Next Article in Journal
Cauchy Problem with Summable Initial-Value Functions for Parabolic Equations with Translated Potentials
Next Article in Special Issue
Evaluation of the Segmental Casting Length of Strongly Restrained Super-Long Mass Concrete Based on Crack Resistance
Previous Article in Journal
A Study on the Nature of Complexity in the Spanish Electricity Market Using a Comprehensive Methodological Framework
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Efficient Linearized Difference Algorithm for a Diffusive Selkov–Schnakenberg System

1
School of Computer Science, Zhengzhou University of Economics and Business, Zhengzhou 451191, China
2
School of Mathematics and Statistics, Zhengzhou University, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
Mathematics 2024, 12(6), 894; https://doi.org/10.3390/math12060894
Submission received: 2 February 2024 / Revised: 5 March 2024 / Accepted: 15 March 2024 / Published: 18 March 2024
(This article belongs to the Special Issue Numerical and Computational Methods in Engineering)

Abstract

:
This study provides an efficient linearized difference algorithm for a diffusive Selkov–Schnakenberg system. The algorithm is developed by using a finite difference method that relies on a three-level linearization approach. The boundedness, existence and uniqueness of the solution of our proposed algorithm are proved. The numerical experiments not only validate the accuracy of the algorithm but also preserve the Turing patterns.

1. Introduction

The occurrence of oscillatory patterns, multiple steady-state solutions and chaotic behaviors is a fascinating phenomenon observed across numerous chemical, biological and physical systems [1,2]. Turing’s theory reveals the principles of the relationship between the patterns produced by convection–diffusion systems and these phenomena [3]. As a famous example related to cellular processes in biochemical systems, the Selkov–Schnakenberg system has attracted the attention of many scholars on the stability and the existence of steady-state solutions [4,5,6,7].
The Selkov–Schnakenberg model, as an extension of the Sel’kov model [8] and Schnakenberg model [9], can describes the limit cycle behavior:
A V , B U , 2 U + V 3 U ,
where A and B are the chemical sources of V and U, respectively. U is the auto-catalyst, and V is the reactant. The mathematical modeling of the process leads to the following Selkov–Schnakenberg system [10]:
{ (2) u t = d 1 Δ u u + u 2 v + b + a v , x Ω , t > 0 , (3) v t = d 2 Δ v + λ u 2 v a v , x Ω , t > 0 ,
where u represents the concentration of the auto-catalyst, and v represents the concentration of the reactant. Their corresponding diffusion coefficients are denoted by d 1 and d 2 , respectively. The variables x and t are time and space variables, respectively. The dimensionless constant rate for the low activity state is given by a, and b and λ represent dimensionless chemical sources, with a , b 0 and λ > 0 .
With the homogeneous Neumann boundary conditions:
u n = v n = 0 , o n Ω × ( 0 , T ) ,
and initial conditions
u ( x , 0 ) = ϕ ( x ) , v ( x , 0 ) = ψ ( x ) , x Ω ,
where Ω represents the boundary of domain Ω , and n denotes the outward unit normal vector of Ω .
Numerous works in the literature are devoted to the study of (2)–(5), including the Sel model and the Schnakenberg model; See [1,2,7,10,11,12,13,14], and the references therein. In 2015, Zhou and Shi [14] investigate stability, instability, time-periodic orbits and spatiotemporal patterns through bifurcation methods and Leray–Schauder degree theory. In 2017, Li and Wang [12] focused on Selkov–Schnakenberg systems, explored steady-state issues, and provided the criteria for the formation of spatial patterns (especially Turing patterns) based on the results of the presence and absence of non-constant steady states. In 2014, Uecker and Wetzel [1] analyzed the patterns of the Selkov–Schnakenberg system in two dimensions by using the pde2path [15] finite element software package, numerically calculating embedded branches, such as hexagons in stripes. In 2021, Al Noufaey [2] used the Galerkin method to study the singularity behavior and stability of the Selkov–Schnakenberg system. In 2023, Wang and Zhou et al. [7] studied the Turing instability (diffusion drive) causing spatial patterns and obtained the conditions for the existence of Turing bifurcations, then the changes in the spatiotemporal pattern that depend on the parameters were theoretically analyzed, and a series of numerical experimental simulations were conducted to verify the analysis results through the finite difference method. To the best of our knowledge, there is relatively limited research on numerical algorithms for the Selkov–Schnakenberg system at present.
In view of the difficulties caused by nonlinear terms in the Selkov–Schnakenberg system, in this paper, we primarily focus on investigating a linearized finite difference algorithm. Taking into account the characteristics of Neumann boundary conditions, we explore the construction of a three-level linearized finite difference algorithm. Then, we provide a rigorous theoretical analysis of the boundedness, existence and uniqueness of the solutions for our proposed algorithm. Numerical results will demonstrate that the algorithm has second-order accuracy in both time and space, and we explore the spatial patterns of the system solutions.
The structure of this paper is as follows: In Section 2, we recall some fundamental notations and lemmas. In Section 3, a three-level finite difference algorithm is constructed in detail. The boundedness, existence and uniqueness of solutions are proved in detail for our algorithm in Section 4. Two numerical examples are provided to verify the efficiency of the algorithm and the theoretical analysis results in Section 5. Finally, we present conclusions in Section 6.

2. Preliminaries

In this section, we review some notations and lemmas that will be used in the remainder of the paper.
We will divide the domain [ L , R ] × [ 0 , T ] . Take positive integers m , n , divide the domain [ L , R ] into m equal parts, and divide the domain [ 0 , T ] into n equal parts. The mesh size and nodes are denoted as follows:
h = ( R L ) / m , x i = i h , 0 i m . τ = T / n , t k = k τ , 0 k n .
Denote
δ x u i + 1 2 = u i + 1 u i h ,           δ x 2 u i = u i 1 2 u i + u i + 1 h 2 , u k ¯ = u k + 1 + u k 2 , Δ t u k = u k + 1 u k 1 2 τ .
For convenience, we introduce the inner product ( · , · ) :
( u , v ) = h ( u 0 v 0 2 + i = 1 m 1 u i v i + u m v m 2 ) ,
then we define norms · , | · | 1 :
u = ( u , u ) , | u | 1 = h i = 1 m ( δ x u i 1 2 ) 2 .
Lemma 1
(See [16]). Let a and b be the given constants, and h > 0 .
(1) If g ( x ) C 3 [ c , c + h ] , then
g ( c ) = 2 h [ g ( c + h ) g ( c ) h g ( c ) ] h 3 g ( η 1 ) , c < η 1 < c + h ,
If g ( x ) C 4 [ c , c + h ] , then
g ( c ) = 2 h [ g ( c + h ) g ( c ) h g ( c ) ] h 3 g ( c ) h 2 12 g ( 4 ) ( η 2 ) , c < η 2 < c + h .
(2) If g ( x ) C 3 [ c h , c ] , then
g ( c ) = 2 h [ g ( c ) g ( c ) g ( c h ) h ] + h 3 g ( η 3 ) , c h < η 3 < c ,
If g ( x ) C 4 [ c h , c ] , then
g ( c ) = 2 h [ g ( c ) g ( c ) g ( c h ) h ] + h 3 g ( c ) h 2 12 g ( 4 ) ( η 4 ) , c h < η 4 < c .
Lemma 2
(See [7]). On the premise that the positive equilibrium
E * ( u * , v * ) = ( λ + b , λ a + ( b + λ ) 2 ) ,
of the corresponding local system is stable, for the Selkov–Schnakenberg system (2)–(5), the following hold:
1.
If λ 2 b 2 a , the positive Equilibrium E * is asymptotically stable for all d 1 , d 2 > 0 .
2.
If λ 2 b 2 > a , we have the following results:
(a)
When d 2 d 1 max { 1 , ( a + ( λ + b ) 2 ) 2 λ 2 b 2 a } , the positive equilibrium E * is asymptotically stable.
(b)
When d 2 d 1 > max { 1 , ( a + ( λ + b ) 2 ) 2 λ 2 b 2 a } , and when a > a T , the positive equilibrium E * is asymptotically stable. And when a < a T , the positive equilibrium E * is unstable. When a = a T with k = k c , the system undergoes Turing bifurcation, where a T , k c are the parameters in literature [7].

3. A Three-Level Linearized Difference Algorithm

In this section, we will present the derivation process of our linearized difference algorithm.
Considering Equations (2) and (3) at the node ( x i , t k ) , we have
{   u t ( x i , t k ) = d 1 u x x ( x i , t k ) u ( x i , t k ) + ( u ( x i , t k ) ) 2 v ( x i , t k ) (10) + b + a v ( x i , t k ) , 0 i m , 1 k n 1 ,   v t ( x i , t k ) = d 2 v x x ( x i , t k ) + λ ( u ( x i , t k ) ) 2 v ( x i , t k ) (11) a v ( x i , t k ) , 0 i m , 1 k n 1 .
Applying numerical differentiation formulas to the above equations, it can be observed that
Δ t U i k = d 1 δ x 2 U i k ¯ U i k ¯ + U i k V i k U i k ¯ + b + a V i k + P i k , 1 i m 1 , 1 k n 1 ,
Δ t V i k = d 2 δ x 2 V i k ¯ + λ ( U i k ) 2 V i k ¯ a V i k ¯ + R i k , 1 i m 1 , 1 k n 1 .
Then, there exist constants c 1 and c 2 such that
| P i k | c 1 ( h 2 + τ 2 ) , 1 i m 1 , 1 k n 1 , | R i k | c 2 ( h 2 + τ 2 ) , 1 i m 1 , 1 k n 1 .
Taking the derivative of both sides of the system of equations with respect to x, we have:
u x t = d 1 u x x x u x + 2 u u x v + u 2 v x + a v x , x Ω , t > 0 , v x t = d 2 v x x x 2 u u x v u 2 v x a v x , x Ω , t > 0 .
Applying boundary conditions (4), we obtain
u x x x ( L , t ) = 0 , u x x x ( R , t ) = 0 , 0 < t T ,
v x x x ( L , t ) = 0 , v x x x ( R , t ) = 0 , 0 < t T .
Applying the boundary conditions (4), (14), (15) and using the Lemma 1, we can deduce
u x x ( L , t ) = 2 h 2 [ u ( x 1 , t ) u ( x 0 , t ) ] + O ( h 2 ) , u x x ( R , t ) = 2 h 2 [ u ( x m , t ) u ( x m 1 , t ) ] + O ( h 2 ) , v x x ( L , t ) = 2 h 2 [ v ( x 1 , t ) v ( x 0 , t ) ] + O ( h 2 ) , v x x ( R , t ) = 2 h 2 [ v ( x m , t ) v ( x m 1 , t ) ] + O ( h 2 ) .
Therefore,
u x x ( L , t k + 1 2 ) = 1 2 [ u x x ( L , t k ) + u x x ( L , t k + 1 ) ] + O ( τ 2 ) = 1 2 [ 2 h 2 ( U 1 k U 0 k ) + 2 h 2 ( U 1 k + 1 U 0 k + 1 ) ] + O ( τ 2 + h 2 ) = 2 h δ x U 1 2 k ¯ + O ( τ 2 + h 2 ) ,
u x x ( R , t k + 1 2 ) = 1 2 [ u x x ( R , t k ) + u x x ( R , t k + 1 ) ] + O ( τ 2 ) = 1 2 [ 2 h 2 ( U m k U m 1 k ) 2 h 2 ( U m k + 1 U m 1 k + 1 ) ] + O ( τ 2 + h 2 ) = 2 h δ x U m 1 2 k ¯ + O ( τ 2 + h 2 ) .
Employing a similar methodology, the result is
v x x ( L , t k + 1 2 ) = 2 h δ x V 1 2 k ¯ + O ( τ 2 + h 2 ) ,
            v x x ( R , t k + 1 2 ) = 2 h δ x V m 1 2 k ¯ + O ( τ 2 + h 2 ) .
Substituting i = 0 into Equation (10) and combining it with Equation (16), we derive
Δ t U 0 k = d 1 2 h δ x U 1 2 k ^ U 0 k ¯ + U 0 k V 0 k U 0 k ¯ + b + a V 0 k + P 0 k , 1 k n 1 .
Similarly, letting i = m in Equation (10) and combining it with Equation (17), we obtain
Δ t U m k = d 1 2 h δ x U m 1 2 k ^ U m k ¯ + U m k V m k U m k ¯ + b + a V m k + P m k , 1 k n 1 ,
where there exists a constant c 3 such that
| P 0 k | c 3 ( h 2 + τ 2 ) , 1 k n 1 , | P m k | c 3 ( h 2 + τ 2 ) , 1 k n 1 .
Similar to the treatment of u, by substituting i = 0 into Equation (11) and combining it with Equation (18), we obtain
Δ t V 0 k = d 2 2 h δ x V 1 2 k ^ + λ U 0 k V 0 k U 0 k ¯ a V 0 k ¯ + R 0 k , 1 k n 1 .
Letting i = m in Equation (11) and combining it with Equation (19), we have
Δ t V m k = d 2 2 h δ x V m 1 2 k ^ + λ ( U m k ) 2 V m k ¯ a V m k ¯ + R m k , 1 k n 1 ,
where there exists a constant c 4 such that
| R 0 k | c 4 ( h 2 + τ 2 ) , 1 k n 1 , | R m k | c 4 ( h 2 + τ 2 ) , 1 k n 1 .
From Equations (2) and (5), it follows that
u t ( x , 0 ) = d 1 ϕ x x ( x ) ϕ ( x ) + ϕ 2 ( x ) ψ ( x ) + b + a ψ ( x ) .
Then, we can obtain
U i 1 = ϕ ( x i ) + τ u t ( x i , 0 ) + P i 0 , 0 i m ,
where there exists a constant c 5 such that
| P i 0 | c 5 τ 2 , 0 i m , | δ x P i + 1 2 0 | c 5 τ 2 , 0 i m .
Deriving from Equations (3) and (5), we arrive
v t ( x , 0 ) = d 2 ψ x x ( x ) + λ ϕ 2 ( x ) ψ ( x ) a ψ ( x ) .
Then, we can obtain
V i 1 = ψ ( x i ) + τ v t ( x i , 0 ) + R i 0 , 0 i m ,
where there exists a constant c 6 such that
| R i 0 | c 6 τ 2 , 0 i m , | δ x R i + 1 2 0 | c 6 τ 2 , 0 i m .
Neglecting the infinitesimal terms in Equations (12), (13) and (20)–(25), we can establish the following three-level linearized difference algorithm for the Sel–Schnakenberg system (2)–(5):
{ (26) Δ t u 0 k = d 1 2 h δ x u 1 2 k ^ u 0 k ¯ + u 0 k v 0 k u 0 k ¯ + b + a v 0 k , 1 k n 1 , (27) Δ t v 0 k = d 2 2 h δ x v 1 2 k ^ + λ ( u 0 k ) 2 v 0 k ¯ a v 0 k ¯ , 1 k n 1 , (28) Δ t u i k = d 1 δ x 2 u i k ¯ u i k ¯ + u i k v i k u i k ¯ + b + a v i k , 1 i m 1 , 1 k n 1 , (29) Δ t v i k = d 2 δ x 2 v i k ¯ + λ ( u i k ) 2 v i k ¯ a v i k ¯ , 1 i m 1 , 1 k n 1 , (30) Δ t u m k = d 1 2 h δ x u m 1 2 k ^ u m k ¯ + u m k v m k u m k ¯ + b + a v m k , 1 k n 1 , (31) Δ t v m k = d 2 2 h δ x v m 1 2 k ^ + λ ( u m k ) 2 v m k ¯ a v m k ¯ , 1 k n 1 , (32) u i 0 = ϕ ( x i ) , v i 0 = ψ ( x i ) , 0 i m , (33) u i 1 = ϕ ( x i ) + τ u t ( x i , 0 ) , 0 i m , (34) v i 1 = ψ ( x i ) + τ v t ( x i , 0 ) , 0 i m .

4. Theoretical Analysis

In this section, we will give a strict theoretical analysis of the boundedness, existence and uniqueness of the solutions to our proposed algorithm, respectively.

4.1. The Boundedness of Our Algorithm Solutions

Theorem 1.
Let { u i k , v i k | 0 i m , 0 k n } be solutions of the system (26)–(34). Then, there exist two constants c 7 and c 8 such that
v k c 7 , u k c 8 , 0 k n .
Proof. 
From Equations (32)–(34), there exists a constant c 9 such that
u 0 c 9 , u 1 c 9 ,
v 0 c 9 , v 1 c 9 .
Subsequently, Equations (26)–(31) can be collectively expressed as follows:
Δ t u i k = d 1 δ x 2 u i k ¯ u i k ¯ + u i k v i k u i k ¯ + b + a v i k , 0 i m , 1 k n 1 ,
Δ t v i k = d 2 δ x 2 v i k ¯ + λ ( u i k ) 2 v i k ¯ a v i k ¯ , 0 i m , 1 k n 1 .
Taking the inner product of v i k ¯ with Equation (39), we obtain
1 4 τ ( v k + 1 2 v k 1 2 ) = d 2 | v k ¯ | 1 2 + ( λ , v k ¯ ) ( ( u k ) 2 v k ¯ , v k ¯ ) a v k ¯ 2 ( λ , v k ¯ ) λ v k ¯ λ v k + 1 + v k 1 2 , 1 k n 1 .
Thus,
v k + 1 v k 1 2 τ λ , 1 k n 1 .
Through recursion, it can be obtained
v k 2 τ λ k + v 0 2 T λ + v 0 , 1 k n .
Noting inequality (37), it can be derived that
v k c 7 , 0 k n .
Taking the inner product of u i k ¯ with Equation (38), we obtain
1 4 τ ( u k + 1 2 u k 1 2 ) = d 1 | u k ¯ | 1 2 u k ¯ 2 + ( u k v k u k ¯ , u k ¯ ) + ( b , u k ¯ ) + a ( v k , u k ¯ ) ( u k v k u k ¯ , u k ¯ ) + ( b , u k ¯ ) + a ( v k , u k ¯ ) u k v k u k ¯ u k ¯ + b u k ¯ + a v k u k ¯ ( u k v k u k ¯ + b + a v k ) u k + 1 + u k 1 2 , 1 k n 1 .
Combining inequality (40), we have
u k + 1 u k 1 2 τ ( u k v k u k ¯ + b + a c 7 ) , 1 k n 1 .
From (36) and (37), we can infer the validity of (35) for k = 0 , 1 .
Assuming that Equation (35) holds when k = n 1 , namely
v n 1 c 7 , u n 1 c 8 .
Let k = n 1 , and we have
u n u n 2 c 8 2 τ ( u n + u n 2 ) + 2 τ ( b + a c 7 ) .
Taking τ 1 c 8 2 1 2 , (41) becomes
u n ( 1 + 4 τ ) u n 2 + 4 τ c ( b + a c 7 ) .
By the Gronwall inequality, it can be derived that
u n e 2 T ( u 0 + b + a c 7 4 ) c 10 e 2 T ,
and thus, the inequality holds when k = n .
By the induction method, we can obtain that (35) holds. □

4.2. The Existence and Uniqueness of Solutions for Our Algorithm

Theorem 2.
The finite difference algorithm (26)–(34) is uniquely solvable.
Proof. 
From Equations (32)–(34), u 0 , u 1 , v 0 and v 1 are uniquely determined. Assuming u k 1 , u k , v k 1 and v k are determined, then the system (26)–(31) yields a system of linear equations for u k + 1 and v k + 1 . Consider the homogeneous system of (26)–(31) as follows:
{ (42) 1 2 τ u 0 k + 1 = d 1 1 h δ x u 1 2 k + 1 1 2 u 0 k + 1 + 1 2 u 0 k v 0 k u 0 k + 1 , (43) 1 2 τ v 0 k + 1 = d 2 1 h δ x v 1 2 k + 1 1 2 ( u 0 k ) 2 v 0 k + 1 1 2 a v 0 k + 1 , (44) 1 2 τ u i k + 1 = 1 2 d 1 δ x 2 u i k + 1 1 2 u i k + 1 + 1 2 u i k v i k u i k + 1 , 1 i m 1 , (45) 1 2 τ v i k + 1 = 1 2 d 2 δ x 2 v i k + 1 1 2 ( u i k ) 2 v i k + 1 1 2 a v i k + 1 , 1 i m 1 , (46) 1 2 τ u m k + 1 = d 1 1 h δ x u m 1 2 k + 1 1 2 u m k + 1 + 1 2 u m k v m k u m k + 1 , (47) 1 2 τ v m k + 1 = d 2 1 h δ x v m 1 2 k + 1 1 2 ( u m k ) 2 v m k + 1 1 2 a v m k + 1 .
Multiplying Equations (42), (44), and (46) by 1 2 h u 0 k + 1 , h u i k + 1 and 1 2 h u m k + 1 respectively, and then summing them up, we obtain
1 2 τ u k + 1 2 = 1 2 d 1 | u k + 1 | 1 2 1 2 u k + 1 2 + 1 2 h ( 1 2 u 0 k v 0 k | u 0 k + 1 | 2 + i = 1 m 1 u i k v i k | u i k + 1 | 2 + 1 2 u m k v m k | u m k + 1 | 2 ) 1 2 h ( 1 2 u 0 k v 0 k | u 0 k + 1 | 2 + i = 1 m 1 u i k v i k | u i k + 1 | 2 + 1 2 u m k v m k | u m k + 1 | 2 ) .
When u and v have opposite signs, (48) becomes
1 2 τ u k + 1 2 0 .
Then,
u k + 1 2 = 0 ,
that is
u k + 1 = 0 .
When u and v have the same sign, in conjunction with Theorem 1, Equation (48) becomes
1 2 τ u k + 1 2 c 11 2 u k + 1 2 .
If τ < 1 c 11 ,
u k + 1 2 = 0 ,
then,
u k + 1 = 0 .
Applying similar techniques to v, we obtain
v k + 1 = 0 .
To sum up, the finite difference algorithm (26)–(31) is uniquely solvable with respect to u k + 1 and v k + 1 . □

5. Numerical Results and Discussion

In this section, we use two numerical examples to verify the accuracy and efficiency of our proposed algorithm. All our tests were carried out using MATLAB 2017b running on a Lenovo desktop with 12 GB of RAM and 3.60 GHz CPU.

5.1. Example 1

In this example, let { U i k , V i k | 0 i m , 0 k n } be the exact solutions to the Selkov–Schnakenberg system (2)–(5), and { u i k , v i k | 0 i m , 0 k n } be the numerical solutions of the finite difference algorithm (26)–(34), then,
U i k = U i k u i k , V i k = V i k v i k , 0 i m , 0 k n .
We denote errors E r r and E r r 1 as follows:
E r r = U 2 + V 2 ,
and
E r r 1 = U 2 + V 2 + | U | 1 2 + | V | 1 2 .
Consider the following Selkov–Schnakenberg system:
{ (49) u t = d 1 Δ u u + u 2 v + b + a v + f , (50) v t = d 2 Δ v + λ u 2 v a v + g .
where f and g are source term functions of x and t.
We choose the domain Ω = ( 0 , 1 ) and take a = b = 0.01 , d 1 = d 2 = 10 , λ = 5 . Select the solutions as
u ( x , t ) = ( x 2 x ) t 2 , v ( x , t ) = ( x 2 x ) t 3 .
The source term functions f and g can be obtained by bringing the above information into Equations (49) and (50).
We take M = 1 / h from 4 to 1024, Δ t = 10 3 with T = 1 ; the numerical results are shown in Table 1. Then, we choose T = 1 ,   N t = 1 / τ , varying from 4 to 512 with h = 1 / 10 , 000 ; the numerical results are shown in Table 2. From Table 1 and Table 2, it is evident that our proposed algorithm exhibits second-order accuracy both in space and time.

5.2. Example 2

In this example, we will choose two sets of parameters corresponding to two cases to investigate the spatial patterns of solutions for the system (2)–(5).
For convenience, we can define the average concentration u ¯ n and v ¯ n
u ¯ n = 1 m i m u i n , v ¯ n = 1 m i m v i n .
(Case 1)
Taking the parameters as follows:
λ = 1 , b = 1 , a = 1 , d 1 = 0.1 , d 2 = 0.2 ,
it is easy to verify that this aligns with Lemma 2 (1).
Selecting two difference initial value conditions
u 0 ( x ) = 1 + sin ( x ) , v 0 ( x ) = 1 + cos ( x ) .
u 0 ( x ) = 2 , x = 25 , u 0 ( x ) = 0 , x 25 , v 0 ( x ) = 2 , x = 75 , v 0 ( x ) = 0 , x 75 .
We choose the spatial domain Ω = [ 0 , 100 ] , the time step τ = 0.005 , and the spatial step size h = 0.5 . Considering two different initial values, we can obtain the snapshots of the concentration at N t = 0 , 100, 10 , 000 and 50 , 000 , respectively, as shown in Figure 1a–h. From Figure 1d,h, we can clearly see that our proposed algorithm is stable and the concentration value of each point is the same.
To investigate the equilibrium points, we calculate the average concentrations u ¯ , v ¯ , u * , and v * . We present the concentration evolution diagrams, the average concentration (Figure 2a,c) and the concentration at point x = 50 (Figure 2b,d), under two initial value conditions. From Figure 2a,c, we can see that the average concentration quickly reaches the equilibrium point ( u * , v * ) . At point x = 50 , the concentration changes drastically after the beginning.
Finally, for a clearer presentation of the concentration variations, we provide numerical solutions in Figure 3a–d at all time steps under two different initial conditions.
(Case 2)
We consider the following parameters:
λ = 1.3 , b = 0.01 , a = 0.01 , d 1 = 1.5 , d 2 = 25 .
It can be verified that this conforms to the conditions stated in Lemma 2 (2a).
We choose the spatial domain Ω = [ 0 , 100 ] and the time step τ = 0.005 , and the space step is chosen as h = 0.5 . Considering two different initial values, we can obtain the snapshots of the concentration at N t = 0 , 100, 10 , 000 and 100 , 000 , respectively, as shown in Figure 4a–h. From Figure 4d–h, we can clearly see that our proposed algorithm is stable. Under the initial conditions (51), the solutions tend to reach an equilibrium state more rapidly. The solutions exhibit periodicity in the spatial domain.
Similar to case 1, we calculate the average concentrations u ¯ , v ¯ , u * , and v * . We present concentration evolution diagrams, the average concentration (Figure 5a,c) and the concentration at point x = 50 (Figure 5b,d), under two initial value conditions. From Figure 5a,c, it is evident that the solution eventually reaches the equilibrium point ( u * , v * ) . From Figure 5b,d, it can be observed that the time taken for the solutions to reach the equilibrium point is relatively longer compared to case 1.
Finally, for a clearer presentation of concentration variations, we provide numerical solutions in Figure 6a–d at all time steps under two different initial conditions.

6. Conclusions

An efficient linearized difference algorithm is developed to solve a diffusive Selkov–Schnakenberg system. We provide a detailed construction process of the algorithm. Proofs for the boundedness of the algorithm solutions, the existence of the algorithm solution and its uniqueness are all provided. Numerical examples validate the efficiency of the algorithm and are consistent with the theoretical analysis results. In the future, we will extend the algorithm to complex nonlinear problems in two and three dimensions and investigate the stability and convergence of the algorithm.

Author Contributions

Y.W.: Conceptualization, Methodology, Validation, Formal analysis, Software, Writing—Original Draft Preparation; X.B.: Conceptualization, Investigation, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank the anonymous referees for many constructive comments and suggestions which led to an improved presentation of this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Uecker, H.; Wetzel, D. Numerical Results for Snaking of Patterns over Patterns in Some 2D Selkov–Schnakenberg Reaction-Diffusion Systems. SIAM J. Appl. Dyn. Syst. 2014, 13, 94–128. [Google Scholar] [CrossRef]
  2. Al Noufaey, K. Stability analysis for Selkov-Schnakenberg reaction-diffusion system. Open Math. 2021, 19, 46–62. [Google Scholar] [CrossRef]
  3. Turing, A.M. The chemical basis of morphogenesis. Bull. Math. Biol. 1990, 52, 153–197. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, Y.H.; Iqbal, M.S.; Baber, M.Z.; Inc, M.; Ahmed, M.O.; Khurshid, H. On traveling wave solutions of an autocatalytic reaction–diffusion Selkov–Schnakenberg system. Results Phys. 2023, 44, 106129. [Google Scholar] [CrossRef]
  5. Goldbeter, A. Biochemical Oscillations and Cellular Rhythms: The Molecular Bases of Periodic and Chaotic Behaviour; Cambridge University Press: Cambridge, UK, 1997. [Google Scholar]
  6. Iqbal, M.S.; Seadawy, A.R.; Baber, M.Z. Demonstration of unique problems from Soliton solutions to nonlinear Selkov–Schnakenberg system. Chaos Solitons Fractals 2022, 162, 112485. [Google Scholar] [CrossRef]
  7. Wang, Y.; Zhou, X.; Jiang, W.; Qi, L. Turing instability and pattern formation in a diffusive Sel’kov–Schnakenberg system. J. Math. Chem. 2023, 61, 1036–1062. [Google Scholar] [CrossRef]
  8. Sel’Kov, E. Self-Oscillations in Glycolysis 1. A Simple Kinetic Model. Eur. J. Biochem. 1968, 4, 79–86. [Google Scholar] [CrossRef] [PubMed]
  9. Schnakenberg, J. Simple chemical reaction systems with limit cycle behaviour. J. Theor. Biol. 1979, 81, 389–400. [Google Scholar] [CrossRef] [PubMed]
  10. Li, B.; Wang, F.; Zhang, X. Analysis on a generalized Sel’kov–Schnakenberg reaction–diffusion system. Nonlinear Anal. Real World Appl. 2018, 44, 537–558. [Google Scholar] [CrossRef]
  11. Li, Y.; Zhou, Y. Turing–Hopf bifurcation in a general Selkov–Schnakenberg reaction–diffusion system. Chaos Solitons Fractals 2023, 171, 113473. [Google Scholar] [CrossRef]
  12. Li, B.; Zhang, X. Steady states of a sel’kov-schnakenberg reactin-diffusion system. Discret. Contin. Dyn. Syst.-Ser. S 2017, 10. [Google Scholar]
  13. Hepson, O.E.; Dag, I. Finite Element Method for Schnakenberg Model. In Mathematical Methods in Engineering: Applications in Dynamics of Complex Systems; Springer: Berlin/Heidelberg, Germany, 2019; pp. 41–51. [Google Scholar]
  14. Zhou, J.; Shi, J. Pattern formation in a general glycolysis reaction–diffusion system. IMA J. Appl. Math. 2015, 80, 1703–1738. [Google Scholar] [CrossRef]
  15. Uecker, H.; Wetzel, D.; Rademacher, J.D. pde2path-A Matlab package for continuation and bifurcation in 2D elliptic systems. Numer. Math. Theory, Methods Appl. 2014, 7, 58–106. [Google Scholar] [CrossRef]
  16. Sun, Z.Z.; Zhang, Q.; Gao, G.h. Finite Difference Methods for Nonlinear Evolution Equations; Walter de Gruyter GmbH & Co KG: Berlin, Germany, 2023; Volume 8. [Google Scholar]
Figure 1. Concentration snapshots at different times, (ad) under the conditions (51), (eh) under the conditions (52), respectively, in case 1.
Figure 1. Concentration snapshots at different times, (ad) under the conditions (51), (eh) under the conditions (52), respectively, in case 1.
Mathematics 12 00894 g001
Figure 2. Concentration evolution: (a) average concentration, (b) concentration at x = 50 under condition (51); (c) average concentration, (d) concentration at x = 50 under condition (52), in case 1.
Figure 2. Concentration evolution: (a) average concentration, (b) concentration at x = 50 under condition (51); (c) average concentration, (d) concentration at x = 50 under condition (52), in case 1.
Mathematics 12 00894 g002
Figure 3. Numerical solutions: (a,b) under the conditions (51); (c,d) under the conditions (52), in case 1.
Figure 3. Numerical solutions: (a,b) under the conditions (51); (c,d) under the conditions (52), in case 1.
Mathematics 12 00894 g003
Figure 4. Concentration snapshots at different times, (ad) under the conditions (51), (eh) under the conditions (52), respectively, in case 2.
Figure 4. Concentration snapshots at different times, (ad) under the conditions (51), (eh) under the conditions (52), respectively, in case 2.
Mathematics 12 00894 g004
Figure 5. Concentration evolution: (a) average concentration, (b) concentration at x = 50 under condition (51); (c) average concentration, (d) concentration at x = 50 under condition (52), in case 2.
Figure 5. Concentration evolution: (a) average concentration, (b) concentration at x = 50 under condition (51); (c) average concentration, (d) concentration at x = 50 under condition (52), in case 2.
Mathematics 12 00894 g005
Figure 6. Numerical solutions: (a,b) under the conditions (51); (c,d) under the conditions (52), in case 2.
Figure 6. Numerical solutions: (a,b) under the conditions (51); (c,d) under the conditions (52), in case 2.
Mathematics 12 00894 g006
Table 1. Errors and convergence rates in space for the proposed algorithm.
Table 1. Errors and convergence rates in space for the proposed algorithm.
MErrOrderErr1Order
45.1655 × 10−1-5.2581 × 10−1-
81.2189 × 10−12.081.2447 × 10−12.08
162.9515 × 10−22.053.0190 × 10−22.04
327.2559 × 10−32.027.4280 × 10−32.02
641.7985 × 10−32.011.8418 × 10−32.01
1284.4769 × 10−42.014.5853 × 10−42.01
2561.1171 × 10−42.001.1437 × 10−42.00
5122.7931 × 10−52.002.8547 × 10−52.00
10247.0136 × 10−61.997.1261 × 10−62.00
Table 2. Errors and convergence rates in time for the proposed algorithm.
Table 2. Errors and convergence rates in time for the proposed algorithm.
N t ErrOrderErr1Order
45.7535 × 10−3-3.0047 × 10−2-
81.3799 × 10−32.067.2819 × 10−32.04
163.3867 × 10−42.031.7980 × 10−32.02
328.3680 × 10−52.024.4489 × 10−42.01
642.0720 × 10−52.011.0985 × 10−42.02
1285.1796 × 10−62.002.7270 × 10−52.01
2561.3224 × 10−61.976.7907 × 10−62.01
5123.6489 × 10−71.861.6929 × 10−62.00
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

Wang, Y.; Bai, X. An Efficient Linearized Difference Algorithm for a Diffusive Selkov–Schnakenberg System. Mathematics 2024, 12, 894. https://doi.org/10.3390/math12060894

AMA Style

Wang Y, Bai X. An Efficient Linearized Difference Algorithm for a Diffusive Selkov–Schnakenberg System. Mathematics. 2024; 12(6):894. https://doi.org/10.3390/math12060894

Chicago/Turabian Style

Wang, Yange, and Xixian Bai. 2024. "An Efficient Linearized Difference Algorithm for a Diffusive Selkov–Schnakenberg System" Mathematics 12, no. 6: 894. https://doi.org/10.3390/math12060894

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