Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Exact solutions and conservation laws for the time-fractional nonlinear dirac system: A study of classical and nonclassical lie symmetries

  • Time-fractional Dirac-type systems arise in quantum field theory, plasma physics, and condensed matter systems where fractional calculus captures nonlocal interactions. In this study, we employ classical and nonclassical Lie symmetry methods to analyze the underlying symmetry structure of the system. By deriving infinitesimal generators and performing similarity reductions, we transform the governing fractional partial differential equations (FPDEs) into fractional ordinary differential equations (FODEs). Exact solutions are constructed using the power series method. Furthermore, we establish conservation laws in the fractional setting, ensuring the physical consistency of the system. Our findings offer new insights into the interplay among symmetry, conservation principles, and exact solutions in fractional quantum field models, expanding the analytical toolkit for studying nonlinear relativistic wave equations.

    Citation: Farzaneh Alizadeh, Samad Kheybari, Kamyar Hosseini. Exact solutions and conservation laws for the time-fractional nonlinear dirac system: A study of classical and nonclassical lie symmetries[J]. AIMS Mathematics, 2025, 10(5): 11757-11782. doi: 10.3934/math.2025532

    Related Papers:

    [1] Huizhang Yang, Wei Liu, Yunmei Zhao . Lie symmetry reductions and exact solutions to a generalized two-component Hunter-Saxton system. AIMS Mathematics, 2021, 6(2): 1087-1100. doi: 10.3934/math.2021065
    [2] Miao Yang, Lizhen Wang . Lie symmetry group, exact solutions and conservation laws for multi-term time fractional differential equations. AIMS Mathematics, 2023, 8(12): 30038-30058. doi: 10.3934/math.20231536
    [3] A. Tomar, H. Kumar, M. Ali, H. Gandhi, D. Singh, G. Pathak . Application of symmetry analysis and conservation laws to a fractional-order nonlinear conduction-diffusion model. AIMS Mathematics, 2024, 9(7): 17154-17170. doi: 10.3934/math.2024833
    [4] Tamara M. Garrido, Rafael de la Rosa, Elena Recio, Almudena P. Márquez . Conservation laws and symmetry analysis of a generalized Drinfeld-Sokolov system. AIMS Mathematics, 2023, 8(12): 28628-28645. doi: 10.3934/math.20231465
    [5] Amjad Hussain, Muhammad Khubaib Zia, Kottakkaran Sooppy Nisar, Velusamy Vijayakumar, Ilyas Khan . Lie analysis, conserved vectors, nonlinear self-adjoint classification and exact solutions of generalized $ \left(N+1\right) $-dimensional nonlinear Boussinesq equation. AIMS Mathematics, 2022, 7(7): 13139-13168. doi: 10.3934/math.2022725
    [6] Yuqiang Feng, Jicheng Yu . Lie symmetry analysis of fractional ordinary differential equation with neutral delay. AIMS Mathematics, 2021, 6(4): 3592-3605. doi: 10.3934/math.2021214
    [7] Ziying Qi, Lianzhong Li . Lie symmetry analysis, conservation laws and diverse solutions of a new extended (2+1)-dimensional Ito equation. AIMS Mathematics, 2023, 8(12): 29797-29816. doi: 10.3934/math.20231524
    [8] Youness Chatibi, El Hassan El Kinani, Abdelaziz Ouhadan . Lie symmetry analysis of conformable differential equations. AIMS Mathematics, 2019, 4(4): 1133-1144. doi: 10.3934/math.2019.4.1133
    [9] Alessandra Jannelli, Maria Paola Speciale . On the numerical solutions of coupled nonlinear time-fractional reaction-diffusion equations. AIMS Mathematics, 2021, 6(8): 9109-9125. doi: 10.3934/math.2021529
    [10] Yanxia Hu, Qian Liu . On traveling wave solutions of a class of KdV-Burgers-Kuramoto type equations. AIMS Mathematics, 2019, 4(5): 1450-1465. doi: 10.3934/math.2019.5.1450
  • Time-fractional Dirac-type systems arise in quantum field theory, plasma physics, and condensed matter systems where fractional calculus captures nonlocal interactions. In this study, we employ classical and nonclassical Lie symmetry methods to analyze the underlying symmetry structure of the system. By deriving infinitesimal generators and performing similarity reductions, we transform the governing fractional partial differential equations (FPDEs) into fractional ordinary differential equations (FODEs). Exact solutions are constructed using the power series method. Furthermore, we establish conservation laws in the fractional setting, ensuring the physical consistency of the system. Our findings offer new insights into the interplay among symmetry, conservation principles, and exact solutions in fractional quantum field models, expanding the analytical toolkit for studying nonlinear relativistic wave equations.



    Nonlinear fractional partial differential equations (NLFPDEs) are equations in which the derivatives are of fractional order. Due to their non-local characteristics and specific complexities, these equations have applications in many scientific fields such as physics, engineering, chemistry, and biology. FPDEs are used to model systems that exhibit delay effects or history-dependent behavior. These equations are commonly observed in nonlinear systems and complex phenomena, such as fluid dynamics, disease models, and social dynamics. One of the advantages of using FPDEs is their greater accuracy in describing natural phenomena and the ability to model behaviors that cannot be represented by ordinary differential equations. Specifically, these equations can explain phenomena where the system's memory and history-dependent effects are influential.

    Solving NLFPDEs often requires complex numerical and analytical methods because the properties of these equations make traditional solution techniques, which depend on integer-order derivatives, ineffective. As a result, researchers are increasingly seeking methods that can effectively solve these equations and provide a better understanding of the phenomena being studied. In this regard, NLFPDEs have been studied using various methods. One common method is the Laplace transform, which converts fractional equations into algebraic equations, enabling their solution.

    For example, the authors of [1] applied the Laplace transform to solve families of fractional differential equations. They extended the classical Frobenius method and derived explicit particular solutions using binomial series expansions. Vatsala and Sambandham [2] developed a Laplace transform method to solve sequential Caputo fractional differential equations. They addressed both initial and boundary value problems, providing solutions in terms of Mittag–Leffler functions. Their approach generalizes classical methods and offers a framework for analyzing fractional systems with memory effects. In [3], the authors investigated Laplace transforms with respect to functions and their applications to fractional differential equations. They established fundamental properties, including an inversion formula, and demonstrate how these transforms can be used to solve fractional equations efficiently.

    Additionally, series methods, such as Maclaurin or Taylor series, are used to expand the solution as a series of power functions, especially when the exact solution of the equation is not accessible. Meanwhile, Cang et al. [4] applied the homotopy analysis method to derive series solutions for nonlinear Riccati differential equations of fractional order. Further extending power series techniques, Angstmann and Henry [5] developed a generalized fractional power series method for solving fractional differential equations. This approach refines traditional methods by incorporating the complexities of fractional calculus, offering a versatile tool for obtaining analytical solutions across a wide range of fractional differential equations. Ali, Kalim, and Khan [6] employed the fractional power series method (FPSM) to solve FPDEs. They demonstrated that the FPSM effectively constructs series solutions for a variety of FPDEs, providing a systematic approach to handling the complexities introduced by fractional derivatives. Tashtoush et al. [7] focused on obtaining exact solutions to the space-time conformable fractional (4+1)-dimensional Fokas equation with Kerr law nonlinearity. The authors of [8] applied a fractional nonlinear dispersive model to describe wave propagation in Murnaghan's rods using β-fractional and M-truncated derivatives. They presented exact solutions and phase portraits to analyze the system's dynamic behavior and singularities.

    In more complex cases, semi-analytical algorithms and numerical methods are employed to approximate fractional derivatives and solve the equations. For instance, Kheybari et al. [9] presented a novel semi-analytical algorithm designed to solve time-fractional modified anomalous sub-diffusion equations. In [10], the author applied pseudospectral methods based on different fractional derivative operator matrices for solving time-space FPDEs characterized by variable coefficients and governed by the Caputo derivative. In [11], Hashemi, Mirzazadeh, and Baleanu proposed an innovative method for computing approximate solutions to non-homogeneous wave equations featuring generalized fractional derivatives. Javeed et al. [12] analyzed the homotopy perturbation method for solving FPDEs.

    Some other numerical approaches for solving FPDEs have been proposed in the literature, such as those in [13,14,15,16]. Furthermore, symmetry methods and Lie group analysis are used to obtain exact solutions for FPDEs, particularly in physics and engineering. Lie group analysis is a powerful analytical method used for studying fractional and nonlinear differential equations. In this method, Lie groups and their principles are employed to find analytical solutions to differential equations. Lie groups are particularly useful for nonlinear differential equations because they can identify the symmetries of the equation and, through them, obtain general solutions [17,18,19]. This method is especially applicable to equations that have spatial or temporal symmetries. In fact, Lie group analysis allows for the modeling of the behavior of complex nonlinear equations using a simpler and more precise symmetry structure, thus enabling the extraction of both specific and general solutions to these equations [20,21,22].

    In the present work, the time-fractional nonlinear Dirac system (TFNLDS), expressed as follows, is investigated:

    Λ1:Dαtp=12qxxp2qq3,Λ2:Dαtq=12pxx+pq2+p3, (1.1)

    where p and q are functions of (t,x), and Dαt(.) represents the time-fractional Riemann–Liouville (RL) derivative of order α, where α(0,1). By setting α=1, the classical type of the nonlinear Dirac system can be recovered from the system (1.1) [23,24,25]. In the original Dirac system, as described by Frolov [26] and Schratz et al. [27], the functions p(x,t) and q(x,t) evolve according to the coupled nonlinear equations given in [28]. This system, which describes the motion of relativistic spin-12 particles in external electromagnetic fields, has significant applications in applied sciences. However, to account for more complex dynamics and long-range interactions, the system can be generalized into a fractional form. In this work, we extend the nonlinear Dirac system by introducing fractional derivatives of order α(0,1) to model the system with non-local effects. The fractional form of the Dirac system provides a more accurate representation of phenomena exhibiting memory effects, such as the self-interaction of nonlinear particles and the influence of long-range forces. By transforming the system into a fractional setting, we can explore new solutions and behaviors that arise from the fractional order, opening up avenues for further research in both theoretical and applied contexts.

    This paper investigates the Lie symmetries and conservation laws of the TFNLDS. Section 2 presents the necessary preliminaries and mathematical framework. Section 3 is dedicated to the Lie symmetry analysis of the TFNLDS, identifying both classical symmetries (Section 3.1.1) and nonclassical symmetries (Section 3.1.2). Utilizing these symmetries, exact solutions are derived, and their implications are explored. Section 4 constructs the conservation laws associated with the TFNLDS, providing insights into the fundamental invariants of the system. Finally, Section 5 summarizes the findings and discusses their significance.

    Some fundamental definitions and properties of fractional order derivatives are presented in this section. Interested readers are referred to [29,30,31] for their definitions and properties.

    Definition 1. [29] Let αR+. The operator Jα defined by

    Jαf(t,x)=t0(tw)α1Γ(α)f(w,x)dw,

    where Γ() is the gamma function, is called the RL fractional integral operator of order α. When α=0, Jα=I is the identity operator.

    Definition 2. [30] Let αR+ and n=α, the operator Dα formulated as

    Dα=DnJnα,

    which referred to as the RL fractional differential operator of order α. When α=0, Dα=I is the identity operator. Therefore, the fractional RL derivative of order α of the function f(t,x) is given by

    Dαtf(t,x)={1Γ(1α)tt0(tw)αf(w,x)dw,0<α<1,tf(t,x),α=1,

    and the RL fractional derivative of tβ is represented by

    Dαttβ={Γ(β+1)Γ(βα+1)tβα,(αβN),β>1,0,(αβN).

    It is evident that when αβN, the right-hand side is the α-th derivative of the classical polynomial of degree α(αβ){0,1,...,α1}, where shows the ceiling function.

    The fractional integral and the RL fractional derivative possess the following properties:

    1) Jαtβ=Γ(β+1)Γ(α+β+1)tα+β,α>0,β>1.

    2) Jα1Jα2f(t)=Jα2Jα1f(t)=Jα1+α2f(t),α1,α20.

    3) DαJαf(t)=f(t),α0.

    4) Dα(c1f(t)+c2g(t))=c1Dαf(t)+c2Dαg(t),c1,c2R,α>0.

    5) Dα[fg](t)=αk=0(αk)(Dαf)(t)(Dαkg)(t)+k=α+1(αk)(Dkf)(t)(Jkαg)(t),α>0,

    where shows the flooring function.

    Definition 3. [30] Let α,β>0. The two-parameter Mittag–Leffler function Eα,β is defined by

    Eα,β(z)=j=0zjΓ(jα+β).

    Definition 4. [31] The fractional integral operator of Erdélyi–Kober for f(t) is

    (Kν,αβf)(t)={1(w1)α1Γ(α)w(ν+α)f(tw1β)dw,α>0,f(t),α=0.

    Definition 5. [31] The fractional derivative operator of Erdélyi–Kober for f(t) is

    (Pν,αβf)(t)=m1i=0(ν+i1βtddt)(Kν+α,mαβf)(t),
    m={α+1,αN,α,αN.

    In this section, we examine the Lie group and both the classical and nonclassical symmetries of the main system (1.1). To this end, we present the relevant concepts of Lie group analysis for the system of time-fractional partial differential equations (STFPDEs), which will be used later.

    In this subsection, a general description is provided of how the Lie group method is applied to STFPDEs of order α, where α(0,1), expressed as follows:

    Ξ1:Dαtpσ1(t,x,p,q,px,qx,pxx,qxx,)=0,Ξ2:Dαtqσ2(t,x,p,q,px,qx,pxx,qxx,)=0. (3.1)

    The system (3.1) consists of p and q as dependent variables, while t and x serve as independent variables. Additionally, the subscripts denote integer-order derivatives. Considering that the system (3.1) remains invariant under the following one-parameter Lie group transformations

    ˘x=x+ϵζ1(t,x,p,q)+O(ϵ2),˘t=t+ϵζ2(t,x,p,q)+O(ϵ2),˘p=u+ϵϖ1(t,x,p,q)+O(ϵ2),˘q=v+ϵϖ2(t,x,p,q)+O(ϵ2),Dαt˘p=Dαtp+ϵϖα,t1(t,x,p,q)+O(ϵ2),Dαt˘q=Dαtq+ϵϖα,t2(t,x,p,q)+O(ϵ2),j˘p˘xj=jpxj+ϵϖj,x1(t,x,p,q)+O(ϵ2),j=1,2,3,,j˘q˘xj=jqxj+ϵϖj,x2(t,x,p,q)+O(ϵ2),j=1,2,3,,

    where ϵ is the group parameter. Furthermore, the vector field corresponding to these transformations is given as follows:

    W=ζ1(t,x,p,q)x+ζ2(t,x,p,q)t+ϖ1(t,x,p,q)u+ϖ2(t,x,p,q)v.

    For the system (3.1), the infinitesimal generator admits a symmetry precisely when the following conditions are satisfied:

    r(α,k1,h1)W(Ξ1)|Ξ1=0=0,r(α,k2,h2)W(Ξ2)|Ξ2=0=0,

    where kl and hl for l=1,2, represent the leading orders in the l-th equation in the system (3.1), and it is also important to note that the fractional prolongation operator r(α,k,h)W is expressed as follows:

    r(α,k,h)W=W+ϖα,t1(Dαxp)+ϖ1,x1px++ϖk,x1pkx+ϖα,t2(Dαxq)+ϖ1,x2qx++ϖh,x2qhx,

    where ϖj,x1 and ϖj,x2 represent the extensions of the infinitesimals in the integer-order framework, as follows:

    ϖj,x1=Dxϖj1,x1(Dxζ1)jpxj(Dxζ2)t(j1pxj1),ϖj,x2=Dxϖj1,x2(Dxζ1)jqxj(Dxζ2)t(j1qxj1),jN,

    where the symbol Dx signifies the total derivative, i.e.,

    Dx=x+pxp+pxxpx++qxq+qxxqx+.

    Therefore, ϖα,t1 represents the α-order extensions of the infinitesimal operator as

    ϖα,t1=Dαtϖ1+ζ1Dαt(px)Dαt(ζ1px)+Dαt(Dt(ζ2)p)Dα+1t(ζ2p)+ζ2Dα+1t(p),

    where the symbol Dαt denotes the total α-order fractional derivative. Utilizing the generalized Leibnitz formula and the chain rule[29,32], ϖα,t1 can be defined as follows:

    ϖα,t1=αϖ1tα+(ϖ1pαDt(ζ2))αptαpαϖ1ptα+(ϖ1qαqtαqαϖ1qtα)+j=1[(αj)jϖ1ptj(αj+1)Dj+1t(ζ2)]Dαjt(p)+j=1(αi)jϖ1qtjDαjt(q)j=1(αj)Djt(ζ1)Dαjt(px)+λ1,

    where

    λ1=j=2jk=2kl=2l1m=0(αj)(jk)(lm)(1)mtjαl!Γ(jα+1)(pmk(plm)tkjk+lϖ1tjkpl+qmk(qlm)tkjk+lϖ1tjkql).

    Similarly, ϖα,t2 can be written as follows:

    ϖα,t2=αϖ2tα+(ϖ2qαDt(ζ2))αqtαqαϖ2qtα+(ϖ2pαptαpαϖ2ptα)+j=1[(αj)jϖ2qtj(αj+1)Dj+1t(ζ2)]Dαjt(q)+j=1(αj)jϖ2utjDαjt(p)j=1(αj)Djt(ζ1)Dαjt(qx)+λ2,

    where

    λ2=j=2jk=2kl=2l1m=0(αj)(jk)(lm)(1)mtjαl!Γ(jα+1)(pmk(plm)tkjk+lϖ2tjkpl+qmk(qlm)tkjk+lϖ2tjkql).

    Since ϖ1 and ϖ2 depend linearly on the variables p and q, their partial derivatives iϖ1pi and iϖ2qi vanish for all iN{1}. Consequently, it follows that λ1=0 and λ2=0.

    Based on the proposed Lie group method, the invariance condition of the system (3.1) is that the following relations hold:

    {r(α,k1,h1)W(Dαtpσ1(t,x,p,q,px,qx,pxx,qxx,))|System(3.1)=0,r(α,k2,h2)W(Dαtqσ2(t,x,p,q,px,qx,pxx,qxx,))|System(3.1)=0. (3.2)

    Thus, on the basis of the relation (3.2) and system (1.1), we have

    {r(α,2,2)W(Dαtp12qxx+p2q+q3)|System(1.1)=0,r(α,2,2)W(Dαtq+12pxxpq2p3)|System(1.1)=0.

    Therefore, the following determining equations are obtained:

    3p2ϖ1+q2ϖ1p3ϖ2qpq2ϖ2q+αp3ζ2t+2pqϖ2αϖ2tα+qαϖ2qtα+αpq2ζ2t12ϖ1xx=0,q3ϖ1pp2ϖ23q2ϖ2+p2qϖ1pαq3ζ2t2pqϖ1αp2qζ2tαϖ1tα+pαϖ1ptα+12ϖ2xx=0,α(α22α+1)ζ2ttt3α(α1)ϖ2ttq=0,α(α22α+1)ζ2ttt3α(α1)ϖ1ttp=0,(αj)jϖ1ptj(αj+1)Dj+1t(ζ2)=0,(αj)jϖ2qtj(αj+1)Dj+1t(ζ2)=0,12ϖ1p12αζ2t+ζ1x12ϖ2q=0,12ϖ2q12αζ2t+ζ1x12ϖ1p=0,ϖ1q=ϖ2p=ϖ1pp=ϖ2qq=0,α(1α)ζ2tt+2αϖ1tp=0,α(1α)ζ2tt+2αϖ2tq=0,ζ2x=ζ2p=ζ2q=0,ζ1p=ζ1q=ζ1t=0,ϖ1xp12ζ1xx=0,12ζ1xxϖ2xq=0,jϖ1ptj=0,jϖ2qtj=0,Djt(ζ1)=0.

    The vector fields are derived on the basis of the determining equations as follows:

    W1=x,W2=4tt+2αxx+αpp+αqq,W3=pp+qq,W4=h(t,x)q,W5=k(t,x)p.

    Considering the vector field W1=x, the corresponding characteristic equation is

    dt0=dx1=dp0=dq0.

    By analyzing the given equation, one can derive the corresponding invariant solutions related to the associated vector field, expressed as p(t,x)=ω(t) and q(t,x)=ϑ(t). By inserting the derived functions into the main equation, the resulting system takes the following form:

    {Dαtω(t)=ϑ(t)ω2(t)ϑ3(t),Dαtϑ(t)=ω(t)ϑ2(t)+ω3(t). (3.3)

    If we suppose that ω(t)=iϑ(t), the system (3.3) can be written as

    {iDαtϑ(t)=0,Dαtϑ(t)=0,

    and the exact solutions of the given system, dependent on time, can be formulated as follows:

    {ω(t)=ic1tα1,ϑ(t)=c2tα1,

    where c1 and c2 are constants. Therefore the final solutions of the main system are p(t,x)=ic1tα1 and q(t,x)=c2tα1.

    The characteristic equation for the vector field W2 is given by

    dt4t=dx2αx=dpαp=dzαq.

    The invariant solutions corresponding to the vector field W2 are

    p(t,x)=tα4f(ε),q(t,x)=tα4g(ε),ε=xtα2. (3.4)

    Theorem 1. By applying the solutions (3.4), the system (1.1) simplifies to the following system of FODEs:

    {(P13α4,α2αf)(ε)12g(ε)=0,(P13α4,α2αg)(ε)+12f(ε)=0,g(ε)f2(ε)+g3(ε)=0,f(ε)g2(ε)+f3(ε)=0. (3.5)

    Proof. Let nN and α(n1,n). Taking the α-order temporal RL derivative of Eq (3.4) yields

    αptα=ntn[1Γ(nα)t0(tw)nα1wα4f(xwα2)dw]. (3.6)

    By using the change in the variable ν=tw, the relation (3.6) can be written as follows:

    αptα=ntn[tn3α4Γ(nα)1(ν1)nα1ν3α4n1f(xtα2να2)dν],αptα=ntn[tn3α4(K1α4,nα2αf)(ε)].

    Furthermore, if we let ε=xtα2, 0<ρ<, then

    ttρ(ε)=tεtdρ(ε)dε=α2εdρ(ε)dε.

    Therefore

    ntn[tn3α4(K1+α4,nα2αf)(ε)]=n1tn1[t(tn3α4(K1+α4,nα2αf)(ε))]
    =n1tn1[tn3α41(n3α4α2εddε)(K1+α4,nα2αf)(ε)]=t3α4n1j=0(13α4+jα2εddε)(K1+α4,nα2αf)(ε)=t3α4(P13α4,α2αf)(ε).

    Thus

    αptα=t3α4(P13α4,α2αf)(ε).

    Similarly

    αqtα=t3α4(P13α4,α2αf)(ε),

    and thus the proof is completed.

    Applying a power series approach to obtain exact solutions

    To find an exact solution to the system (1.1), let g(ε)=if(ε). Consequently, it is enough to solve the following equation:

    (P13α4,α2αf)(ε)i2f(ε)=0. (3.7)

    By employing a power series approach, we assume that f(ε) can be expanded as follows:

    f(ε)=j=0ajεj,(Pτ,αβf)(ε)=k=0akΓ(τkβ+1)Γ(τkβ+1α)εk, (3.8)

    and

    f(ε)=j=1jajεj1,f(ε)=j=2j(j1)ajεj2. (3.9)

    Substituting Eqs (3.8) and (3.9) into Eq (3.7) yields the following equation:

    j=0ajΓ(23α4αj2)Γ(27α4αj2)εji2j=0(j+1)(j+2)bj+2εj=0. (3.10)

    If we compare the coefficients in Eq (3.10), for j=0, we have

    a2=2Γ(23α4)iΓ(27α4)a0,

    and for j1, we have

    aj+2=2ajΓ(23α4αj2)i(j+1)(j+2)Γ(27α4+αj2).

    Therefore, by inserting the obtained coefficients into the series (3.8) we have

    f(ε)=a0+a1εi(2Γ(23α4)Γ(27α4)a0ε2j=12bjΓ(23α4αj2)(j+1)(j+2)Γ(27α4+αj2)εj). (3.11)

    Thus

    g(ε)=i(a0+a1ε)+(2Γ(23α4)Γ(27α4)a0ε2j=12bjΓ(23α4αj2)(j+1)(j+2)Γ(27α4+αj2)εj). (3.12)

    Through the application of (3.11), (3.12), and (3.4), the following exact solutions are obtained for the governing system (1.1):

    p(t,x)=tα4(a0+a1xtα2ij=02ajΓ(23α4αj2)(j+1)(j+2)Γ(27α4+αj2)(xtα2)j),q(t,x)=tα4(i(a0+a1xtα2)+j=02ajΓ(23α4αj2)(j+1)(j+2)Γ(27α4+αj2)(xtα2)j). (3.13)

    To illustrate that the plots are consistent with the solutions in Eq (3.13), the series are truncated at N=30. Figures 1 and 2 present the real and imaginary parts, as well as the absolute values of p(t,x) and q(t,x) for a0=a1=0.2 and different fractional orders α at t=20.

    Figure 1.  The plots derived from the classical vector field W2 show (a) the real part, (b) the imaginary part, and (c) the absolute value of p(t,x) for various values of α at t=20. These plots indicate that the obtained solutions consistently exhibit convergent behavior.
    Figure 2.  The plots derived from the classical vector field W2 show (a) the real part, (b) the imaginary part, and (c) the absolute value of q(t,x) for various values of α at t=20. These plots demonstrate that the obtained solutions exhibit convergence behavior.

    For the vector field W1+W3, the following invariant solutions are obtained:

    p(t,x)=exf(t),q(t,x)=exg(t). (3.14)

    Utilizing the transformation (3.14) to the system (1.1), the following fractional and integer order ODE system is obtained:

    {Dαtf(t)12g(t)=0,Dαtg(t)+12f(t)=0,g(t)f2(t)+g3(t)=0,f(t)g2(t)+f3(t)=0. (3.15)

    To obtain an exact solution of the system (3.15), let g(t)=if(t). Thus, it is sufficient to determine the solution of the following equation:

    Dαtf(t)i2f(t)=0. (3.16)

    The solutions to Eq (3.16) are obtained using the fractional Laplace transforms (FLTs) [33,34,35] as follows:

    f(t)=λt2α2Eα,2α1(i2tα),g(t)=iλt2α2Eα,2α1(i2tα), (3.17)

    where λ=kΓ(1α) and k is a constant. Substituting the relations (3.17) into the system (3.14) yields the exact solutions of the system (1.1) as follows:

    p(t,x)=λext2α2Eα,2α1(i2tα),q(t,x)=iλext2α2Eα,2α1(i2tα).

    Due to definition of the Mittage–Leffler function, we have

    p(t,x)=λext2α2(j=0(1)j(i2tα)2jΓ(2jα+2α1)+ij=0(1)j+1(i2tα)2j+1Γ(2jα+3α1)),q(t,x)=λext2α2(ij=0(1)j(i2tα)2jΓ(2jα+2α1)j=0(1)j+1(i2tα)2j+1Γ(2jα+3α1)).

    In Figures 3 and 4, the three-dimensional (3D) and two-dimensional (2D) plots of the real and imaginary parts of the solutions p(t,x) and q(t,x), obtained from the classical vector field W1+W3, are presented for two different values of α.

    Figure 3.  Subfigures {(a),(d)} and {(b),(e)} display the real and imaginary parts of p(t,x), respectively, for α=0.95 and the classical case α=1, with λ=1. Subfigures (c) and (f) present 2D plots of the real and imaginary parts at x=10 for both values of α. These plots reveal that as α approaches 1, the amplitude of the oscillations increases noticeably.
    Figure 4.  Subfigures {(a),(d)} and {(b),(e)} represent the real and imaginary parts of q(t,x) for α=0.95 and the classical case α=1, with λ=1. The 2D plots in (c) and (f), evaluated at x=10, illustrate the effect of α on the oscillatory behavior, showing that as α approaches 1, the amplitude of the oscillations increases.

    To obtain new exact solutions for the heat equation, Bluman and Cole proposed a nonclassical reduction method [36]. The essence of this approach lies in incorporating a fixed surface condition, meaning that applying this condition along with the primary determining equations leads to a system of nonlinear determining equations for infinitesimals. In the analysis of the nonclassical scenario within Lie's symmetry theory, beyond the condition r(α,k,h)W=0, the invariant surface conditions must also be satisfied. In the nonclassical symmetry method, by applying the invariant surface condition along with the governing differential equations, a nonlinear system of partial differential equations is obtained. This system yields the infinitesimals that characterize the nonclassical symmetries of the original problem. Unlike the classical Lie symmetry method, which requires the invariance of the entire differential equation under prolonged vector fields, the nonclassical approach imposes a more restrictive criterion by requiring invariance on a solution manifold defined by both the differential equation and the invariant surface condition.

    As a result, the number of determining equations in the nonclassical framework is generally fewer than those in the classical method. This reduction in the determining system, however, comes at the cost of increased complexity due to its nonlinearity. Despite this, the nonclassical symmetry method is capable of uncovering a wider class of symmetry reductions and exact solutions that are not accessible through classical methods. Consequently, the overall solution set in the nonclassical case is typically more extensive, providing deeper insights into the structure and integrability of nonlinear differential equations.

    In this work, we aim to apply this method to our system to derive new exact solutions. Consider the following invariant surface conditions:

    Ω1ζ1(t,x,p,q)px+ζ2(t,x,p,q)ptϖ1=0,Ω2ζ1(t,x,p,q)qx+ζ2(t,x,p,q)qtϖ2=0.

    Assume that, ζ1=1 and ζ2=0. Thus, the surface conditions are as follows:

    px=ϖ1,qx=ϖ2.

    Therefore

    pxx=ϖ1x+ϖ1ϖ1p+ϖ2ϖ1q,qxx=ϖ2x+ϖ1ϖ2p+ϖ2ϖ2q.

    By substituting these relations into the system (1.1), we obtain ϖ2=p and ϖ1=q. Consequently, we derive the vector field W6=xqp+pq, whose corresponding invariant solutions for this case are given as follows:

    p(t,x)=f(t)cos(x)+g(t)sin(x),q(t,x)=g(t)cos(x)+f(t)sin(x). (3.18)

    Moreover, by applying (3.18), the reduced system takes the following form:

    {Dαtf(t)12g(t)f2(t)g(t)g3(t)=0,Dαtg(t)+12f(t)+g2(t)g(t)+f3(t)=0. (3.19)

    To compute a set of solutions for the system (3.19), it suffices to assume g(t)=if(t) and solve the following equation:

    Dαtf(t)i2f(t)=0. (3.20)

    The solution to Eq (3.20) can be obtained using the FLTs as follows:

    f(t)=λt2α2Eα,2α1(i2tα), (3.21)

    where λ=kΓ(1α) and k is a constant. Substituting Eq (3.21) to Eq (3.18), the exact solutions of system (1.1) are derived as follows:

    p(t,x)=λt2α2(Eα,2α1(i2tα)cos(x)+iEα,2α1(i2tα)sin(x)),q(t,x)=λt2α2(Eα,2α1(i2tα)sin(x)iEα,2α1(i2tα)cos(x)).

    Hence,

    p(t,x)=λt2α2[cos(x)(j=0(1)j(i2tα)2jΓ(2jα+2α1))sin(x)(j=0(1)j+1(i2tα)2j+1Γ(2jα+3α1))]+iλt2α2[sin(x)(j=0(1)j(i2tα)2jΓ(2jα+2α1))+cos(x)(j=0(1)j+1(i2tα)2j+1Γ(2jα+3α1))],q(t,x)=λt2α2[sin(x)(j=0(1)j(i2tα)2jΓ(2jα+2α1))+cos(x)(j=0(1)j+1(i2tα)2j+1Γ(2jα+3α1))]iλt2α2[cos(x)(j=0(1)j(i2tα)2jΓ(2jα+2α1))sin(x)(j=0(1)j+1(i2tα)2j+1Γ(2jα+3α1))].

    Figures 5 and 6 provide a detailed visualization of the real and imaginary components of the solutions p(t,x) and q(t,x), computed for two distinct values of α. These plots illustrate the spatiotemporal behavior of the solutions across the specified domain, with a focus on the effects of varying α on the oscillatory characteristics of the solutions. The representations in both 3D and 2D formats help highlight the differences in the behavior of the solutions as α changes, offering a clear insight into the dynamics of the system.

    Figure 5.  The solution p(t,x), derived from the nonclassical vector field W6, is displayed for α=0.85 and α=1 with λ=1. Subfigures (a), (b), (d), and (e) present 3D views of the real and imaginary parts over space and time. The corresponding 2D profiles at t=2 are shown in (c) and (f). As observed in the plots, increasing the value of α significantly amplifies the oscillatory behavior of the solution.
    Figure 6.  Real and imaginary parts of the solution q(t,x), associated with the nonclassical vector field W6, are visualized for α=0.85 and α=1 with λ=1. The 3D plots in subfigures (a), (b), (d), and (e) show the real and imaginary parts over space and time, while the 2D plots in (c) and (f), taken at t=2, illustrate the effect of increasing α on the oscillations' amplitude.

    In the context of nonclassical symmetries, we consider ζ20 and ζ1=1, and assume that ζ2p=ζ2q=0. Under these conditions, the corresponding surface constraints are given as follows:

    px=ϖ1ζ2pt,qx=ϖ2ζ2qt.

    In this case, we obtain

    W7=x+ct.

    According to the vector field W7, the following invariant solutions can be derived:

    {p(t,x)=f(ε),q(t,x)=g(ε),ε=tcx,

    and these variables reduce the system (1.1) to the following system of FODEs:

    {Dαtf(ε)12c2g(ε)+f2(ε)g(ε)+g3(ε)=0,Dαtg(ε)+12c2f(ε)g2(ε)f(ε)f3(ε)=0.

    Let g(ε)=if(ε). It is enough to solve the following equation:

    Dαtf(ε)i2c2f(ε)=0. (3.22)

    For Eq (3.22), by using the FLTs, we have

    f(ε)=2λc2iεαE2α,α+1(2c2iε2α)+λ1E2α,1(2c2iε2α)+λ2εE2α,2(2c2iε2α),

    where λ=λ1Γ(1α), λ1, andλ2 are constants. Thus, the exact solution is given by

    p(t,x)=2λc2i(tcx)αE2α,α+1(2c2i(tcx)2α)+λ1E2α,1(2c2i(tcx)2α)+λ2(tcx)E2α,2(2c2i(tcx)2α),

    and

    q(t,x)=2λc2(tcx)αE2α,α+1(2c2(tcx)2α)+iλ1E2α,1(2c2i(tcx)2α)+iλ2(tcx)E2α,2(2c2i(tcx)2α).

    By separating the real and imaginary parts of p(t,x) and p(t,x), the following exact solutions are obtained:

    p(t,x)=2λc2(tcx)α(j=0(1)j+1(2(tcx)2α)2j+1c4j+2Γ(4j2jα+3))+(λ1+λ2(tcx))(j=0(1)j(2(tcx)2α)2jc4jΓ(4j2jα+α+1))+i[2λc2(tcx)α(j=0(1)j(2(tcx)2α)2jc4jΓ(4j2jα+α+1))+(λ1+λ2(tcx))(j=0(1)j+1(2(tcx)2α)2j+1c4j+2Γ(4j2jα+3))],q(t,x)=2λc2(tcx)α(j=0(1)j(2(tcx)2α)2jc4jΓ(4j2jα+α+1))(λ1+λ2(tcx))(j=0(1)j+1(2(tcx)2α)2j+1c4j+2Γ(4j2jα+3))+i[2λc2(tcx)α(j=0(1)j+1(2(tcx)2α)2j+1c4j+2Γ(4j2jα+3))+(λ1+λ2(tcx))(j=0(1)j(2(tcx)2α)2jc4jΓ(4j2jα+α+1))]. (3.23)

    To illustrate that the plots are consistent with the solutions (3.23), the series are truncated at N=30. Figures 7 and 8 present the real and imaginary parts, as well as the absolute values of p(t,x) and q(t,x), for c=λ=λ1=λ2=1 and different fractional orders α at x=2.

    Figure 7.  The plots, derived from the classical vector field W7, show (a) the real part, (b) the imaginary part, and (c) the absolute value of p(t,x) for various values of α at x=2. These plots demonstrate that the obtained solutions exhibit convergence behavior.
    Figure 8.  The plots, derived from the classical vector field W7, show (a) the real part, (b) the imaginary part, and (c) the absolute value of q(t,x) for various values of α at x=2. These plots indicate that the obtained solutions consistently exhibit convergent behavior.

    The conservation laws (CLs) of the system (1.1) are derived in this section through Ibragimov's method [37], a well-established approach for obtaining CLs in time-fractional nonlinear differential systems. This methodology has been extensively utilized in the study of time-fractional equations [38,39,40,41,42,43]. The aim is to identify CLs corresponding to both the classical and nonclassical vector fields. A CL for the system (1.1) is expressed as follows:

    Dx(Vx)+Dt(Vt)|(3.1)=0,

    where Vt represents the time flow and Vx represents the space flow. As observed by Ibragimov [37], the formal Lagrangian of the system (1.1) can be expressed as follows:

    H=μ1(t,x)(Dαtp12qxx+p2q+q3)+μ2(t,x)(Dαtq+12pxxpq2p3), (4.1)

    where the variables μ1(t,x) and μ2(t,x) are treated as dependent, and the corresponding adjoint equations for the formal Lagrangian operator (4.1) are derived as follows:

    {M1δHδp=0,M2δHδq=0, (4.2)

    where δδp and δδq represent the Euler–Lagrange operators, which are defined as follows:

    δδp=p+(Dαt)(Dαtp)+k1(1)kDx...Dxpkx,

    and

    δδq=q+(Dαt)(Dαtq)+k1(1)kDx...Dxqkx,

    where the adjoint operator of Dαt is denoted by (Dαt). By taking the RL fractional differential operators into account, the following expression is obtained:

    (Dαt)=(1)mJmαT(nt)=(DαT)Ct,JmαTs(t,x)=Tts(τ,x)(τt)m(1+α)Γ(mα)dτ,m=[α]+1,

    where the operator (DαT)Ct represents the right-sided Caputo fractional derivative. By replacing Eq (4.1) in Eq (4.2), the adjoint equations for the system (1.1) are obtained:

    {M1=μ1(2pq)μ2(q2+3p2)+(Dαt)μ1+12μ2xx,M2=μ1(p2+3q2)μ2(2pq)+(Dαt)μ212μ2xx. (4.3)

    Referring to [37], the system (1.1) admits the following CL:

    DxVxi+DtVti=0, (4.4)

    where the specified equations represent the conserved vectors Vi=(Vti,Vxi)

    Vxi=(pWiδHδpx+j1Dx...Dx(pWi)Hp(j+1)x)+(qWiδHδqx+j1Dx...Dx(qWi)Hq(j+1)x),Vti=n1j=0(1)j[Dα1jt(pWi)Djt(H(Dαtp))+Dα1jt(qWi)Djt(H(Dαtq))](1)n[J(pWi,Dnt(H(Dαtp)))+J(qWi,Dnt(H(Dαtq)))],n=[α]+1, (4.5)

    where pWi=ϖ1iζ1ipxζ2ipt, qWi=ϖ2iζ1iqxζ2iqt, and the notation J denotes the following integral:

    J(h,s)=t0Tth(η,x)s(σ,x)(ση)n(α+1)Γ(nα)dσdη. (4.6)

    If we have the following relation for the time-fractional nonlinear Eq (4.1), then we can say that the system (1.1) is self-adjoint

    MδHδp=λ1Λ1+λ2Λ2,MδHδp=λ3Λ1+λ4Λ2,

    where λ1,λ2,λ3,λ4 are unknown and are to be determined. Thus, we can write the nonlinear self adjoint condition as follows:

    λ1=λ2=λ3=λ4=0,μ1(x,t,p)=A,μ2(x,t,q)=B,A,BR.

    Hence, if we suppose that A=B=1, then

    H=Dαtp+Dαtq+12(pxxqxx)+p2q+q3pq2p3. (4.7)

    Drawing on the previous analysis and the generators of both classical and nonclassical Lie symmetries, the conserved vectors (CVs) for the TFNLDS are derived as follows. Initially, for the classical vector fields, the CVs are computed as follows.

    CVs for (W1): In this case, W1 is associated with the Lie characteristic functions given by

    pW1=px,qW1=qx. (4.8)

    Using these functions, the CVs corresponding to W1 are determined as outlined below:

    Vx1=pW1(DxHpxx)+Dx(pW1)Hpxx+qW1(DxHqxx)+Dx(qW1)Hqxx,Vt1=J1α(pW1)J1α(qW1),

    and

    Vx1=12(qxxpxx),Vt1=J1αpxJ1αqx.

    CVs for (W2): For the generator W2=4tt+2αxx+αpp+αqq, the following Lie characteristic functions are concluded:

    pW2=αp2αxpx4tpt,qW2=αq2αxqx4tqt, (4.9)

    then

    Vx2=12(αpx2αxpxx4tptx)12(αqx2αxqxx4tqtx),Vt2=J1α(αp2αxpx4tpt)J1α(αq2αxqx4tqt).

    CVs for (W=W1+W3): For the generator W=x+pp+qq, we have

    pW=ppx,qW=qqx, (4.10)

    and thus

    Vx3=12(pxpxx)12(qxqxx),Vt3=J1α(ppx)J1α(qqx).

    Now, the investigation of the nonclassical vector fields proceeds as follows.

    CVs for (W6): In the case where ζ1=1 and ζ2=0, the vector field is expressed as W6=xqp+pq. From this, the following Lie characteristic functions are derived:

    pW6=qpx,qW6=pqx. (4.11)

    Therefore

    Vx6=12(qxpxxpx+qxx),Vt6=J1α(qpx)J1α(pqx).

    CVs for (W7): In this case, ζ1=1 and ζ20, and the vector field is expressed as W7=x+ct. From this, the following Lie characteristic functions are derived:

    pW7=pxcpt,qW7=qxcqt, (4.12)

    and

    Vx7=12(qxxpxx+cqtxcptx),Vt7=J1α(pxcpt)J1α(qxcqt).

    In this paper, we studied the classical and nonclassical Lie symmetry methods and CLs of the TFNLDS. This study represents the first exploration of exact solutions for the TFNLDS equation incorporating the RL fractional derivative. The significance of this investigation lies in its contribution to the analytical study of nonlinear fractional systems, which are essential for modeling memory-dependent and nonlocal physical phenomena. By applying Lie group analysis, we determined the system's symmetries and employed both classical and nonclassical Lie symmetry methods to derive similarity reductions, transforming the original equation into reduced forms on the basis of the obtained vector fields and corresponding invariant solutions. This methodological framework allowed for systematic reductions and the construction of exact forms, providing new insights into the structure of such fractional systems.

    Additionally, we constructed exact solutions using various approaches, including the power series method for the derived generators. The use of different solution strategies emphasized the robustness and flexibility of our analytical treatment.

    Our analysis, supported by Figures 1, 2, 7, and 8, demonstrated that the obtained solutions exhibited convergence behavior in both classical and nonclassical cases. These results underscore the reliability of the derived solutions and confirm the effectiveness of the symmetry-based reductions.

    Furthermore, on the basis of the Lie symmetry generators, we systematically constructed CLs for the corresponding classical and nonclassical vector fields of the TFNLDS. These conservation laws reflect the underlying invariance properties and provide a deeper physical understanding of the model.

    These results highlighted the effectiveness of the proposed approach in finding analytical solutions to a broad class of FPDEs, making it a promising tool for further studies in fractional systems. Therefore, this work not only offers exact solutions for a specific fractional model but also establishes a general pathway for tackling nonlinear fractional PDEs through symmetry and conservation law techniques. Extending this approach to higher-dimensional systems or using other fractional derivative operators (such as Caputo or Hadamard) presents an important challenge for future work. In addition, future research could explore the applicability of this method to coupled fractional equations and systems with more complex boundary conditions, as well as its use in modeling real-world phenomena in fields such as viscoelasticity, fluid dynamics, and anomalous diffusion.

    Farzaneh Alizadeh: Conceptualization, writing original draft, methodology, investigation, formal analysis, writing review and editing; Kamyar Hosseini: Writing original draft, methodology, formal analysis; Samad Kheybari: Formal analysis, supervision, writing review and editing.

    All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are very grateful to the respected reviewers for their valuable suggestions to improve the quality of the paper.

    Kamyar Hosseini is a Guest Editor of special issue "Emerging Trends in Algebra, Geometry, and Topology of Soliton Systems" for AIMS Mathematics. Kamyar Hosseini was not involved in the editorial review and the decision to publish this article.

    All authors declare no conflicts of interest in this paper.



    [1] S. D. Lin, C. H. Lu, Laplace transform for solving some families of fractional differential equations and its applications, Adv. Differ. Equ. , 2013 (2013), 137. http://doi.org/10.1186/1687-1847-2013-137 doi: 10.1186/1687-1847-2013-137
    [2] A. S. Vatsala, B. Sambandham, Laplace transform method for sequential Caputo fractional differential equations, Math. Eng. Sci. Aerospace, 2016.
    [3] H. M. Fahad, M. Ur Rehman, A. Fernandez, On Laplace transforms with respect to functions and their applications to fractional differential equations, Math. Method. Appl. Sci. , 46 (2023), 8304–8323. https://doi.org/10.1002/mma.7772 doi: 10.1002/mma.7772
    [4] J. Cang, Y. Tan, H. Xu, S. Liao, Series solutions of non-linear Riccati differential equations with fractional order, Chaos Soliton. Fract. , 40 (2009), 1–9. https://doi.org/10.1016/j.chaos.2007.04.018 doi: 10.1016/j.chaos.2007.04.018
    [5] C. N. Angstmann, B. I. Henry, Generalized fractional power series solutions for fractional differential equations, Appl. Math. Lett. , 102 (2020), 106107. https://doi.org/10.1016/j.aml.2019.106107 doi: 10.1016/j.aml.2019.106107
    [6] A. I. Ali, M. Kalim, A. Khan, Solution of fractional partial differential equations using fractional power series method, Int. J. Differ. Equ. , 2021 (2021), 6385799. https://doi.org/10.1155/2021/6385799 doi: 10.1155/2021/6385799
    [7] M. A. Tashtoush, I. A. Ibrahim, W. M. Taha, M. H. Dawi, A. F. Jameel, E. A. Az-Zo'bi, Various closed-form solitonic wave solutions of conformable higher-dimensional Fokas model in fluids and plasma physics, Iraqi J. Comput. Sci. Math. , 5 (2024), 18. https://doi.org/10.52866/ijcsm.2024.05.03.027 doi: 10.52866/ijcsm.2024.05.03.027
    [8] R. Ur Rahman, Z. Hammouch, A. S. A. Alsubaie, K. H. Mahmoud, A. Alshehri, E. A. Az-Zo'bi, et al., Dynamical behavior of fractional nonlinear dispersive equation in Murnaghan's rod materials, Results Phys. , 56 (2024), 107207. https://doi.org/10.1016/j.rinp.2023.107207 doi: 10.1016/j.rinp.2023.107207
    [9] S. Kheybari, M. T. Darvishi, M. S. Hashemi, A semi-analytical approach to Caputo type time-fractional modified anomalous sub-diffusion equations, Appl. Numer. Math. , 158 (2020), 103–122. https://doi.org/10.1016/j.apnum.2020.07.023 doi: 10.1016/j.apnum.2020.07.023
    [10] S. Kheybari, Numerical algorithm to Caputo type time-space fractional partial differential equations with variable coefficients, Math. Comput. Simulat. , 182 (2021), 66–85. https://doi.org/10.1016/j.matcom.2020.10.018 doi: 10.1016/j.matcom.2020.10.018
    [11] M. S. Hashemi, M. Mirzazadeh, D. Baleanu, Innovative method for computing approximate solutions of non-homogeneous wave equations with generalized fractional derivatives, Contemp. Math. , 4 (2023), 1026–1047. https://doi.org/10.37256/cm.4420233593 doi: 10.37256/cm.4420233593
    [12] S. Javeed, D. Baleanu, A. Waheed, M. S. Khan, H. Affan, Analysis of homotopy perturbation method for solving fractional order differential equations, Mathematics, 7 (2019), 40. https://doi.org/10.3390/math7010040 doi: 10.3390/math7010040
    [13] J. Li, Explicit and structure-preserving exponential wave integrator fourier pseudo-spectral methods for the Dirac equation in the simultaneously massless and nonrelativistic regime, Calcolo, 61 (2024), 3. https://doi.org/10.1007/s10092-023-00554-0 doi: 10.1007/s10092-023-00554-0
    [14] J. Li, L. Zhu, A uniformly accurate exponential wave integrator fourier pseudo-spectral method with structure-preservation for long-time dynamics of the Dirac equation with small potentials, Numer. Algor. , 92 (2023), 1367–1401. https://doi.org/10.1007/s11075-022-01345-4 doi: 10.1007/s11075-022-01345-4
    [15] J. Li, T. Wang, Optimal point-wise error estimate of two conservative fourth-order compact finite difference schemes for the nonlinear Dirac equation, Appl. Numer. Math. , 162 (2021), 150–170. https://doi.org/10.1016/j.apnum.2020.12.010 doi: 10.1016/j.apnum.2020.12.010
    [16] W. Alhejaili, E. A. Az-Zo'bi, R. Shah, S. A. El-Tantawy, On the analytical soliton approximations to fractional forced korteweg–de vries equation arising in fluids and plasmas using two novel techniques, Commun. Theor. Phys. , 76 (2024), 085001. https://doi.org/10.1088/1572-9494/ad53bc doi: 10.1088/1572-9494/ad53bc
    [17] J. Yu, Y. Feng, On the generalized time fractional reaction-diffusion equation: Lie symmetries, exact solutions and conservation laws, Chaos Soliton. Fract. , 182 (2024), 114855. https://doi.org/10.1016/j.chaos.2024.114855 doi: 10.1016/j.chaos.2024.114855
    [18] J. Yu, Y. Feng, Lie symmetry analysis, power series solutions and conservation laws of (2+1)-dimensional time fractional modified Bogoyavlenskii–Schiff equations, J. Nonlinear Math. Phys. , 31 (2024), 27. https://doi.org/10.1007/s44198-024-00195-z doi: 10.1007/s44198-024-00195-z
    [19] J. Yu, Y. Feng, Group classification of time fractional Black–Scholes equation with time-dependent coefficients, Fract. Calc. Appl. Anal. , 27 (2024), 2335–2358. https://doi.org/10.1007/s13540-024-00339-4 doi: 10.1007/s13540-024-00339-4
    [20] P. J. Olver, Applications of Lie groups to differential equations, New York: Springer, 1993. https://doi.org/10.1007/978-1-4612-4350-2
    [21] G. W. Bluman, S. C. Anco, Symmetry and integration methods for differential equations, New York: Springer, 2002. https://doi.org/10.1007/b97380
    [22] N. H. Ibragimov, CRC handbook of lie group analysis of differential equations, New York: CRC press, 1993. https://doi.org/10.1201/9781003419808
    [23] W. Ma, Binary nonlinearization for the Dirac systems, 1995. arXiv: solv-int/9512002v1. https://doi.org/10.48550/arXiv.solv-int/9512002
    [24] W. Ma, K. Li, Virasoro symmetry algebra of Dirac soliton hierarchy, Inverse Probl. , 12 (1996), L25. https://doi.org/10.1088/0266-5611/12/6/001 doi: 10.1088/0266-5611/12/6/001
    [25] E. G. Fan, N-fold Darboux transformation and soliton solutions for a nonlinear Dirac system, J. Phys. A: Math. Gen. , 38 (2005), 1063. https://doi.org/10.1088/0305-4470/38/5/008 doi: 10.1088/0305-4470/38/5/008
    [26] I. S. Frolov, Inverse scattering problem for a Dirac system on the whole axis, Dokl. Akad. Nauk, 207 (1972), 44–47.
    [27] K. Schratz, Y. Wang, X. Zhao, Low-regularity integrators for nonlinear Dirac equations, Math. Comp, 90 (2021), 189–214. https://doi.org/10.1090/mcom/3557 doi: 10.1090/mcom/3557
    [28] F. Zhang, X. Xin, Y. Zhang, Nonlocal symmetries, exact solutions, and conservation laws for the nonlinear Dirac system, Comput. Appl. Math. , 44 (2025), 99. https://doi.org/10.1007/s40314-024-03067-w doi: 10.1007/s40314-024-03067-w
    [29] A. A. Kilbas, H. M. Srivastava, J. J. Trujillo, Theory and applications of fractional differential equations, Amsterdam: Elsevier, 2006.
    [30] K. Diethelm, The analysis of fractional differential equations, New York: Springer, 2010.
    [31] S. G. Samko, Fractional integrals and derivatives, Yverdon: Gordon and Breach, 1993.
    [32] K. S. Miller, B. Ross, An introduction to the fractional calculus and fractional differential equations, Wiley-Interscience, 1993.
    [33] S. Kazem, Exact solution of some linear fractional differential equations by Laplace transform, Int. J. Nonlinear Sci. , 16 (2013), 3–11.
    [34] G. D. Medina, N. R. Ojeda, J. H. Pereira, L.G. Romero, Fractional Laplace transform and fractional calculus, Int. Math. Forum, 12 (2017), 991–1000. https://doi.org/10.12988/imf.2017.71194 doi: 10.12988/imf.2017.71194
    [35] C. A. Monje, Y. Chen, B. M. Vinagre, D. Xue, V. Feliu, Fractional-order systems and controls: Fundamentals and applications, New York: Springer, 2010.
    [36] G. W. Bluman, J. D. Cole, The general similarity solution of the heat equation, J. Math. Mech. , 18 (1969), 1025–1042.
    [37] N. H. Ibragimov, A new conservation theorem, J. Math. Anal. Appl. , 333 (2007), 311–328. https://doi.org/10.1016/j.jmaa.2006.10.078 doi: 10.1016/j.jmaa.2006.10.078
    [38] R. K. Gazizov, N. H. Ibragimov, S. Y. Lukashchuk, Nonlinear self-adjointness, conservation laws and exact solutions of time-fractional Kompaneets equations, Commun. Nonlinear Sci. , 23 (2015), 153–163. https://doi.org/10.1016/j.cnsns.2014.11.010 doi: 10.1016/j.cnsns.2014.11.010
    [39] N. H. Ibragimov, Conservation laws and non-invariant solutions of anisotropic wave equations with a source, Nonlinear Anal.-Real, 40 (2018), 82–94. https://doi.org/10.1016/j.nonrwa.2017.08.005 doi: 10.1016/j.nonrwa.2017.08.005
    [40] M. S. Hashemi, A. Haji-Badali, F. Alizadeh, M. Inc, Classical and non-classical lie symmetry analysis, conservation laws and exact solutions of the time-fractional Chen–Lee–Liu equation, Comput. Appl. Math. , 42 (2023), 73. https://doi.org/10.1007/s40314-023-02217-w doi: 10.1007/s40314-023-02217-w
    [41] M. S. Hashemi, A. Haji-Badali, F. Alizadeh, X. J. Yang, Non-classical lie symmetries for nonlinear time-fractional Heisenberg equations, Math. Method. Appl. Sci. , 45 (2022), 10010–10026. https://doi.org/10.1002/mma.8353 doi: 10.1002/mma.8353
    [42] F. Alizadeh, M. S. Hashemi, A. H. Badali, Lie symmetries, exact solutions, and conservation laws of the nonlinear time-fractional Benjamin–Ono equation, Comput. Method. Differ. Equ. , 10 (2022), 608–616. https://doi.org/10.22034/cmde.2021.45436.1911 doi: 10.22034/cmde.2021.45436.1911
    [43] F. Alizadeh, E. Hincal, K. Hosseini, M. S. Hashemi, A. Das, The (2+1)-dimensional generalized time-fractional Zakharov Kuznetsov Benjamin Bona Mahony equation: its classical and nonclassical symmetries, exact solutions, and conservation laws, Opt. Quant. Electron. , 55 (2023), 1061. https://doi.org/10.1007/s11082-023-05387-3 doi: 10.1007/s11082-023-05387-3
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(308) PDF downloads(40) Cited by(0)

Figures and Tables

Figures(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog