Research article

Traveling wave solutions of the nonlinear Gilson-Pickering equation in crystal lattice theory

  • A.T. Nguyen , a, b ,
  • O. Nikan , c ,
  • Z. Avazzadeh , d, *
Expand
  • a Division of Applied Mathematics, Science and Technology Advanced Institute, Van Lang University, Ho Chi Minh City, Vietnam
  • b Faculty of Technology, Van Lang University, Ho Chi Minh City, Vietnam
  • c School of Mathematics, Iran University of Science and Technology, Narmak, Tehran, Iran
  • d Department of Applied Mathematics, Xi'an Jiaotong-Liverpool University, Suzhou 215123, China
* E-mail addresses: (A.T. Nguyen), (O. Nikan), (Z. Avazzadeh).

Received date: 2022-05-10

  Revised date: 2022-05-30

  Accepted date: 2022-06-04

  Online published: 2024-08-17

Abstract

This paper focuses on obtaining the traveling wave solutions of the nonlinear Gilson-Pickering equation (GPE), which describes the prorogation of waves in crystal lattice theory and plasma physics. The solution of the GPE is approximated via the finite difference technique and the localized meshless radial basis function generated finite difference. The association of the technique results in a meshless approach that does not require linearizing the nonlinear terms. At the first step, the PDE is converted to a system of nonlinear ODEs with the help of the radial kernels. In the second step, a high-order ODE solver is adopted to discretize the nonlinear ODE system. The global collocation techniques pose a considerable computationl burden due to the calculation of the dense algebraic system. The proposed method approximates differential operators over the local support domain, leading to sparse differentiation matrices and decreasing the computational burden. Numerical results and comparisons are provided to confirm the efficiency and accuracy of the method.

Cite this article

A.T. Nguyen , O. Nikan , Z. Avazzadeh . Traveling wave solutions of the nonlinear Gilson-Pickering equation in crystal lattice theory[J]. Journal of Ocean Engineering and Science, 2024 , 9(1) : 40 -49 . DOI: 10.1016/j.joes.2022.06.009

1. Introduction

Nonlinear partial differential equations (NPDEs) are a significant tool for the analysis of nonlinear physical processes and natural phenomena. Indeed, NPDEs play a major role in the description of the physical behavior of real-world processes and dynamical phenomena such as in ocean engineering, physics, fluid mechanics, geochemistry, plasma physics, optical fibers, geophysics, and many other scientific areas [1], [2], [3], [4], [5], [6], [7], [8]. Nonlinear processes are a field of interest to researchers in modern times. They have focused on finding the analytical or exact solutions to problems due to their contribution to the analysis of the actual system characteristics. For example, Helal [9] carried out a comprehensive overview of the soliton solutions for some common PDEs and introduced various analytical and numerical treatment techniques. According to [9], solitons represent a special nonlinear type of localized wave. In fact, a “ soliton” describes any solution of a nonlinear equation or system that: (i) represents a permanent wave; (ii) is localized, decaying or becoming constant at infinity, and (iii) may interact strongly with other solitons such that after the interaction it retains its form, almost obeying the principle of superposition [9].
In this work, we numerically study the traveling wave solutions of the following Gilson–Pickering equation (GPE) as a nonlinear third-order PDE as
u t ϵ u x x t + 2 κ u x u u x x x α u u x β u x u x x = 0 , ( x , t ) ( a , b ) × ( 0 , T ] ,
along with the initial and boundary conditions as u ( x , 0 ) = f ( x ) , a x b , u ( a , t ) = u ( b , t ) = 0 , t [ 0 , T ] ,where u is unknown function should be determined, subscripts denote the partial derivative, the parameters ϵ, κ, α, and β are non-zero real numbers, the function f ( x ) represents a continuous function, and notation T is a final time.
Gilson and Pickering [10] first introduced the GPE in 1995. There are three types of special cases for the nonlinear GPE based on specific choices of its parameters as follows. When ϵ = 1, α = 1, β = 3, and κ = 0.5, the GPE (1) converts to the Fornberg–Whitham model, which was developed to analyze the qualitative characteristics of wave breakage and admits a wave of the highest height [11]. For ϵ = 1, α = 3 and β = 2, the GPE (1) corresponds to the Fuchssteiner–Fokas–Camassa–Holm model, which is a completely integrable nonlinear NPDE that arises at various levels of approximation in shallow water theory [12] and when ϵ = 0, α = 1, β = 3, and κ = 0, the GPE (1) becomes the Rosenau–Hyman model, which occurs in the study of the influence of nonlinear dispersion on the structure of patterns in liquid drops [13]. The Camassa–Holm equation (CHE) constitutes the main form of the GPE [12]. The CHE is an NPDE capable of modeling waves in shallow water [12]. This PDE was introduced by Camassa and Holm [12] and has been demonstrated to have a robust mathematical structure. A significant property of this PDE is its acceptance of non-smooth and smooth solitary wave solutions that are solitons. One can enforce non-smooth or smooth solutions by twisting a parameter in the CHE. The “Peakonsǥ is the name given to the non-smooth solutions, which are solitons that have sharp cusps (or peaks). This leads to a discontinuous derivative of the soliton. Hence, these peakons are solutions merely in the distributional or weak sense. Interested readers can find further analyses regarding the physical and mathematical background of the CHE [12], [14].
To our best knowledge, some analytical and numerical methods have been cited for finding the solutions of the GPE. Irshad and Tauseef [15] employed the tanh-coth method for the numerical solution of GPE. Fan et al. [16] applied the G / G-expansion scheme for solving the GPE. Chen et al. [17] adopted the qualitative theory of polynomial differential system to study travelling wave solutions of the GPE, whereas Khakzad and Garshasbi [18] combined a meshfree technique with the Crank–Nicolson scheme to simulate the CHE. Saffarian and Zabihi [19] used a not-a-knot meshfree technique to approximate the GPE, while Ali and Mehanna [20] implemented the finite difference (FD) method to solve the GPE. Bilal et al. [21] developed the G / G 2-expansion and expansion function methods to derive new exact wave structures of the GPE. Kamal Ali et al. [22] considered the 1 / G -expansion and generalized exponential rational function approaches based on a homogeneous balance technique to construct solitary wave solutions of the GPE. Yokuş et al. [23] constructed the soliton solutions of the GPE with the help of the sinh-Gordon function. Rezazadeh et al. [24] considered the exponential rational function and the Jacobi elliptic functions schemes to find new wave surfaces of the GPE. Samir et al. [25] implemented a modified extended mapping technique to obtain solitary wave solutions of the GPE.
A meshless (meshfree) method is a numerical approach used to generate an algebraic system of equations over the whole problem domain without a predefined mesh to discretize the domain or boundary. Meshless methods require data points within the domain as opposed to being involved with node connectivity. The radial basis function (RBF) technique dose not depend on a grid and, hence, constitute a category of meshless methods [26], [27]. The RBF is formed by generating a univariate basic function with a norm that is usually Euclidean. This makes the problem dimension-independent and virtually one-dimensional. Over time, researchers have demonstrated the RBF interpolation to work where polynomial interpolation has failed [26], [28]. The RBF technique overcomes the limitations of polynomial interpolation since they do not require to be applied to tensor product grids or over rectangular domains. The RBF technique is frequently employed to express topographical surfaces and other complicated three-dimensional shapes [29]. They have been successfully implemented in diverse areas, such as medical imaging, ocean floor mapping, automotive and aircraft design, facial recognition, and climate modeling. The RBF techniques have been under active development over the last four decades, with the RBF research field remaining highly active as many questions in this area are still open. The RBF technique can be applied globally and locally. This work compares these two methods in order to demonstrate how the local method can lead to a similar accuracy as the global method while using only a small subset of the available points and, hence, considerably less computer memory. Some authors have tried localized RBF-based strategies, such as the localized RBF-generated FD (LRBF-FD) [30], [31], [32], [33], [34], [35], [36], [37], [38] and the localized RBF partition of unity (LRBF-PU) [39], [40], [41], [42], [43], which produce well-conditioned systems. These methods are advantageous in that they allow us to handle large-scale problems at a relatively small computational burden [26]. Based on [30], [32], the LRBF-FD, which is used in this paper, has an inherent similarity to FD although it has support domains that are built on scattered nodal points instead of grid points and is derived from an RBF-based interpolation instead of a polynomial-based one.
The main motivation of this paper is to construct a meshless LRBF-FD technique for finding the traveling wave solutions of the one dimensional nonlinear GPE. The FD approximates a linear operator differential at a specific point via a linear combination of the function’s values at the closest nodes (called a support domain or stencil). On the other hand, the LRBF-FD uses an RBF with local collocation similar to FD, and it decreases the number of links (stencil or support domain) for each point (center), resulting in a sparse matrix. In every support domain, we require to resolve a small-sized linear system of algebraic equations including a conditionally positive definite interpolation matrix. Indeed, the major advantage of the LRBF-FD lies in its ability to approximate differential operators over the local support domain, leading to sparse differentiation matrices while remarkably decreasing the computational burden.
The organizational structure of the article is as follows. Section 2 gives a description of the RBF and LRBF-FD collocation methods and presents a meshless approach of lines numerical technique by means of the LRBF-FD collocation technique to discretize the space direction of the governing equations. This results in a nolinear system of ordinary differential equations (ODEs) to which either a direct solver or a numerical time-stepping method must be applied to move in the time dimension. Section 3 reports the two numerical examples to confirm the computational efficiency and applicability of the LRBF-FD collocation method. In addition, we show that the obtained numerical experiments are in a good agreement with the results reported in literature. Finally, Section 4 provides some concluding remarks.

2. The RBF collocation method

Suppose that X = { x 1 , x 2 , , x N } R d, is a finite set of distinct data points belong to a bounded domain including with corresponding values f i for i = 1 , 2 , , N.

2.1. The RBF interpolation

According to the Kansa method [26], the solution u ( x ) can be approximated through a linear combination of the RBFs at discrete nodes as
u ( x ) s ( x , ε ) = j = 1 N a j ϕ j ( x ) ,
where λ j are the unknown expansion parameters to be determined, N the total number of the collocation nodal points, ϕ represents the radial function, ε is the shape parameter and ϕ j ( x ) = ϕ ( x x j , ε ), . denotes the Euclidean distance.
s ( x i , ε ) = f i , i = 1 , , N .
Inserting Eq. (3) in Eq. (2), we obtain following linear system:
A ϕ Λ = f ,
with A ϕ , i j = ϕ j ( x i ) , f = [ f 1 f 2 f N ] , Λ = [ a 1 a 2 a N ] , i , j = 1 , , N ,where the collocation method can be used for finding the unknown vector Λ.

2.2. The LRBF-FD method

Suppose that I i = { x i 1 , x i 2 , , x i n s } be a support domain of x i and we consider L as a linear differential operator. In other words, we have, I i = I ( x i ) = { j { 1 , , N } | x j x i R } ,where I ( x i ) represents the family of indices and R denotes known as the radius of support domain [30]. The differential operator L can be estimated by the weighted linear sum of function values as follows:
L u ( x i ) = j = 1 n s w j u ( x i j ) ,
where n s is the number of nearest neighboring points I i on x i surrounding collocation point x i containing the collocation node itself. Fig. 1 shows the distributed nodes in the computational domain with a support domain.
The weighting coefficients w 1 , w 2 , , w n s at the support domain are the unknown coefficients to be calculated. The stencil weights arise by enforcing the linear constraint (5) within the space spanned through the RBF { ϕ j ( x ) } | j = 1 n s that are centered at stencil node locations, so that
L u ( x i ) = j = 1 n s w j ϕ j ( x i j ) .
This expression (6) can be simplified to
A ϕ w = Ψ ,
where
w = [ w 1 w 2 w n s ] , Ψ = [ L ϕ i 1 ( x ) | x = x i L ϕ i 2 ( x ) | x = x i L ϕ i n s ( x ) | x = x i ] , A ϕ , r s = ϕ i r ( x i s ) , r , s = 1 , , n s .
The weighting coefficients w 1 , w 2 , , w n s can be calculated from the aforesaid system. For non-singularity of the coefficient matrix in the system (7), we can refer to [44].

2.3. Discretization of the GPE

Here, we discretize spatial derivatives of the GPE by using the LRBF-FD method. For this aim, the first, second and third order derivatives of u ( x , t ) can be estimated with the help of the function values at a set of n s nodes (including x i) in the support domain of x i. That is, we can write
u x ( x i , t ) = j = 1 n s w i , j x , 1 u ( x i j , t ) = W x u ( t ) ,
u x x ( x i , t ) = j = 1 n s w i , j x , 2 u ( x i j , t ) = W x x u ( t ) ,
u x x x ( x i , t ) = j = 1 n s w i , j x , 3 u ( x i j , t ) = W x x x u ( t ) ,
where u ( t ) = [ u 1 ( t ) , , u N ( t ) ] and the symbol w i , j x , l represents the weighted differences at the support domain for the order derivatives l = { 1 , 2 , 3 } . The structure of matrices W x, W x x and W x x x depends on the number of points in each support domain. For instance, if we choose three nodes in each support domain, then the matrices W x, W x x and W x x x are tridiagonal matrices.
Inserting Eqs. (9), (10) and (11) in Eq. (1) and collocating the nodal points in it, we obtain semi-discrete system in the following system of ODEs:
d ( I ϵ W xx ) u ( t ) d t = 2 κ W x u ( t ) + u ( t ) . * W x x x u ( t ) + α u ( t ) . * W x u ( t ) + β W x u ( t ) . * W x x u ( t ) .
Now, we use an ODE solver to solve the ODE system (12) in the temporal direction. We can implement for this aim, the ODE solver “command ode15s” that is acceptable for stiff ODEs and other traditional time-stepping methods. The method of lines constitutes an approach that utilizes FD in the temporal direction to determine the solution for the system of ODEs. Generally, this method may be considered stable provided each eigenvalue of the space discretization operator, scaled by using the temporal step (denoted by δ t), lies in the stability domain of the operator that discretizes time. In what follows, the Algorithm 1 illustrates the full discretization for the one dimensional nonlinear GPE.
Algorithm 1. Full discretization of the GPE.

3. Numerical results and discussion

The current section illustrates the simplicity, applicability and superiority of the proposed strategy by using two numerical examples on the GPE. We calculate the L , L 2, and L rms norm errors in order to check the accuracy and adaptability of the LRBF-FD as:
L = max 1 i N | u i exact u i numerical | , L 2 = i = 1 N ( u i exact u i numerical ) 2 , L rms = i = 1 N ( u i exact u i numerical ) 2 N ,
where N denotes the number of collocation nodes and u i exact and u i numerical represent the exact and numerical solutions, respectively.
In the numerical illustrations, we apply the multiquadric (MQ) RBF ϕ ( r ) = 1 + ε 2 r 2 as basis function with the shape parameter ε. Acceptable conditioning results need small shape parameters ε and large separation distances between the centers. Additionally, we apply the algorithm provided by Sarra [45] in this paper to calculate an optimal shape parameter (see Algorithm 2), where M represents a coefficient matrix, “svd” denotes the singular value decomposition (SVD), and σ max and σ min are the largest and smallest singular values obtained from the SVD, respectively. In this algorithm, we choose K min = 10 e + 02, K max = 10 e + 07 and ε Increment = 1 N · All numerical simulations were performed by using MATLAB R2016a in a computer system having a configuration with memory 8.00G RAM. The time needed to obtain the solutions of the GPE is denoted by “CPUTime” in seconds and the condition number (CN) of the coefficient matrix is evaluated based on the MATLAB command condest.
Example 1 Let us consider the nonlinear GPE (1) on the space interval [ 8 , 8 ] with the exact solution of the form u ( x , t ) = 3 ( c 2 κ ) ϵ c α ϵ c c + 2 κ ( 1 + tanh [ 1 2 2 κ c ϵ c ( x c t ) ] 2 ) ,and the following parameters in two cases
Case I ϵ = 1.50 , κ = 1.5 , α = 1.750 , c = 0.500 , β = 2,
Case II ϵ = 0.75 , κ = 1.5 , α = 0.875 , c = 0.125 , β = 2,
where the initial condition can be obtained from the aforementioned exact solution [15], [19].
Algorithm 2. Optimal shape parameter [45].
This problem is examined by means of the LRBF-FD method with various parameters of δ t, N and n s at the final time T. Table 1 lists the L norm errors, CN and CPU run times (in seconds) for Case I with δ t = 0.001 when T = 3 at various values of stencil sizes n s. Table 2 compares the L , L 2, and L rms norm errors and CPU run times with method described in Zabihi and Saffarian [19] for Case I by taking δ t = 0.001, N = 400 and n s = 135 at several final times. Table 3 makes the comparison of the L , L 2, and L rms norm errors and CPU run times with method introduced in Zabihi and Saffarian [19] for Case II by choosing δ t = 0.001, N = 300 and n s = 125 at several final times. In view of Tables 2 and 3, we can see that the LRBF-FD collocation technique is more accurate than the scheme presented in Zabihi and Saffarian [19]. Fig. 2 plots the motion of the single solitary wave with N = 400 , n s = 145 and δ t = 0.001 at different final times T { 0 , 1 , 2 , 3 , 4 , 5 , 6 } in Case I over spatial domain [ 8 , 8 ]. Fig. 3 represents the motion of the single solitary wave with N = 400 , n s = 145 and δ t = 0.001 at various final times T { 0 , 3 , 4 , 6 , 8 , 9 , 12 } in Case II over spatial domain [ 8 , 8 ]. As seen in Figs. 2 and 3, the single solitons move to the left at a constant speed preserving their amplitude and shape. Fig. 4 portraits the soliton profiles in Case I when δ t = 0.001, N = 600 and n s = 251 at various temporal intervals. Fig. 5 presents the soliton profiles in Case II by letting δ t = 0.001, N = 700 and n s = 451 at various temporal intervals. Fig. 6 depicts the contour of solutions and associated absolute errors in Case I by taking δ t = 0.001, N = 500 and n s = 335 at temporal interval [0,4]. Fig. 7 represents the contour of solutions and associated absolute errors in Case II by choosing δ t = 0.001, N = 400 and n s = 121 at temporal interval [0,2]. Finally, Fig. 8 displays the sparsity patterns of the coefficient matrix M for two different stencil sizes n s with N = 100 .
Example 2 We consider the nonlinear GPE (1) on the space interval [ 8 , 8 ] with the exact solution of the form u ( x , t ) = c exp ( | x c t | ) ,and the following parameters
ϵ = 1 , κ = 0 , α = 3 , c = 1 × 10 4 , β = 2 , where the initial condition can be computed from the aforesaid exact solution [15], [19].
Table 1. L norm errors, CN and CPUTime of Example 1 for Case I with δ t = 0.001 when T = 1 at various stencil sizes n s.
N n s L CN CPUTime
400 113 9.5404 e 05 4.2971 e + 05 2.341001
400 119 7.9713 e 05 4.1670 e + 05 2.380281
400 139 2.3485 e 05 3.5491 e + 05 2.420148
400 155 1.4817 e 05 5.8869 e + 05 2.556396
400 175 1.4021 e 05 4.9684 e + 05 2.603521
Table 2. L , L 2, and L rms norm errors and CPUTime when δ t = 0.001, N = 400 and n s = 135 at different final times T in Case I over the spatial domain [ 8 , 8 ] for Example 1.
T Method [19] LRBF-FD
L L 2 L rms L L 2 L rms CPUTime
0.5 1.8425 e 06 2.0140 e 05 5.0304 e 07 7.4506 e 07 3.5181 e 06 2.0312 e 07 0.325940
1.0 7.2685 e 06 9.3918 e 05 2.3458 e 06 1.3576 e 06 6.7522 e 06 3.8984 e 07 0.423046
1.5 2.2759 e 05 2.8776 e 04 7.1872 e 06 1.8230 e 06 1.0367 e 05 5.9856 e 07 0.524313
2.0 6.9942 e 05 8.9111 e 04 2.2257 e 06 3.4505 e 06 1.6473 e 05 9.5108 e 07 0.631450
Table 3. L , L 2, and L rms norm errors and CPUTime by considering δ t = 0.001, N = 300 and n s = 125 at different final times T in Case II over spatial domain [ 8 , 8 ] for Example 1.
T Method [19] LRBF-FD
L L 2 L rms L L 2 L rms CPUTime
1 9.0137 e 07 5.6183 e 06 1.4033 e 07 9.8143 e 09 3.9591 e 08 2.2858 e 09 0.295395
2 3.6118 e 06 3.4581 e 05 8.6371 e 07 2.9988 e 08 1.4725 e 07 8.5016 e 09 0.327970
3 1.1994 e 05 1.1297 e 04 2.8216 e 06 7.0717 e 08 3.9377 e 07 2.2735 e 08 0.357182
4 3.2036 e 05 2.9399 e 04 7.3428 e 06 5.5848 e 08 9.6731 e 07 1.5208 e 07 0.381461
5 7.7520 e 05 7.0524 e 04 1.7614 e 05 3.7515 e 07 2.1518 e 06 1.2424 e 07 0.406676
Fig. 2. Motion of the single solitary wave with N = 400 , n s = 145 and δ t = 0.001 at different final times T { 0 , 1 , 2 , 3 , 4 , 5 , 6 } in Case I over spatial domain [ 8 , 8 ] for Example 1.
Fig. 3. Motion of the single solitary wave with N = 400 , n s = 145 and δ t = 0.001 at various final times T { 0 , 3 , 4 , 6 , 8 , 9 , 12 } in Case II over spatial domain [ 8 , 8 ] for Example 1.
Fig. 4. The soliton profiles of the GPE when δ t = 0.001, N = 600 and n s = 251 in Case I over spatial domain [ 8 , 8 ] for Example 1.
Fig. 5. The soliton profiles of the GPE with δ t = 0.001, N = 700 and n s = 451 in Case II over spatial domain [ 8 , 8 ] for Example 1.
Fig. 6. The contour of solutions and associated absolute errors when δ t = 0.001, N = 500 and n s = 335 in Case I over spatial domain [ 8 , 8 ] for Example 1.
Fig. 7. The contour of solutions and associated absolute errors with δ t = 0.001, N = 400 and n s = 121 in Case II over the spatial domain [ 8 , 8 ] for Example 1.
Fig. 8. The sparsity patterns of the coefficient matrix M for two different stencil sizes n s with N = 100.
Here, we simulate this example by using the LRBF-FD method with various parameters of δ t, N and n s at the final time T. Due to the numerical solutions in this example are small, therefore, we use the relative error L re to show the effectiveness of the method as:
L re = L u exact ·
Table 4 reports the L norm errors, CN and CPUTime with δ t = 0.001 when T = 10 for different values of stencil sizes n s. Table 5 compares the L , L 2, L rms and L re norm errors and CPU run times with technique introduced in Zabihi and Saffarian [19] by letting δ t = 0.001, N = 600 and n s = 425 at various values of final times. From Table 5, we can observe that the LRBF-FD collocation technique is more accurate than the method described in Zabihi and Saffarian [19]. Fig. 9 shows the soliton profiles with δ t = 0.001, N = 650 and n s = 371 at various temporal intervals. Finally, Fig. 10 portraits the contour of solutions and associated absolute errors when δ t = 0.001, N = 800 and n s = 401 at temporal interval [0,6].
Table 4. L norm errors, CN and CPUTime of Example 1 with δ t = 0.01 when T = 10 for different stencil sizes.
N n s L CN CPUTime
800 191 6.8288 e 08 1.7910 e + 06 0.520603
800 219 6.8284 e 08 1.7910 e + 06 0.528558
800 299 6.7364 e 08 2.3978 e + 06 0.529682
800 609 6.7167 e 08 4.1260 e + 06 0.743456
800 685 6.7013 e 08 1.6611 e + 06 0.697687
Table 5. L , L 2, L rms, and L re norm errors and CPUTime when δ t = 0.001, N = 600 and n s = 425 at various final times T over the spatial domain [ 8 , 8 ] for Example 2.
T Method [19] LRBF-FD
L L 2 L rms L re L L 2 L rms L re CPUTime
1 1.7273 e 08 2.4865 e 07 6.2104 e 09 1.7274 e 04 3.8678 e 09 5.1417 e 09 2.2994 e 10 4.0174 e 05 0.431546
2 3.3891 e 08 4.8857 e 07 1.2203 e 08 3.3897 e 04 7.6964 e 09 1.0266 e 08 4.5913 e 10 8.0849 e 05 0.437426
3 4.9899 e 08 7.2176 e 07 1.8027 e 08 4.9914 e 04 1.1587 e 08 1.5530 e 08 6.9454 e 10 1.2169 e 04 0.444489
4 6.4994 e 08 9.4356 e 07 2.3567 e 08 1.5020 e 03 1.5370 e 08 2.0675 e 08 9.2461 e 10 1.6280 e 04 0.466018
5 8.2218 e 08 1.1909 e 06 1.7614 e 08 8.2260 e 03 1.8929 e 08 2.5521 e 08 1.1413 e 09 2.0419 e 04 0.467160
Fig. 9. The soliton profiles of the GPE when δ t = 0.001, N = 650 and n s = 371 over the spatial domain [ 8 , 8 ] in Example 2.
Fig. 10. The contour of solutions and associated absolute errors with δ t = 0.001, N = 800 and n s = 401 over spatial domain [ 8 , 8 ] in Example 2.

4. Conclusions

This paper proposed an efficient and robust computational algorithm to simulate the one dimensional nonlinear GPE without using the linearization. The structure of the method involved two stages. First, the spatial dimension was approximated by virtue of the LRBF-FD collocation method. Second, the time dimension was discretized using the method of lines that yields a system of nonlinear ODEs. Furthermore, the obtained nonlinear system of ODEs was solved using an ODE solver to derive high-numerical results. The LRBF-FD collocation method efficiently handles the computational load unlike the case with the dense and ill-conditioned matrices required using global RBF collocation techniques. Two numerical examples demonstrated that the proposed method is efficient and accurate. The results obtained in this paper showed that the proposed method has great potential for simulating nonlinear evaluations.

Declaration of Competing Interest

Authors declare that they have no conflict of interest.

Acknowledgment

The authors are very grateful to the three anonymous referees for their valuable comments and suggestions that improved the original manuscript greatly. In addition, Anh Tuan Nguyen is supported by Van Lang University.
[1]
S.R. Islam, S.Y. Arafat, H. Wang. J. Ocean Eng. Sci. (2022), DOI:10.1016/j.joes.2022.01.012

[2]
A. Tripathy, S. Sahoo. J. Ocean Eng. Sci., 6 (2021), pp. 405-409

[3]
L. Akinyemi, M. Åženol, M. Osman. J. Ocean Eng. Sci., 7 (2021), pp. 143-154

[4]
A. Ali, A.R. Seadawy. J. Ocean Eng. Sci., 6 (2021), pp. 85-98

[5]
L. Ravi, S.S. Ray, S. Sahoo. J. Ocean Eng. Sci., 2 (2017), pp. 34-46

[6]
M.M. Khater, D. Kumar. J. Ocean Eng. Sci., 2 (2017), pp. 223-228

[7]
S. Arora, R. Jain, V. Kukreja. J. Ocean Eng. Sci. (2022), DOI:

[8]
V. Kukreja. J. Ocean Eng. Sci., 7 (2022), pp. 99-111

[9]
M. Helal. Chaos, Solitons Fractals, 13 (2002), pp. 1917-1929

[10]
C. Gilson, A. Pickering. J. Phys. , 28 (1995), p. 2871 DOI: 10.1088/0305-4470/28/10/017

[11]
G.B. Whitham. Linear and nonlinear waves. John Wiley & Sons (2011)

[12]
R. Camassa, D.D. Holm. Phys. Rev. Lett., 71 (1993), p. 1661

[13]
P. Rosenau, J.M. Hyman. Phys. Rev. Lett., 70 (1993), p. 564

[14]
X. Raynaud, On a shallow water wave equation, phd thesis, norwegian university of science and technology, 2006.

[15]
A. Irshad, S.T. Mohyud. Stud. Nonlinear Sci., 3 (2012), pp. 24-48

[16]
X. Fan, S. Yang, D. Zhao. Int. J. Nonlinear Sci., 8 (2009), pp. 368-373

[17]
A. Chen, W. Huang, S. Tang. Nonlinear Anal., 10 (2009), pp. 2659-2665

[18]
M. Garshasbi, M. Khakzad. J. Adv. Res. Dyn. Control Syst., 4 (2015), pp. 65-83

[19]
F. Zabihi, M. Saffarian. Eng. Comput., 34 (2018), pp. 37-44 DOI: 10.1007/s00366-017-0519-9

[20]
K.K. Ali, M. Mehanna. Results Phys., 28 (2021), p. 104596

[21]
M. Bilal, A.R. Seadawy, M. Younis, S. Rizvi, K. El-Rashidy, S.F. Mahmoud. Results Phys., 23 (2021), p. 103959

[22]
K.K. Ali, H. Dutta, R. Yilmazer, S. Noeiaghdam. Front. Phys., 8 (2020), p. 54

[23]
A. YokuÅŸ, H. Durur, K.A. Abro, D. Kaya. Eur. Phys. J. Plus., 135 (2020), pp. 1-19 DOI: 10.1109/sensors47125.2020.9278818

[24]
H. Rezazadeh, A. Jhangeer, E. Tala-Tebue, M.S. Hashemi, S. Sharif, H. Ahmad, S.W. Yao. Results Phys. 24 (2021), p. 104192

[25]
I. Samir, N. Badra, H.M. Ahmed, A.H. Arnous, A.S. Ghanem. Results Phys., 36 (2022), p. 105427

[26]
W. Chen, Z.-J. Fu, C.S. Chen. Recent advances in radial basis function collocation methods. Springer (2014)

[27]
Y. Gu. Int. J. Comput. hods, 2 (2005), pp. 477-515

[28]
S.A. Sarra. Eng. Anal. Bound. Elem., 35 (2011), pp. 68-76

[29]
T. Sauer, Numerical analysis with CD-Rom, 2006.

[30]
V. Bayona, M. Moscoso, M. Carretero, M. Kindelan. J. Comput. Phys., 229 (2010), pp. 8281-8295

[31]
O. Nikan, Z. Avazzadeh, J.T. Machado. J. Comput. Sci., 53 (2021), p. 101394

[32]
J. Rashidinia, M.N. Rasoulizadeh. Wave Motion, 90 (2019), pp. 152-167

[33]
M.N. Rasoulizadeh, J. Rashidinia. J. King Saud Univ. Sci., 32 (2020), pp. 2277-2283

[34]
J. Rashidinia, M. Rasoulizadeh. TWMS J. Appl. Eng. Math., 11 (2021), p. 893

[35]
O. Nikan, A. Golbabai, T. Nikazad. Eur. Phys. J. Plus., 134 (2019), p. 367

[36]
M. Rasoulizadeh, M. Ebadi, Z. Avazzadeh, O. Nikan. Eng. Anal. Bound. Elem., 131 (2021), pp. 258-268

[37]
Z. Avazzadeh, O. Nikan, J.A.T. Machado. Mathematics, 8 (2020), p. 1601 DOI: 10.3390/math8091601

[38]
M. Rasoulizadeh, O. Nikan, Z. Avazzadeh. Math. Sci., 15 (2021), pp. 365-376 DOI: 10.1007/s40096-021-00375-8

[39]
O. Nikan, Z. Avazzadeh, J.T. Machado. Int. Commun. Heat Mass Transf., 132 (2022), p. 105892

[40]
G. Garmanjani, R. Cavoretto, M. Esmaeilbeigi. Comput. Math. Appl., 75 (2018), pp. 4066-4090

[41]
R. Cavoretto, A. De Rossi, E. Perracchione. J. Sci. Comput., 74 (2018), pp. 1-22 DOI: 10.1007/s10915-017-0418-7

[42]
O. Nikan, Z. Avazzadeh. Eng. Anal. Bound. Elem., 130 (2021), pp. 268-285

[43]
O. Nikan, Z. Avazzadeh, M. Rasoulizadeh. Nonlinear Dyn., 106 (2021), pp. 783-813 DOI: 10.1007/s11071-021-06822-4

[44]
G.E. Fasshauer. Meshfree approximation methods with MATLAB. 6, World Scientific (2007)

[45]
S.A. Sarra. Comput. Math. Appl., 218 (2012), pp. 9853-9865

Outlines

/