删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

Analytical and approximate solutions of (2【-逻*辑*与-】plus;1)-dimensional time-fractional Burgers-Kadom

本站小编 Free考研考试/2022-01-02

Mehmet Senol,Department of Mathematics, Nevsehir Haci Bektas Veli University, Nevsehir, Turkey

Received:2019-12-11Revised:2020-02-6Accepted:2020-02-16Online:2020-04-27


Abstract
In this paper, we applied the sub-equation method to obtain a new exact solution set for the extended version of the time-fractional Kadomtsev-Petviashvili equation, namely Burgers-Kadomtsev-Petviashvili equation (Burgers-K-P) that arises in shallow water waves. Furthermore, using the residual power series method (RPSM), approximate solutions of the equation were obtained with the help of the Mathematica symbolic computation package. We also presented a few graphical illustrations for some surfaces. The fractional derivatives were considered in the conformable sense. All of the obtained solutions were replaced back in the governing equation to check and ensure the reliability of the method. The numerical outcomes confirmed that both methods are simple, robust and effective to achieve exact and approximate solutions of nonlinear fractional differential equations.
Keywords: fractional partial differential equations;Burgers-Kadomtsev-Petviashvili equation;conformable fractional derivative;sub-equation method;residual power series method


PDF (842KB)MetadataMetricsRelated articlesExportEndNote|Ris|BibtexFavorite
Cite this article
Mehmet Senol. Analytical and approximate solutions of (2+1)-dimensional time-fractional Burgers-Kadomtsev-Petviashvili equation. Communications in Theoretical Physics, 2020, 72(5): 055003- doi:10.1088/1572-9494/ab7707

1. Introduction

Fractional analysis, including arbitrary derivatives and integrals, has gained much attention in both theoretical and applied sciences in recent years. In fact, fractional analysis is not a new mathematical theory. The basics are almost as old as the classical analysis.

There have been many studies on the theory of fractional analysis, but in recent years these studies have gained a great deal of momentum thanks to programs such as Mathematica, Matlab and Maple, which are used for symbolic calculations. Thus, the equations that are too complicated to be calculated manually are easily solved. In addition, problems in mathematics, physics and many engineering fields are modeled in more detail using fractional calculus.

Grünwald, Letnikov, Riemann, Liouville and many others have contrived very substantial studies related to the theory of the subject, but it is thought that it cannot be applied to real life problems. However, recent studies in the fields of viscoelasticity and control have shown that the concepts of fractional derivative and integration are more convenient and reliable than the classical derivative and integral concepts that we have been using for modeling and solving many problems. There have been many applications using fractional calculus in real models see [13] and the references therein.

Recently, several definitions of fractional derivatives are taken part in the literature. One of them is fractional derivative with exponential kernel defined by Caputo and Fabrizio [4]. Another one is fractional derivative with Mittag-Leffler kernel defined by Atangana and Baleanu [58]. Some of the studies which exist in the literature are, Atangana, Aguilar and Jarad et al [911] and Al-Refai and Abdeljawad et al [12, 13] studied a fractional operator with exponential kernel for some modeling problems and their solution methods.

In this study, the definition of conformable derivative proposed by Khalil et al [14] has been used instead of the Riemann-Liouville and Caputo fractional derivative definitions, which are frequently used in the solution of fractional differential equations in recent years. This is because the definition of conformable derivative is different from other derivative definitions and allows to obtain the exact solution of the equation. In this direction, the new analytical solutions of the fractional Burgers-K-P equation was first obtained by the sub-equation method and the solutions were compared with the approximate solutions obtained by residual power series method (RPSM) proposed by Arqub [15, 16] by the help of tables and graphs.

After this introductory section, section 2 is devoted to introducing some basic definitions, section 3 is reserved for a brief review of the governing equation, and in section 4 a basic description of the implemented method is illustrated. In section 5, the analytical and approximate solutions of the Burgers-K-P equation are presented. Finally the paper ends with a conclusion in section 6.

2. Basic definitions

The conformable fractional derivative of an $f$ function for an arbitrary order of $\mu $ is given by

$\begin{eqnarray}{T}_{\mu }(f)(t)=\mathop{\mathrm{lim}}\limits_{\varepsilon \to 0}\displaystyle \frac{f(t+\varepsilon {t}^{1-\mu })-(f)(t)}{\varepsilon },\end{eqnarray}$for $f\,:[0,\infty )\to R$ and for all t>0, μ ∈ (0,1).

The next theorem includes some properties of this new definition [14].

Let $f$ and $g$ are $\mu $-differentiable functions at point $t\gt 0$ for $0\lt \mu \leqslant 1$. Then

${T}_{\mu }({mf}+{ng})={{mT}}_{\mu }(f)+{{nT}}_{\mu }(g)$ for all $m,n\in {\mathbb{R}}$
${T}_{\mu }({t}^{p})={{pt}}^{p-\mu }$ for all p
${T}_{\mu }(f.g)={{fT}}_{\mu }(g)+{{gT}}_{\mu }(f)$
${T}_{\mu }(\tfrac{f}{g})=\tfrac{{{gT}}_{\mu }(f)-{{fT}}_{\mu }(g)}{{g}^{2}}$
${T}_{\mu }(c)=0$ for all $f(t)=c$ constant functions
Moreover, ${T}_{\mu }(f)(t)={t}^{1-\mu }\tfrac{{df}(t)}{{dt}},$ if f is differentiable.


Let $f$ function is defined with $n$ variables ${x}_{1},\,\ldots ,\,{x}_{n}.$ The conformable partial derivatives of $f$ of order $\mu \in (0,1]$ in ${x}_{i}$ is given by [17]$\begin{eqnarray}\begin{array}{l}\displaystyle \frac{{{\rm{d}}}^{\mu }}{{\rm{d}}{x}_{i}^{\mu }}f({x}_{1},\,\ldots ,\,{x}_{n})\\ =\,\mathop{\mathrm{lim}}\limits_{\varepsilon \to 0}\displaystyle \frac{f({x}_{1},\,\ldots ,\,{x}_{i-1},{x}_{i}+\varepsilon {x}_{i}^{1-\mu },\,\ldots ,\,{x}_{n})-f({x}_{1},\,\ldots ,\,{x}_{n})}{\varepsilon }.\end{array}\end{eqnarray}$

The conformable integral of an $f$ function for $\mu \geqslant 0$ is defined as [18]:$\begin{eqnarray}{I}_{\mu }^{a}(f)(s)=\mathop{\mathop{\int }\limits^{s}}\limits_{a}\displaystyle \frac{f(t)}{{t}^{1-\mu }}{\rm{d}}t.\end{eqnarray}$

Next, some necessary theorems and definitions about fractional power series will be presented.

Assume that $f$ is an infinitely $\mu -$ differentiable function at a neighborhood of a point ${t}_{0}$ for some $0\lt \mu \leqslant 1,$ then $f$ has the fractional power series expansion of the form:$\begin{eqnarray}\begin{array}{rcl}f(t) & = & \mathop{\sum _{k=0}}\limits^{\infty }\displaystyle \frac{{\left({T}_{\mu }^{{t}_{0}}f\right)}^{(k)}({t}_{0}){\left(t-{t}_{0}\right)}^{k\mu }}{{\mu }^{k}k!},\\ & & \times \ {t}_{0}\lt t\lt {t}_{0}+{R}^{\tfrac{1}{\mu }},\ R\gt 0.\end{array}\end{eqnarray}$Here ${\left({T}_{\mu }^{{t}_{0}}f\right)}^{(k)}({t}_{0})$ represents the application of the fractional derivative $k-$ times [12].

$\mathop{\sum _{n=0}}\limits^{\infty }{f}_{n}(x){\left(t-{t}_{0}\right)}^{n\mu }$ is called a multiple fractional power series at ${t}_{0}=0,$ where $0\leqslant m-1\lt \mu \,\lt m,t$ is a variable and ${f}_{n}(x)$ are functions termed the coefficients of the series [19, 20].

Assume that $u(x,t)$ has a multiple fractional power series representation at ${t}_{0}=0$ of the form [19]$\begin{eqnarray}\begin{array}{rcl}u(x,t) & = & \mathop{\sum _{n=0}}\limits^{\infty }{f}_{n}(x){t}^{n\mu },\\ & & 0\leqslant m-1\lt \mu \lt m,\ x\in I,\ 0\leqslant t\leqslant {R}^{\tfrac{1}{\mu }}.\end{array}\end{eqnarray}$If ${u}_{t}^{(n\mu )}(x,t),$ $n=0,1,2,\ldots $ are continuous on $I\times (0,{R}^{\tfrac{1}{\mu }})$, then ${f}_{n}(x)=\tfrac{{u}_{t}^{(n\mu )}(x,0)}{{\mu }^{n}n!}.$

Many studies have been carried out using the above definitions by combining them with different methods for obtaining analytical and approximate solutions of various fractional partial differential equations. In contrast to the definition of Caputo and Riemann-Liouville, the definition of conformable derivative fulfills the chain rule [12]. Thus, using this definition, reliable solutions can be calculated for fractional differential equations.

3. Governing equation

Long waves, also known as shallow water waves, are waves moving at a depth of less than 1/20 of their wavelength in shallow waters of the sea or ocean. Many flow types can be modeled with the help of shallow water equations that are extensively used in flood analysis. These equations offer an attractive and productive field of research. They are very suitable for wave propagation phenomena and describe the interaction of two long waves with different dispersion relations. Eventually, such equations provide strong solutions to the problems arising from hydrodynamic systems.

Kadomtsev and Petviashvili relaxed the restriction that the waves be strictly one-dimensional, so they extended the well-known KdV equation to the (2+1)-dimensional K-P equation given by$\begin{eqnarray}{\left({u}_{t}+6{{uu}}_{x}+{u}_{{xxx}}\right)}_{x}+\lambda {u}_{{yy}}=0,\end{eqnarray}$where u=u(x, y, t). The K-P equation is used to model shallow water waves with weak nonlinear restoring forces.

On the other hand, the Burgers equation is a fundamental second-order partial differential equation arising in fluid mechanics given by$\begin{eqnarray}{u}_{t}+{{uu}}_{x}+\nu {u}_{{xx}}=0,\end{eqnarray}$where ν is the kinematic viscosity constant.

In this study, we examine the extension sense of the K-P equation, that is a completely integrable dissipative equation, namely the Burgers-K-P equation, with a conformable time-fractional derivative as$\begin{eqnarray}{T}_{x}[{T}_{t}^{\mu }u+{{uT}}_{x}u+\nu {T}_{x}^{2}u]+\lambda {T}_{y}^{2}u=0.\end{eqnarray}$In addition to these developments in fractional analysis, in the literature, there exists a number of techniques that have been developed and implemented to acquire analytical and approximate solutions of fractional differential equations, especially in the case of the Burgers-K-P equation. These techniques include the exp-function method [21], Lie symmetry approach [22], $({G}^{{\prime} }/G)$-expansion methods [23], modified direct Clarkson-Kruskal's method [24], the first integral method [25] and Hirota's bilinear method and tanh-coth method [26], the Schwarz function method [27] and q-homotopy analysis method [28]. To the author's knowledge, the conformable fractional version of the Burgers-K-P equation has not been studied. See [2933] for more methods and applications.

4. Description of the methods

4.1. Sub-equation method

Here, we identify the first implemented method called sub-equation method [3436] which is built on the Riccati equation given as in the following:$\begin{eqnarray}{\varphi }^{{\prime} }(\xi )=\sigma +{\left(\varphi (\xi \right)}^{2}.\end{eqnarray}$

Step 1:Suppose the general form of the nonlinear time-fractional partial differential equation is given as$\begin{eqnarray}P\left(u,{T}_{t}^{\mu }u,{T}_{x}u,{T}_{y}u,{T}_{x}^{2}u,{T}_{y}^{2}u,\ldots \right)=0,\end{eqnarray}$where ${T}_{t}^{\mu }$ implies a conformable derivative operator with arbitrary order.

Step 2:Allow the fractional wave transformation [37]$\begin{eqnarray}u(x,y,t)=U(\xi ),\ \xi ={kx}+{wy}+c\displaystyle \frac{{t}^{\mu }}{\mu },\end{eqnarray}$where $c,k,w$ are arbitrary constants to be calculated afterwards.

Step 3:Applying the chain rule given in [12], equation ((10) reduces to an integer order nonlinear ordinary differential equation of the form$\begin{eqnarray}G(U(\xi ),{U}^{{\prime} }(\xi ),{U}^{{\prime\prime} }(\xi ),\ \ldots )=0.\end{eqnarray}$

Step 4:Let the solution of equation (12) be in the following form$\begin{eqnarray}U(\xi )=\sum _{i=0}^{N}{a}_{i}{\varphi }^{i}(\xi ),\ {a}_{N}\ne 0,\end{eqnarray}$where ai are stable coefficients for $(0\leqslant i\leqslant N)$ and to be calculated. N is a positive integer to be calculated by the balancing principle [38] in equation (12) and φ(ξ) is the solution of the Riccati equation (9).

Step 5:A set of the solutions which confirm the equation (9)) is given below.$\begin{eqnarray}\varphi (\xi )=\left\{\begin{array}{c}-\sqrt{-\sigma }\tanh \left(\sqrt{-\sigma }\xi \right),\ \sigma \lt 0\\ -\sqrt{-\sigma }\coth \left(\sqrt{-\sigma }\xi \right),\ \ \sigma \lt 0\\ \sqrt{\sigma }\tan \left(\sqrt{\sigma }\xi \right),\ \ \ \ \ \ \ \ \ \sigma \gt 0\\ -\sqrt{\sigma }\cot \left(\sqrt{\sigma }\xi \right),\ \ \ \ \ \ \ \ \ \sigma \gt 0\\ -\tfrac{1}{\xi +\varpi },\ \varpi \ {is}\ a\ \mathrm{cons}.,\ \ \ \sigma =0.\end{array}\right.\end{eqnarray}$

Step 6:Combining all of the obtained results, we achieve a polynomial in terms of $\varphi (\xi )$.

Step 7:Vanishing the coefficients of ${\varphi }^{i}(\xi )$ for $(i=0,1,\,\ldots ,\,N)$ produces a nonlinear algebraic equation system in ${a}_{i},c,k,w,$ $(i=0,1,\,\ldots ,\,N)$.

Step 8:Computing the solution of these nonlinear algebraic equations, we attain the values for ${a}_{i},c,k,w,$ $(i\,=0,1,\,\ldots ,\,N)$.

Step 9:Subrogating all of the acquired values in the formulas of (14), we get the exact solutions for the equation (10).

4.2. Residual power series method (RPSM)

To illustrate the basic idea of RPSM [15, 16, 3941], let us consider the following nonlinear fractional differential equation:$\begin{eqnarray}\begin{array}{l}{T}_{\mu }u(x,y,t)+N[x,y]u(x,y,t)\\ \qquad +\ L[x,y]u(x,y,t)=g(x,y,t),\end{array}\end{eqnarray}$where $n-1\lt n\mu \leqslant n,$ $t\gt 0,$ $x\in {\mathbb{R}}.$ The initial condition of the equation is$\begin{eqnarray}{f}_{0}(x,y)=u(x,y,0)=f(x,y).\end{eqnarray}$Here, $N[x,y]$ is a nonlinear and $L[x,y]$ is a linear operator given with the g(x, y, t) continuous functions.

Step 1: The RPSM method consists of specifying the solution of (15), which is subject to (16) as a fractional power series expansion around t=0.$\begin{eqnarray}{f}_{(n-1)}(x,y)={T}_{t}^{(n-1)\mu }u(x,y,0)=h(x,y).\end{eqnarray}$

Step 2: The solution can be expressed by a series expansion as$\begin{eqnarray}u(x,y,t)=f(x,y)+\mathop{\sum _{n=1}}\limits^{\infty }{f}_{n}(x,y)\displaystyle \frac{{t}^{n\mu }}{{\mu }^{n}n!}.\end{eqnarray}$Thus, the kth truncated series of u(x, y, t), that is ${u}_{k}(x,y,t)$ can be represented as$\begin{eqnarray}{u}_{k}(x,y,t)=f(x,y)+\mathop{\sum _{n=1}}\limits^{k}{f}_{n}(x,y)\displaystyle \frac{{t}^{n\mu }}{{\mu }^{n}n!}.\end{eqnarray}$

Step 3: Since the 1st approximate solution u1(x, y, t) is$\begin{eqnarray}{u}_{1}(x,y,t)=f(x,y)+{f}_{1}(x,y)\displaystyle \frac{{t}^{\mu }}{{\mu }^{n}},\end{eqnarray}$uk(x, y, t) might be reformulated as$\begin{eqnarray}\begin{array}{rcl}{u}_{k}(x,y,t) & = & f(x,y)+{f}_{1}(x,y)\displaystyle \frac{{t}^{\mu }}{{\mu }^{n}}\\ & & +\ \mathop{\sum _{n=2}}\limits^{k}{f}_{n}(x,y)\displaystyle \frac{{t}^{n\mu }}{{\mu }^{n}n!},\ k=2,3,4,\ldots \end{array}\end{eqnarray}$for $0\leqslant t\lt {{\mathbb{R}}}^{\tfrac{1}{v}},$ $x\in I$ and $0\lt \mu \leqslant 1$.

Step 4: Initially we express the residual function and the kth residual function$\begin{eqnarray}\begin{array}{rcl}\mathrm{Res}(x,y,t) & = & {T}_{\mu }u(x,y,t)+N[x,y]u(x,y,t)\\ & & +\ L[x,y]u(x,y,t)-g(x,y,t),\end{array}\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{\mathrm{Res}}_{k}(x,t) & = & {T}_{\mu }{u}_{k}(x,y,t)+N[x,y]{u}_{k}(x,y,t)\\ & & +L[x,y]{u}_{k}(x,y,t)-g(x,y,t),\\ k & = & 1,2,3,\ldots \end{array}\end{eqnarray}$respectively.

Step 5: Obviously, $\mathrm{Res}(x,y,t)=0$ and $\mathop{\mathrm{lim}}\limits_{k\to \infty }{\mathrm{Res}}_{k}(x,y,t)\,=\mathrm{Res}(x,y,t)$ for each $x\in I$ and $t\geqslant 0.$ Indeed this brings about $\tfrac{{\partial }^{(n-1)\mu }}{\partial {t}^{(n-1)\mu }}{\mathrm{Res}}_{k}(x,y,t)=0$ for $n=1,2,3,\,\ldots ,\,k$ because, the fractional derivative of a constant is zero in the conformable sense [17, 42].

Step 6: Integrating the differential equation $\tfrac{{\partial }^{(n-1)\mu }}{\partial {t}^{(n-1)\mu }}{\mathrm{Res}}_{k}(x,y,0)=0$ produces the needed ${f}_{n}(x,y)$ parameters. Therefore, using these parameters one can achieve the ${u}_{n}(x,y,t)$ approximate solutions in this fashion, respectively.

5. Results and discussion

5.1. Analytical solutions of the Burgers-K-P equation

Consider the time fractional Burgers-K-P equation$\begin{eqnarray}{T}_{x}[{T}_{t}^{\mu }u+{{uT}}_{x}u+\nu {T}_{x}^{2}u]+\lambda {T}_{y}^{2}u=0,\end{eqnarray}$where ${T}_{t}^{\mu }$ is conformable derivative of function u(x, y, t) with arbitrary order. Performing the chain rule [12] and wave transform (11) to equation (24) and integrating the resulting ODE once, we find out the reduced form can be expressed as follows:$\begin{eqnarray}{{kcU}}^{{\prime} 2}{{UU}}^{{\prime} 3}\nu {U}^{{\prime} ^{\prime} 2}{U}^{{\prime} }=0.\end{eqnarray}$Assume that the solution of equation (25) is in terms of $\varphi (\xi ),$ while the analytical solutions of equation (9) are given as follows:$\begin{eqnarray}U(\xi )=\sum _{i=0}^{N}{a}_{i}{\varphi }^{i}(\xi ),\ {a}_{N}\ne 0.\end{eqnarray}$With the aid of the balancing principle [38], we have N=1. Bringing together all the obtained data in equation (25), an algebraic equation system comes to exist with respect to $k,w,c,{a}_{0},{a}_{1}$. Solving the obtain system yields following solution set:$\begin{eqnarray}{a}_{1}=-2k\nu ,c=\displaystyle \frac{-{a}_{0}{k}^{2}-\lambda {w}^{2}}{k},\end{eqnarray}$and ${a}_{0},k,w$ are free constants. When $\sigma \lt 0$, using (14) and (11) the traveling wave solutions of equation (24) can be deducted:$\begin{eqnarray}\begin{array}{rcl}{u}_{1}(x,y,t) & = & {a}_{0}+2k\nu \sqrt{-\sigma }\tanh \\ & & \times \left(\sqrt{-\sigma }\left(\displaystyle \frac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }+{kx}+{wy}\right)\right),\end{array}\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{2}(x,y,t) & = & {a}_{0}+2k\nu \sqrt{-\sigma }\coth \\ & & \times \left(\sqrt{-\sigma }\left(\displaystyle \frac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }+{kx}+{wy}\right)\right).\end{array}\end{eqnarray}$In a similar way, for σ>0, we obtain$\begin{eqnarray}\begin{array}{rcl}{u}_{3}(x,y,t) & = & {a}_{0}-2k\nu \sqrt{\sigma }\tan \\ & & \times \left(\sqrt{\sigma }\left(\displaystyle \frac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }+{kx}+{wy}\right)\right),\end{array}\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{4}(x,y,t) & = & {a}_{0}+2k\nu \sqrt{\sigma }\cot \\ & & \times \left(\sqrt{\sigma }\left(\displaystyle \frac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }+{kx}+{wy}\right)\right).\end{array}\end{eqnarray}$Lastly, when σ=0$\begin{eqnarray}{u}_{5}(x,y,t)={a}_{0}+\displaystyle \frac{2k\nu }{\tfrac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }+{kx}+l+{wy}}.\end{eqnarray}$

5.2. Approximate solutions of the Burgers-K-P equation

Consider the nonlinear time-fractional Burgers-K-P equation as follows:$\begin{eqnarray}{T}_{x}[{T}_{t}^{\mu }u+{{uT}}_{x}u+\nu {T}_{x}^{2}u]+\lambda {T}_{y}^{2}u=0,\end{eqnarray}$where u=u(x, y, t), t≥0, 0< μ≤1.

In this case, take the initial condition$\begin{eqnarray}f(x,y)={a}_{0}+\displaystyle \frac{2k\nu }{1+{kx}+{wy}},\end{eqnarray}$that is deduced from the exact solution equation (32). For residual power series

$\begin{eqnarray}u(x,y,t)=f(x,y)+\mathop{\sum _{n=1}}\limits^{\infty }{f}_{n}(x,y)\displaystyle \frac{{t}^{n\mu }}{{\mu }^{n}n!},\end{eqnarray}$and kth truncated series of u(x, y, t)$\begin{eqnarray}{u}_{k}(x,y,t)=f(x,y)+\mathop{\sum _{n=1}}\limits^{k}{f}_{n}(x,y)\displaystyle \frac{{t}^{n\mu }}{{\mu }^{n}n!},\ k=1,2,3,\ldots \end{eqnarray}$the general form of the kth residual functions of the time-fractional Burgers-K-P equation can be expressed by$\begin{eqnarray}\begin{array}{rcl}\mathrm{Res}{u}_{k}(x,y,t) & = & {\left({\partial }_{t}^{\mu }{u}_{k}\right)}_{x}+\lambda {\left({u}_{1}(x,y,t\right)}_{{yy}}\\ & & +\ \nu {\left({u}_{1}(x,y,t\right)}_{{xxx}}+{\left({u}_{1}(x,y,t\right)}_{x}^{2}\\ & & +\ {u}_{1}(x,y,t){\left({u}_{1}(x,y,t\right)}_{{xx}}.\end{array}\end{eqnarray}$

To determine the parameter f1(x, y), in u1(x, y, t), one should replace the 1st truncated series ${u}_{1}(x,y,t)\,=f(x,y)+{f}_{1}(x,y)\tfrac{t\mu }{\mu }$ into the 1st truncated residual function to obtain$\begin{eqnarray}\begin{array}{rcl}\mathrm{Res}{u}_{1}(x,y,t) & = & {\left({f}_{1}(x,y\right)}_{x}+\lambda \left({\left(f(x,y\right)}_{{yy}}\right.\\ & & +\ \left.\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{{yy}}}{\mu }\right)\\ & & +\ \nu \left({\left(f(x,y\right)}_{{xxx}}+\displaystyle \frac{{t}^{\mu }f{\left({}_{1}(x,y\right)}_{{xxx}}}{\mu }\right)\\ & & +\ \left({\left(f(x,y\right)}_{x}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{x}}{\mu }\right){}^{2}\\ & & +\ \left(f(x,y)+\displaystyle \frac{{t}^{\mu }{f}_{1}(x,y)}{\mu }\right)\\ & & \times \ \left({\left(f(x,y\right)}_{{xx}}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{{xx}}}{\mu }\right).\end{array}\end{eqnarray}$Next, by substitution of t=0 by $\mathrm{Res}{u}_{1}(x,y,t)$, we obtain$\begin{eqnarray}\begin{array}{rcl}{\left({f}_{1}(x,y\right)}_{x} & = & -\lambda {\left(f(x,y\right)}_{{yy}}-\nu {\left(f(x,y\right)}_{{xxx}}\\ & & -\ {\left(f(x,y\right)}_{x}^{2}-f(x,y){\left(f(x,y\right)}_{{xx}}.\end{array}\end{eqnarray}$Solving this differential equation and evaluating with the initial condition $f(x,y)$ gives the first unknown parameter as$\begin{eqnarray}{f}_{1}(x,y)=\displaystyle \frac{2\nu \left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}{{\left({kx}+l+{wy}\right)}^{2}}.\end{eqnarray}$Thus, we calculate the first RPSM solution of the time-fractional Burgers-K-P equation as$\begin{eqnarray}{u}_{1}(x,y,t)={a}_{0}+\displaystyle \frac{2k\nu }{{kx}+l+{wy}}+\displaystyle \frac{2\nu {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}{\mu {\left({kx}+l+{wy}\right)}^{2}}.\end{eqnarray}$

Similarly, to calculate f2(x, y), which is the second unknown coefficient, we replace the second truncated series solution ${u}_{2}(x,y,t)=f(x,y)+{f}_{1}(x,y)\tfrac{{t}^{\mu }}{\mu }+{f}_{2}(x,y)\tfrac{{t}^{2\mu }}{2{\mu }^{2}}$ into the second truncated residual function to attain$\begin{eqnarray}\begin{array}{l}\mathrm{Res}{u}_{2}(x,y,t)={\left({f}_{1}(x,y\right)}_{x}+\displaystyle \frac{{t}^{\mu }{\left({f}_{2}(x,y\right)}_{x}}{\mu }\\ \quad +\,\lambda \left({\left(f(x,y\right)}_{{yy}}+\displaystyle \frac{{t}^{2\mu }{\left({f}_{2}(x,y\right)}_{{yy}}}{2{\mu }^{2}}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{{yy}}}{\mu }\right)\\ \quad +\,\nu \left({\left(f(x,y\right)}_{{xxx}}+\displaystyle \frac{{t}^{2\mu }{\left({f}_{2}(x,y\right)}_{{xxx}}}{2{\mu }^{2}}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{{xxx}}}{\mu }\right)\\ \quad +\,\left({\left(f(x,y\right)}_{x}+\displaystyle \frac{{t}^{2\mu }{\left({f}_{2}(x,y\right)}_{x}}{2{\mu }^{2}}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{x}}{\mu }\right){}^{2}\\ \quad +\,\left(f(x,y)+\displaystyle \frac{{t}^{2\mu }{f}_{2}(x,y)}{2{\mu }^{2}}+\displaystyle \frac{{t}^{\mu }{f}_{1}(x,y)}{\mu }\right)\\ \quad \times \,\left({\left(f(x,y\right)}_{{xx}}+\displaystyle \frac{{t}^{2\mu }{\left({f}_{2}(x,y\right)}_{{xx}}}{2{\mu }^{2}}+\displaystyle \frac{{t}^{\mu }{\left({f}_{1}(x,y\right)}_{{xx}}}{\mu }\right).\end{array}\end{eqnarray}$

Next, applying the conformable derivative Tμ on both sides of $\mathrm{Res}{u}_{2}(x,y,t)$ and solving for t=0 yields$\begin{eqnarray}\begin{array}{l}{\left({f}_{2}(x,y\right)}_{x}=-2{\left(f(x,y\right)}_{x}{\left({f}_{1}(x,y\right)}_{x}-{f}_{1}(x,y){\left(f(x,y\right)}_{{xx}}\\ \ \ -\ \lambda {\left({f}_{1}(x,y\right)}_{{yy}}-\nu {\left({f}_{1}(x,y\right)}_{{xxx}}-f(x,y){\left({f}_{1}(x,y\right)}_{{xx}}.\end{array}\end{eqnarray}$

Solving this differential equation and evaluating with the initial condition $f(x,y)$ and first unknown parameter f1(x, y) gives$\begin{eqnarray}{f}_{2}(x,y)=\displaystyle \frac{4\nu {\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}}{k{\left({kx}+l+{wy}\right)}^{3}}.\end{eqnarray}$So, the second RPSM approximation of the equation is$\begin{eqnarray}\begin{array}{rcl}{u}_{2}(x,y,t) & = & {a}_{0}+\displaystyle \frac{2k\nu }{{kx}+l+{wy}}+\displaystyle \frac{2\nu {t}^{2\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}}{k{\mu }^{2}{\left({kx}+l+{wy}\right)}^{3}}\\ & & +\displaystyle \frac{2\nu {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}{\mu {\left({kx}+l+{wy}\right)}^{2}}.\end{array}\end{eqnarray}$

Similarly, following the above scheme for n=3 and 4 gives the subsequent outcomes.$\begin{eqnarray}{f}_{3}(x,y)=\displaystyle \frac{12\nu {\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}}{{k}^{2}{\left({kx}+l+{wy}\right)}^{4}}.\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{3}(x,y,t) & = & {a}_{0}+\displaystyle \frac{2k\nu }{{kx}+l+{wy}}+\displaystyle \frac{2\nu {t}^{3\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}}{{k}^{2}{\mu }^{3}{\left({kx}+l+{wy}\right)}^{4}}\\ & & +\displaystyle \frac{2\nu {t}^{2\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}}{k{\mu }^{2}{\left({kx}+l+{wy}\right)}^{3}}+\displaystyle \frac{2\nu {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}{\mu {\left({kx}+l+{wy}\right)}^{2}}.\end{array}\end{eqnarray}$$\begin{eqnarray}{f}_{4}(x,y)=\displaystyle \frac{48\nu {\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{4}}{{k}^{3}{\left({kx}+l+{wy}\right)}^{5}}.\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{4}(x,y,t) & = & {a}_{0}+\displaystyle \frac{2k\nu }{{kx}+l+{wy}}+\displaystyle \frac{2\nu {t}^{3\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}}{{k}^{2}{\mu }^{3}{\left({kx}+l+{wy}\right)}^{4}}\\ & & +\displaystyle \frac{2\nu {t}^{2\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}}{k{\mu }^{2}{\left({kx}+l+{wy}\right)}^{3}}+\displaystyle \frac{2\nu {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}{\mu {\left({kx}+l+{wy}\right)}^{2}}\\ & & +\ \displaystyle \frac{2\nu {t}^{4\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{4}}{{k}^{3}{\mu }^{4}{\left({kx}+l+{wy}\right)}^{5}}.\end{array}\end{eqnarray}$

In figures 13, the obtained solutions and exact solutions are compared graphically.$\begin{eqnarray}u(x,y,t)={a}_{0}+\displaystyle \frac{2k\nu }{1+{kx}+{wy}+\tfrac{{t}^{\mu }(-{a}_{0}{k}^{2}-\lambda {w}^{2})}{k\mu }},\end{eqnarray}$where a0, k, w, λ and ν are constants.

Figure 1.

New window|Download| PPT slide
Figure 1.(Case 5.1) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, t=0.1 and μ=0.35.


As seen, the differences between approximate and exact solutions in figures 1(a), (b), 2(a), (b), 3(a) and (b) in different cases is not clearly obvious. Therefore, the errors should be computed for the accurate comparison of the numerical and exact solutions. Error results for μ=0,35, μ=0,50 and μ=0,75 in the numerical solutions obtained by exact and the fourth-order RPSM solutions are presented in table 1. The following absolute error formula is used for the numerical solutions;$\begin{eqnarray*}\mathrm{Error}=\left|u(x,y,t)-{u}_{n}(x,y,t)\right|.\end{eqnarray*}$In the above formula, u(x, y, t) and un(x, y, t) represent the exact solution and the numerical solution at the point $(x,y,t)$ respectively.

Figure 2.

New window|Download| PPT slide
Figure 2.(Case 5.1) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, t=0.1 and μ=0.50.


Figure 3.

New window|Download| PPT slide
Figure 3.(Case 5.1) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, t=0.1 and μ=0.75.



Table 1.
Table 1.(Case 5.1) RPSM approximate results and comparison with the exact solutions by absolute errors for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1 and t=0.1.
μ=0.35μ=0.50μ=0.75
xRPSMExactAbs. ErrorRPSMExactAbs. ErrorRPSMExactAbs. Error
0.00.010 731 70.010 761 32.959 80E-50.010 490 10.010 490 65.70129E-70.010 402 70.010 402 73.465 43E-9
0.10.010 729 20.010 758 42.921 92E-50.010 488 90.010 489 45.63592E-70.010 401 90.010 401 93.427 19E-9
0.20.010 726 70.010 755 52.884 61E-50.010 487 70.010 488 25.57142E-70.010 401 10.010 401 13.389 45E-9
0.30.010 724 20.010 752 72.847 87E-50.010 486 50.010 487 15.50778E-70.010 400 30.010 400 33.352 19E-9
0.40.010 721 70.010 749 92.811 67E-50.010 485 30.010 485 95.44499E-70.010 399 50.010 399 53.315 41E-9
0.50.010 719 30.010 747 12.776 03E-50.010 484 20.010 484 75.38304E-70.010 398 70.010 398 73.279 10E-9
0.60.010 716 90.010 744 32.740 91E-50.010 483 00.010 483 55.32190E-70.010 397 90.010 397 93.243 25E-9
0.70.010 714 50.010 741 52.706 33E-50.010 481 80.010 482 45.26158E-70.010 397 10.010 397 13.207 86E-9
0.80.010 712 10.010 738 82.672 26E-50.010 480 70.010 481 25.20206E-70.010 396 30.010 396 33.172 92E-9
0.90.010 709 70.010 736 12.638 69E-50.010 479 50.010 480 05.14332E-70.010 395 50.010 395 53.138 42E-9
1.00.010 707 30.010 733 42.605 63E-50.010 478 40.010 478 95.08535E-70.010 394 80.010 394 83.104 36E-9

New window|CSV

For μ=0.35 at x=0, error is about $2.9E-5$ and that result is decreasing to $2.6E-5$ when x is increasing to 1.

For μ=0.50, the results are about 200 times better than the results obtained when μ=0.35. For μ=0.75 at x=0, error is about $3.4E-9$ and that results is decreasing to $3.1E-9$ when x is increasing to 1.

Now, consider the initial condition$\begin{eqnarray}f(x,y)={a}_{0}+2k\nu \sqrt{-\sigma }\tanh \left({kx}+{wy})\sqrt{-\sigma }\right),\end{eqnarray}$that is obtained from the exact solution equation (28). Applying the similar procedure we obtain$\begin{eqnarray}{f}_{1}(x,y)=2\nu \sigma \left({a}_{0}{k}^{2}+\lambda {w}^{2}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right),\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{1}(x,y,t) & = & \displaystyle \frac{2\nu \sigma {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{\mu }\\ & & +{a}_{0}+2k\nu \sqrt{-\sigma }\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right),\end{array}\end{eqnarray}$$\begin{eqnarray}{f}_{2}(x,y)=-\displaystyle \frac{4\nu {\sigma }^{2}{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{k\sqrt{-\sigma }},\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{2}(x,y,t) & = & \displaystyle \frac{2\nu \sigma {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right){\sec }^{2}\left(\sqrt{\sigma }({kx}+{wy}\right)\left(k\mu -\sqrt{\sigma }{t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)\tan \left(\sqrt{\sigma }({kx}+{wy}\right)\right)}{k{\mu }^{2}}\\ & & +{a}_{0}-2k\nu \sqrt{\sigma }\tan \left(\sqrt{\sigma }({kx}+{wy}\right),\end{array}\end{eqnarray}$

$\begin{eqnarray}{f}_{3}(x,y)=-\displaystyle \frac{4\nu {\sigma }^{2}{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}\left(\cosh \left(2\sqrt{-\sigma }({kx}+{wy}\right)-2\right){{\rm{sech}} }^{4}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{{k}^{2}},\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{3}(x,y,t) & = & -\displaystyle \frac{2\nu {\sigma }^{2}{t}^{3\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}\left(\cosh \left(2\sqrt{-\sigma }({kx}+{wy}\right)-2\right){{\rm{sech}} }^{4}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{3{k}^{2}{\mu }^{3}}\\ & & -\displaystyle \frac{2\nu {\sigma }^{2}{t}^{2\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{k{\mu }^{2}\sqrt{-\sigma }}\\ & & +\displaystyle \frac{2\nu \sigma {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{\mu }\,+{a}_{0}+2k\nu \sqrt{-\sigma }\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right),\end{array}\end{eqnarray}$$\begin{eqnarray}{f}_{4}(x,y)=\displaystyle \frac{8\nu {\sigma }^{3}{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{4}\left(\cosh \left(2\sqrt{-\sigma }({kx}+{wy}\right)-5\right)\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right){{\rm{sech}} }^{4}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{{k}^{3}\sqrt{-\sigma }},\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}{u}_{4}(x,y,t) & = & -\displaystyle \frac{2\nu {\sigma }^{2}{t}^{3\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{3}\left(\cosh \left(2\sqrt{-\sigma }({kx}+{wy}\right)-2\right){{\rm{sech}} }^{4}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{3{k}^{2}{\mu }^{3}}\\ & & -\displaystyle \frac{2\nu {\sigma }^{2}{t}^{2\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{2}\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{k{\mu }^{2}\sqrt{-\sigma }}\\ & & +\displaystyle \frac{2\nu \sigma {t}^{\mu }\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right){{\rm{sech}} }^{2}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{\mu }\\ & & +\displaystyle \frac{\nu {\sigma }^{3}{t}^{4\mu }{\left({a}_{0}{k}^{2}+\lambda {w}^{2}\right)}^{4}\left(\cosh \left(2\sqrt{-\sigma }({kx}+{wy}\right)-5\right)\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right){{\rm{sech}} }^{4}\left(\sqrt{-\sigma }({kx}+{wy}\right)}{3{k}^{3}{\mu }^{4}\sqrt{-\sigma }}\\ & & +{a}_{0}+2k\nu \sqrt{-\sigma }\tanh \left(\sqrt{-\sigma }({kx}+{wy}\right).\end{array}\end{eqnarray}$Besides, in figures 46, the graphical illustrations of the fourth order RPSM and exact solution$\begin{eqnarray}\begin{array}{rcl}u(x,y,t) & = & 2k\nu \sqrt{-\sigma }\tanh \left(\sqrt{-\sigma }\left(\displaystyle \frac{{t}^{\mu }\left(-{a}_{0}{k}^{2}-\lambda {w}^{2}\right)}{k\mu }\right.\right.\\ & & +\ \left.\left.{kx}+{wy}\right)\right)+{a}_{0},\end{array}\end{eqnarray}$where a0, k, w, λ , σ and ν are constants. It is clear that they are almost identical in terms of accuracy and are in perfect agreement with each other.

Figure 4.

New window|Download| PPT slide
Figure 4.(Case 5.2) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, σ=−0.5, $t=0.1$ and μ=0.45.


Figure 5.

New window|Download| PPT slide
Figure 5.(Case 5.2) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, σ=−0.5, $t=0.1$ and μ=0.55.


Figure 6.

New window|Download| PPT slide
Figure 6.(Case 5.2) Comparison for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, σ=−0.5, $t=0.1$ and μ=0.75.


Also, in table 2, the absolute errors are offered for μ=0,45, μ=0,55 and μ=0,75 respectively. The outcomes reveal that absolute errors decrease with increasing x values, as they decrease with increasing μ values.


Table 2.
Table 2.(Case 5.2) RPSM approximate results and comparison with the exact solutions by absolute errors for a0=0.01, k=0.1, w=0.5, $\nu =0.01,$ λ=0.9, l=5, y=1, σ=−0.5 and t=0.1.
μ=0.45μ=0.55μ=0.75
xRPSMExactAbs. ErrorRPSMExactAbs. ErrorRPSMExactAbs. Error
0.00.009 361 20.008 986 03.752 06E-40.009 428 70.009 389 33. 93727E-50.009 966 80.009 966 25.890 58E-7
0.10.009 358 70.008 990 83.678 53E-40.009 435 60.009 397 53.81430E-50.009 976 80.009 976 25.569 00E-7
0.20.009 356 10.008 995 83.602 89E-40.009 442 60.009 405 73.68946E-50.009 986 80.009 986 25.246 58E-7
0.30.009 353 30.009 000 83.525 22E-40.009 449 60.009 414 03.56291E-50.009 996 70.009 996 24.923 68E-7
0.40.009 350 30.009 005 83.445 62E-40.009 456 60.009 422 33.43478E-50.010 006 70.010 006 34.600 65E-7
0.50.009 347 30.009 010 93.364 17E-40.009 463 70.009 430 63.30521E-50.010 016 70.010 016 34.277 82E-7
0.60.009 344 10.009 016 03.280 97E-40.009 470 80.009 439 03.17437E-50.010 026 70.010 026 33.955 54E-7
0.70.009 340 80.009 021 23.196 12E-40.009 477 90.009 447 53.04238E-50.010 036 60.010 036 33.634 15E-7
0.80.009 337 40.009 026 43.109 72E-40.009 485 10.009 456 02.90941E-50.010 046 60.010 046 33.313 97E-7
0.90.009 333 90.009 031 73.021 85E-40.009 492 30.009 464 52.77559E-50.010 056 50.010 056 22.995 31E-7
1.00.009 330 30.009 037 02.932 61E-40.009 499 50.009 473 12.64106E-50.010 066 50.010 066 22.678 50E-7

New window|CSV

6. Conclusion

In this paper, two reliable methods, namely the residual power series method and the sub-equation method, have been implemented to the time-fractional Burgers-K-P partial differential equation which arise in the shallow water waves. Using conformable derivative definition, numerous new exact solutions of the equation have been obtained which do not exist in the literature. Then approximate solutions obtained by RPSM have been compared with these exact solutions by surface plots and tables. Thus, it is shown that the methods are very effective and could be used for different types of FPDEs arising in different areas of mathematical physics.

Acknowledgments

The author received no financial support for the research, authorship, and/or publication of this article.


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

Iyiola O S 2014 A fractional diffusion equation model for cancer tumor
AIP Adv. 4 107 121

DOI:10.1063/1.4898331 [Cited within: 1]

Caputo M 1999 Diffusion of fluids in porous media with memory
Geothermics 28 113 130

DOI:10.1016/S0375-6505(98)00047-9

Iyiola O S 2017 Exponential integrator methods for nonlinear fractional reaction-diffusion models
Theses and Dissertations 1644 1-181

[Cited within: 1]

Caputo M Fabrizio M 2015 A new definition of fractional derivative without singular kernel
Prog. Fract. Differ. Appl. 1 1 13

DOI:10.12785/pfda/010201 [Cited within: 1]

Al-Refai M Hajji M A 2019 Analysis of a fractional eigenvalue problem involving Atangana-Baleanu fractional derivative: a maximum principle and applications
Chaos 29 013135

DOI:10.1063/1.5083202 [Cited within: 1]

Atangana A Baleanu D 2016 New fractional derivatives with non-local and nonsingular kernel: theory and application to heat transfer model
Therm. Sci. 20 763 769

DOI:10.2298/TSCI160111018A

Saad K M Khader M M Gómez-Aguilar J F Baleanu D 2019 Numerical solutions of the fractional Fisher’s type equations with Atangana-Baleanu fractional derivative by using spectral collocation methods
Chaos 29 023116

DOI:10.1063/1.5086771

Sene N 2019 Analytical solutions of Hristov diffusion equations with non-singular fractional derivatives
Chaos 29 023112

DOI:10.1063/1.5082645 [Cited within: 1]

Gómez-Aguilar J F Morales-Delgado V F Taneco-Hernández M A Baleanu D Escobar-Jiménez R F Al Qurashi M M 2016 Analytical solutions of the electrical RLC circuit via Liouville-Caputo operators with local and non-local kernels
Entropy 18 402

DOI:10.3390/e18080402 [Cited within: 1]

Atangana A Baleanu D 2017 Caputo-Fabrizio derivative applied to groundwater flow within confined aquifer
J. Eng. Mech. 143 D4016005

DOI:10.1061/(ASCE)EM.1943-7889.0001091

Jarad F Uğurlu E Abdeljawad T Baleanu D 2017 On a new class of fractional operators
Adv. Differ. Equ. 1 247

DOI:10.1186/s13662-017-1306-z [Cited within: 1]

Abdeljawad T 2015 On conformable fractional calculus
Journal of Computational and Applied Mathematics, J. Comput. Appl. Math. 279 57-66

DOI:10.1016/j.cam.2014.10.016 [Cited within: 5]

Al-Refai M Abdeljawad T 2017 Analysis of the fractional diffusion equations with fractional derivative of non-singular kernel
Adv. Differ. Equ. 1 315

DOI:10.1186/s13662-017-1356-2 [Cited within: 1]

Khalil R M Al Horani Yousef A Sababheh M 2014 A new definition of fractional derivative
J. Comput. Appl. Math. 264 65

DOI:10.1016/j.cam.2014.01.002 [Cited within: 2]

Arqub O A 2013 Series solution of fuzzy differential equations under strongly generalized differentiability
J. Adv. Res. Appl. Math. 5 31

DOI:10.5373/jaram.1447.051912 [Cited within: 2]

Arqub O A El-Ajou A Bataineh A S Hashim I 2013 A representation of the exact solution of generalized Lane-Emden equations using a new analytical method
Abstr. Appl. Anal. 2013 1-10

DOI:10.1155/2013/378593 [Cited within: 2]

Atangana A Baleanu D Alsaedi A 2015 New properties of conformable derivative
Open Math. 13 889

DOI:10.1515/math-2015-0081 [Cited within: 2]

Tasbozan O Cenesiz Y Kurt A 2016 New solutions for conformable fractional Boussinesq and combined KdV-mKdV equations using Jacobi elliptic function expansion method
Eur. Phys. J. Plus 131 244

DOI:10.1140/epjp/i2016-16244-x [Cited within: 1]

Alabsi T Y 2017 Solution of conformable fractional navier-stokes equation
M.S. Thesis Zarqa University

[Cited within: 2]

El-Ajou A Arqub O A Zhour Z A Momani S 2013 New results on fractional power series: theories and applications
Entropy 15 5305

DOI:10.3390/e15125305 [Cited within: 1]

Chun C Sakthivel R 2010 New soliton and periodic solutions for two nonlinear physical models
Z. Naturforsch. A 65 1049

DOI:10.1515/zna-2010-1205 [Cited within: 1]

Kumar S Zerrad E Yildirim A Biswas A 2013 Topological solitons and Lie symmetry analysis for the Kadomtsev-Petviashvili-Burgers equation with power law nonlinearity in dusty plasmas
Proc. Romanian Acad. A 14 204

[Cited within: 1]

Mhlanga I E Khalique C M 2017 Travelling wave solutions and conservation laws of a generalized (2 + 1)-dimensional Burgers-Kadomtsev-Petviashvili equation
AIP Conf. Proc. 1863 280006

DOI:10.1063/1.4992437 [Cited within: 1]

Ren B Wang J Y 2014 Full symmetry groups and exact solutions to BKP and GKP equations
Phys. Res. Int. 2014 1-4

DOI:10.1155/2014/786965 [Cited within: 1]

Taghizadeh N Mirzazadeh M Farahrooz F 2011 Exact soliton solutions of the modified KdV-KP equation and the Burgers-KP equation by using the first integral method
Appl. Math. Model. 35 3991

DOI:10.1016/j.apm.2011.02.001 [Cited within: 1]

Wazwaz A M 2008 Multiple-front solutions for the Burgers-Kadomtsev-Petviashvili equation
Appl. Math. Comput. 200 437

DOI:10.1016/j.amc.2007.11.032 [Cited within: 1]

Akinyemi L Savina T V Nepomnyashchy A A 2017 Exact solutions to a Muskat problem with line distributions of sinks and sources
Complex Analysis and Dynamical Systems VII, Contemp. Math. 699 19

[Cited within: 1]

Akinyemi L 2019 q-Homotopy analysis method for solving the seventh-order time-fractional Lax’s Korteweg–de Vries and Sawada-Kotera equations
Comput. Appl. Math 38 191

DOI:10.1007/s40314-019-0977-3 [Cited within: 1]

Iyiola O S Zaman F 2016 A note on analytical solutions of nonlinear fractional 2D heat equation with non-local integral terms
Pramana 87 51

DOI:10.1007/s12043-016-1239-1 [Cited within: 1]

Iyiola O S 2016 Exact and approximate solutions of fractional diffusion equations with fractional reaction terms
Progr. Fract. Differ. Appl. 2 21 30

DOI:10.18576/pfda/020103

Iyiola O S Wade B 2018 Exponential integrator methods for systems of non-linear space-fractional models with super-diffusion processes in pattern formation
Comput. Math. Appl. 75 3719 3736

DOI:10.1016/j.camwa.2018.02.027

Khater M M Lu D C Attia R A Inc M 2019 Analytical and approximate solutions for complex nonlinear schrodinger equation via generalized auxiliary equation and numerical schemes
Commun. Theor. Phys. 11 1267

DOI:10.1088/0253-6102/71/11/1267

Lin J Jin X W Gao X L Lou S Y 2018 Solitons on a periodic wave background of the modified KdV-Sine-Gordon equation
Commun. Theor. Phys. 70 119

DOI:10.1088/0253-6102/70/2/119 [Cited within: 1]

Yépez-Martínez H Gómez-Aguilar J F 2019 Fractional sub-equation method for Hirota-Satsuma-coupled KdV equation and coupled mKdV equation using the Atangana’s conformable derivative
Waves Random Complex Media 29 678-93

DOI:10.1080/17455030.2018.1464233 [Cited within: 1]

Yépez-Martínez H Gómez-Aguilar J F Baleanu D 2018 Beta-derivative and sub-equation method applied to the optical solitons in medium with parabolic law nonlinearity and higher order dispersion
Optik 155 357

DOI:10.1016/j.ijleo.2017.10.104

Zhang S Zhang H Q 2011 Fractional sub-equation method and its applications to nonlinear fractional PDEs
Phys. Lett. A 375 1069

DOI:10.1016/j.physleta.2011.01.029 [Cited within: 1]

Cenesiz Y Kurt A 2016 New fractional complex transform for conformable fractional partial differential equations
J. Appl. Math. Stat. Inform. 12 41

DOI:10.1515/jamsi-2016-0007 [Cited within: 1]

Malfliet W 1992 Solitary wave solutions of nonlinear wave equations
Amer. J. Phys. 60 650

DOI:10.1119/1.17120 [Cited within: 2]

Alquran M 2014 Analytical solutions of fractional foam drainage equation by residual power series method
Math. Sci. 8 153

DOI:10.1007/s40096-015-0141-1 [Cited within: 1]

Alquran M 2015 Analytical solution of time-fractional two-component evolutionary system of order 2 by residual power series method
J. Appl. Anal. Comput. 5 589



Arqub O A El-Ajou A Momani S 2015 Constructing and predicting solitary pattern solutions for nonlinear time-fractional dispersive partial differential equations
J. Comput. Phys. 293 385

DOI:10.1016/j.jcp.2014.09.034 [Cited within: 1]

Jaradat H M Al-Shara S Khan Q J Alquran M Al-Khaled K 2016 Analytical solution of time-fractional Drinfeld-Sokolov-Wilson system using residual power series method
IAENG Int. J. Appl. Math 46 64

[Cited within: 1]

相关话题/Analytical approximate solutions