ξ\Xi involves arranging the eigenfunctions 𝒴1,𝒴2,,𝒴n\mathcal{Y}_{1},\;\mathcal{Y}_{2},\;...,\;\mathcal{Y}_{n}, where each 𝒴i\mathcal{Y}_{i} represents an eigenfunction of the Lax operators Γ\Gamma and Δ\Delta, presented as a 2×22\times 2 matrix. Similarly, assembling the matrix ZZ involves the eigenfunctions 𝒵1,𝒵2,,𝒵n\mathcal{Z}_{1},\;\mathcal{Z}_{2},\;...,\;\mathcal{Z}_{n}, with 𝒵i\mathcal{Z}_{i} serving as eigenfunctions of the adjoint Lax operators Γ\Gamma^{\dagger} and Δ\Delta^{\dagger}. The matrix Υ(Ξ,Z)\Upsilon(\Xi,\;Z) is a matrix 2n×2n2n\times 2n, its entries being scalar components (1×1)(1\times 1). As we proceed to discuss the non-commutative case, we will consider every component of 𝒴i\mathcal{Y}_{i} and 𝒵i\mathcal{Z}_{i} as a matrix. Presenting the matrix QQ, a constant matrix 2n×2n2n\times 2n, we define ZZ as the outcome of multiplying Ξ\Xi by the Hermitian adjoint of QQ, denoted as Z=ΞQZ=\Xi Q^{\dagger}, allowing us to express the equation (8.4) as

𝒲=QΞ𝒥Ξ𝑑x+I2n=QΘ+I2n,\mathcal{W}=Q\int\Xi^{{\dagger}}\mathcal{J}\Xi dx+I_{2n}=Q\Theta+I_{2n}, (8.5)

where

Θ=(ı˙xφ1φ1𝑑x02ı˙xφ1φn𝑑x0202ı˙xχ1χ1𝑑x02ı˙χ1χn𝑑xı˙xφnφ1𝑑x02ı˙xφnφn𝑑x0202ı˙xχnχ1𝑑x02ı˙xχnχn𝑑x).\Theta=\left(\begin{array}[]{ccccc}-\dot{\imath}\int^{x}\varphi_{1}^{\ast}\varphi_{1}dx&0_{2}&\ldots&-\dot{\imath}\int^{x}\varphi_{1}^{\ast}\varphi_{n}dx&0_{2}\\ 0_{2}&\dot{\imath}\int^{x}\chi_{1}^{\ast}\chi_{1}dx&\ldots&0_{2}&\dot{\imath}\int\chi_{1}^{\ast}\chi_{n}dx\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ -\dot{\imath}\int^{x}\varphi_{n}^{\ast}\varphi_{1}dx&0_{2}&\ldots&-\dot{\imath}\int^{x}\varphi_{n}^{\ast}\varphi_{n}dx&0_{2}\\ 0_{2}&\dot{\imath}\int^{x}\chi_{n}^{\ast}\chi_{1}dx&\ldots&0_{2}&\dot{\imath}\int^{x}\chi_{n}^{\ast}\chi_{n}dx\\ \end{array}\right). (8.6)

To derive one soliton (n=1,u[2]u1n=1,\;u_{[2]}\equiv u_{1}) solution for the commutative Hirota equation. We choose Q=(q1q2q2q1)Q=\left(\begin{smallmatrix}q_{1}&q_{2}\\ q_{2}&q_{1}\end{smallmatrix}\right), one soliton solution is given by

u1=8Ic1q2λI2e2Iξ14IλIc1q1cosh(ξ2)4λI2c12+q12q22u_{1}=\frac{8\,\mathrm{I}c_{1}q_{\mathit{2}}\lambda_{I}^{2}{\mathrm{e}}^{2\,\mathrm{I}\xi_{1}}}{4\,\mathrm{I}\lambda_{I}c_{1}q_{\mathit{1}}\cosh\!\left(\xi_{2}\right)-4\lambda_{I}^{2}c_{1}^{2}+q_{\mathit{1}}^{2}-q_{\mathit{2}}^{2}} (8.7)

where

ξ1=[8(λR46λR2λI2+λI4)γ+2(λR2+λI2)α2+4(λR3+3λRλI2)α1]tλRx,\displaystyle\xi_{1}=[8\left(\lambda_{R}^{4}-6\lambda_{R}^{2}\lambda_{I}^{2}+\lambda_{I}^{4}\right)\gamma+2\left(-\lambda_{R}^{2}+\lambda_{I}^{2}\right)\alpha_{2}+4\left(-\lambda_{R}^{3}+3\lambda_{R}\lambda_{I}^{2}\right)\alpha_{1}]t-\lambda_{R}x, (8.8)
ξ2=8[8(λR3+λRλI2)γ+λRα2+(λI2+3λR2)α1]λIt+2xλI,\displaystyle\xi_{2}=8[8\left(-\lambda_{R}^{3}+\lambda_{R}\lambda_{I}^{2}\right)\gamma+\lambda_{R}\alpha_{2}+\left(-\lambda_{I}^{2}+3\lambda_{R}^{2}\right)\alpha_{1}]\lambda_{I}t+2x\lambda_{I},

To visualize this soliton solution, we plot the propagation of the u1u_{1} soliton in the commutative case with a velocity of 8(λR3+λRλI2)γ+λRα2+(λI2+3λR2)α18\left(-\lambda_{R}^{3}+\lambda_{R}\lambda_{I}^{2}\right)\gamma+\lambda_{R}\alpha_{2}+\left(-\lambda_{I}^{2}+3\lambda_{R}^{2}\right)\alpha_{1}, where λ=λR+iλI\lambda=\lambda_{R}+i\lambda_{I}. Fig. 2 illustrates the behavior of the soliton over time.

Refer to caption
(a)
Refer to caption
(b)
Figure 2. Evolution of the solution (8.7) with parameters α1=1.5,α2=γ=1,c1=0.5,q1=2,q2=1,(a)λ=0.1+0.5ı˙,(b)λ=0.5ı˙\alpha_{1}=1.5,\;\alpha_{2}=\gamma=1,\;c_{1}=0.5,\;q_{1}=2,\;q_{2}=-1,\;(a)\;\lambda=0.1+0.5\dot{\imath},\;(b)\;\lambda=0.5\dot{\imath}.

8.1. Noncommutative case

We now discuss the noncommutative case. It has been shown in [10] that the behavior of matrix solitons differs from their scalar counterparts. Unlike scalar solitons, which maintain their amplitudes unchanged during interactions, matrix solitons undergo transformations that depend on certain rules. These transformations affect the amplitudes, which are determined by vectors rather than individual values in the noncommutative setting. When considering the case of n=1n=1, we choose the solutions φ\varphi and χ\chi of the Lax pair to be 2×22\times 2 matrices, given by

φj=eı˙ζjI2,χj=eı˙ζjI2,ζj=λjx+2λj2(4γλj22λjα1α2)t,\varphi_{j}=e^{\dot{\imath}\zeta_{j}}I_{2},\;\quad\chi_{j}=e^{-\dot{\imath}\zeta_{j}}I_{2},\quad\zeta_{j}=-\lambda_{j}x+2\lambda_{j}^{2}(4\gamma\,\lambda_{j}^{2}-2\lambda_{j}\alpha_{1}-\alpha_{2})t, (8.9)

Each entry in φ\varphi and χ\chi and constant matrix QQ is a 2×22\times 2 matrix, so that these matrices are given by

φ=(φ10000φ100),χ=(00χ10000χ1),Q=(q11q12q13q14q12q11q14q13q13q14q33q34q14q13q34q33),\displaystyle\varphi=\left(\begin{array}[]{cccc}\varphi_{1}&0&0&0\\ 0&\varphi_{1}&0&0\end{array}\right),\;\chi=\left(\begin{array}[]{cccc}0&0&\chi_{1}&0\\ 0&0&0&\chi_{1}\end{array}\right),\;Q=\left(\begin{array}[]{cccc}q_{11}&q_{12}&q_{13}&q_{14}\\ q_{12}&q_{11}&q_{14}&q_{13}\\ q_{13}&q_{14}&q_{33}&q_{34}\\ q_{14}&q_{13}&q_{34}&q_{33}\end{array}\right), (8.18)

Therefore, the quasi-Gramian expression for u2u_{2} (which we will now denote as u1u^{1} for the non-commutative case) can be expressed as follows

u1\displaystyle u^{1} =\displaystyle= 2ı˙(|Υ(Ξ,Z)χ11φ10000||Υ(Ξ,Z)χ12φ10000||Υ(Ξ,Z)χ110φ1000||Υ(Ξ,Z)χ120φ1000|),\displaystyle{2\dot{\imath}}\left(\begin{array}[]{cc}\left|\begin{array}[]{ccccc}&&&&\\ &&\Upsilon(\Xi,\;Z)&&\chi^{11}\\ &&&&\\ \varphi_{1}&0&0&0&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{ccccc}&&&&\\ &&\Upsilon(\Xi,\;Z)&&\chi^{12}\\ &&&&\\ \varphi_{1}&0&0&0&\framebox{$0$}\end{array}\right|\\ \left|\begin{array}[]{ccccc}&&&&\\ &&\Upsilon(\Xi,\;Z)&&\chi^{11}\\ &&&&\\ 0&\varphi_{1}&0&0&\framebox{$0$}\end{array}\right|&\left|\begin{array}[]{ccccc}&&&&\\ &&\Upsilon(\Xi,\;Z)&&\chi^{12}\\ &&&&\\ 0&\varphi_{1}&0&0&\framebox{$0$}\end{array}\right|\end{array}\right), (8.37)
=\displaystyle= 2ı˙(u11u12u21u22),say,\displaystyle{2\dot{\imath}}\left(\begin{array}[]{cc}u_{11}&u_{12}\\ u_{21}&u_{22}\end{array}\right),\;\text{say}, (8.40)

where χ11=(q13χ1q14χ1q33χ1q34χ1),χ12=(q14χ1q13χ1q34χ1q33χ1)\chi^{11}=(q_{13}\chi_{1}^{\ast}\;q_{14}\chi_{1}^{\ast}\;q_{33}\chi_{1}^{\ast}\;q_{34}\chi_{1}^{\ast})^{{\dagger}},\;\chi^{12}=(q_{14}\chi_{1}^{\ast}\;q_{13}\chi_{1}^{\ast}\;q_{34}\chi_{1}^{\ast}\;q_{33}\chi_{1}^{\ast})^{{\dagger}} and Υ\Upsilon is the potential defined in (8.4) with each entry being a 2×22\times 2 matrix.

In the context of non-commutative systems, the soliton solution (8.37) is intricately influenced by both the spectral parameter λ\lambda and the elements composing the matrix QQ. When specific entries, such as q13q_{13} and q14q_{14}, are deliberately set to zero, the resulting solutions for u11u_{11}, u12u_{12}, u21u_{21}, and u22u_{22} appear trivial. And, where q13=q14=2q_{13}=q_{14}=-2, the graphical representations of solutions u11u_{11}, u12u_{12}, u21u_{21}, and u22u_{22} manifesting as a single consolidated plot instead of the originally intended four. Noteworthy is the fact that under this condition, all solitons propagate with a consistent amplitude of 0.21630.2163 units. Unlike this symmetry, when q13q14q_{13}\neq q_{14}, the resulting graphs exhibit a variety of double- and single-peaked patterns for each component of the matrix u1u^{1} (see Fig. 4). It is worth highlighting that solitons u12u_{12} and u21u_{21} advance with an amplitude of 0.3860 units of large peak and 0.2317 units of small peak, while the solitons u11u_{11} and u22u_{22} display amplitude of 1.9797 units. Additionally, an intriguing situation unfolds when we choose q12=q14=2q_{12}=-q_{14}=-2 and q11=q13=q33=q34=0q_{11}=q_{13}=q_{33}=q_{34}=0. Under these conditions, we notice a consistent single-peaked soliton for u11u_{11} and u22u_{22} with an amplitude of 0.2214 units, while simultaneously observing a kink pattern in u12u_{12} and u21u_{21} of maximum height 0.8858 units (as seen in Fig. 5).

Refer to caption
Figure 3. Evolution of the solution (8.37) with parameters α1=1.5,α2=γ=1,c1=0.5,q11=0.5,q12=0,q13=q14=1,q33=0.2,q34=0.1,λ=0.1+0.5ı˙\alpha_{1}=1.5,\;\alpha_{2}=\gamma=1,\;c_{1}=0.5,\;q_{11}=0.5,\;q_{12}=0,\;q_{13}=q_{14}=-1,\;q_{33}=-0.2,\;q_{34}=-0.1,\;\lambda=-0.1+0.5\dot{\imath}.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 4. Evolution of the solution (8.37) with parameters α1=1.5,α2=γ=1,c1=0.5,q11=0.5,q12=0.2,q13=1,q14=0.6,q33=0,q34=0.1,λ=0.1+0.5ı˙\alpha_{1}=1.5,\;\alpha_{2}=\gamma=1,\;c_{1}=0.5,\;q_{11}=0.5,\;q_{12}=-0.2,\;q_{13}=-1,\;q_{14}=-0.6,\;q_{33}=0,\;q_{34}=0.1,\;\lambda=-0.1+0.5\dot{\imath}.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 5. Evolution of the solution (8.37) with parameters α1=1.5,α2=γ=1,c1=0.5,q11=0,q12=2,q13=0,q14=2,q33=0,q34=0,λ=0.1+0.5ı˙\alpha_{1}=1.5,\;\alpha_{2}=\gamma=1,\;c_{1}=0.5,\;q_{11}=0,\;q_{12}=-2,\;q_{13}=0,\;q_{14}=2,\;q_{33}=0,\;q_{34}=0,\;\lambda=-0.1+0.5\dot{\imath}.

The solution (8.37) includes several particular cases. In the case where both α2\alpha_{2} and γ\gamma are zero, the solution (8.37) takes a different form. It becomes a solution of a non-commutative generalization of the complex modified Korteweg-de Vries (KdV) equation and is further reduced to the standard modified KdV equation when the variable uu is real-valued. Additionally, if we set γ\gamma to zero, we obtain the solution of the non-commutative extension of the Hirota equation. On the other hand, when we simultaneously set α1\alpha_{1} and α2\alpha_{2} to zero, we have the solution of the non-commutative Lakshmanan–Porsezian–Daniel (LPD) equation. Lastly, if we set α1\alpha_{1} and γ\gamma to zero, we get the solution of non-commutative generalization of the nonlinear Schrödinger (NLS) equation.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 6. The profiles of u11u_{11} and u12u_{12} with ν=α2=0\nu=\alpha_{2}=0 in (a) and (b) and, with ν=0\nu=0 in (c) and (d). All other parameters are the same as in Fig. 5.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 7. The profiles of u11u_{11} and u12u_{12} with α1=α2=0\alpha_{1}=\alpha_{2}=0 in (a) and (b) and, with α1=ν=0\alpha_{1}=\nu=0 in (c) and (d). All other parameters are the same as in Fig. 5.

In summary, studying the non-commutative version is important because it gives us different choices for arranging solitons. These arrangements depend not only on the spectral parameter λ\lambda but also on values in a matrix. Similarly, two soliton solutions for the non-commutative case are depicted in Figs. 8-10

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 8. The profiles of uij,i,j=1, 2u_{ij},\;i,\;j=1,\;2 with α1=0.5,α2=ν=1,λ1=0.5i,λ2=0.10.1i\alpha_{1}=0.5,\;\alpha_{2}=\nu=1,\;\lambda_{1}=0.5i,\;\lambda_{2}=-0.1-0.1i. All other parameters are the same as in Figure 5.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 9. The profiles of uij,i,j=1, 2u_{ij},\;i,\;j=1,\;2 with α1=0.5,α2=ν=1,λ1=0.6i,λ2=1.11.1i\alpha_{1}=0.5,\;\alpha_{2}=\nu=1,\;\lambda_{1}=0.6i,\;\lambda_{2}=-1.1-1.1i.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 10. The profiles of uij,i,j=1, 2u_{ij},\;i,\;j=1,\;2 with α1=0.5,α2=ν=1,λ1=0.1+0.5i,λ2=0.5i\alpha_{1}=0.5,\;\alpha_{2}=\nu=1,\;\lambda_{1}=0.1+0.5i,\;\lambda_{2}=-0.5i.

From Fig. 8, it can be seen that solitons with breather-like structures can be seen in uij,i,j=1, 2u_{ij},i,\;j=1,\;2 components. This is because when the force between the two solitons is strong enough, bound-state solitons can merge into breather-like solitons. This process is known as soliton fusion, and it occurs when two solitons combine to form a single soliton.

Figs. 9-10 show two-soliton solutions of nc-HNLS equation, which correspond to the interaction of two individual solitons, that is, two distinct lumps of energy moving at different velocities that interact and scatter without changing their shape. Other multisoliton expressions can be obtained by repeatedly applying the Darboux transformation to the seed solution. The three-soliton configuration represents three distinct amplitudes of soliton scattering. Note that, we have omitted the explicit expression of soliton solutions for non-commutative cases as it is long and cumbersome.

Acknowledgment: We acknowledge the support of the National Natural Science Foundation of China, Nos. 11835011 and 12375006.

Data Availability: Not Applicable.

9. Concluding remarks

This study explored the non-commutative extension of the higher-order nonlinear Schrödinger equation. Through Darboux and binary Darboux transformations, we expressed in quasi-Wronskian and quasi-Gramian forms. These solutions were intricately linked to the non-commutative higher-order nonlinear schrödidnger equation and its associated Lax pair. Importantly, we demonstrated single-, double-peaked, kink, and bright solitons in non-commutative settings. Our approach offers a powerful tool not only for understanding the non-commutative higher order nonlinear Schrödinger equation but also for constructing multisolitons in related non-commutative integrable systems.

References

  • [1] Ablowitz, M. J., Prinari, B., and Trubatch, A. D. (2004). Discrete and continuous nonlinear Schrödinger systems (Vol. 302). Cambridge University Press.
  • [2] Kivshar, Y. S., and Agrawal, G. P. (2003). Optical solitons: from fibers to photonic crystals. Academic press.
  • [3] Agrawal, G. P. (2000). Nonlinear fiber optics. In Nonlinear Science at the Dawn of the 21st Century (pp. 195-211). Berlin, Heidelberg: Springer Berlin Heidelberg.
  • [4] Daniel, M., and Deepamala, K. (1995). Davydov soliton in alpha helical proteins: higher order and discreteness effects. Physica A: Statistical Mechanics and its Applications, 221(1-3), 241-255.
  • [5] Lakshmanan, M., Porsezian, K., and Daniel, M. (1988). Effect of discreteness on the continuum limit of the Heisenberg spin chain. Physics Letters A, 133(9), 483-488.
  • [6] Porsezian, K., Daniel, M., andLakshmanan, M. (1992). On the integrability aspects of the one‐dimensional classical continuum isotropic biquadratic Heisenberg spin chain. Journal of mathematical physics, 33(5), 1807-1816.
  • [7] Ankiewicz, A., Wang, Y., Wabnitz, S., and Akhmediev, N. (2014). Extended nonlinear Schrödinger equation with higher-order odd and even terms and its rogue wave solutions. Physical Review E, 89(1), 012907.
  • [8] Ankiewicz, A., and Akhmediev, N. (2014). Higher-order integrable evolution equation and its soliton solutions. Physics Letters A, 378(4), 358-361.
  • [9] Nimmo, J. J. C., Gilson, C. R., and Ohta, Y. (2000). Applications of Darboux transformations to the self-dual Yang-Mills equations. Theoretical and Mathematical Physics, 122(2), 239-246.
  • [10] Goncharenko, V. M. (2001). Multisoliton solutions of the matrix KdV equation. Theoretical and Mathematical Physics, 126(1), 81-91.
  • [11] Lakshmanan, M. (1977). Continuum spin system as an exactly solvable dynamical system. Physics Letters A, 61(1), 53-54.
  • [12] Hirota, R. (1973). Exact envelope‐soliton solutions of a nonlinear wave equation. Journal of Mathematical Physics, 14(7), 805-809.
  • [13] Minwalla, S., Van Raamsdonk, M., and Seiberg, N. (2000). Noncommutative perturbative dynamics. Journal of High Energy Physics, 2000(02), 020.
  • [14] Furuta, K., and Inami, T. (2000). Ultraviolet property of noncommutative Wess–Zumino–Witten model. Modern Physics Letters A, 15(15), 997-1002.
  • [15] Seiberg, N., and Witten, E. (1999). String theory and noncommutative geometry. Journal of High Energy Physics, 1999(09), 032.
  • [16] Dimakis, A., and Muller-Hoissen, F. (2000). Bicomplexes, integrable models, and noncommutative geometry. arXiv preprint hep-th/0006005.
  • [17] Lechtenfeld, O., and Popov, A. D. (2001). Noncommutative multi-solitons in 2+ 1 dimensions. Journal of High Energy Physics, 2001(11), 040.
  • [18] Lechtenfeld, O., Mazzanti, L., Penati, S., Popov, A. D., and Tamassia, L. (2005). Integrable noncommutative sine-Gordon model. Nuclear Physics B, 705(3), 477-503.
  • [19] Gilson, C. R., and Nimmo, J. J. C. (2007). On a direct approach to quasideterminant solutions of a noncommutative KP equation. Journal of Physics A: Mathematical and Theoretical, 40(14), 3839.
  • [20] Gilson, C. R., and Macfarlane, S. R. (2009). Dromion solutions of noncommutative Davey–Stewartson equations. Journal of Physics A: Mathematical and Theoretical, 42(23), 235202.
  • [21] Etingof, P., Gelfand, I., and Retakh, V. (1997). Factorization of differential operators, quasideterminants, and nonabelian Toda field equations. arXiv preprint q-alg/9701008.
  • [22] Riaz, H. W. A., and ul Hassan, M. (2018). Multisoliton solutions of integrable discrete and semi-discrete principal chiral equations. Communications in Nonlinear Science and Numerical Simulation, 54, 416-427.