Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article Special Issues

Solving the reaction-diffusion Brusselator system using Generalized Finite Difference Method

  • In this paper, we investigate the numerical solution of the Brusselator system using a meshless method. A numerical scheme is derived employing the formulas of the Generalized Finite Difference Method, and the convergence of the approximate solution to the exact solution is examined. In order to demonstrate the applicability and accuracy of the method, several examples are proposed.

    Citation: Ángel García, Francisco Ureña, Antonio M. Vargas. Solving the reaction-diffusion Brusselator system using Generalized Finite Difference Method[J]. AIMS Mathematics, 2024, 9(5): 13211-13223. doi: 10.3934/math.2024644

    Related Papers:

    [1] Chunting Ji, Hui Liu, Jie Xin . Random attractors of the stochastic extended Brusselator system with a multiplicative noise. AIMS Mathematics, 2020, 5(4): 3584-3611. doi: 10.3934/math.2020233
    [2] Bin He . Developing a leap-frog meshless methods with radial basis functions for modeling of electromagnetic concentrator. AIMS Mathematics, 2022, 7(9): 17133-17149. doi: 10.3934/math.2022943
    [3] Mourad S. Semary, M. T. M. Elbarawy, Aisha F. Fareed . Discrete Temimi-Ansari method for solving a class of stochastic nonlinear differential equations. AIMS Mathematics, 2022, 7(4): 5093-5105. doi: 10.3934/math.2022283
    [4] Aisha F. Fareed, Emad A. Mohamed, Mokhtar Aly, Mourad S. Semary . A novel numerical method for stochastic conformable fractional differential systems. AIMS Mathematics, 2025, 10(3): 7509-7525. doi: 10.3934/math.2025345
    [5] Leqiang Zou, Yanzi Zhang . Numerical solutions of multi-term fractional reaction-diffusion equations. AIMS Mathematics, 2025, 10(1): 777-792. doi: 10.3934/math.2025036
    [6] Din Prathumwan, Thipsuda Khonwai, Narisara Phoochalong, Inthira Chaiya, Kamonchat Trachoo . An improved approximate method for solving two-dimensional time-fractional-order Black-Scholes model: a finite difference approach. AIMS Mathematics, 2024, 9(7): 17205-17233. doi: 10.3934/math.2024836
    [7] Eunjung Lee, Dojin Kim . Stability analysis of the implicit finite difference schemes for nonlinear Schrödinger equation. AIMS Mathematics, 2022, 7(9): 16349-16365. doi: 10.3934/math.2022893
    [8] Cagnur Corekli . The SIPG method of Dirichlet boundary optimal control problems with weakly imposed boundary conditions. AIMS Mathematics, 2022, 7(4): 6711-6742. doi: 10.3934/math.2022375
    [9] Yanhua Gu . High-order numerical method for the fractional Korteweg-de Vries equation using the discontinuous Galerkin method. AIMS Mathematics, 2025, 10(1): 1367-1383. doi: 10.3934/math.2025063
    [10] Mubashara Wali, Sadia Arshad, Sayed M Eldin, Imran Siddique . Numerical approximation of Atangana-Baleanu Caputo derivative for space-time fractional diffusion equations. AIMS Mathematics, 2023, 8(7): 15129-15147. doi: 10.3934/math.2023772
  • In this paper, we investigate the numerical solution of the Brusselator system using a meshless method. A numerical scheme is derived employing the formulas of the Generalized Finite Difference Method, and the convergence of the approximate solution to the exact solution is examined. In order to demonstrate the applicability and accuracy of the method, several examples are proposed.



    In this paper we focus on the numerical discretization of the non-linear Brusselator system, which describes oscillatory chemical behavior in autocatalytic reactions. The Brusselator equations read as:

    {Ut=U2V2U+14ΔU,(x,y)Ω,t>0,Vt=UU2V+14ΔV,(x,y)Ω,t>0, (1.1)

    where xΩ, time t>0, and U and V are the concentrations of the chemical species (the system is also valid to model biological species). Throughout this paper ΩR2 is a bounded domain with a regular boundary.

    For its importance in biological or chemical processes we consider solving system (1.1) numerically using the meshless method of Generalized Finite Differences. Recently, several numerical methods have been used for solving reaction-diffusion Brusselator system. For instance, in [1] authors used the dual-reciprocity boundary element method (DRBEM) for the numerical solution. Also, a method of lines implemented as a meshless method based on spatial trial spaces spanned by the Newton basis functions have been applied in [2]. A non-standard finite difference method has been applied in [3], weighted residual methods have been used in [4] and the polynomial based differential quadrature method has been considered in [5] for solving the Brusselator system.

    By the flexibility of our method due to the easy choice of the clouds of points and the steps of time and space, we derive a discretization of the system by means of the Generalized Finite Difference Method (GFDM) and we prove the convergence of the discrete solution to the analytical one. Several numerical examples on the applications of this meshless method over regular and irregular domains are presented in order to illustrate our results. Very recently, a high order accuracy of time discretization technique has been combined with the generalized finite difference method (see [6] and [7]).

    The paper is organized as follows: In Section 2 we recall the analytical foundations of the GFDM and obtain the explicit formulas of the spatial derivatives. Section 3 is devoted to the analytical study of the explicit GFD scheme where we prove the main result of this section, enclosed in Theorem 3.1. In Section 4 we present several numerical examples over regular and irregular domains which show the applicability of the method. Finally, some conclusions are obtained in Section 5.

    The aim of this section is to obtain explicit linear expressions for the approximation of partial derivatives in the points (nodes) of the domain. First of all, an irregular grid or cloud of nodes is generated in the domain. For each one of the nodes of the domain, a star is defined as a set of selected nodes x0,x1,,xs, where x0 is denominated central node of star. In order to select the other nodes of star different criteria can be used, see [8,9].

    Let x0=(x0,y0) be the central node and hi=xix0,ki=yiy0, where (xi,yi) are the coordinates of the ith node of the star. Let us put U0=U(x0) and Ui=U(xi), then by the Taylor series expansion, for i=1,...,s, we have

    Ui=U0+hiU0x+kiU0y+12(h2i2U0x2+k2i2U0y2+2hiki2U0xy)+. (2.1)

    Let us call

    ciT={hi,ki,h2i2,k2i2,hiki} (2.2)

    and the spatial derivatives

    DT5={u0x,u0y,2u0x2,2u0y2,2u0xy}. (2.3)

    If in (2.1) we don't consider the higher than second-order terms, we may obtain a second-order approximation of Ui, which we shall denote by ui. We may then define:

    B(u)=si=1[(u0ui)+hiu0x+kiu0y++12(h2i2u0x2+k2i2u0y2+2hiki2u0xy)]2φ2i, (2.4)

    where φi=φ(hi,ki) are positive symmetrical weighting functions which decrease in magnitude as the distance to the center increases. Some weighting functions such as potentials or exponentials can be used (see [10] for more details).

    In order to minimize the error, the norm function given by (2.4) is minimized with respect to the partial derivatives and the following linear system is obtained:

    A(hi,ki,φi)D5=b(hi,ki,φi,u0,ui). (2.5)

    By solving (2.5), the partial derivatives can be obtained as a function of the values (u0,ui,hi,ki,φi).

    Remark 2.1. Matrix A defined in (2.5) is a positive definite matrix (see [10] for a complete proof) and the approximation of the spatial derivatives is of second order O(h2i,k2i).

    If in accordance with [10] we define

    A1=QQT, (2.6)

    then

    D5=QQTb (2.7)

    and (2.7) can be rewritten as

    D5=u0QQTsi=1φ2ici+QQTsi=1uiφ2ici (2.8)

    or

    D5=QQTΦ(uu01), (2.9)

    where 1={1,1,,1}T. Thus, spatial derivatives using GFD as in [11] are approximated by

    2u(x0,nΔt)x2+2u(x0,nΔt)y2=m0un0+si=1miuni+O(h2i,k2i), (2.10)

    Finally, time derivative is approximated as follows

    ut(x0,nΔt)=u(x0,(n+1)Δt)u(x0,nΔt)Δt+O(Δt).

    Let us consider a bounded domain ΩR2. Then, the GFD scheme for system (1.1) is:

    {un+10un0Δt=(un0)2vn02un0+14[m0un0+si=1miuni]+O(Δt,h2i,k2i),vn+10vn0Δt=un0(un0)2vn0+14[m0un0+si=1mivni]+O(Δt,h2i,k2i), (3.1)

    Thus,

    {un+10=un0[12Δtm0Δt4]+Δt[(un0)2vn0+si=1miuni4]+O(Δt,h2i,k2i),vn+10=Δt[un0(un0)2vn0+si=1mivni4]+vn0[1m0Δt4]+O(Δt,h2i,k2i), (3.2)

    In order to prove the main result of the paper concerning the conditional convergence of the GFD scheme for solving system (1.1), we need the following basic results:

    Lemma 3.1. Let AMn×n(R). If there exists some matrix norm such that A<1, then

    lim

    Lemma 3.2. Assuming A\in\mathfrak{M}_{n\times n}(\mathbb{R}) , then the following are equivalent:

    (i) \lim\limits_{k\to\infty}A^k = \textbf{0},

    (ii) \rho(A) < 1 ,

    where \rho(\cdot) stands for the spectral radius.

    Our main result with respect to the proposed numerical scheme is as stated below

    Theorem 3.1. Let (U, V) be the exact solution to system (1.1). Then, the GFD explicit scheme given by (3.2) is convergent if the following holds:

    \begin{equation} \Delta t\leq\frac{1}{1+\frac{m_0}{4}-v_0^n\left(u_0^n+U_0^n \right) }. \end{equation} (3.3)

    Proof.

    Notice that, since U, V are the exact solution of system (1.1), they also solve the discrete equation (3.2). Let us denote by U^n_i the exact U -solution at time n and node i (respetively V^n_i ). We take the difference between (3.2) and the same expression for the exact solution. We call \tilde{u}^n_i = u^n_i-U^n_i (similarly for \tilde{v}^n_i ) and notice the following relations:

    \begin{equation} \begin{split} \left(u_0^n \right)^2v_0^n-\left(U_0^n \right)^2V_0^n = \\ &\left(u_0^n \right)^2v_0^n-\left(U_0^n \right)^2v_0^n+\left(U_0^n \right)^2v_0^n -\left(U_0^n \right)^2V_0^n = \\ &\tilde{u}_0^n\left(u_0^n+U_0^n \right)v_0^n +\tilde{v}_0^n\left(U_0^n\right)^2 . \end{split} \end{equation} (3.4)

    Then,

    \begin{equation} \left\lbrace \begin{aligned}{} &\tilde{u}_0^{n+1} = \tilde{u}_0^n\left[1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right) \right] +\\ &+\Delta t\frac{ \sum\limits_{i = 1}^sm_iu^n_i}{4}+\Delta t\tilde{v}_0^n\left(U_0^n \right) ^2 +\mathcal{O}(\Delta t, h_i^2, k_i^2), \\ &\tilde{v}_0^{n+1} = \Delta t\tilde{u}_0^n\left[1-v_0^n\left(u_0^n+U_0^n \right) \right]+\tilde{v}_0^n\left[1-\frac{m_0\Delta t}{4}-\Delta t\left(U_0^n \right)^2 \right] +\\ &+\Delta t\frac{ \sum\limits_{i = 1}^sm_i\tilde{v}^n_i}{4} +\mathcal{O}(\Delta t, h_i^2, k_i^2), \\ \end{aligned} \right. \end{equation} (3.5)

    Let us call \tilde{u} = \max_{i}\{|\tilde{u}^n_i|\} (similarly for \tilde{v} ). Then, we write (3.5) as

    \begin{equation} \left\lbrace \begin{aligned}{} &\tilde{u}^{n+1} = \tilde{u}^n\left[\left| 1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right)\right| +\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}\right] +\\ &+\Delta t\tilde{v}^n\left(U_0^n \right) ^2 +\mathcal{O}(\Delta t, h_i^2, k_i^2), \\ &\tilde{v}^{n+1} = \Delta t\tilde{u}^n\left| 1-v_0^n\left(u_0^n+U_0^n \right)\right| +\\ &+\tilde{v}^n\left[\left| 1-\frac{m_0\Delta t}{4}-\Delta t\left(U_0^n \right)^2\right|+\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4} \right] +\mathcal{O}(\Delta t, h_i^2, k_i^2), \\ \end{aligned} \right. \end{equation} (3.6)

    Let us rewrite (3.6) as

    \begin{equation} \left( \begin{array}{c} \tilde{u}^{n+1} \\ \tilde{v}^{n+1}\\ \end{array} \right)\leq \left( \begin{array}{cc} C_1&\Delta t\left(U_0^n \right) ^2 \\ \Delta t\left[\left| 1-v_0^n\left(u_0^n+U_0^n \right)\right| \right]&C_2 \\ \end{array} \right)\left( \begin{array}{c} \tilde{u}^{n} \\ \tilde{v}^{n}\\ \end{array} \right). \end{equation} (3.7)

    where

    C_1 = \left| 1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right)\right|+ \Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}
    C_2 = \left| 1-\frac{m_0\Delta t}{4}-\Delta t\left(U_0^n \right)^2\right|+\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}.

    Consider the matrix

    \begin{equation} \mathfrak{A} = \left( \begin{array}{cc} C_1&\Delta t\left(U_0^n \right) ^2 \\ \Delta t\left[\left| 1-v_0^n\left(u_0^n+U_0^n \right)\right| \right]&C_2 \\ \end{array} \right), \end{equation} (3.8)

    Let us consider the matrix norm N_{1}(\mathfrak{A}) = \max_{i = 1, 2}\{\sum_{j = 1}^2|a_{ij}|\} . From their definition it is clear that

    \begin{equation} N_{1}(\mathfrak{A}) = C_1+\Delta t\left(U_0^n \right) ^2. \end{equation} (3.9)

    Then,

    \begin{equation} C_1+\Delta t\left(U_0^n \right) ^2 < 1\Rightarrow \left| 1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right)\right| < 1-\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}-\Delta t\left(U_0^n \right) ^2, \end{equation} (3.10)

    inequality (3.10) is equivalent to

    \begin{equation} \left\lbrace \begin{aligned}{} &-1+\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}+\Delta t\left(U_0^n \right) ^2 < 1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right), \\ &1-2 \Delta t-\frac{m_0\Delta t}{4}-\Delta t v_0^n\left(u_0^n+U_0^n \right) < 1-\Delta t\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}-\Delta t\left(U_0^n \right) ^2, \\ \end{aligned} \right. \end{equation} (3.11)

    and inequalities (3.11) are equivalent to

    \begin{equation} \left\lbrace \begin{aligned}{} &\Delta t < \frac{1}{1+\frac{m_0}{4}- v_0^n\left(u_0^n+U_0^n \right)}, \\ &\frac{ \sum\limits_{i = 1}^s\left| m_i\right|}{4}+\left(U_0^n \right) ^2+v_0^n\left(u_0^n+U_0^n \right) < 1+\frac{m_0}{4}.\\ \end{aligned} \right. \end{equation} (3.12)

    Applying Lemma 3.1, we have that \lim_{k\to\infty}\mathfrak{A}^k = \textbf{0} . Now, by Lemma 3.2, this is equivalent to \rho(\mathfrak{A}) < 1, that is, the greatest absolute value of all eigenvalues of matrix \mathfrak{A} is bounded by 1, which implies the convergence of the explicit scheme under the condition (3.3).

    In this section we illustrate the application of the GFDM for solving the rection-diffusion Brusselator system given by (1.1). We test the method using the regular and irregular clouds of points of Figure 1.

    Figure 1.  Clouds of points used: 1 to 4.

    In all the cases considered, the errors has been calculated according to the norms

    \begin{equation} \left\{\begin{array}{l} l_2 = \sqrt{\dfrac{ \sum\limits_{i = 1}^{NI} \bigg( sol(i) - exac(i) \bigg)^2}{NI}}\\ \\ l_\infty = max|sol(i)-exac(i)| \end{array}\right. \end{equation} (4.1)

    where sol(i) is the GFD solution in node i , exac(i) is the exact value of the solution at node i , and NI is the number of nodes of the domain \Omega .

    For all the numerical results shown in the following sections, we choose \Delta t as stated, which is to say, the condition of convergence is chosen as the minimum value computated using formulaes Eq (3.3) for each star of the domain \Omega . Notice that, therefore, each \Delta t may vary from one cloud of points to another since its value depends on the star of nodes. We choose \Delta t = 0.001 , fulfilling the assumption made in Theorem 3.1. The distance criterion has been used, the number of nodes per star is 8 plus the central node and the weighting function is the inverse of the distance squared, \varphi(d_i) = \frac{1}{d_i^2} with d_i being the distance from node i to the central node of the star.

    We present two cases with different boundary conditions: Dirichlet and Neumann.

    Case 1: Dirichlet boundary conditions

    We consider for this first case the clouds of points 1 and 2 of Figure 1.

    \begin{equation} \left\lbrace \begin{aligned}{} &\frac{\partial{U}}{\partial{t}} = U^2V-2U+\dfrac{1}{4}\Delta{U}, & (x, y)\in \Omega, \; t > 0, \\ &\frac{\partial V}{\partial t} = U-U^2V+\dfrac{1}{4}\Delta V , &(x, y)\in \Omega, \; t > 0 , \\ &U(x, y, 0) = e^{-(x+y)}, \, V(x, y, 0) = e^{(x+y)}, &(x, y)\in \Omega .\\ \end{aligned} \right. \end{equation} (4.2)

    By a direct check, the exact solution is (the Dirichlet boundary conditions are chosen so that the equation is fulfilled):

    \begin{equation} U(x, y, t) = e^{-\left(\frac{t}{2}+x+y \right) };\, V(x, y, t) = e^{\left(\frac{t}{2}+x+y \right) } \end{equation} (4.3)

    Table 1 shows the error norms in the clouds of points.

    Table 1.  Error norms l_2 and l_{\infty} for case 1 in clouds of nodes of the Figure 1.
    Cloud of nodes l_2(U) l_{\infty}(U) l_2(V) l_{\infty}(V)
    Cloud 1 6.7004\times 10^{-6} 1.7156\times 10^{-5} 8.7630\times 10^{-5} 1.7048\times 10^{-4}
    Cloud 2 2.6218\times 10^{-4} 5.4143\times 10^{-4} 4.6584\times 10^{-4} 1.4191\times 10^{-4}

     | Show Table
    DownLoad: CSV

    Figures 2 and 3 show the plots of the analytical, (U, V) , and approximate, (u, v) , solutions of case 1 in the clouds of points 1 and 2.

    Figure 2.  Analytical and approximate solutions of case 1 in the cloud of points 1.
    Figure 3.  Analytical and approximate solutions of case 1 in the cloud of points 2.

    Convergence test

    In order to perform a convergence test, we use the clouds of points of Figure 4: 5 (347 nodes), 6 (491 nodes) and 7 (573 nodes). The values of the l_2(U) and l_2(V) errors, and \Delta t are collected in Table 2.

    Figure 4.  Clouds of points used: 5 to 7.
    Table 2.  Error norms l_2 and \Delta t for case 1 in clouds of nodes of Figure 4.
    Cloud of nodes l_2(U) l_2(V) \Delta t
    Cloud 5 2.7430\times 10^{-4} 3.0255\times 10^{-4} 0.001\, s
    Cloud 6 1.6243\times 10^{-4} 1.9572\times 10^{-4} 0.001\, s
    Cloud 7 1.1902\times 10^{-4} 1.5702\times 10^{-4} 0.001\, s

     | Show Table
    DownLoad: CSV

    We plot in Figure 5 the l_2(U) and l_2(V) error norms, respectively, for T = 1\, s , \Delta t = 0.001\, s and three clouds of points in Figure 4.

    Figure 5.  Case 1 on clouds of points 5, 6, and 7 of Figure 5, l_2 error norms U and V versus number of nodes.

    Case 2: Neumann boundary conditions

    We consider for this first case the clouds of points 3 and 4 of Figure 1.

    \begin{equation} \left\lbrace \begin{aligned}{} &\frac{\partial{U}}{\partial{t}} = U^2V-2U+\dfrac{1}{4}\Delta{U}, & (x, y)\in \Omega, \; t > 0, \\ &\frac{\partial V}{\partial t} = U-U^2V+\dfrac{1}{4}\Delta V , &(x, y)\in \Omega, \; t > 0 , \\ &U(x, y, 0) = e^{-(x+y)}, \, V(x, y, 0) = e^{(x+y)}, &(x, y)\in \Omega .\\ \end{aligned} \right. \end{equation} (4.4)

    By a direct check, the exact solution is (the Neumann boundary conditions are chosen so that the equation is fulfilled):

    \begin{equation} U(x, y, t) = e^{-\left(\frac{t}{2}+x+y \right) };\, V(x, y, t) = e^{\left(\frac{t}{2}+x+y \right) }. \end{equation} (4.5)

    Table 3 shows the error norms in the clouds of points.

    Table 3.  Error norms l_2 and l_{\infty} for case 2 in clouds of nodes of the Figure 1.
    Cloud of nodes l_2(U) l_{\infty}(U) l_2(V) l_{\infty}(V)
    Cloud 3 2.8835\times 10^{-5} 1.5345\times 10^{-4} 6.1822\times 10^{-4} 3.4847\times 10^{-3}
    Cloud 4 2.3402\times 10^{-4} 1.4582\times 10^{-3} 5.8804\times 10^{-4} 4.0817\times 10^{-3}

     | Show Table
    DownLoad: CSV

    Figures 4 and 5 show the plots of the analytical, (U, V) , and approximate, (u, v) , solutions of case 2 in the clouds of points 3 and 4.

    Figure 6.  Analytical and approximate solutions of case 2 in the cloud of points 3.
    Figure 7.  Analytical and approximate solutions of case 2 in the cloud of points 5.

    We have studied the convergence and obtained the discretization of system (1.1) using a Generalized Finite Difference Method explicit scheme and the order of convergence of the method is shown. The conditional convergence has been obtained for the numerical scheme. We have presented several numerical cases over different domains. The main advantages of the method are its flexibility in discretizing irregular and complex domains and the simplicity in calculations, saving time and computational resources. As a future line of research, there is the possibility of implementing space-time methods, such as STCM, and comparing them with the current method.

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

    Prof. Dr. Francisco Ureña is the Guest Editor of special issue “Applications of Partial Differential Equations to Science and Engineering Problems: Numerical Resolution” for AIMS Mathematics. Prof. Dr. Francisco Ureña was not involved in the editorial review and the decision to publish this article.

    The authors declare that they have no conflicts of interest.



    [1] A. Whye-Teong, The two-dimensional reaction-diffusion Bruselator system: A dual-reciprocity boundary element solution, Eng. Anal. Bound. Elem., 27 (2003), 897–903. https://doi.org/10.1016/S0955-7997(03)00059-6 doi: 10.1016/S0955-7997(03)00059-6
    [2] M. Mohammadi, R. Mokhtart, R. Schaback, A Meshless Method for Solving the 2D Brusselator Reaction-Diffusion System, CMES, 101 (2014), 113–138.
    [3] Z. Zafar, K. Rehan, M. Mushtaq, M. Rafiq, Numerical treatment for nonlinear Brusselator chemical model, J. Differ. Equ. Appl., 23 (2017), 521–538, http://doi.org/10.1080/10236198.2016.1257005 doi: 10.1080/10236198.2016.1257005
    [4] R. K. Saeed, S. A. Manaa, F. H. Easif, Numerical Solution of Brusselator Model by Expansion Methods, Aust. J. Basic Appl. Sci., 4 (2010), 3389–3403.
    [5] R. C. Mittal, R. Jiwari, Numerical solution of two-dimensional reaction-diffusion Brusselator system, Appl. Math. Comput., 217 (2011), 5404–5415. https://doi.org/10.1016/j.amc.2010.12.010 doi: 10.1016/j.amc.2010.12.010
    [6] W. Sun, W. Qu, Y. Gu, P. W. Li, An arbitrary order numerical framework for transient heat conduction problems, Int. J. Heat Mass Tran., 218 (2024), 124798. https://doi.org/10.1016/j.ijheatmasstransfer.2023.124798 doi: 10.1016/j.ijheatmasstransfer.2023.124798
    [7] W. Sun, H. Ma, W. Qu, A hybrid numerical method for non-linear transient heat conduction problems with temperature-dependent thermal conductivity, Appl. Math. Lett., 146 (2024), 108868. https://doi.org/10.1016/j.aml.2023.108868 doi: 10.1016/j.aml.2023.108868
    [8] J. Benito, F. Ureña, L. Gavete, B. Alonso, Application of the Generalized Finite Difference Method to improve the approximated solution of PDEs, CMES, 38 (2009), 39–58.
    [9] J. Benito, F. Ureña, L. Gavete, Influence of several factors in the generalized finite difference method, Appl. Math. Modell., 25 (2001), 1039–1053. https://doi.org/10.1016/S0307-904X(01)00029-4 doi: 10.1016/S0307-904X(01)00029-4
    [10] L. Gavete, F. Ureña, J. Benito, A. Garcia, M. Ureña, E. Salete, Solving second order non-linear elliptic partial differential equations using generalized finite difference method, J. Comput. Appl. Math., 318 (2017), 378–387. https://doi.org/10.1016/j.cam.2016.07.025 doi: 10.1016/j.cam.2016.07.025
    [11] F. Ureña, L. Gavete, A. Garcia, J. Benito, A. M. Vargas, Solving second order non-linear parabolic PDEs using generalized finite difference method (GFDM), J. Comput. Appl. Math., 354 (2019), 221–241. https://doi.org/10.1016/j.cam.2018.02.016 doi: 10.1016/j.cam.2018.02.016
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(1266) PDF downloads(79) Cited by(0)

Figures and Tables

Figures(7)  /  Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog