Next Article in Journal
A Computational Methodology Based on Maximum Overlap Discrete Wavelet Transform and Autoencoders for Early Prediction of Sudden Cardiac Death
Previous Article in Journal
Dynamics of an Intraguild Predation Food Web Cooperation Model Under the Influence of Fear and Hunting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solution of Coupled Systems of Reaction–Diffusion Equations Using Explicit Numerical Methods with Outstanding Stability Properties

1
Institute of Physics and Electrical Engineering, University of Miskolc, 3515 Miskolc, Hungary
2
Institute of Mathematics, University of Miskolc, 3515 Miskolc, Hungary
*
Author to whom correspondence should be addressed.
Computation 2025, 13(6), 129; https://doi.org/10.3390/computation13060129
Submission received: 15 May 2025 / Revised: 27 May 2025 / Accepted: 29 May 2025 / Published: 1 June 2025

Abstract

Recently, new and nontrivial analytical solutions that contain the Kummer functions have been found for an equation system of two diffusion–reaction equations. The equations are coupled by two different types of linear reaction terms which have explicit time-dependence. We first make some corrections to these solutions in the case of two different reaction terms. Then, we collect eight efficient explicit numerical schemes which are unconditionally stable if the reaction terms are missing, and apply them to the system of equations. We show that they severely outperform the standard explicit methods when low or medium accuracy is required. Using parameter sweeps, we thoroughly investigate how the performance of the methods depends on the coefficients and parameters such as the length of the examined time interval. We obtained that, similarly to the single-equation case, the leapfrog–hopscotch method is usually the most efficient to solve these problems.

1. Introduction

Diffusion and diffusion–reaction equations are among the most widespread partial differential equations (PDEs) in science and technology. Therefore, analytical solutions have been sought and found not only in historical research [1,2], but recently as well. For example, Simpson et al. [3] constructed analytical solutions of coupled linear reaction–diffusion equations where the space domain grows in time. Escorcia and Suazo [4] very recently showed how to construct solutions explicitly via similarity transformations and a Ricatti system for coupled reaction–diffusion equations where the coefficients depend on time. The solutions of a reaction–diffusion equation where the reaction term explicitly depends on time were examined by Covei [5]. Mátyás and Barna found nontrivial solutions [6] a few years ago to regular and irregular diffusion equations containing Kummer or Whittaker functions using similarity transformations. Last year, they extended their investigations to systems of diffusion–reaction equations [7] as well.
These solutions not only provide insight into the diffusion process itself, but make the investigation of the properties and performance of numerical algorithms possible. There are a large number of numerical techniques that can be used to solve these and similar equations. Typical methods can be classified as explicit or implicit algorithms, and both categories have advantages and disadvantages. Explicit algorithms usually have a seriously confined stability domain [8], which means that when the time step size exceeds a certain threshold, sometimes referred to as the CFL (Courant–Friedrichs–Lewy) limit, the solution sooner or later explodes, even in the case of the simple linear diffusion equation. On the other hand, the execution of one time step of the explicit scheme is relatively quick, even for larger systems, and, in principle, the calculations are easily parallelizable. This characterization also refers to the method of lines approach, where the time integration is based on the explicit Runge–Kutta (RK) scheme, or the multistep Adams–Bashforth (AB) scheme. There are explicit methods with better qualitative properties, such as the so-called exact and nonstandard finite difference (NSFD) algorithms [9], which have been proposed for single reaction–diffusion equations [10,11,12] and for cross-diffusion equations [13,14]. However, the favorable properties of these approaches manifest themselves typically only for short time step sizes.
Implicit methods have much better stability properties, and the time step size is limited generally by the accuracy requirements, with some exceptions [15]. The largest drawback of the implicit methods is the need to solve a system of algebraic equations in each time step, which can be time- and memory-consuming if the number of mesh elements is large, which is frequent in the case of two- and three-space-dimensional problems. Still, many scientists consider them to be superior and frequently employ them [16] for these and other related equations. The Crank–Nicolson method is maybe the most widely applied implicit approach [17,18]. For instance, Manaa and Sabawi [19] used the Crank–Nicolson method, as well as the explicit Euler time integration. Their finding was that the former is more accurate and does not have significant stability problems, but the latter has shorter running times, as expected. Pargaei and Kumar [20] applied the Haar wavelet method for systems of degenerate reaction–diffusion PDEs with nonlinear sources. Anjuman et al. proposed a shifted airfoil collocation method [21] for nonlinear drift-diffusion–reaction equations. Rufai et al. [22] developed an implicit hybrid block algorithm with Newton iterations for nonlinear reaction–diffusion equations where the coefficients depend on the time variable. The so-called proper orthogonal decomposition method was combined with the unconditionally positive finite difference scheme by Ndou et al. [23] for advection–diffusion–reaction equations, which can be linear and nonlinear. A Legendre–Gauss–Lobatto spectral collocation approach was used by Heidari et al. [24]. An implicit numerical algorithm preserving the non-negativity of the solutions was presented by Kolev et al. [25].
Certainly, it is not true that either the whole system is treated explicitly or fully implicitly. Another typical approach is that one term, most frequently the diffusion term, is treated implicitly and the reaction term is treated explicitly. These implicit–explicit, or IMEX, methods [26] try to balance stability and computational cost. For example, Calvo et al. [27] combined L-stable linearly implicit RK schemes with explicit ones, all of them with high order, for advection–reaction–diffusion equations. A reaction–diffusion equation system of the Lotka–Volterra type in two space dimensions is solved by Settanni and Sgura [28]. They also treated the diffusion term implicitly, but the reaction term was handled not only by traditional explicit schemes, but the alternating-direction implicit schemes as well. Other similar examples are the IMEX RK-type methods recently developed by Yadav et al. [29] for stiff convection–diffusion–reaction systems, where the diffusion, as well as the reaction terms, are handled implicitly.
Semi-implicit or semi-explicit schemes, developed to solve ODE systems, can also be applied after the spatial semi-discretization of the PDE. Time-efficient variants of these schemes based on the BDF (backward differentiation formula) and AB schemes were developed by Beuken et al. [30]. A variable step size controller was introduced in the paper of Fedoseev et al. [31] to make the performance of semi-implicit composition integration schemes better. Methods utilizing the Gauss–Legendre quadrature, the generalized Padé approximation, and explicit RK methods were developed by Ji and Xing [32].
A further step is treating the diffusion term exactly [33], which can be advised in very stiff cases. For example, semilinear and nonlinear advection–reaction–diffusion equations were examined by Jiang and Zhang [34] using Krylov implicit integration factor methods, where the effect of the linear terms was calculated exactly. On the other hand, it was reported [35] that, although tiny time step sizes below the CFL limit are used, explicit methods can beat implicit schemes in the case of large systems.
Nevertheless, these achievements do not mean that the above-mentioned shortages of the explicit and implicit methods are finally fixed. Due to the reasons explained so far, it seems logical to invest time to construct and test explicit methods with enhanced stability. A first-order member of this family is the so-called unconditionally positive finite difference (UPFD) method, proposed by Chen-Charpentier and Kojouharov [36], and further examined by other scholars [37,38]. The UPFD scheme has very good qualitative properties, but, on the other hand, the error decreases much slower with the time step size than in the case of other first-order methods. Other, more accurate methods also exist with unconditional stability for the diffusion equation. Al-Bayati et al. [39] compared the performance of the Saulyev-type alternating-direction explicit (ADE) method, the alternating-direction implicit (ADI) method, and the odd–even hopscotch method to solve a system of diffusion–reaction equations. They found that the implicit method is more accurate but slower than the explicit methods. Nonlinear PDEs were also solved by Pourghanbar et al. [40] using the ADE scheme, which was faster than using major implicit solvers.
Our research group proposed novel explicit methods in the last few years to solve diffusion and diffusion–reaction equations. We proved analytically and presented in our previous work that these algorithms are unconditionally stable for the single linear diffusion equation. The performance of these schemes was extensively examined for the diffusion equation in which the diffusion coefficient is a constant [41] and depends on time [42] or space [43], or both at the same time [44]. Furthermore, they were tested for different kinds of reaction–diffusion equations, for example, with a space- and time-dependent coefficient of the reaction term [45]. It turned out that the new methods can frequently outperform the conventional methods, either explicit RK types or implicit solvers. We are now continuing these investigations with the inclusion of equation systems. This is the first time that these methods are applied to mutually coupled reaction–diffusion equations.
This paper has the following structure. After the presentation of the studied problem in Section 2, we display the analytical solution constructed in paper [7], analyze the mistake in it, and propose a way to correct it. The examined numerical methods and their properties are described briefly in Section 3. Two case studies using ten numerical methods are displayed in Section 4. Section 5 presents parameter sweeps to discover how the numerical errors of the methods depend on some parameters of the problems to be solved, such as the coefficients of the reaction term and the duration of the simulated time interval. Finally, the conclusions are presented at the end of this paper.

2. The Equations and the Analytical Solutions

It is well known that the simplest form of the diffusion equation (also called the heat equation) is the following:
u ( x , t ) t = D 2 u ( x , t ) x 2 ,
where x , t are the space and time variables, the u ( x , t ) concentration function meets the required smoothness conditions, and the non-negative number D is called the diffusivity.

2.1. Coupled Diffusion Equations

The authors of paper [7] examine the following coupled system of PDEs, numbered as Equation (19) in [7]:
u ( x , t ) t = D 2 u ( x , t ) x 2 + F ( u ( x , t ) , w ( x , t ) ) , w ( x , t ) t = D 2 w ( x , t ) x 2 + G ( u ( x , t ) , w ( x , t ) ) ,
where F and G are real functions. The authors of [7] introduced the so-called reduced variable η = x t , and applied the self-similar Ansatz
u ( x , t ) = t α f η   and   w x , t = t γ g η ,
where f η and g η are the shape functions, while α and γ are independent real parameters. They constructed analytical solutions for two different versions of the reaction term. The first one, which will be called simple coupling, has the following form (Equation (20) in [7]):
F ( u , w ) = a w t ,   G ( u , w ) = b u t ,
where a , b . The transformation (3) is supposed to yield the ODEs
α f ( η ) 1 2 η d f d η = D d 2 f d η 2 + a g ( η ) , γ g ( η ) 1 2 η d g d η = D d 2 g d η 2 + b f ( η ) ,
which, according to Maple 12 and Equations (21) and (22) in [7], have the solution
f η = η e η 2 4 D   c 1 M p q , 3 2 , η 2 4 D + c 2 M p + q , 3 2 , η 2 4 D + c 3 U p q , 3 2 , η 2 4 D + c 4 U p + q , 3 2 , η 2 4 D ,
and
g η = η 2 a e η 2 4 D   c 1 M p q , 3 2 , η 2 4 D + c 3 U p q , 3 2 , η 2 4 D γ α + 2 q + c 2 M p + q , 3 2 , η 2 4 D + c 4 U p + q , 3 2 , η 2 4 D γ α 2 q .
The abbreviations p = 1 γ 2 α 2 and q = 4 a b 2 α γ + γ 2 + α 2 / 2 were introduced. It is clear that if a b 0 , then q is real regardless of the values of α and γ.
The other type of reaction term (Equation (26) in [7]) is the following:
F ( u , w ) = k t u ( x , t ) w ( x , t ) ,   G ( u , w ) = k t u ( x , t ) w ( x , t ) ,
where k and k is usually positive. This case will be called time-decaying isomerization based on [7], where Equation (27) claims that the transformation yields the following ODE system:
α f ( η ) 1 2 η d f d η = D d 2 f d η 2 k f ( η ) g ( η ) , γ g ( η ) 1 2 η d g d η = D d 2 g d η 2 + k f ( η ) g ( η ) ,
which has the following solution (Equations (28) and (29) in [7]):
f η = η e η 2 4 D   c 1 M p q + k , 3 2 , η 2 4 D + c 2 M p + q + k , 3 2 , η 2 4 D + c 3 U p q + k , 3 2 , η 2 4 D + c 4 U p + q + k , 3 2 , η 2 4 D
g η = η 2 k e η 2 4 D   c 1 M p q + k , 3 2 , η 2 4 D + c 3 U p q + k , 3 2 , η 2 4 D γ α + 2 q + c 2 M p + q + k , 3 2 , η 2 4 D + c 4 U p + q + k , 3 2 , η 2 4 D γ α 2 q .
where q = 4 k 2 2 α γ + γ 2 + α 2 / 2 .
However, after performing the transformation, we discovered that the ODE systems (5) and (9) above cannot be the transformed versions of the original equations with the reaction terms (4) and (8) after the general Ansatz (3) is used. The reason for this is that α and γ are independent parameters and γ can be eliminated from the first equation of (5) and (9), and α can be eliminated from the second equation of (5) and (9) only if they are equal. The authors probably accidentally substituted γ with α and α with γ in the first and second equations, respectively; therefore, their solutions are valid only in the α = γ case. This does not mean that the solutions in [7] are not valid for the examined equations, only that they contain only one free parameter and their form is simpler than given in [7]. We note that the paper [7] presented analytical results for some other equations which we did not check.
Our first goal in this paper is to save these valuable solutions in the general case. For this, we slightly change the PDE systems which are to be solved.

2.2. The Modified Equations and Their Solution for Simple Coupling

The first system of equations will be
u t = D 2 u x 2 + a w t A , w t = D 2 w x 2 + b u t A 2 .
where A is an arbitrary real parameter. If now we perform the transformation (3), we obtain
α t α 1 f ( η ) 1 2 t α x t 3 / 2 d f d η = D t α 1 t d 2 f d η 2 + a t γ g ( η ) t A , γ t γ 1 g ( η ) 1 2 t γ x t 3 / 2 d g d η = D t γ 1 t d 2 g d η 2 + b t α f ( η ) t A 2 .
Multiplying the first equation by t α + 1 and the second one by t γ + 1 , we obtain
α f ( η ) 1 2 η d f d η = D d 2 f d η 2 + a t A + α + 1 γ g ( η ) , γ g ( η ) 1 2 η d g d η = D d 2 g d η 2 + b t γ 1 A α f ( η ) .
This gives back ODE system (5) if the exponents of t in the last terms are zero, i.e., γ = A + α + 1 , where α is still arbitrary. One can easily see that we obtain the symmetric case for A = 1 . However, now (6) and (7) are valid solutions for arbitrary values of A. This means that the solution of equation system (12) is the following:
u x , t = x t α + 1 / 2 e x 2 4 D t   c 1 M p q , 3 2 , x 2 4 D t + c 2 M p + q , 3 2 , x 2 4 D t + c 3 U p q , 3 2 , x 2 4 D t + c 4 U p + q , 3 2 , x 2 4 D t
and
w x , t = x 2 a t A + α + 3 / 2 e x 2 4 D t   c 1 M p q , 3 2 , x 2 4 D t + c 3 U p q , 3 2 , x 2 4 D t γ α + 2 q + c 2 M p + q , 3 2 , x 2 4 D t + c 4 U p + q , 3 2 , x 2 4 D t γ α 2 q .

2.3. The Time-Decaying Isomerization

On the other hand, let us start with the modified PDE system:
u t = D 2 u x 2 k t u w t B , w t = D 2 w x 2 k t w u t B .
Now, the transformation (3) yields
α t α 1 f ( η ) 1 2 t α x t 3 / 2 d f d η = D t α 1 t d 2 f d η 2 k t t α f ( η ) t γ g ( η ) t B , γ t γ 1 g ( η ) 1 2 t γ x t 3 / 2 d g d η = D t γ 1 t d 2 g d η 2 k t t γ g ( η ) t α f ( η ) t B .
In the same way as above, we obtain
α f ( η ) 1 2 η d f d η = D d 2 f d η 2 k f ( η ) t γ + α + B g ( η ) γ g ( η ) 1 2 η d g d η = D d 2 g d η 2 k g ( η ) t α B + γ f ( η ) .
This is equivalent to the ODE system (9) for an arbitrary α only if γ = B + α , thus the symmetric case is obtained for B = 0 . Now, (10) and (11) are valid solutions for arbitrary values of B, hence the solution of equation system (15) is the following:
u x , t = x t α + 1 / 2 e x 2 4 D t   c 1 M p q + k , 3 2 , x 2 4 D t + c 2 M p + q + k , 3 2 , x 2 4 D t + c 3 U p q + k , 3 2 , x 2 4 D t + c 4 U p + q + k , 3 2 , x 2 4 D t
and
w x , t = x 2 k t B + α + 1 / 2 e x 2 4 D t   c 1 M p q + k , 3 2 , x 2 4 D t + c 3 U p q + k , 3 2 , x 2 4 D t γ α + 2 q + c 2 M p + q + k , 3 2 , x 2 4 D t + c 4 U p + q + k , 3 2 , x 2 4 D t γ α 2 q .
In the rest of this paper, we will numerically reproduce the solutions (13) and (14) and (16) and (17) for several parameter sets.

3. The Numerical Methods and Their Properties

3.1. The Spatial and Temporal Discretization

We solve the equations in the one-dimensional space interval x x 0 , x 0 + L x . This is discretized by creating N x \ 0 nodes equidistantly as usual: x i = i Δ x ,   i = 1 , , N x . The time interval is t t in , t fin , where t fin = t in + TIME , and it is also discretized uniformly, t n = t in + n Δ t ,   n 0 , 1 , 2 , , T , and the numerical solution at x i and t n is denoted by u i n . We define the mesh ratio by the widely applied formula r = α Δ t Δ x 2 .

3.2. The Algorithms

We now provide essential details regarding the algorithms which solve the system of equations. The formulas of the algorithms are given for a general reaction term, denoted by Ru and Rw for the first, and the last member of the system of equations, respectively. Then, the specific forms of the reaction terms are given.
1. First, the FTCS formula is applied to Equation (18), which is based on the explicit Euler time discretization
u i n + 1 = u i n + r u i + 1 n 2 u i n + u i - 1 n + Δ t R u   and     w i n + 1 = w i n + r w i + 1 n 2 w i n + w i - 1 n + Δ t R w
where
R u = a w i n t A   and   R w = b u i n t ( A 2 ) for   simple   coupling , R u = k ( u i n w i n t B ) t   and   R w = k ( u i n t B w i n ) t for   isomerization .
2. The Adams–Bashforth method is an explicit multistep numerical technique. We use its two-step second order variant, which predicts u i n + 1 using information from the previous two points, making it more accurate than single-step methods like Euler’s method. Hence, this method is not a self-starter, and usually an RK method is utilized to start the calculation. In our case, the first step is performed using Euler’s method in the same way as above. Now, we have u and w values for two time levels, and the values in the old time levels will be denoted by u 2 i n and w 2 i n . Now, the formulas of the Adams–Bashforth method are the following:
f u i n + 1 = 2 r u i n + r u i + 1 n + u i 1 n + Δ t R u   and   f w i n + 1 = 2 r w i n + r w i + 1 n + w i 1 n + Δ t R w   ,   where R u = a w i n t A   and R w = b u i n t ( A 2 ) ,   R u = k ( u i n w i n t B ) t   and   R w = k ( u i n t B w i n ) t f u o l d i n + 1 = 2 r u 2 i n + r u 2 i + 1 n + u 2 i 1 n + Δ t R u   and   f w o l d i n + 1 = 2 r w 2 i n + r w 2 i + 1 n + w 2 i 1 n + Δ t R w   ,   where R u = a w 2 i n t A   and   R w = b u 2 i n t ( A 2 ) ,   R u = k u 2 i n w 2 i n t B t   and   R w = k u 2 i n t B w 2 i n t .
Finally, the updated values are calculated as follows:
u i n + 1 = u i n + ( 3 f u i n + 1 f u o l d i n + 1 ) 2   and   w i n + 1 = w i n + ( 3 f w i n + 1 f w o l d i n + 1 ) 2 .
3. The alternating-direction explicit (ADE) scheme is a known [46,47] but non-conventional method for which the condition of consistency is also known. We include it here for comparison purposes. In a one-dimensional equidistant mesh, one splits the calculation, i.e., one first sweeps the mesh from left to right, and then vice versa, independently. This means that the following equations are solved from left to right and from right to left, respectively:
u r i g h t i n + 1 = 1 r u i n + r u i + 1 n + u r i g h t i 1 n + 1 + Δ t R u 1 + r   and   u l e f t i n + 1 = 1 r u i n + r u l e f t i + 1 n + 1 + u i 1 n + Δ t R u 1 + r
w r i g h t i n + 1 = 1 r w i n + r w i + 1 n + w r i g h t i 1 n + 1 + Δ t R w 1 + r   and   w l e f t i n + 1 = 1 r w i n + r w l e f t i + 1 n + 1 + w i 1 n + Δ t R w 1 + r ,
and (19) is used to calculate the reaction terms. The final values are the simple averages of the two half-sided terms:
u i n + 1 = ( u r i g h t i n + 1 + u l e f t i n + 1 ) / 2 ,   and   w i n + 1 = ( w r i g h t i n + 1 + w l e f t i n + 1 ) / 2 .
We note that for non-uniform meshes, the ADE method loses its fully explicit character, and matrix calculations would be necessary.
4. The Dufort–Frankel (DF) method is a classic example of explicit and stable methods [48] (p. 313) and is second-order in time. We adapted it to the case of Equation (2), where the following formula must be used
u i n + 1 = 1 2 r u i n 1 + 2 r u i + 1 n + u i - 1 n + 2 Δ t R u 1 + 2 r ,   w i n + 1 = 1 2 r w i n 1 + 2 r w i + 1 n + w i - 1 n + 2 Δ t R w 1 + 2 r ,
and in (19), the latest u i n and w i n values are used. As can be seen, it is a two-step, one-stage method (the formula includes u i n 1 ). Since it is not self-starting, the calculation u i 1 must be performed by another method. We use the UPFD formula for this purpose
u i 1 = u i 0 + r u i + 1 0 + u i 1 0 + Δ t R u 1 + 2 r ,   w i 1 = w i 0 + r w i + 1 0 + w i 1 0 + Δ t R w 1 + 2 r
where, of course, the initial values are used in Equation (19) for the reaction terms.
5. The original odd–even hopscotch (OOEH) method was discovered more than 50 years ago [49]. It requires a special spatial and temporal structure as shown in Figure 1. In essence, the mesh must be divided into two parts, the so-called odd and even nodes (or cells), where the closest neighbors of the even cells are odd, and vice versa. First, the FTCS formula given in Equation (18) is applied, but only to the odd cells. Then, it is followed by the BTCS (backward-time central-space) formula, which is based on implicit Euler time discretization with the equations
u i n + 1 = u i n + r u i + 1 n + 1 + u i 1 n + 1 + Δ t R u 1 + 2 r         w i n + 1 = w i n + r w i + 1 n + 1 + w i 1 n + 1 + Δ t R w 1 + 2 r
and with the application of (19). Since the values of the neighbors u i + 1 n + 1 , etc., are already calculated, this is an explicit method. After every time step, the odd and even labels are switched, as is illustrated in Figure 1.
6. The CpC method [50] is the organization of the so-called CNe scheme into a two-stage algorithm. The first stage is a fractional time step with half time step size employing the CNe formula:
u i pred = u i n e r + u i + 1 n + u i 1 n + Δ t R u / r 2 1 e r ,   w i pred = w i n e r + w i + 1 n + w i 1 n + Δ t R w / r 2 1 e r
where Equation (19) is used for the reaction terms. These predictor values are used at the second stage, which is a full time step size corrector step; thus, the final values are as follows:
u i n + 1 = u i n e 2 r + u i + 1 pred + u i 1 pred + Δ t R u / r 2 1 e 2 r ,   w i n + 1 = u i n e 2 r + w i + 1 pred + w i 1 pred + Δ t R w / r 2 1 e 2 r R u = a w i pred t A   and   R w = b u i pred t ( A 2 ) ,   R u = k t u i pred w i pred t B   and   R w = k t u i pred t B w i pred
7. The next method is called the CCL algorithm, which is an abbreviation of Constant–Constant–Linear neighbor. It is a one-step but three-stage method, recently published by our group [51], with third-order accuracy. The CNe formula is applied in the first and second stages and the so-called linear-neighbor (LNe) formula in the last stage. The first stage is a predictor with a Δ t / 3 -sized time step
a u 0 i = u i + 1 n + u i 1 n + Δ t R u / r 2 ,   a w 0 i = w i + 1 n + w i 1 n + Δ t R w / r 2 u i C = u i n e 2 r / 3 + 1 e 2 r / 3 a u 0 i ,   w i C = w i n e 2 r / 3 + 1 e 2 r / 3 a w 0 i   ,
where Equation (19) is used again for the R terms. Then, the first corrector stage is performed:
a u 1 i = u i + 1 C + u i 1 C + Δ t R u / r 2 ,   a w 1 i = w i + 1 C + w i 1 C + Δ t R w / r 2 R u = a w i C t A   and   R w = b u i C t A 2 R u = k u i C w i C t B t   and   R w = k u i C t B w i C t u i C C = u i n e 4 r / 3 + 1 e 4 r / 3 a u 1 i ,   w i C C = w i n e 4 r / 3 + 1 e 4 r / 3 a w 1 i
The last stage is as follows:
a u 2 i = u i + 1 C C + u i 1 C C + Δ t R u / r 2 ,   a w 2 i = w i + 1 C C + w i 1 C C + Δ t R w / r 2 R u = a w i C C t A   and   R w = b u i C C t ( A 2 ) , R u = k ( u i C C w i C C t B ) t   and   R w = k ( u i C C t B w i C C ) t u i n + 1 = u i n e 2 r + 1 e 2 r a u 0 i + 3 ( a u 2 i a u 0 i ) 2 1 e 2 r 2 r w i n + 1 = w i n e 2 r + 1 e 2 r a w 0 i + 3 ( a w 2 i a w 0 i ) 2 1 e 2 r 2 r
8. The four-stage CLQ2 (Constant–Linear–Quadratic–Quadratic) neighbor method [52] is a bit more complicated than the previous ones, since it has four stages and it employs the quadratic-neighbor approximation as well. Its first stage utilizes the CNe formula
a u i C = u i + 1 n + u i 1 n + Δ t R u / r 2                             a w i C = w i + 1 n + w i 1 n + Δ t R w / r 2 u i C = u i n e 2 r + a u i C ( 1 e 2 r ) ,                 w i C = w i n e 2 r + a w i C ( 1 e 2 r )
where Equation (19) holds again. After this, the LNe formula is applied in the second stage with not only a full time step, but simultaneously a half-length time step as well, to obtain the sets u i CL ,   w i CL and u i CL 1 2 ,   w i CL 1 2 , respectively:
a u i C L = u i + 1 C + u 1 i 1 C + Δ t R u / r 2               and       a w i C L = w i + 1 C + w i 1 C + Δ t R w / r 2 R u = a w i C t A   and   R w = b u i C t ( A 2 ) ,   or   R u = k ( u i C w i C t B ) t   and   R w = k ( u i C t B w i C ) t u i C L = u i n e 2 r + a u i C L ( 1 e 2 r ) + ( a u i C L a u i C ) 1 e 2 r 2 r u i C L m i d = u i n e r + a u i C L ( 1 e r ) + a u i C L a u i C 2 1 e r r w i C L = w i n e 2 r + a w i C L ( 1 e 2 r ) + ( a w i C L a w i C ) 1 e 2 r 2 r w i C L m i d = w i n e r + a w i C L ( 1 e r ) + a w i C L a w i C 2 1 e r r
The third stage contains the calculations below:
a u i C L Q = u i + 1 C L + u i + 1 C L + Δ t R u / r 2   ,     a w i C L Q = w i + 1 C L + w i 1 C L + Δ t R w / r 2 a u i , m i d C L Q = u i + 1 C L m i d + u i + 1 C L m i d + Δ t R a u 1 / r 2   ,     a w i , m i d C L Q = w i + 1 C L m i d + w i 1 C L m i d ) + Δ t R a w 1 / r 2 R u = a w i C L t A   and   R w = b u i C L t ( A 2 )             or           R u = a w i C L t A   and   R w = b u i C L t ( A 2 ) R a u 1 = k ( u i C L w i C L t A ) t   and   R a w 1 = k ( u i C L t A w i C L ) t s u = 4 a u i , m i d C L Q a u i C L Q 3 a u i C   ,   g u = 2 ( a u i C L Q 2 a u i , m i d C L Q + a u i C ) s w = 4 a w i , m i d C L Q a w i C L Q 3 a w i C   ,   g w = 2 ( a w i C L Q 2 a w i , m i d C L Q + a w i C ) u i C L Q = u i n e 2 r + ( 1 e 2 r ) ( g u / 2 / r 2 s u / 2 / r + a u i C ) + g u g u / r + s u w i C L Q = w i n e 2 r + ( 1 e 2 r ) ( g w / 2 / r 2 s w / 2 / r + a w i C ) + g w g w / r + s w u i ,   m i d C L Q = u i n e r + ( 1 e r ) ( g u / 2 / r 2 s u / 2 / r + a u i C ) + g u / 4 g u / 2 / r + s u / 2 w i ,   m i d C L Q = w i n e r + ( 1 e r ) ( g w / 2 / r 2 s w / 2 / r + a w i C ) + g w / 4 g w / 2 / r + s w / 2
There is no need to mention that the su, sw, gu, and gw auxiliary quantities also depend on the space and time level, but since they are used only immediately after their evaluation, they can be defined as scalar variables instead of arrays. We add the fourth stage by using the last four values u i C L Q , …, w i ,   m i d C L Q . Using these, one can calculate a u i C L Q Q ,…, a w i , m i d C L Q Q values in the same manner as the previous formulas. Then, one repeats the calculations of the third stage, including those of su, , gw, and finally obtains the u i C L Q Q and w i C L Q Q sets as final results at the end of the actual time step. With these, one has four stages altogether, which constitute the CLQ2 algorithm [52].
9. The pseudo-implicit (PI) two-stage method was developed in [53] (Algorithm 5 there, applied to the pure diffusion Equation (1)). A half time step is taken to obtain the predictor values and then a full time step for the corrector values. The first-stage formulas are the following:
u i pred = u i n + r / 2 u i + 1 n + u i 1 n + Δ t R u / 2 1 + r                         w i pred = w i n + r / 2 w i + 1 n + w i 1 n + Δ t R w / 2 1 + r
The second stage is as follows
u i n + 1 = 1 r u i n + r u i + 1 pred + u i 1 pred + Δ t R u 1 + r                   w i n + 1 = 1 r w i n + r w i + 1 pred + w i 1 pred + Δ t R w 1 + r
and (19) is used in both stages, but in the second stage, the predictor u and w values are substituted.
10. The recently invented leapfrog–hopscotch (LH) approach [42] also requires the odd–even space structure. Moreover, it has a structure made up of many full time steps and two half time steps, as can be seen in Figure 1. Using the initial values, the calculation begins by taking a half-sized time step for the odd nodes. Full time steps are then taken strictly alternately for the even and odd nodes (marked by light purple diamonds in Figure 1) until the last time step is reached, which should be halved for odd nodes to reach the same final time point as the even nodes. Since the first stage’s time step is halved, the following general formula is applied:
u i o d d = u i n + r / 2 u i + 1 n + u i 1 n + Δ t R u / 2 1 + r   and   w i o d d = w i n + r / 2 w i + 1 n + w i 1 n + Δ t R w / 2 1 + r
Next, for the even nodes, a full time step is made using
u i e v e n = 1 r u i n + r u i + 1 n + u i 1 n + Δ t R u 1 + r                 w i e v e n = 1 r w i n + r w i + 1 n + w i 1 n + Δ t R w 1 + r
When the calculations are performed, the latest available values of the left and right neighbors are always used. This is not true for the reaction terms, where the values at the beginning of the actual time step are substituted to Equation (19) even if newer values are available for ui.

3.3. Properties of the Algorithms

The methods have well-established properties when they are applied to the simple Equation (1). First of all, they have the following theoretical order of temporal accuracy. The FTCS is first, the AB2, ADE, DF, OOEH, CpC, CCL, PI and LH methods are second, the CCL method is third, and finally the CLQ2 scheme is fourth in order of time step size. With the exception of the FTCS and AB2, the methods are unconditionally stable for the simple diffusion Equation (1), i.e., the previously mentioned CFL limit does not hold for them. We need to stress again that among explicit methods, unconditional stability is not the rule but the exception.
We note that the CpC, CCL, CLQ2, PI, and LH methods are recently invented by our research group, and the verifications, analytical proofs, etc., are typically presented in those original papers.
Furthermore, the CpC and the CLQ2 methods have an extra property that the calculated u i n + 1 values are the convex combination of the old u i n ,   u i 1 n ,   u i + 1 n , etc., values in the case of Equation (1), thus the maximum and minimum principles are trivially fulfilled and unphysical oscillations cannot appear, even for extremely large time step sizes. Based on this, the positivity-preserving property can be easily proved for the CpC scheme when applied to the system of Equation (12) with simple coupling for non-negative coupling coefficients.
Proposition 1.
Assume that the CpC method is applied to Equation (12). If the initial u and w values are non-negative, than the new values are also non-negative for arbitrary  A  and   D , a , b , Δ t 0 .
Proof. 
In the Formulas (24) and (25), the coefficient of each term u i n ,   u i 1 n ,   u i + 1 n and R u , R w lies in the unit interval. Since the u j n values are non-negative, j = 1 , , N x , the reaction terms R u , R w are also non-negative, which implies the statement. □
Remark 1.
We note that the goal of this work is the numerical investigation of the ten listed methods. A detailed analytical examination of their properties when applied to the studied system of equations is out of scope of this work.

4. Verification in 1D

The L2 error norm is determined by taking the square root of the sum of the squared differences between the numerical solution and the analytical (reference) solution at each point:
E r r o r L 2 u = i = 1 N x ( u i n u r e f ( x 1 , t n ) ) 2 1 / 2 , E r r o r L 2 w = i = 1 N x ( w i n w r e f ( x 1 , t n ) ) 2 1 / 2 .
We perform two case studies, one for the simple coupling and one for isomerization using the analytical solutions (13), (14), (16), and (17), with different parameter values. We always use the same coefficients c 1 = 1 ,   c 2 = 0 ,   c 3 = 0 ,   c 4 = 0 in each case study, including the rest of this paper.
Case Study 1: Diffusion with simple coupling
In this experiment, we used the analytical solutions (13) and (14) with the following parameters: x 0 = 2 ,   L x = 4 , N x = 1000 ,   t in = 0.04 ,   TIME = 0.1   ,   a = 4 ,   b = 9 ,   A = 0.6 ,   α = 0.87 . First, we present the graphs of the initial function u ( x , t = 0.04 ) , the reference analytical solution u e x a c t ( x , t f i n = 0.14 ) , and two corresponding numerical solutions for Δ t = 10 4 in Figure 2. Then, in Figure 3, we display the errors calculated by Formula (36) as a function of the time step size for both u and w for each method. We consider the verification successful, since for decreasing Δ t , the errors are decreasing as expected based on our similar experiments conducted for single diffusion-type equations.
Case study 2: Isomerization for large coupling coefficient
In this experiment, we used the analytical solutions (16) and (17) with the following parameters:
x 0 = 2 ,   L x = 4 , N x = 800 ,   t in = 0.04 , TIME = 0.24 ,   k = 3 ,   B = 0.26 ,   α = 0.5 . We present the graphs of the initial function u ( x , t = 0.04 ) . The reference analytical solution u e x a c t ( x , t f i n ) , and two corresponding numerical solutions with Δ t = 1.25 10 4 in Figure 4. Then, in Figure 5, we display the errors as a function of the time step size for both u and w for each method.

5. Parameter Sweeps for Further Testing of the Methods

In this section, we conduct parameter sweep experiments to explore how the performance of the methods depends on the parameters. In the case of simple coupling and isomerization, the analytical solutions (13), (14), (16), and (17) will be used for reference purposes, and we fix all the parameters except those we sweep. So unless it is stated explicitly otherwise, the PDEs are discretized using x 0 = 1 ,   L x = 0.8 , N x = 100 ,   t in = 0.1 , TIME = 0.1 , while the coefficients are a = 3 ,   b = 2 a ,   k = 0.3 ,   A = 0.9 ,   α = 0.6 in each simulation.
We calculate the average of the logarithm of the maximum or L errors, i.e., the following formula is used:
AgErr ( L u ) = 1 S s = 1 S log max i u i n u r e f ( x 1 , t n ) ,
which will be called the aggregated error. Our goal with this aggregated error AgErr is to assess the overall accuracy of the methods for small, medium, and large time step sizes. Note that when the AgErr is negative with a large absolute value, the method is very accurate. In this section, we calculated the error for S = 9 different time step sizes for all the examined methods. Since the FTCS and AB2 methods produce “NaN” (not-a-number) values, the aggregated errors cannot be calculated for them; therefore, we proceed only with the remaining eight methods with good stability properties.
Case Study 3: Errors as a function of parameter a for the simple coupling case
The following series is used for the values of the coefficient: a 0.1 ,   0.3 ,   1 ,   3 ,   6 ,   10 ,   20 ,   30 ,   100 ,   300 ; b = 2 a holds, which means that the two coupling coefficients a and b grow together. The aggregated errors are displayed in Table 1 and Figure 6.
Case Study 4: Errors as a function of the b parameter, simple coupling case
The series used for the values of the coefficient is b 0.1 ,   0.3 ,   1 ,   3 ,   10 ,   30 ,   100 ,   300 , while a = 3 is kept fixed, thus only the second is increased gradually. The aggregated errors are displayed in Table 2 and Figure 7.
Case Study 5: Errors as a function of the A parameter, simple coupling case
The following series is used for the values of the coefficient: A 10 ,   3 , 1 ,   0 ,   1 ,   3 ,   10 . The aggregated errors are displayed in Table 3 and Figure 8.
Case Study 6: Errors as a function of the TIME parameter, simple coupling case
The following series is used for the values of the coefficient: T I M E 0.0001 ,   0.0003 ,   0.001 ,   0.003 ,   0.01 ,   0.03 ,   0.1 ,   0.3 ,   1 ,   3 ,   10 ,     30 ,   100 ,   300 . The aggregated errors are displayed in Table 4 and Figure 9.
Case Study 7: Errors as a function of the TIME parameter, isomerization case
The following series is used for the values of the coefficient: T I M E 0.0001 ,   0.0003 ,   0.001 ,   0.003 ,   0.01 ,   0.03 ,   0.1 ,   0.3 ,   1 ,   3 ,   10 ,     30 ,   100 ,   300 . The aggregated errors are displayed in Table 5 and Figure 10.
Case Study 8: Errors as a function of the k parameter, isomerization case
The following series is used for the values of the coefficient: k 0.01 ,   0.03 ,   0.1 ,   0.3 ,   1 ,   3 ,     10 ,     30 ,   60 ,   100 ,   170 ,   300 . The used coefficients are a = 3 ,   b = 2 a ,   A = 0.9 and α = 0.6 . The aggregated errors are displayed in Table 6 and Figure 11.

6. Conclusions and Summary

We studied coupled sets of reaction–diffusion equations in one space dimension with two different, time-dependent reaction terms. We recited a set of recently published analytical solutions, which are highly nontrivial since they contain the Kummer functions. We showed that there is a mistake in these solutions, which means that they are valid only in the simple case, when α = γ , as they contain one less free parameter. We modified the time-dependent terms in the original PDEs, and now they contain a free parameter in the exponent apart from the coupling coefficient. Therefore, the solutions are now valid for a more general set of reaction–diffusion equations and in the general case when α γ as well.
We then reproduced the analytical solutions by 10 explicit numerical methods including the most standard FTCS and Adams–Bashforth methods. The rest of the methods are unconditionally stable for the diffusion equation without the reaction term, and some of them have been constructed by our research group in the past few years. First, we verified the algorithms by showing that the errors properly decrease with the time step size for both types of reaction term. Then, we performed sweeps for some key parameters to study how the methods behave as a function of these parameters.
The first conclusion is that the methods are not unconditionally stable for the studied diffusion–reaction equation system. If the coefficients of the reaction term are large, the errors start to grow as time elapses. However, the methods that are stable for the diffusion equation are still more stable for the case when a reaction term is present than the traditional FTCS and AB2 schemes. The least stable among these methods is the OOEH method and the most stable is the CpC method, which has the best qualitative properties. Nevertheless, the error is usually the smallest for the LH method, but the DF, OOEH, and ADE schemes also perform well if they are not close to their stability limits. The OOEH scheme is the best if the simulated time interval is very short, but as time elapses, its error drastically increases.
After this study’s publication, we plan to analytically investigate the properties of the methods when applied to this system of equations. We also wish to repeat some numerical experiments to the other terms in the analytical solutions, i.e., when the coefficients c 2 ,   c 3 ,     and   c 4 are nonzero.

Author Contributions

Conceptualization, resources, methodology, and supervision: E.K. investigation and visualization: H.K. and A.Z.; software, writing—original draft preparation, review and editing: H.K., A.Z. and E.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Two codes are available at the following link https://github.com/Drendre/Coupled-diffusion-reaction-PDEs-codes, accessed on 30 May 2025.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jacobs, M.H. Diffusion Processes; Springer: Berlin/Heidelberg, Germany, 1935; ISBN 978-3-642-86414-8. [Google Scholar]
  2. Lienhard, J.H., IV; Lienhard, J.H., V. A Heat Transfer Textbook, 4th ed.; Phlogiston Press: Cambridge, MA, USA, 2017; ISBN 978-0-9713835-2-4. [Google Scholar]
  3. Simpson, M.J.; Sharp, J.A.; Morrow, L.C.; Baker, R.E. Exact Solutions of Coupled Multispecies Linear Reaction–Diffusion Equations on a Uniformly Growing Domain. PLoS ONE 2015, 10, e0138894. [Google Scholar] [CrossRef]
  4. Escorcia, J.M.; Suazo, E. On explicit solutions for coupled reaction-diffusion and Burgers-type equations with variable coefficients through a Riccati system. Discret. Contin. Dyn. Syst. S 2025. [Google Scholar] [CrossRef]
  5. Covei, D.-P. A nonlinear diffusion equation of the Gurtin-MacCamy type: Existence, uniqueness, and numerical simulations. arXiv 2025. [Google Scholar] [CrossRef]
  6. Mátyás, L.; Barna, I.F. General Self-Similar Solutions of Diffusion Equation and Related Constructions. Rom. J. Phys. 2022, 67, 101. [Google Scholar]
  7. Barna, I.F.; Mátyás, L. Diffusion Cascades and Mutually Coupled Diffusion Processes. Mathematics 2024, 12, 3298. [Google Scholar] [CrossRef]
  8. Jejeniwa, O.A.; Gidey, H.H.; Appadu, A.R. Numerical Modeling of Pollutant Transport: Results and Optimal Parameters. Symmetry 2022, 14, 2616. [Google Scholar] [CrossRef]
  9. Köroğlu, C.; Aydin, A. Exact and nonstandard finite difference schemes for the Burgers equation B(2,2). Turk. J. Math. 2021, 45, 647–660. [Google Scholar] [CrossRef]
  10. Agbavon, K.M.; Appadu, A.R. Construction and analysis of some nonstandard finite difference methods for the FitzHugh–Nagumo equation. Numer. Methods Partial. Differ. Equ. 2020, 36, 1145–1169. [Google Scholar] [CrossRef]
  11. Agbavon, K.M.; Appadu, A.R.; Khumalo, M. On the numerical solution of Fisher’s equation with coefficient of diffusion term much smaller than coefficient of reaction term. Adv. Differ. Equ. 2019, 2019, 146. [Google Scholar] [CrossRef]
  12. Appadu, A.R.; Inan, B.; Tijani, Y.O. Comparative study of some numerical methods for the Burgers-Huxley equation. Symmetry 2019, 11, 1333. [Google Scholar] [CrossRef]
  13. Marc, E. Songolo a Positivity-Preserving Nonstandard Finite Difference Scheme for Parabolic System with Cross-Diffusion Equations and Nonlocal Initial Conditions. Am. Sci. Res. J. Eng. Technol. Sci. 2016, 18, 252–258. [Google Scholar]
  14. Chapwanya, M.; Lubuma, J.M.S.; Mickens, R.E. Positivity-Preserving nonstandard finite difference schemes for Cross-Diffusion equations in biosciences. Comput. Math. Appl. 2014, 68, 1071–1082. [Google Scholar] [CrossRef]
  15. Bansal, S.; Natesan, S. A novel Higher-Order efficient computational method for pricing European and Asian options. Numer. Algorithms 2024. [Google Scholar] [CrossRef]
  16. Rieth, Á.; Kovács, R.; Fülöp, T. Implicit numerical schemes for generalized heat conduction equations. Int. J. Heat. Mass. Transf. 2018, 126, 1177–1182. [Google Scholar] [CrossRef]
  17. Britz, D.; Baronas, R.; Gaidamauskaitė, E.; Ivanauskas, F. Further comparisons of finite difference schemes for computational modelling of biosensors. Nonlinear Anal. Model. Control 2009, 14, 419–433. [Google Scholar] [CrossRef]
  18. Mbroh, N.A.; Munyakazi, J.B. A robust numerical scheme for singularly perturbed parabolic Reaction-Diffusion problems via the method of lines. Int. J. Comput. Math. 2021, 99, 1139–1158. [Google Scholar] [CrossRef]
  19. Manaa, S.; Sabawi, M. Numerical Solution and Stability Analysis of Huxley Equation. AL-Rafidain J. Comput. Sci. Math. 2005, 2, 85–97. [Google Scholar] [CrossRef]
  20. Pargaei, M.; Kumar, B.V.R. A 3D Haar wavelet method for a coupled degenerate system of parabolic equations with nonlinear source coupled with Non-Linear ODEs. Appl. Numer. Math. 2023, 185, 141–164. [Google Scholar] [CrossRef]
  21. Anjuman; Leung, A.Y.T.; Das, S. Two-Dimensional Time-Fractional Nonlinear Drift Reaction–Diffusion Equation Arising in Electrical Field. Fractal Fract. 2024, 8, 456. [Google Scholar] [CrossRef]
  22. Rufai, M.A.; Kosti, A.A.; Anastassi, Z.A.; Carpentieri, B. A New Two-Step Hybrid Block Method for the FitzHugh–Nagumo Model Equation. Mathematics 2024, 12, 51. [Google Scholar] [CrossRef]
  23. Ndou, N.; Dlamini, P.; Jacobs, B.A. Enhanced Unconditionally Positive Finite Difference Method for Advection–Diffusion–Reaction Equations. Mathematics 2022, 10, 2639. [Google Scholar] [CrossRef]
  24. Heidari, M.; Ghovatmand, M.; Skandari, M.H.N.; Baleanu, D. Numerical Solution of Reaction–Diffusion Equations with Convergence Analysis. J. Nonlinear Math. Phys. 2023, 30, 384–399. [Google Scholar] [CrossRef]
  25. Kolev, M.K.; Koleva, M.N.; Vulkov, L.G. An Unconditional Positivity-Preserving Difference Scheme for Models of Cancer Migration and Invasion. Mathematics 2022, 10, 131. [Google Scholar] [CrossRef]
  26. Ascher, U.M.; Ruuth, S.J.; Wetton, B.T.R. Implicit-Explicit Methods for Time-Dependent Partial Differential Equations. SIAM J. Numer. Anal. 1995, 32, 797–823. [Google Scholar] [CrossRef]
  27. Calvo, M.P.; de Frutos, J.; Novo, J. Linearly implicit Runge–Kutta methods for Advection–Reaction–Diffusion equations. Appl. Numer. Math. 2001, 37, 535–549. [Google Scholar] [CrossRef]
  28. Settanni, G.; Sgura, I. Devising efficient numerical methods for oscillating patterns in Reaction-Diffusion systems. J. Comput. Appl. Math. 2016, 292, 674–693. [Google Scholar] [CrossRef]
  29. Yadav, V.S.; Singh, A.; Maurya, V.; Rajpoot, M.K. New RK type Time-Integration methods for stiff Convection–Diffusion–Reaction systems. Comput. Fluids 2023, 257, 105865. [Google Scholar] [CrossRef]
  30. Beuken, L.; Cheffert, O.; Tutueva, A.; Butusov, D.; Legat, V. Numerical Stability and Performance of Semi-Explicit and Semi-Implicit Predictor–Corrector Methods. Mathematics 2022, 10, 2015. [Google Scholar] [CrossRef]
  31. Fedoseev, P.; Pesterev, D.; Karimov, A.; Butusov, D. New Step Size Control Algorithm for Semi-Implicit Composition ODE Solvers. Algorithms 2022, 15, 275. [Google Scholar] [CrossRef]
  32. Ji, Y.; Xing, Y. Highly Accurate and Efficient Time Integration Methods with Unconditional Stability and Flexible Numerical Dissipation. Mathematics 2023, 11, 593. [Google Scholar] [CrossRef]
  33. Chou, C.S.; Zhang, Y.T.; Zhao, R.; Nie, Q. Numerical methods for stiff Reaction-Diffusion systems. Discret. Contin. Dyn. Syst.-Ser. B 2007, 7, 515–525. [Google Scholar] [CrossRef]
  34. Jiang, T.; Zhang, Y.T. Krylov implicit integration factor WENO methods for semilinear and fully nonlinear Advection-Diffusion-Reaction equations. J. Comput. Phys. 2013, 253, 368–388. [Google Scholar] [CrossRef]
  35. Essongue, S.; Ledoux, Y.; Ballu, A. Speeding up mesoscale thermal simulations of powder bed additive manufacturing thanks to the forward Euler Time-Integration scheme: A critical assessment. Finite Elem. Anal. Des. 2022, 211, 103825. [Google Scholar] [CrossRef]
  36. Chen-Charpentier, B.M.; Kojouharov, H.V. An unconditionally positivity preserving scheme for Advection-Diffusion reaction equations. Math. Comput. Model. 2013, 57, 2177–2185. [Google Scholar] [CrossRef]
  37. Appadu, A.R. Performance of UPFD scheme under some different regimes of advection, diffusion and reaction. Int. J. Numer. Methods Heat. Fluid. Flow. 2017, 27, 1412–1429. [Google Scholar] [CrossRef]
  38. Savović, S.; Drljača, B.; Djordjevich, A. A comparative study of two different finite difference methods for solving Advection–Diffusion reaction equation for modeling exponential traveling wave in heat and mass transfer processes. Ric. Mat. 2022, 71, 245–252. [Google Scholar] [CrossRef]
  39. Al-Bayati, A.; Manaa, S.; Al-Rozbayani, A. Comparison of Finite Difference Solution Methods for Reaction Diffusion System in Two Dimensions. AL-Rafidain J. Comput. Sci. Math. 2011, 8, 21–36. [Google Scholar] [CrossRef]
  40. Pourghanbar, S.; Manafian, J.; Ranjbar, M.; Aliyeva, A.; Gasimov, Y.S. An efficient alternating direction explicit method for solving a nonlinear partial differential equation. Math. Probl. Eng. 2020, 2020, 47416. [Google Scholar] [CrossRef]
  41. Nagy, Á.; Majár, J.; Kovács, E. Consistency and Convergence Properties of 20 Recent and Old Numerical Schemes for the Diffusion Equation. Algorithms 2022, 15, 425. [Google Scholar] [CrossRef]
  42. Nagy, Á.; Omle, I.; Kareem, H.; Kovács, E.; Barna, I.F.; Bognar, G. Stable, Explicit, Leapfrog-Hopscotch Algorithms for the Diffusion Equation. Computation 2021, 9, 92. [Google Scholar] [CrossRef]
  43. Saleh, M.; Kovács, E.; Barna, I.F.; Mátyás, L. New Analytical Results and Comparison of 14 Numerical Schemes for the Diffusion Equation with Space-Dependent Diffusion Coefficient. Mathematics 2022, 10, 2813. [Google Scholar] [CrossRef]
  44. Saleh, M.; Kovács, E.; Barna, I.F. Analytical and Numerical Results for the Transient Diffusion Equation with Diffusion Coefficient Depending on Both Space and Time. Algorithms 2023, 16, 184. [Google Scholar] [CrossRef]
  45. Askar, A.H.; Nagy, Á.; Barna, I.F.; Kovács, E. Analytical and Numerical Results for the Diffusion-Reaction Equation When the Reaction Coefficient Depends on Simultaneously the Space and Time Coordinates. Compututation 2023, 11, 127. [Google Scholar] [CrossRef]
  46. Barakat, H.Z.; Clark, J.A. On the solution of the diffusion equations by numerical methods. J. Heat. Transf. 1966, 88, 421–427. [Google Scholar] [CrossRef]
  47. Liu, H.; Leung, S. An Alternating Direction Explicit Method for Time Evolution Equations with Applications to Fractional Differential Equations. Methods Appl. Anal. 2020, 26, 249–268. [Google Scholar] [CrossRef]
  48. Hirsch, C. Numerical Computation of Internal and External Flows, Volume 1: Fundamentals of Numerical Discretization; John Wiley & Sons: Hoboken, NJ, USA, 1988. [Google Scholar]
  49. Gourlay, A.R.; McGuire, G.R. General Hopscotch Algorithm for the Numerical Solution of Partial Differential Equations. IMA J. Appl. Math. 1971, 7, 216–227. [Google Scholar] [CrossRef]
  50. Kovács, E.; Nagy, Á.; Saleh, M. A set of new stable, explicit, second order schemes for the Non-Stationary heat conduction equation. Mathematics 2021, 9, 2284. [Google Scholar] [CrossRef]
  51. Kovács, E.; Nagy, Á. A new stable, explicit, and generic Third-Order method for simulating conductive heat transfer. Numer. Methods Partial. Differ. Equ. 2023, 39, 1504–1528. [Google Scholar] [CrossRef]
  52. Kovács, E.; Majár, J.; Saleh, M. Unconditionally Positive, Explicit, Fourth Order Method for the Diffusion- and Nagumo-Type Diffusion–Reaction Equations. J. Sci. Comput. 2024, 98, 39. [Google Scholar] [CrossRef]
  53. Jalghaf, H.K.; Kovács, E.; Majár, J.; Nagy, Á.; Askar, A.H. Explicit stable finite difference methods for Diffusion-Reaction type equations. Mathematics 2021, 9, 3308. [Google Scholar] [CrossRef]
Figure 1. The structures of original odd–even hopscotch (OOEH), leapfrog–hopscotch (LH), and Dufort–Frankel (DF) methods.
Figure 1. The structures of original odd–even hopscotch (OOEH), leapfrog–hopscotch (LH), and Dufort–Frankel (DF) methods.
Computation 13 00129 g001
Figure 2. The concentration of u in terms of x in the case of the initial function u 0 ; for u e x a c t analytical solution, CLL and LH are the numerical methods.
Figure 2. The concentration of u in terms of x in the case of the initial function u 0 ; for u e x a c t analytical solution, CLL and LH are the numerical methods.
Computation 13 00129 g002
Figure 3. L2 errors as a function of the time step size Δ t for the simple coupling case.
Figure 3. L2 errors as a function of the time step size Δ t for the simple coupling case.
Computation 13 00129 g003
Figure 4. The concentration of u in terms of x in the case of the initial function u 0 , for u e x a c t analytical solution, and the numerical solutions by the CLQ2 and LH schemes.
Figure 4. The concentration of u in terms of x in the case of the initial function u 0 , for u e x a c t analytical solution, and the numerical solutions by the CLQ2 and LH schemes.
Computation 13 00129 g004
Figure 5. Errors as a function of the time step size Δ t for the simple coupling case.
Figure 5. Errors as a function of the time step size Δ t for the simple coupling case.
Computation 13 00129 g005
Figure 6. Aggregated errors as a function of the number of a for the 8 methods in Case Study 3.
Figure 6. Aggregated errors as a function of the number of a for the 8 methods in Case Study 3.
Computation 13 00129 g006
Figure 7. Aggregated errors as a function of the coefficient b in Case Study 4.
Figure 7. Aggregated errors as a function of the coefficient b in Case Study 4.
Computation 13 00129 g007
Figure 8. Aggregated errors as a function of the parameter value of A in Case Study 5.
Figure 8. Aggregated errors as a function of the parameter value of A in Case Study 5.
Computation 13 00129 g008
Figure 9. Aggregated errors as a function of the duration of the simulation TIME in Case Study 6.
Figure 9. Aggregated errors as a function of the duration of the simulation TIME in Case Study 6.
Computation 13 00129 g009
Figure 10. Aggregated errors as a function of the duration of the simulation TIME in Case Study 7.
Figure 10. Aggregated errors as a function of the duration of the simulation TIME in Case Study 7.
Computation 13 00129 g010
Figure 11. Aggregated errors as a function of the coupling coefficient k in Case Study 8.
Figure 11. Aggregated errors as a function of the coupling coefficient k in Case Study 8.
Computation 13 00129 g011
Table 1. The aggregated errors (AgErr) for the values of a for simple coupling in Case Study 3.
Table 1. The aggregated errors (AgErr) for the values of a for simple coupling in Case Study 3.
Numerical Methods0.10.3136102030100300
ADEu−20.45−18.70−12.14−8.56−2.214.3324.6938.61166.61420.26
ADEw−16.91−16.05−10.02−6.51−0.166.3626.6940.61168.60422.25
DFu−20.29−16.63−13.09−11.74−5.26−0.7913.5223.0587.91215.24
DFw−16.92−14.68−10.84−9.73−3.221.2215.5625.0889.93217.23
OOEHu−18.71−17.06−10.68−4.496.1018.1150.6575.71267.83676.56
OOEHw−15.08−14.04−8.44−2.198.5820.8653.9879.51273.61684.90
CpCu−9.40−6.12−4.96−2.95−1.26−1.360.293.4133.03123.56
CpCw−6.54−3.79−2.69−0.840.800.692.315.4435.03125.56
CCLu−10.27−7.05−6.38−4.30−1.89−1.143.819.3160.57208.27
CCLw−7.42−4.71−4.17−2.220.160.905.8211.3362.57210.27
CLQ2u−10.46−7.23−6.61−4.51−1.96−1.054.3310.2365.33228.88
CLQ2w−7.61−4.90−4.41−2.430.090.986.3312.2567.33230.88
PIu−9.80−6.58−5.78−3.67−1.68−1.381.976.3247.37165.17
PIw−6.95−4.24−3.57−1.570.370.663.988.3449.37167.16
LHu−23.69−19.97−17.01−15.08−6.31−0.8617.2529.56113.61278.26
LHw−20.43−18.05−14.78−12.98−4.291.1419.2831.58115.61280.25
Table 2. The aggregated errors (AgErr) for the values of b for simple coupling in Case Study 4.
Table 2. The aggregated errors (AgErr) for the values of b for simple coupling in Case Study 4.
Numerical Methods0.10.3131030100300
ADEu−16.03−12.90−9.03−8.85−5.46−3.174.5612.58
ADEw−21.48−16.69−10.41−8.08−2.412.0012.0722.22
DFu−12.74−9.49−12.40−8.31−8.02−6.01−1.465.60
DFw−18.70−13.19−13.79−7.64−4.96−0.836.0615.30
OOEHu−14.57−11.11−7.42−6.09−1.325.1421.0942.46
OOEHw−19.50−14.86−8.60−4.922.0210.8229.4853.50
CpCu−2.93−0.99−2.88−1.18−3.12−4.28−5.87−7.54
CpCw−8.62−4.53−4.21−0.44−0.020.921.672.09
CCLu−3.89−2.15−3.82−1.92−3.74−4.51−4.72−4.03
CCLw−9.59−5.68−5.20−1.19−0.650.682.805.59
CLQ2u−4.08−2.35−4.01−2.08−3.87−4.53−4.52−3.52
CLQ2w−9.77−5.88−5.40−1.35−0.780.663.016.10
PIu−3.42−1.64−3.36−1.53−3.51−4.56−5.53−5.84
PIw−9.11−5.17−4.73−0.80−0.420.642.003.78
LHu−16.09−12.95−16.06−10.53−9.53−6.72−0.289.53
LHw−22.09−16.70−17.45−9.87−6.48−1.557.2319.21
Table 3. The aggregated errors (AgErr) for the values of A for simple coupling in Case Study 5.
Table 3. The aggregated errors (AgErr) for the values of A for simple coupling in Case Study 5.
Numerical
Methods
−10−3−101310
ADEu−8.43−10.59−8.10−13.86−22.27−38.65−88.89
ADEw−67.24−23.25−6.75−5.48−6.57−8.04−9.96
DFu−8.09−7.60−10.85−18.43−26.66−43.72−97.15
DFw−65.49−19.45−9.50−9.45−10.23−12.91−14.55
OOEHu−9.87−6.78−4.01−9.94−17.95−32.74−80.60
OOEHw−67.03−18.38−2.38−1.55−2.83−4.04−5.52
CpCu−2.64−0.58−2.25−9.92−19.80−37.71−92.86
CpCw−60.87−13.13−0.90−0.85−2.57−4.43−6.22
CCLu−3.47−1.45−3.59−10.94−20.54−37.94−92.41
CCLw−61.37−13.91−2.24−2.05−3.74−5.56−7.96
CLQ2u−3.64−1.62−3.84−11.10−20.69−38.05−92.39
CLQ2w−61.51−14.06−2.49−2.22−3.90−5.73−8.15
PIu−3.11−1.02−2.96−10.54−20.24−37.70−91.99
PIw−61.07−13.50−1.60−1.62−3.36−5.17−7.40
LHu−10.00−10.08−14.72−19.77−27.89−44.40−100.32
LHw−67.32−21.82−13.37−11.13−12.03−14.53−20.20
Table 4. The aggregated errors (AgErr) for the values of A for simple coupling in Case Study 6.
Table 4. The aggregated errors (AgErr) for the values of A for simple coupling in Case Study 6.
Numerical Methods0.00010.00030.0010.0030.010.030.10.3131030100300
ADEu−52.56−47.12−40.45−32.62−22.69−13.64−8.56−7.20−9.39−9.53−8.55−7.80−6.43−5.91
ADEw−50.28−44.85−38.18−30.35−20.43−11.45−6.51−5.31−7.30−7.54−6.51−5.48−4.14−3.61
DFu−48.82−42.92−36.21−29.69−22.20−15.01−11.74−6.28−4.83−4.69−2.64−2.06−2.67−3.25
DFw−46.49−40.58−33.91−27.42−19.95−12.82−9.73−4.41−3.13−3.20−0.91−0.11−0.96−1.54
OOEHu−53.20−48.50−41.81−34.40−24.77−14.54−4.496.8318.7334.1052.1763.9773.2079.25
OOEHw−51.96−46.83−39.85−32.11−22.22−11.89−2.198.9720.5935.5653.2164.5873.3478.96
CpCu−47.82−39.68−30.37−21.53−12.55−5.79−2.95−0.40−8.05−4.40−5.53−4.64−4.50−4.52
CpCw−45.48−37.32−27.99−19.24−10.27−3.56−0.841.52−5.94−3.30−2.81−2.26−2.21−2.23
CCLu−51.52−44.14−35.15−26.33−16.16−7.85−4.30−0.74−8.23−4.99−6.37−5.54−5.59−5.73
CCLw−49.20−41.80−32.78−24.04−13.90−5.63−2.221.15−6.20−4.01−3.75−3.27−3.37−3.49
CLQ2u−52.89−45.88−36.93−28.14−17.98−8.68−4.51−0.77−8.43−5.01−6.37−5.54−5.59−5.73
CLQ2w−50.56−43.52−34.56−25.83−15.70−6.46−2.431.12−6.21−4.02−3.75−3.27−3.37−3.50
PIu−47.69−39.67−30.54−21.83−13.03−6.41−3.67−0.66−7.59−4.95−6.24−5.41−5.41−5.43
PIw−45.35−37.31−28.15−19.55−10.76−4.18−1.571.23−5.98−3.93−3.65−3.15−3.20−3.21
LHu−50.12−44.52−38.13−31.76−24.23−17.10−15.08−9.11−8.07−8.73−6.69−4.02−3.19−3.73
LHw−47.79−42.19−35.85−29.50−22.01−14.92−12.98−7.29−6.41−7.14−5.11−2.26−1.40−1.85
Table 5. The aggregated errors (AgErr) for the values of TIME for isomerization in Case Study 7.
Table 5. The aggregated errors (AgErr) for the values of TIME for isomerization in Case Study 7.
Numerical Method0.00010.00030.0010.0030.010.030.10.3131030100300
ADEu−68.44−62.54−55.36−47.34−37.54−28.60−20.34−17.65−15.43−14.64−12.99−11.14−9.63−9.00
ADEw−81.73−76.07−68.39−60.20−50.42−41.13−31.54−26.54−24.12−22.05−21.29−20.08−19.59−19.58
DFu−62.42−55.65−47.85−40.71−32.86−25.84−19.77−14.45−9.78−7.44−3.44−1.75−0.93−0.64
DFw−71.05−63.66−55.17−47.40−39.11−32.38−29.13−21.52−19.26−17.15−12.74−11.96−11.40−11.55
OOEHu−69.91−64.44−57.03−49.07−39.32−29.95−19.48−15.66−9.79−5.810.645.7110.4014.62
OOEHw−82.76−76.75−68.90−60.71−50.28−40.29−29.67−22.47−16.36−8.601.3310.1619.0527.14
CpCu−61.06−51.54−40.12−30.35−21.57−14.87−8.71−4.57−2.22−1.46−1.23−1.19−1.19−1.19
CpCw−69.20−58.62−46.87−37.18−28.23−21.67−16.29−13.01−11.39−11.62−12.87−14.84−16.25−17.81
CCLu−66.24−57.47−46.83−36.66−25.40−17.05−9.91−5.33−2.84−2.13−2.01−2.02−2.05−2.06
CCLw−76.31−66.46−54.48−43.21−31.79−23.54−17.17−13.63−12.20−12.97−14.14−15.73−16.92−18.29
CLQ2u−67.91−59.61−49.04−38.79−27.70−17.91−10.12−5.39−2.86−2.14−2.01−2.03−2.05−2.07
CLQ2w−78.83−69.35−57.48−46.18−33.82−24.36−17.36−13.67−12.22−13.00−14.15−15.74−16.92−18.29
PIu−60.84−51.36−40.12−30.60−22.03−15.48−9.38−5.17−2.78−2.10−1.98−2.00−2.03−2.04
PIw−68.88−58.42−46.89−37.36−28.55−22.06−16.69−13.46−12.12−12.90−14.13−15.70−16.85−18.04
LHu−63.78−57.23−49.73−42.79−35.23−28.59−22.32−17.01−13.13−11.22−7.89−3.91−1.63−0.94
LHw−73.22−66.26−58.30−50.91−42.77−35.93−30.48−24.46−22.75−20.90−16.49−13.42−11.13−10.88
Table 6. The aggregated errors (AgErr) for the values of k for isomerization in Case Study 8.
Table 6. The aggregated errors (AgErr) for the values of k for isomerization in Case Study 8.
Numerical Method0.010.030.10,313103060100170300
ADEu−18.3−18.5−19.1−20.34−22.07−22.84−22.48−18.89−5.3921.3779.39198.79
ADEw−42.81−38.67−34.48−31.54−29.95−29.51−28.8−24.59−10.6616.0774.48193.67
DFu−19.43−19.80−20.07−19.77−18.52−16.88−11.1418.8557.48101.26165.45262.00
DFw−38.57−34.60−31.13−29.13−25.89−25.22−18.4212.1351.2695.02159.50256.10
OOEHu−16.86−17.07−17.77−19.48−23.34−24.15−22.45−12.483.9123.4263.60142.68
OOEHw−41.34−37.16−32.86−29.67−27.69−27.19−25.72−16.33−0.2319.0459.13137.85
CpCu−10.17−10.88−12.10−8.71−6.85−6.71−7.56−8.790−9.56−7.146.0848.02
CpCw−27.72−23.57−19.35−16.29−14.62−14.63−15.94−17.71−18.17−15.27−1.5540.72
CCLu−11.09−11.82−13.69−9.91−7.99−8.02−9.14−10.56−11.43−3.5416.3365.85
CCLw−28.57−24.42−20.20−17.17−15.63−15.94−17.77−19.58−20.00−11.378.9158.87
CLQ2u−11.28−12.02−13.93−10.12−8.20−8.24−9.37−10.79−11.69−12.30−12.203.89
CLQ2w−28.75−24.60−20.38−17.36−15.83−16.17−18.03−19.80−20.28−20.57−19.75−2.89
PIu−10.59−11.32−13.07−9.38−7.45−7.42−8.51−9.97−10.91−11.57−11.88−1.10
PIw−28.11−23.96−19.70−16.69−15.10−15.31−17.04−19.04−19.61−19.90−19.58−7.98
LHu−22.80−23.13−23.41−22.32−20.46−18.91−19.46−19.56−19.67−19.20−16.47−6.27
LHw−41.86−37.82−33.82−30.48−27.98−27.00−27.90−27.81−26.83−25.96−22.92−12.67
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khayrullaev, H.; Zain, A.; Kovács, E. Solution of Coupled Systems of Reaction–Diffusion Equations Using Explicit Numerical Methods with Outstanding Stability Properties. Computation 2025, 13, 129. https://doi.org/10.3390/computation13060129

AMA Style

Khayrullaev H, Zain A, Kovács E. Solution of Coupled Systems of Reaction–Diffusion Equations Using Explicit Numerical Methods with Outstanding Stability Properties. Computation. 2025; 13(6):129. https://doi.org/10.3390/computation13060129

Chicago/Turabian Style

Khayrullaev, Husniddin, Andicha Zain, and Endre Kovács. 2025. "Solution of Coupled Systems of Reaction–Diffusion Equations Using Explicit Numerical Methods with Outstanding Stability Properties" Computation 13, no. 6: 129. https://doi.org/10.3390/computation13060129

APA Style

Khayrullaev, H., Zain, A., & Kovács, E. (2025). Solution of Coupled Systems of Reaction–Diffusion Equations Using Explicit Numerical Methods with Outstanding Stability Properties. Computation, 13(6), 129. https://doi.org/10.3390/computation13060129

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop