删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

Nonclassical Lie symmetry and conservation laws of the nonlinear time-fractional Korteweg【-逻*辑*与-】nd

本站小编 Free考研考试/2022-01-02

Mir Sajjad Hashemi1, Ali Haji-Badali1, Farzaneh Alizadeh1, Mustafa Inc,2,3,41Department of Mathematics, Basic Science Faculty, University of Bonab, Bonab, Box 55513-95133, Iran
2Department of Computer Engineering, Biruni University, Istanbul, Turkey
3Department of Mathematics, Science Faculty, Firat University, 23119 Elazig, Turkey
4Department of Medical Research, China Medical University Hospital, China Medical University, Taichung, Taiwan, China

Received:2021-04-6Revised:2021-06-9Accepted:2021-06-10Online:2021-08-05


Abstract
In this paper, we use the symmetry of the Lie group analysis as one of the powerful tools that deals with the wide class of fractional order differential equations in the Riemann–Liouville concept. In this study, first, we employ the classical and nonclassical Lie symmetries (LS) to acquire similarity reductions of the nonlinear fractional far field Korteweg–de Vries (KdV) equation, and second, we find the related exact solutions for the derived generators. Finally, according to the LS generators acquired, we construct conservation laws for related classical and nonclassical vector fields of the fractional far field KdV equation.
Keywords: fractional equation;Lie symmetry analysis;classical and non-classical symmetries


PDF (711KB)MetadataMetricsRelated articlesExportEndNote|Ris|BibtexFavorite
Cite this article
Mir Sajjad Hashemi, Ali Haji-Badali, Farzaneh Alizadeh, Mustafa Inc. Nonclassical Lie symmetry and conservation laws of the nonlinear time-fractional Korteweg–de Vries equation. Communications in Theoretical Physics, 2021, 73(9): 095006- doi:10.1088/1572-9494/ac09df

1. Introduction

Hyperelastic materials such as rubber and rubber-like materials have great properties. For instance, some of these properties are the large deformation, high elasticity, and typical characteristics of nonlinearity. Their products are used in transportation, aerospace, petrochemicals, etc frequently. These materials and their products are usually used in specific environments, and all encounter the problems of instability, deformation, destruction, etc. Hence, the dynamic stability of these materials and structures has always been considered [13]. Since then, a lot of remarkable work has been done. For hyperelastic materials the main aspects of nonlinear elastodynamics can be found in Chou-Wang et al [49]. Cohen et al [10] studied the propagation of nonlinear traveling waves in a cylindrical rod composed of a class of compressible Mooney–Rivlin materials and obtained a system of nonlinear evolution equations for describing the symmetrical motion of the rod. By applying the reduced perturbation method, the authors reduced the governing equation to the Korteweg–de Vries (KdV) equation and proved that there exist finite-amplitude waves [11].

Lie group transformations were first put forward by Sophus Lie, who was an influential Norwegian mathematician in the early 19th century. These transformations play a fundamental role in the analysis of different kind of differential equations [1214]. The classical Lie symmetry method (LSM) is utilized for fractional differential equations in [1519]. Various branches of LSM are introduced in order to find additional generators. Some of the most important ones are nonclassical Lie symmetries [2023], $\lambda -$ symmetries [2426], and so on.

Tian [27] efficiently applied Lie symmetry analysis to obtain exact solutions for a fourth-order nonlinear generalized Boussinesq water wave equation. An interesting application of Lie symmetry analysis to solve the time-fractional Fordy–Gibbons equation has been presented by Feng et al [28]. Mao et al [29] used the Lie group method to extract analytical solutions for the (2 + 1)-dimensional chiral nonlinear Schrödinger equation. They employed this method to find the optimal system of the equation. Furthermore, in their work, the similarity reductions were examined based upon the obtained optimal system. Recently, Wang et al [30] studied the KdV equation in (2+1) dimensions by employing Bell’s polynomial to obtain a bilinear representation of the equation. They also constructed the general lump solutions of the equation by considering the resulting bilinear form. Some new analytical solutions of the extended Zakharov–Kuznetsov equation have been evaluated by Wael et al [31] via symmetry reduction. This equation was employed to model the nonlinear plasma dust acoustic waves in a magnetized dusty plasma. Jahangeer et al [32, 33] employed the Lie symmetry analysis technique as the most effective method to obtain some exact explicit solutions for time-fractional partial differential equations. It is necessary to note that they used some similarity reductions for the governing equations, which convert them to ordinary differential equations. Fractional calculus theory has received much attention from researchers and scientists in applied mathematics and physics because of its remarkable properties to accurately describe different abnormal physical events and complex processes in engineering and applied science such that classical calculations were insufficient to explain them [3437]. To investigate fractional calculation theory, several abnormal events have been sampled through integral or functional differential equations in the real world. There is an essential fundamental difference between differential operators of integer order and the Riemann–Liouville (RL) fractional derivatives; the former are local operators, the latter are not. In recent centuries, the theory of fractional calculus and providing various fractional order derivative and integral operators have gained considerable importance among researchers. In fact, they found the fractional order operators to be a powerful instrument which can be used to perform more accurate descriptions of some nonlocal phenomena than the classical integer-order models in the different fields of applied science and engineering. Now, the well-known fields of fractional calculus include control, porous media, electrical networks, electrochemistry, signal processing, and spread of disease [36, 3840].

We investigate the Lie symmetries of the valid asymptotic far field equation for axis symmetry longitudinal strain waves passing through a cylindrical rod obtained by using the reduction perturbation method in [11]. The mentioned cylindrical rod is composed of general noncompressible materials such as isotropic rods that are nonlinear in the real world. This extraction was performed by expanding radial and axial displacement asymptotically in [11]. The longitudinal strain waves explained by the KdV equation of the far field [11] are given by the following equation:$\begin{eqnarray}{ \mathcal F }:\,\displaystyle \frac{{\partial }^{\alpha }}{\partial {t}^{\alpha }}u+\left(\displaystyle \frac{\tau }{4{\pi }_{1}}-1\right){{uu}}_{x}+\left(\displaystyle \frac{\nu }{16\varepsilon }\right){u}_{{xxx}}=0,\end{eqnarray}$where $\tau ,{\pi }_{1}$ and $\nu ,\varepsilon $ are material and structural coefficients, respectively. Assume that in this equation ${\beta }_{1}=\displaystyle \frac{\tau }{4{\pi }_{1}}-1$ and ${\beta }_{2}=\displaystyle \frac{\nu }{16\varepsilon }$. Therefore, β1 is an arbitrary value and β2 is not negative. In the cylindrical rod, the structural constants and combined material are introduced as general incompressible materials, and the waves that travel through such rods are longitudinal strain waves. Equation (1) is obtained by applying the reductive perturbation method in [11]. Lou's direct method and symmetry transformation were used to find the exact solution of equation (1) in [41]. The problems extracted from solid theory are often explained as nonlinear partial differential equations, and various numerical and analytical methods have been employed to study and solve for the unknowns. Here we intend to obtain the exact solutions of equation (1) with one of the most powerful analytical methods (the Lie symmetry analysis method, LSAM).

2. Preliminaries

In this part, we present some basic definitions for the fractional operators and LSAM of fractional partial differential equations (FPDEs). We advise readers who wish to read more on these definitions and other properties of fractional differentiation to see [3436].

2.1. Fractional operators

RL derivative operator of order α, which is described in [36] as:$\begin{eqnarray*}\begin{array}{l}\displaystyle \frac{{\partial }^{\alpha }}{\partial {t}^{\alpha }}f(x,t)\\ =\ \left\{\begin{array}{ll}\displaystyle \frac{{\partial }^{n}}{\partial {t}^{n}}{\displaystyle \int }_{0}^{t}\displaystyle \frac{{\left(t-\sigma \right)}^{n-\alpha -1}f(x,\sigma )}{{\rm{\Gamma }}(n-\alpha )}{\rm{d}}\sigma & n-1\lt \alpha \lt n,\\ \displaystyle \frac{{\partial }^{n}}{\partial {t}^{n}}f(x,t) & \alpha =n,\qquad n\in {\mathbb{N}},\end{array}\right.\end{array}\end{eqnarray*}$where Γ(·) denotes the well-known gamma function.

Erdélyi–Kober fractional integral operators for ${ \mathcal F }(\zeta )$$\begin{eqnarray*}\begin{array}{l}\left({{ \mathcal K }}_{\beta }^{\tau ,\alpha }{ \mathcal F }\right)\ := \\ \left\{\begin{array}{c}\displaystyle \frac{1}{{\rm{\Gamma }}(\alpha )}{\displaystyle \int }_{1}^{\infty }{\left(s-1\right)}^{\alpha -1}{s}^{-(\tau +\alpha )}{ \mathcal F }\left(\zeta {s}^{\tfrac{1}{\beta }}\right){\rm{d}}s,\qquad \alpha \gt 0,\\ { \mathcal F }(\zeta ),\qquad \alpha =0.\end{array}\right.\end{array}\end{eqnarray*}$

Erdélyi–Kober fractional derivative operators for ${ \mathcal F }(\zeta )$$\begin{eqnarray*}\left({{ \mathcal P }}_{\beta }^{\tau ,\alpha }{ \mathcal F }\right):= \displaystyle \prod _{j=0}^{n-1}\left(\tau +j-\frac{1}{\beta }\zeta \frac{{\rm{d}}}{{\rm{d}}\zeta }\right)\left({{ \mathcal K }}_{\beta }^{\tau +\alpha ,n-\alpha }{ \mathcal F }\right)(\zeta ),\end{eqnarray*}$$\begin{eqnarray*}n=\left\{\begin{array}{l}[\alpha ]+1,\qquad \alpha \, {/}\!\!\!\!\!{\in }{\mathbb{N}},\\ \alpha ,\,\qquad \qquad \alpha \in {\mathbb{N}}.\end{array}\right.\end{eqnarray*}$

2.2. LSAM for FPDEs

The following FPDE of order $\alpha \in (0,1)$ is considered as:$\begin{eqnarray}{\rm{\Delta }}\equiv \displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}-\delta (x,t,u,{u}_{x},{u}_{{xx}},{u}_{{xxx}})=0,\end{eqnarray}$where x, t are independent variables, and u is dependent on x, t.

Suppose that equation (31) is invariant with respect to the following one parameter Lie symmetry transformations:$\begin{eqnarray}\begin{array}{rcl}\hat{x} & = & x+\varepsilon \xi (x,t,u)+O({\varepsilon }^{2}),\\ \hat{t} & = & t+\varepsilon \tau (x,t,u)+O({\varepsilon }^{2}),\\ \hat{u} & = & u+\varepsilon \eta (x,t,u)+O({\varepsilon }^{2}),\\ \displaystyle \frac{{\partial }^{\alpha }\hat{u}}{\partial {\hat{t}}^{\alpha }} & = & \displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}+\varepsilon {\eta }^{\alpha ,t}(x,t,u)+O({\varepsilon }^{2}),\\ \displaystyle \frac{\partial \hat{u}}{\partial \hat{x}} & = & \displaystyle \frac{\partial u}{\partial x}+\varepsilon {\eta }^{x}(x,t,u)+O({\varepsilon }^{2}),\\ \displaystyle \frac{{\partial }^{2}\hat{u}}{\partial \hat{{x}^{2}}} & = & \displaystyle \frac{{\partial }^{2}u}{\partial {x}^{2}}+\varepsilon {\eta }^{{xx}}(x,t,u)+O({\varepsilon }^{2}),\\ \displaystyle \frac{{\partial }^{3}\hat{u}}{\partial \hat{{x}^{3}}} & = & \displaystyle \frac{{\partial }^{3}u}{\partial {x}^{3}}+\varepsilon {\eta }^{{xxx}}(x,t,u)+O({\varepsilon }^{2}),\end{array}\end{eqnarray}$where the group parameter is shown by ϵ, and its corresponding infinitesimal generator is$\begin{eqnarray}V=\xi (x,t,u)\displaystyle \frac{\partial }{\partial x}+\tau (x,t,u)\displaystyle \frac{\partial }{\partial t}+\eta (x,t,u)\displaystyle \frac{\partial }{\partial u}.\end{eqnarray}$This vector field must create a symmetry of (31) iff the following is true in invariance conditions:$\begin{eqnarray}{\left.{{\Pr }}^{(\alpha ,3)}V({\rm{\Delta }})\right|}_{{\rm{\Delta }}=0}=0,\end{eqnarray}$where 3 show the highest integer orders of extended infinitesimals in equation (2), and it is worth noting that the extended operator of fractional prolongation ${{\Pr }}^{(\alpha ,3)}V$ is introduced by [15]:$\begin{eqnarray}\begin{array}{l}{{\Pr }}^{(\alpha ,3)}V=V+{\eta }^{\alpha ,t}\displaystyle \frac{\partial }{\partial ({\partial }_{t}^{\alpha }u)}\\ +{\eta }^{x}\displaystyle \frac{\partial }{\partial {u}_{x}}+{\eta }^{{xx}}\displaystyle \frac{\partial }{\partial {u}_{{xx}}}+{\eta }^{{xxx}}\displaystyle \frac{\partial }{\partial {u}_{{xxx}}},\end{array}\end{eqnarray}$where ${\eta }^{x},{\eta }^{{xx}}$ and ${\eta }^{{xxx}}$ represent integer-order extended infinitesimals which are defined as:$\begin{eqnarray}\begin{array}{rcl}{\eta }^{x} & = & {D}_{x}\eta -{D}_{x}(\xi ){u}_{x}-{D}_{x}(\tau ){u}_{t},\\ {\eta }^{{xx}} & = & {D}_{x}{\eta }^{x}-{D}_{x}(\xi ){u}_{{xx}}-{D}_{x}(\tau ){u}_{{xt}},\\ {\eta }^{{xxx}} & = & {D}_{x}{\eta }^{{xx}}-{D}_{x}(\xi ){u}_{{xxx}}-{D}_{x}(\tau ){u}_{{xxt}},\end{array}\end{eqnarray}$and the operator of the total derivative along x, i.e, Dx, is:$\begin{eqnarray}{D}_{x}=\displaystyle \frac{\partial }{\partial x}+{u}_{x}\displaystyle \frac{\partial }{\partial u}+{u}_{{xx}}\displaystyle \frac{\partial }{\partial {u}_{x}}+\cdots .\end{eqnarray}$Moreover, the α-order extended infinitesimal operator ${\eta }^{\alpha ,t}$ is (see [42, 43])$\begin{eqnarray}\begin{array}{rcl}{\eta }^{\alpha ,t} & = & {D}_{t}^{\alpha }\eta +{\xi }_{1}{D}_{t}^{\alpha }({u}_{x})-{D}_{t}^{\alpha }(\xi {u}_{x})\\ & & +{D}_{t}^{\alpha }\left({D}_{t}(\tau )u\right)-{D}_{t}^{\alpha +1}(\tau u)+\tau {D}_{t}^{\alpha +1}(u),\end{array}\end{eqnarray}$where ${D}_{t}^{\alpha }$ represents the total α-order fractional derivative in a course of time. We recall the Leibnitz formula that is given by the following formula [35, 36]:$\begin{eqnarray}{D}_{t}^{\alpha }\left(f(t)g(t)\right)=\sum _{n=0}^{\infty }\left(\begin{array}{c}\alpha \\ n\end{array}\right){D}_{t}^{\alpha -n}f(t){D}_{t}^{n}g(t),\qquad \alpha \gt 0,\end{eqnarray}$where$\begin{eqnarray*}\left(\begin{array}{c}\alpha \\ n\end{array}\right)=\displaystyle \frac{{\left(-1\right)}^{n-1}\alpha {\rm{\Gamma }}(n-\alpha )}{{\rm{\Gamma }}(1-\alpha ){\rm{\Gamma }}(n+1)}.\end{eqnarray*}$Using equation (10), equation (9) can be written as$\begin{eqnarray}\begin{array}{rcl}{\eta }^{\alpha ,t} & = & {D}_{t}^{\alpha }\eta -\alpha {D}_{t}(\tau )\displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}\\ & & \displaystyle -\displaystyle \sum _{n=1}^{\infty }\left(\begin{array}{c}\alpha \\ n\end{array}\right){D}_{t}^{\alpha -n}({u}_{x}){D}_{t}^{n}(\xi )\\ & & -\displaystyle \sum _{n=1}^{\infty }\left(\begin{array}{c}\alpha \\ n+1\end{array}\right){D}_{t}^{\alpha -n}(u){D}_{t}^{n+1}(\tau ).\end{array}\end{eqnarray}$We want to be reminded that the generalization of the well-known chain rule for a composite function [34] is given by$\begin{eqnarray}\frac{{d}^{m}g\left(y(t)\right)}{{{dt}}^{m}}=\displaystyle \sum _{k=0}^{m}\displaystyle \sum _{r=0}^{k}\left(\begin{array}{c}k\\ r\end{array}\right)\frac{1}{k!}{[-y(t)]}^{r}\frac{{{\rm{d}}}^{m}}{{\rm{d}}{t}^{m}}[y{\left(t\right)}^{k-r}]\frac{{{\rm{d}}}^{k}g(y)}{{\rm{d}}{y}^{k}}.\end{eqnarray}$Using the chain rule (12) along with the generalized Leibnitz rule (10), it is assumed that f(t)=1, and we can write the first term ${D}_{t}^{\alpha }(\eta )$ on the right hand side of (11) as$\begin{eqnarray}\begin{array}{l}{D}_{t}^{\alpha }(\eta )=\displaystyle \frac{{\partial }^{\alpha }\eta }{\partial {t}^{\alpha }}+{\eta }_{u}\displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}-u\displaystyle \frac{{\partial }^{\alpha }{\eta }_{u}}{\partial {t}^{\alpha }}\\ +\displaystyle \sum _{n=1}^{\infty }\left(\begin{array}{c}\alpha \\ n\end{array}\right)\displaystyle \frac{{\partial }^{n}{\eta }_{u}}{\partial {t}^{n}}{D}_{t}^{\alpha -n}(u)+\mu ,\end{array}\end{eqnarray}$where$\begin{eqnarray*}\begin{array}{l}\mu =\displaystyle \sum _{n=2}^{\infty }\displaystyle \sum _{m=2}^{n}\displaystyle \sum _{k=2}^{m}\displaystyle \sum _{r=0}^{k-1}\left(\begin{array}{c}\alpha \\ n\end{array}\right)\left(\begin{array}{c}n\\ m\end{array}\right)\left(\begin{array}{c}k\\ r\end{array}\right)\\ \times \ \displaystyle \frac{{t}^{n-\alpha }}{k!{\rm{\Gamma }}(n-\alpha +1)}{\left(-1\right)}^{r}{u}^{r}\displaystyle \frac{{\partial }^{m}({u}^{k-r})}{\partial {t}^{m}}\displaystyle \frac{{\partial }^{n-m+k}\eta }{\partial {t}^{n-m}\partial {u}^{k}},\end{array}\end{eqnarray*}$Then, the α-order extension of infinitesimal ${\eta }^{\alpha ,t}$ can be obtained as$\begin{eqnarray}\begin{array}{rcl}{\eta }^{\alpha ,t} & = & \displaystyle \frac{{\partial }^{\alpha }\eta }{\partial {t}^{\alpha }}+\left({\eta }_{u}-\alpha {D}_{t}(\tau )\right)\displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}-u\displaystyle \frac{{\partial }^{\alpha }{\eta }_{u}}{\partial {t}^{\alpha }}\\ & & +\displaystyle \sum _{n=1}^{\infty }\left[\left(\begin{array}{c}\alpha \\ n\end{array}\right)\displaystyle \frac{{\partial }^{n}{\eta }_{u}}{\partial {t}^{n}}-\left(\begin{array}{c}\alpha \\ n+1\end{array}\right){D}_{t}^{n+1}(\tau )\right]{D}_{t}^{\alpha -n}(u)\\ & & -\displaystyle \sum _{n=1}^{\infty }\left(\begin{array}{c}\alpha \\ n\end{array}\right){D}_{t}^{n}(\xi ){D}_{t}^{\alpha -n}({u}_{x})+\mu .\end{array}\end{eqnarray}$

Furthermore, since η is linear wrt variable u, then derivatives $\displaystyle \frac{{\partial }^{n}\eta }{\partial {u}^{n}}$ for $n\geqslant 2$ are vanished and immediately we conclude that μ=0. By substituting the above equations in (6) and solving the determining equations we obtain the classical symmetry generators by comparing the coefficients of ${u}_{x},{u}_{{xx}},{u}_{{xt}},\cdots $. In this method we lose some invariance. Bluman and Cole presented the nonclassical method of reduction to obtain new exact solutions of the heat equation [44]. The main idea of this method is based on inserting an invariant surface condition such that combination of the original determining equation and this condition yield a nonlinear determining equation for infinitesimals. We can solve this new nonlinear determining equation for infinitesimals too and get more symmetries.

To obtain the solutions u=u(x, t) of equation (2) which are invariant under the group transformations (3), we must consider$\begin{eqnarray}{\rm{\Omega }}:\xi {u}_{x}+\tau {u}_{t}-\eta =0.\end{eqnarray}$For the nonclassical method, we are looking for invariability of the two following equations simultaneously:$\begin{eqnarray}{\left.{{\Pr }}^{(\alpha ,3)}V({\rm{\Delta }})\right|}_{{\rm{\Delta }}=0,{\rm{\Omega }}=0}=0,\end{eqnarray}$$\begin{eqnarray}{\left.{{\Pr }}^{(\alpha ,3)}V({\rm{\Omega }})\right|}_{{\rm{\Delta }}=0,{\rm{\Omega }}=0}=0.\end{eqnarray}$In order that equation (17) is satisfied for every ξ, τ and η, it is sufficient that Ω=0. Therefore, we consider only the invariant surface condition$\begin{eqnarray}{\rm{\Omega }}:\xi (x,t,u){u}_{x}+\tau (x,t,u){u}_{t}-\eta (x,t,u)=0,\end{eqnarray}$with equation (16). There is no loss of generality in assuming ξ=1 and ξ=0. When ξ=1 we have two conditions: τ=0 and τu=0. In the case ξ=0, where τ is opposite to zero, the acquired system of determining equations is usually complex and difficult to calculate.

3. Implementation of LSAM for time-fractional KdV equation

The time-fractional far field KdV equation is in the following form:$\begin{eqnarray}{ \mathcal F }:\,\displaystyle \frac{{\partial }^{\alpha }}{\partial {t}^{\alpha }}u+{\beta }_{1}{{uu}}_{x}+{\beta }_{2}{u}_{{xxx}}=0,\end{eqnarray}$where ${\beta }_{1}=\displaystyle \frac{\tau }{4{\pi }_{1}}-1$ and ${\beta }_{2}=\displaystyle \frac{\nu }{16\varepsilon }$.

3.1. Classical symmetries

If we apply prolongation of the fractional vector field to the this equation, we obtain$\begin{eqnarray}{\eta }^{\alpha }+{\beta }_{1}{u}_{x}\eta +{\beta }_{1}u{\eta }^{x}+{\beta }_{2}{\eta }^{{xxx}}=0,\end{eqnarray}$and by substituting (7) and (14) in (20) and solving the determining equations we acquire the following symmetry generators:$\begin{eqnarray}{X}_{1}=\displaystyle \frac{\partial }{\partial x},{X}_{2}=\alpha x\displaystyle \frac{\partial }{\partial x}+3t\displaystyle \frac{\partial }{\partial t}-2\alpha u\displaystyle \frac{\partial }{\partial u}.\end{eqnarray}$The invariant solution of vector field X1 is $u(x,t)=f(t)$, and then equation (19) is reduced to the following time-fractional ODE equation:$\begin{eqnarray}\frac{{\partial }^{\alpha }}{\partial {t}^{\alpha }}f(t)=0.\end{eqnarray}$Hence, the exact solution of equation (19) is $u(x,t)={c}_{1}{t}^{\alpha -1}$.

According to the vector field X2 we have the following invariant solution for equation (19):$\begin{eqnarray}u(x,t)={t}^{\tfrac{-2\alpha }{3}}f(\zeta ),\zeta ={{xt}}^{\tfrac{-\alpha }{3}},\end{eqnarray}$and by using the above transformation and substituting in equation (19) we obtain the fractional ODE$\begin{eqnarray}\left({{ \mathcal P }}_{\displaystyle \frac{3}{\alpha }}^{1-\tfrac{5\alpha }{3},\alpha }f\right)(\zeta )+{\beta }_{1}f(\zeta )f^{\prime} (\zeta )+{\beta }_{2}f\prime\prime\prime (\zeta )=0.\end{eqnarray}$We can find the exact solution of (24) by using a power series method. Let$\begin{eqnarray}f(\zeta )=\sum _{n=0}^{\infty }{a}_{n}{\zeta }^{n},\end{eqnarray}$then,$\begin{eqnarray}\begin{array}{rcl}f^{\prime} (\zeta ) & = & \displaystyle \sum _{n=0}^{\infty }(n+1){a}_{n+1}{\zeta }^{n},\\ f\prime\prime\prime (\zeta ) & = & \displaystyle \sum _{n=0}^{\infty }(n+2)(n+1){a}_{n+2}{\zeta }^{n},\\ \left({{ \mathcal P }}_{\beta }^{\tau ,\alpha }f\right)(\zeta ) & = & \displaystyle \sum _{n=0}^{\infty }{a}_{n}\displaystyle \frac{{\rm{\Gamma }}(\tau -\tfrac{n}{\beta }+1)}{{\rm{\Gamma }}(\tau -\tfrac{n}{\beta }+1-\alpha )}{\zeta }^{n},\end{array}\end{eqnarray}$and substituting the above relation in (24) we have$\begin{eqnarray}\begin{array}{l}\displaystyle \sum _{n=0}^{\infty }{a}_{n}\displaystyle \frac{{\rm{\Gamma }}(1-\tfrac{5}{3}-\tfrac{n\alpha }{3}+1)}{{\rm{\Gamma }}(1-\tfrac{8\alpha }{3}-\tfrac{n\alpha }{3}+1)}{\zeta }^{n}\\ +{\beta }_{1}\displaystyle \sum _{n=0}^{\infty }\displaystyle \sum _{k=0}^{n}(n-k+1){a}_{k}{a}_{n-k+1}{\zeta }^{n}\\ +{\beta }_{2}\displaystyle \sum _{n=0}^{\infty }(n+1)(n+2)(n+3){a}_{n+3}{\zeta }^{n}=0.\end{array}\end{eqnarray}$According to equation (27), for n≥0 we have the following recursive relation:$\begin{eqnarray}\begin{array}{l}{a}_{n+3}=\displaystyle \frac{-1}{(n+1)(n+2)(n+3){\beta }_{2}}\\ \times \ \left({a}_{n}\displaystyle \frac{{\rm{\Gamma }}(2-\tfrac{5\alpha }{3}-\tfrac{n\alpha }{3})}{{\rm{\Gamma }}(2-\tfrac{8\alpha }{3}-\tfrac{n\alpha }{3})}+{\beta }_{1}\displaystyle \sum _{k=0}^{n}(n-k+1){a}_{k}{a}_{n-k+1}\right),\end{array}\end{eqnarray}$and ${a}_{0},{a}_{1}$ and a2 are arbitrary constants. According to (23), the exact solution of equation (19) is$\begin{eqnarray}u(x,t)={t}^{\tfrac{-2\alpha }{3}}\sum _{n=0}^{\infty }{a}_{n}{\left({{xt}}^{\tfrac{-\alpha }{3}}\right)}^{n}.\end{eqnarray}$In order to illustrate the related plots of the obtained exact solution (29), we truncate its series with N=24. Figure 1 shows the plots of u(x, t) for a0=1, a1=0.5, a2=0.25 and β1=β2=1. Figure 1(a) shows the convergence properties of the obtained exact solutions to the exact solution of the integer-order KdV equation when α tends to 1.

Figure 1.

New window|Download| PPT slide
Figure 1.Related plots of the exact solutions of (29) obtained by the classical generator of X2 for different values of α in 2D and 3D.


3.2. Non-classical symmetries

In this case for ξ=1 and τ=0 we have$\begin{eqnarray}{X}_{3}=x\displaystyle \frac{\partial }{\partial x}+u\displaystyle \frac{\partial }{\partial u},\end{eqnarray}$and the invariant solution according to this vector field is$\begin{eqnarray}u(x,t)={xf}(t).\end{eqnarray}$Substituting (31) into equation (31) yields$\begin{eqnarray}\displaystyle \frac{{\partial }^{\alpha }}{\partial {t}^{\alpha }}f(t)+{\beta }_{1}f{\left(t\right)}^{2}=0,\end{eqnarray}$and therefore, the solution of (32) is$\begin{eqnarray}f(t)=-\displaystyle \frac{{\rm{\Gamma }}(1-\alpha )}{{\beta }_{1}{\rm{\Gamma }}(1-2\alpha )}{t}^{-\alpha },\end{eqnarray}$so the exact solution of (19) is as follows:$\begin{eqnarray}u(x,t)=-\displaystyle \frac{{\rm{\Gamma }}(1-\alpha )}{{\beta }_{1}{\rm{\Gamma }}(1-2\alpha )}{{xt}}^{-\alpha }.\end{eqnarray}$Figure 2 shows the plot of u(x, t) for β1=β2=1 and α=0.75 in 3D.

Figure 2.

New window|Download| PPT slide
Figure 2.Related plot of the exact solution of (34) obtained by the nonclassical generator of X3 for α=0.75 in 3D.


For ξ=1, $\tau \ne 0$ and τu=0 we have the following vector fields:$\begin{eqnarray}\begin{array}{rcl}{X}_{4} & = & (\alpha x+\mu )\displaystyle \frac{\partial }{\partial x}+3t\displaystyle \frac{\partial }{\partial t}-2\alpha u\displaystyle \frac{\partial }{\partial u},\\ {X}_{5} & = & \displaystyle \frac{\partial }{\partial x}+\lambda \displaystyle \frac{\partial }{\partial t}.\end{array}\end{eqnarray}$By the infinitesimal generator X4, we obtain$\begin{eqnarray}u(x,t)={t}^{\tfrac{-2\alpha }{3}}f(\zeta ),\zeta =\displaystyle \frac{(\alpha x+\mu )}{\alpha }{t}^{\tfrac{-\alpha }{3}},\end{eqnarray}$and then, equation (19) can be reduced to the following fractional ODE:$\begin{eqnarray}\left({{ \mathcal P }}_{\displaystyle \frac{3}{\alpha }}^{1-\tfrac{5\alpha }{3},\alpha }f\right)(\zeta )+{\beta }_{1}f(\zeta )f^{\prime} (\zeta )+{\beta }_{2}f\prime\prime\prime (\zeta )=0,\end{eqnarray}$so in the same manner used for the infinitesimal generator X2, we have the following exact solution:$\begin{eqnarray}u(x,t)={t}^{\tfrac{-2\alpha }{3}}\sum _{n=0}^{\infty }{a}_{n}{\left(\displaystyle \frac{\alpha x+\mu }{\alpha }{t}^{\tfrac{-\alpha }{3}}\right)}^{n},\end{eqnarray}$where μ is a constant.

In order to illustrate the related plots of the computed exact solution (38), we truncate its series with N=24. Figure 3 shows the plots of u(x, t) for a0=1, a1=0.5, a2=0.25, μ=0.5 and β1=β2=1. Figure 3 (a) shows the convergence properties of the obtained exact solutions to the exact solution of the integer-order KdV equation when α approaches 1.

Figure 3.

New window|Download| PPT slide
Figure 3.Related plots of the exact solutions of (38) obtained by the nonclassical generator of X4 for different values of α in 2D and 3D.


The invariant solution of the infinitesimal generator X5$\begin{eqnarray}u(x,t)=f(\zeta ),\zeta =x-\lambda t\end{eqnarray}$can reduce the equation (19) to the time-fractional ODE$\begin{eqnarray}{\left(-1\right)}^{\alpha }\frac{{\partial }^{\alpha }}{\partial {\zeta }^{\alpha }}f(\zeta )+{\beta }_{1}f(\zeta )f^{\prime} (\zeta )+{\beta }_{2}f\prime\prime\prime (\zeta )=0,\end{eqnarray}$and it is not easy to find the exact solution of this equation.

4. Conservation laws of the KdV equation

In this part, we turn attention to the nonlinear self-adjointness [45] and new conservation laws (cl) theorem that was first used to construct the cl by Ibragimov in [46]. This theorem uses the LS to find conserved vectors. Here we intend to construct cl for (19) according to this method. The cl for equation (19) is given by$\begin{eqnarray}{D}_{t}{{ \mathcal T }}_{i}^{t}+{D}_{x}{{ \mathcal T }}_{i}^{x}=0,\end{eqnarray}$where ${{ \mathcal T }}_{i}^{t}={{ \mathcal T }}_{i}^{t}(t,x,u,...)$, ${{ \mathcal T }}_{i}^{x}={{ \mathcal T }}_{i}^{x}(t,x,u,...)$ are the conserved vectors$\begin{eqnarray}\begin{array}{rcl}{{ \mathcal T }}_{i}^{x} & = & \left({}^{u}{{ \mathcal W }}_{i}\displaystyle \frac{\delta { \mathcal H }}{\delta {u}_{x}}+\displaystyle \sum _{k\geqslant 1}{D}_{x}...{D}_{x}({}^{u}{{ \mathcal W }}_{i})\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{(k+1)x}}\right),\\ {{ \mathcal T }}_{i}^{t} & = & \displaystyle \sum _{k=0}^{n-1}{\left(-1\right)}^{k}{\partial }_{t}^{\alpha -1-k}({}^{u}{{ \mathcal W }}_{i}){D}_{t}^{k}\left(\displaystyle \frac{\partial { \mathcal H }}{\partial ({\partial }_{t}^{\alpha }u)}\right)\\ & & -{\left(-1\right)}^{n}{ \mathcal J }\left({}^{u}{{ \mathcal W }}_{i},{D}_{t}^{n}\left(\displaystyle \frac{\partial { \mathcal H }}{\partial ({\partial }_{t}^{\alpha }u)}\right)\right),\qquad n=[\alpha ]+1,\end{array}\end{eqnarray}$in which ${}^{u}{{ \mathcal W }}_{i}={\eta }_{i}-{\xi }_{i}{u}_{x}-{\tau }_{i}{u}_{t}$ and the integral ${ \mathcal J }$ is given by$\begin{eqnarray}{ \mathcal J }(h,g)={\int }_{0}^{t}{\int }_{t}^{T}\frac{h(\lambda ,x)g(\mu ,x){\left(\mu -\lambda \right)}^{n-(\alpha +1)}}{{\rm{\Gamma }}(n-\alpha )}{\rm{d}}\lambda {\rm{d}}\mu .\end{eqnarray}$The formal Lagrangian for (19) is given by$\begin{eqnarray}{ \mathcal H }={\rm{\Phi }}(x,t)\left(\displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}+{\beta }_{1}{{uu}}_{x}+{\beta }_{2}{u}_{{xxx}}\right),\end{eqnarray}$where ${\rm{\Phi }}(x,t)={\rm{\Psi }}(x,t,u)$ is a new dependent variable and the adjoint equations of the fractional KdV equation can be specified as follows:$\begin{eqnarray}{{ \mathcal F }}^{* }\equiv \displaystyle \frac{\delta { \mathcal H }}{\delta u}=0,\end{eqnarray}$where the Euler–Lagrange operators are given by$\begin{eqnarray*}\displaystyle \frac{\delta }{\delta u}=\displaystyle \frac{\partial }{\partial u}+{\left({\partial }_{t}^{\alpha }\right)}^{* }\displaystyle \frac{\partial }{\partial ({\partial }_{t}^{\alpha }u)}+\sum _{k\geqslant 1}{\left(-1\right)}^{k}{D}_{x}...{D}_{x}\displaystyle \frac{\partial }{\partial {u}_{{kx}}},\end{eqnarray*}$where ${\left({\partial }_{t}^{\alpha }\right)}^{* }$ denotes the adjoint operator for ${\partial }_{t}^{\alpha }$. According to the RL fractional differential operator$\begin{eqnarray*}\begin{array}{rcc}{\left({\partial }_{t}^{\alpha }\right)}^{* } & = & {\left(-1\right)}^{n}{{ \mathcal I }}_{T}^{n-\alpha }({\partial }_{t}^{n})={\left({\partial }_{T}^{\alpha }\right)}_{t}^{{ \mathcal C }},\\ {{ \mathcal I }}_{T}^{n-\alpha }h(t,x) & = & {\int }_{t}^{T}\frac{h(\tau ,x){\left(\tau -t\right)}^{n-(1+\alpha )}}{{\rm{\Gamma }}(n-\alpha )}{\rm{d}}\tau ,\\ n & = & [\alpha ]+1,\end{array}\end{eqnarray*}$where ${\left({\partial }_{T}^{\alpha }\right)}_{t}^{{ \mathcal C }}$ is the operator of the right-sided Caputo derivative.

If we have the following relation for the time-fractional nonlinear equation (44), then we can say that equation (19) is self-adjoint:$\begin{eqnarray}{{ \mathcal F }}^{* }\equiv \displaystyle \frac{\delta { \mathcal H }}{\delta u}=\mu F,\end{eqnarray}$where μ remains to be determined. Thus, we can write the nonlinear self-adjoint condition (46) as follows:$\begin{eqnarray*}\mu =0,\qquad {\rm{\Psi }}(x,t,u)=A,\qquad A\in {\mathbb{R}}.\end{eqnarray*}$Hence, if we suppose A=1, then$\begin{eqnarray}{ \mathcal H }=\displaystyle \frac{{\partial }^{\alpha }u}{\partial {t}^{\alpha }}+{\beta }_{1}{{uu}}_{x}+{\beta }_{2}{u}_{{xxx}}.\end{eqnarray}$According to the above analysis and Lie symmetry generator, we consider the conserved vectors for classical and nonclassical generators of the time-fractional far field KdV equation. We have the following cases for classical generators:

Case 1: Here for the first classical generator X1, the respective Lie characteristic function is$\begin{eqnarray}{}^{u}{{ \mathcal W }}_{1}=-{u}_{x}.\end{eqnarray}$Substituting (48) into (42) yields the conserved vector as follows:$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{1}^{x} & = & {}^{u}{{ \mathcal W }}_{1}\left(\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{x}}-{D}_{x}\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{{xx}}}+{D}_{x}^{2}\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{{xxx}}}\right)\\ & & +{D}_{x}({}^{u}{{ \mathcal W }}_{1})\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{{xx}}}+{D}_{x}^{2}({}^{u}{{ \mathcal W }}_{1})\displaystyle \frac{\partial { \mathcal H }}{\partial {u}_{{xxx}}},\\ {{ \mathcal T }}_{1}^{t} & = & {{ \mathcal I }}^{1-\alpha }({}^{u}{{ \mathcal W }}_{1}){\rm{\Psi }}+{ \mathcal J }\left({}^{u}{{ \mathcal W }}_{1},{{\rm{\Psi }}}_{t}\right),\end{array}\end{eqnarray*}$then$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{1}^{x} & = & -{u}_{x}({\beta }_{1}u+{\beta }_{2})+{\beta }_{2}{u}_{{xx}}+{\beta }_{2}{u}_{{xxx}}\\ {{ \mathcal T }}_{1}^{t} & = & -{{ \mathcal I }}^{1-\alpha }{u}_{x}{\rm{\Psi }}+{ \mathcal J }\left(-{u}_{x},{{\rm{\Psi }}}_{t}\right).\end{array}\end{eqnarray*}$

Case 2: For the generator ${X}_{2}=\alpha x\displaystyle \frac{\partial }{\partial x}\,+3t\displaystyle \frac{\partial }{\partial t}\,-2\alpha u\displaystyle \frac{\partial }{\partial u}$, the respective Lie characteristic function is obtained by$\begin{eqnarray}{}^{u}{{ \mathcal W }}_{2}=-2\alpha u-\alpha {{xu}}_{x}-3{{tu}}_{t}.\end{eqnarray}$Substituting (49) into (42) yields$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{2}^{x} & = & -\left({\beta }_{1}u+{\beta }_{2}\right)\left(2\alpha u+\alpha {{xu}}_{x}+3{{tu}}_{t}\right)\\ & & +{\beta }_{2}\left(3\alpha {u}_{x}+\alpha {{xu}}_{{xx}}+3{{tu}}_{{xt}}\right)\\ & & -{\beta }_{2}\left(4\alpha {u}_{{xx}}+\alpha {{xu}}_{{xxx}}+3{{tu}}_{{xxt}}\right),\\ {{ \mathcal T }}_{2}^{t} & = & -{{ \mathcal I }}^{1-\alpha }(2\alpha u+\alpha {{xu}}_{x}+3{{tu}}_{t}).\end{array}\end{eqnarray*}$For nonclassical generators:

Case 3: For ξ=1 and τ=0 we have ${X}_{3}=x\displaystyle \frac{\partial }{\partial x}\,+u\displaystyle \frac{\partial }{\partial u}$, and the respective Lie characteristic function is obtained by$\begin{eqnarray}{}^{u}{{ \mathcal W }}_{3}=u-{{xu}}_{x}.\end{eqnarray}$Hence, substituting (50) into (42) yields the conserved vector as follows:$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{3}^{x} & = & \left({\beta }_{1}u+{\beta }_{2}\right)\left(u-{{xu}}_{x}\right)+{\beta }_{2}{{xu}}_{x}-{\beta }_{2}\left({u}_{{xx}}+{{xu}}_{{xxx}}\right)\\ {{ \mathcal T }}_{3}^{t} & = & {{ \mathcal I }}^{1-\alpha }(u-{{xu}}_{x}).\end{array}\end{eqnarray*}$

Case 4: For ξ=1 and $\tau \ne 0$, we have ${X}_{4}=(\alpha x\,+\mu )\displaystyle \frac{\partial }{\partial x}+3t\displaystyle \frac{\partial }{\partial t}-2\alpha u\displaystyle \frac{\partial }{\partial u}$. The respective Lie characteristic function is obtained by$\begin{eqnarray}{}^{u}{{ \mathcal W }}_{4}=-2\alpha u-(\alpha x+\mu ){u}_{x}-3{{tu}}_{t}.\end{eqnarray}$The corresponding conserved vectors are$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{4}^{x} & = & -\left({\beta }_{1}u+{\beta }_{2}\right)\left(2\alpha u+\left(\alpha x+\mu \right){u}_{x}+3{{tu}}_{t}\right)\\ & & +{\beta }_{2}\left(3\alpha {u}_{x}+(\alpha x+\mu ){u}_{{xx}}+3{{tu}}_{{xt}}\right)\\ & & -{\beta }_{2}\left(4\alpha {u}_{{xx}}+(\alpha x+\mu ){u}_{{xxx}}+3{{tu}}_{{xxt}}\right),\\ {{ \mathcal T }}_{4}^{t} & = & -{{ \mathcal I }}^{1-\alpha }(2\alpha u+(\alpha x+\mu ){u}_{x}+3{{tu}}_{t}).\end{array}\end{eqnarray*}$

Case 5: The respective Lie characteristic function for ${X}_{5}=\displaystyle \frac{\partial }{\partial x}+\lambda \displaystyle \frac{\partial }{\partial t}$ is obtained by$\begin{eqnarray}{}^{u}{{ \mathcal W }}_{5}=-{u}_{x}-\lambda {u}_{t}.\end{eqnarray}$The corresponding conserved vectors are$\begin{eqnarray*}\begin{array}{rcl}{{ \mathcal T }}_{5}^{x} & = & -\left({\beta }_{1}u+{\beta }_{2}\right)\left({u}_{x}+\lambda {u}_{t}\right)\\ & & +{\beta }_{2}\left({u}_{{xx}}+\lambda {u}_{{xt}}\right)-{\beta }_{2}\left({u}_{{xxx}}+\lambda {u}_{{xxt}}\right),\\ {{ \mathcal T }}_{5}^{t} & = & -{{ \mathcal I }}^{1-\alpha }({u}_{x}+\lambda {u}_{t}).\end{array}\end{eqnarray*}$

5. Conclusion

In this work, we mainly study the classical Lie symmetry, nonclassical Lie symmetry and conservation laws of the nonlinear time-fractional KdV equation. We use the symmetry of the Lie group analysis for the nonlinear time-fractional far field KdV equation and employ classical and nonclassical Lie symmetry to acquire similarity reductions of this equation according to the obtained vector fields and the corresponding invariant solutions. We also find the related exact solutions by some methods like the power series method for the derived generators. From figures 1 and 3, one may find that our obtained solutions in both the classical and nonclassical cases have convergence behavior. Finally, according to the LS generators acquired, we construct conservation laws for related classical and nonclassical vector fields of the mentioned equation. From these findings, we conclude that the presented method might become a promising one in finding the analytical solutions of different kinds of fractional partial differential equations.

Acknowledgments

The authors would like to thank the respected referees for their valuable comments that led to improvement of this paper.


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

Beatty M F 1987 Topics in finite elasticity: hyper-elasticity of rubber, elastomers, and biological tissues – with examples
Appl. Mech. Rev. 40 1699 1734

DOI:10.1115/1.3149545 [Cited within: 1]

Ou Z Yao X Zhang X Fan X 2014 Wrinkling analysis in a film bonded to a compressible compliant substrate in large deformation
Comput. Mater. Continua 44 205 221



Fu Y Ogden R W 2001 Nonlinear Elasticity: Theory and Applications UK Cambridge University Press
[Cited within: 1]

Chou-Wang M S O Horgan C O 1989 Cavitation in nonlinear elastodynamics for neo-Hookean materials
Int. J. Eng. Sci. 27 967 973

DOI:10.1016/0020-7225(89)90037-2 [Cited within: 1]

Yuan X Zhang R 2006 Nonlinear dynamical analysis in incompressible transversely isotropic nonlinearly elastic materials: Cavity formation and motion in solid spheres
Comput. Mater. Continua 3 119 129

DOI:10.3970/cmc.2006.003.119

Yuan X Zhang H 2009 Effects of constitutive parameters and dynamic tensile loads on radially periodic oscillation of micro-void centered at incompressible hyperelastic spheres
Comput. Model. Eng. Sci. 40 201

DOI:10.3970/cmes.2009.040.201

Yuan X Zhang R Zhang H 2008 Controllability conditions of finite oscillations of hyper-elastic cylindrical tubes composed of a class of Ogden material models
Comput. Mater. Continua 7 155 156



Akbarov S D Ipek C 2012 Dispersion of axisymmetric longtudinal waves in a pre-strained imperfectly bonded bi-layered hollow cylinder
Comput. Mater. Continua 30 99

DOI:10.3970/cmc.2012.030.099

Negin M Akbarov S D Erguven M E 2014 Generalized Rayleigh wave dispersion analysis in a pre-stressed elastic stratified half-space with imperfectly bonded interfaces
Comput. Mater. Continua 42 25 61

[Cited within: 1]

Cohen H Dai H 1993 Nonlinear axisymmetric waves in compressible hyperelastic rods: Long finite amplitude waves
Acta Mech. 100 223 239

DOI:10.1007/BF01174791 [Cited within: 1]

Dai H-H Huo Y 2002 Asymptotically approximate model equations for nonlinear dispersivewaves in incompressible elastic rods
Acta Mech. 157 97 112

DOI:10.1007/BF01182157 [Cited within: 5]

Bluman G Anco S 2008 Symmetry and Integration Methods for Differential Equations New York Springer 154
[Cited within: 1]

Ibragimov N H 1995 CRC Handbook of Lie Group Analysis of Differential Equations Boca Raton, FL CRC Press 3


Olver P J 2000 Applications of Lie Groups to Differential Equations New York Springer 107
[Cited within: 1]

Hashemi M S Baleanu D 2020 Lie Symmetry Analysis of Fractional Differential Equations Boca Raton, FL CRC Press
[Cited within: 2]

Yaşar E Yıldırım Y Khalique C M 2016 Lie symmetry analysis, conservation laws and exact solutions of the seventh-order time fractional sawada-kotera-ito equation
Results Phys. 6 322 328

DOI:10.1016/j.rinp.2016.06.003

Bakkyaraj T Sahadevan R 2015 Invariant analysis of nonlinear fractional ordinary differential equations with Riemann–Liouville fractional derivative
Nonlinear Dyn. 80 447 455

DOI:10.1007/s11071-014-1881-4

Hashemi M S Inc M Bayram M 2019 Symmetry properties and exact solutions of the time fractional Kolmogorov–Petrovskii–Piskunov equation
Rev. Mex. Fís. 65 529 535

DOI:10.31349/RevMexFis.65.529

Wang G W Hashemi M S 2017 Lie symmetry analysis and soliton solutions of time fractional k(m, n) equation
Pramana 88 7

DOI:10.1007/s12043-016-1320-9 [Cited within: 1]

Hashemi M S Nucci M C 2013 Nonclassical symmetries for a class of reaction–diffusion equations: the method of heir-equations
J. Nonlinear Math. Phys. 20 44 60

DOI:10.1080/14029251.2013.792469 [Cited within: 1]

Najafi R Bahrami F Hashemi M S 2017 Classical and nonclassical lie symmetry analysis to a class of nonlinear time-fractional differential equations
Nonlinear Dyn. 87 1785 1796

DOI:10.1007/s11071-016-3152-z

Clarkson P A 1995 Nonclassical symmetry reductions of the Boussinesq equation
Chaos, Solitons Fractals 5 2261 2301

DOI:10.1016/0960-0779(94)E0099-B

Nucci M Clarkson P 1992 The nonclassical method is more general than the direct method for symmetry reductions, an example of the Fitzhugh–Nagumo equation
Phys. Lett. A 164 49 56

DOI:10.1016/0375-9601(92)90904-Z [Cited within: 1]

Gaeta G Morando P 2004 On the geometry of lambda-symmetries and PDE reduction
J. Phys. A: Math. Gen. 37 6955 6975

DOI:10.1088/0305-4470/37/27/007 [Cited within: 1]

Muriel C Romero J Ruiz A 2017 Symmetries and integrability by quadratures
IMA J. Appl. Math. 82 1061 1087

DOI:10.1093/imamat/hxx024

Gaeta G 2005 Lambda and mu-symmetries
Symmetry and Perturbation Theory Singapore World Scientific 99 105

[Cited within: 1]

Tian S F 2020 Lie symmetry analysis, conservation laws and solitary wave solutions to a fourth-order nonlinear generalized Boussinesq water wave equation
Appl. Math. Lett. 100 106056

DOI:10.1016/j.aml.2019.106056 [Cited within: 1]

Feng L L Tian S F Wang X B Zhang T T 2016 Lie symmetry analysis, conservation laws and exact power series solutions for time-fractional Fordy–Gibbons equation
Commun. Theor. Phys. 66 321 329

DOI:10.1088/0253-6102/66/3/321 [Cited within: 1]

Mao J J Tian S F Zhang T T Yang X J 2020 Lie symmetry analysis, conservation laws and analytical solutions for chiral nonlinear Schrödinger equation in (2 + 1)-dimensions
Nonlinear Anal. Model. Control 25 358 377

DOI:10.15388/namc.2020.25.16653 [Cited within: 1]

Wang H Tian Sh F Zhang T T Chen Y Fang Y 2019 General lump solutions, lumpoff solutions, and rogue wave solutions with predictability for the (2+1)-dimensional Korteweg–de Vries equation
Comp. Appl. Math. 38 1 15

DOI:10.1007/s40314-019-0938-x [Cited within: 1]

Wael S Seadawy A R EL-Kalaawy O H Maowad S M Baleanu D 2020 Symmetry reduction, conservation laws and acoustic wave solutions for the extended Zakharov–Kuznetsov dynamical model arising in a dust plasma
Results Phys. 19 103652

DOI:10.1016/j.rinp.2020.103652 [Cited within: 1]

Jhangeer A Hussain A Junaid-U-Rehman M Khan I Baleanu D Sooppy Nisar K 2020 Lie analysis, conservation laws and travelling wave structures of nonlinear Bogoyavlenskii–Kadomtsev–Petviashvili equation
Results Phys. 19 103492

DOI:10.1016/j.rinp.2020.103492 [Cited within: 1]

Jhangeer A Hussain A Junaid-U-Rehman M Baleanu D Riaz M B 2021 Quasi-periodic, chaotic and travelling wave structures of modified Gardner equation
Chaos, Solitons Fractals 143 110578

DOI:10.1016/j.chaos.2020.110578 [Cited within: 1]

Oldham K Spanier J 1974 The Fractional Calculus Theory and Applications of Differentiation and Integration to Arbitrary Order New York Elsevier
[Cited within: 3]

Miller K S Ross B 1993 An Introduction to the Fractional Calculus and Fractional Differential Equations New York Wiley
[Cited within: 1]

Kilbas A Srivastava H M Trujillo J J 2006 Theory and Applications of Fractional Differential Equations Amsterdam Elsevier
[Cited within: 4]

Uchaikin V V 2013 Fractional Derivatives for Physicists and Engineers Berlin Springer 2
[Cited within: 1]

Cui Z Yu P Mao Z 2012 Existence of solutions for nonlocal boundary value problems of nonlinear fractional differential equations
Adv. Dyn. Syst. Appl. 7 31 40

[Cited within: 1]

Gaul L Klein P Kemple S 1991 Damping description involving fractional operators
Mech. Syst. Signal Process 5 81 88

DOI:10.1016/0888-3270(91)90016-X

Khan S A Shah K Zaman G Jarad F 2019 Existence theory and numerical solutions to smoking model under Caputo– Fabrizio fractional derivative
Chaos: Interdisciplinary J. Nonlinear Sci. 29 013128

DOI:10.1063/1.5079644 [Cited within: 1]

Wang R Yuan X Zhang H Zhang J Lv N 2018 Symmetry transformations and exact solutions of a generalized hyperelastic rod equation
Comput. Mater. Continua 55 345 357

DOI:10.3970/cmc.2018.00233 [Cited within: 1]

Gazizov R K Kasatkin A A Lukashchuk S Y 2007 Continuous transformation groups of fractional differential equations
Vestnik Usatu 9 21

[Cited within: 1]

Gazizov R K Kasatkin A A Yu Lukashchuk S 2009 Symmetry properties of fractional diffusion equations
Phys. Scr. 2009 014016

DOI:10.1088/0031-8949/2009/T136/014016 [Cited within: 1]

Bluman G W Cole J D 1969 The general similarity solution of the heat equation
J. Math. Mech. 18 1025 1042

[Cited within: 1]

Ibragimov N H 2011 Nonlinear self-adjointness and conservation laws
J. Phys. A 44 432002

DOI:10.1088/1751-8113/44/43/432002 [Cited within: 1]

Ibragimov N H 2007 A new conservation theorem
J. Math. Anal. Appl. 333 311 328

DOI:10.1016/j.jmaa.2006.10.078 [Cited within: 1]

相关话题/Nonclassical symmetry conservation