Funding: The authors received no specific funding for this work.
ABSTRACT
In this paper, we present a stable numerical scheme for solving two-dimensional
-component reaction–diffusion systems. The proposed approach utilizes the backward Euler method for temporal discretization and the hybridized discontinuous Galerkin (HDG) method for spatial discretization. We analyze the stability of the proposed HDG method for problems with Dirichlet and Neumann boundary conditions, demonstrating that this method is stable, in the sense of the energy method, under certain mild conditions on the stabilization parameters. Several numerical experiments are provided to validate the proposed scheme, with applications to two reaction–diffusion models: the Brusselator and glycolysis systems. Numerical results confirm that the proposed method achieves the expected optimal convergence rate for both the approximate solutions and their first derivatives. To exhibit relevant physical concepts, we demonstrate the convergence behavior of approximate solutions at the stable equilibrium points of the selected reaction–diffusion system with small diffusion coefficients. Furthermore, the nonconvergence behavior is given at unstable equilibrium points.
1 Introduction
Investigating diverse scientific phenomena often relies on applying partial differential equations (PDEs). The inherent complexity of PDEs prompts a keen interest in the development of approximate solutions. Indeed, this pursuit arises from the recognition that, in many cases, analytical solutions may not be readily accessible. Also, calculating numerical approximations is a difficult task in terms of computational cost and convergence. This paper focuses on solving nonlinear reaction–diffusion systems, one of the most widely used and significant types of PDEs. This system finds extensive application across diverse scientific and engineering domains, with a particular emphasis on their utility in modeling chemical systems [1]. Moreover, it serves as a valuable tool for theoretical modeling in biological contexts. Their applications span predictions of morphological mechanisms in tissue and organ formation, modeling chemotaxis and cell polarity in cellular biology, and investigating properties related to cell polarity [2-4]. The initial form of the reaction–diffusion system has been extensively developed by Turing instability [2, 5-8], and despite their various dynamics, it provided suitable information about patterns. Therefore, it failed to model interfaces. On the other hand, the Allen–Cahn equation is a scalar reaction–diffusion model for investigating the interfaces in applied mathematics; see [9, 10]. In this paper, we address the following two-dimensional (2D)
-component reaction–diffusion system [11]
(1)
where
is a positive constant,
is a given function, and
is the usual Laplacian operator over a rectangular-type domain
. To have a well-posed problem, system (1) is supplemented with appropriate initial and boundary conditions. Herein, the initial conditions are given by
and one of the following boundary conditions is imposed:
Dirichlet boundary conditions
(2)
Neumann boundary conditions
(3)
where
and
are given functions and
is the outward unit normal vector. Assuming the positivity of the solution to the reaction–diffusion system (1) is grounded in its representation of essential physical concepts such as density and population sizes, which inherently possess positive attributes.
Although we cannot claim that the HDG method is the best DG method, as shown in [12, 13], we anticipate that the HDG method is an efficient numerical approach for solving (1). The first HDG method was introduced for solving second-order elliptic problems [14, 15]. Then, it was used significantly in solving many known and significant equations, such as the Ito-type coupled KdV system [16], coupled Burgers equations [17], the second-order elliptic equations with Dirac delta source [18], the convection–diffusion equation [19], the Navier–Stokes equations [20], some fractional PDEs [21-23], and many others [24, 25]. While it is possible to formulate a DG system with a reduced number of unknowns, the computational time required to solve these nonlinear systems is not as straightforward. The efficacy of the HDG method hinges on its distinctive approach to defining numerical fluxes, setting it apart from other DG methods, particularly the local discontinuous Galerkin (LDG) methods. In the HDG method, after reformulating the given system into the first-order system of equations along with its corresponding weak formulation, it is essential to define appropriate numerical fluxes. Numerical fluxes are defined in such a way that they introduce global unknowns, known as numerical traces, and combined with the local unknowns. Based on the types of numerical fluxes, a matrix–vector equation with a smaller bandwidth is generated, considering the first benefit of the HDG method. Also, imposing global unknowns in numerical traces reduces the degrees of freedom in the full-discretization system, which makes less computational time for solving the matrix–vector equation compared to other DG methods. Finally, to balance the number of equations and the unknowns, the continuous Galerkin (CG) method is exploited to define sufficient global equations. Regardless of the stated advantages of the HDG method, investigating stability is the main achievement of any DG method. For instance, system (1) with determined boundary conditions has been solved in [26] by using a specific kind of LDG method on the Cartesian grids. The stability of the proposed LDG method in [26] has been investigated by using the matrix method whereas, in this paper, the stability of the proposed HDG method is studied in the context of the energy method. We note that the proposed HDG method can be applied to all numerical examples in [26].
Various applicable systems across diverse fields of science and engineering can derive by selecting different forms of
in (1) for
. Recently, some researchers have made efforts to solve these types of systems using various numerical methods [27-29]. This paper concentrates on validating the HDG method through numerical results, with a specific focus on applying it to the Brusselator and glycolysis systems. Using the modeling of autocatalytic chemical interactions [30], the Brusselator system has recently found extensive applications across various domains of chemistry and biology. Numerous numerical methods were employed to obtain the numerical solution of the Brusselator system. Notable approaches include the finite volume element method [30], the LDG method [31], and several others [32-34]. The glycolysis reaction–diffusion model is another type of the two-component reaction–diffusion system. The glycolysis model represents a biochemical reaction occurring within living cells [35, 36]. This paper presents a discussion of both the Brusselator and glycolysis systems.
The rest of the paper reads as follows. In Section 2, we introduce essential prerequisites, including approximation spaces, their associated inner products, and spatial partitioning. Section 3 is dedicated to the derivation of the HDG scheme. In Section 3, a fully discrete scheme corresponding to
-component reaction–diffusion system (1) is obtained by using the backward Euler method and an HDG method for temporal and spatial discretizations, respectively. Also, the stability analysis of the proposed method is investigated. Additionally, Section 3 provides insights into the implementation, along with instructions regarding the linearization of the nonlinear system. Section 4 illustrates the effectiveness of the proposed HDG method by presenting some numerical experiments. Various tests in this section are conducted on two distinct reaction–diffusion systems: glycolysis and Brusselator systems. The optimal order of convergence for both systems is reported with respect to approximation spaces and generated meshes. Furthermore, this section demonstrates that the obtained approximate solutions exhibit convergence to the equilibrium points of the systems for small diffusion coefficients, especially as the time step size tends to zero. The final section is devoted to the conclusion.
2 Prerequisites
The first step of constructing an HDG method is to consider suitable meshes for the temporal and spatial domains. The time interval
is partitioned into
equal subintervals as
(4)
where
where
is the time step size. The spatial domain
is partitioned into the set
as a collection of triangular elements
where
is the length of the longest edge among the edges of elements, and
. This mesh generation is chosen in such a way that two arbitrary elements do not have any overlap. The set
includes the boundaries of all elements where
denote the boundary of element
. The set of all interior faces and boundary faces of
, respectively, are denoted by
and
. Also, the set
is the collection of all faces. Note that, each member of
is repeated and counted twice in
. The concept of jump for scalar-valued function
on the interior faces
is defined as
and on the boundary faces
is defined as
.
and
denote outward unit normal vectors respect to the adjacent elements
and
with the common face
; see Figure 1.
Outward unit normal vectors
and
respect to the elements
and
. [Colour figure can be viewed at wileyonlinelibrary.com]
In the HDG method, two approximation spaces are needed, that is, for approximating the local unknowns on the elements and for approximating the global unknowns on the faces. Discontinuous finite element space for vector-valued functions is defined as
where
and
is the set of all polynomial of degree at most
over element
. More precisely,
includes all piecewise polynomial functions over the generated mesh
. Skeleton space or trace space is defined as
Also, applicable approximation subspace of
for performing the HDG method is defined as
where
and
is the
projection into the skeleton space. In fact, the subspace
is used to impose the projection of the given boundary data on
. Inner products of the spaces
and
are defined, respectively, as
where “
” is the inner product of two vector functions. Norms induced by the inner products are as follows:
In the following, the
norms over
and
simply are denoted by
.
3 Construction of the Numerical Scheme
This section is dedicated to the derivation of a fully discrete numerical scheme associated with the HDG approximation approach. Using the backward Euler method, a semidiscrete scheme can be achieved by discretization of the temporal variable. Then, a fully discrete form of system (1) is obtained by using the HDG method with defining suitable and relevant numerical fluxes/traces and stabilization parameters. Using the following backward finite difference scheme
(5)
where
, we get the following semidiscrete scheme corresponding to (1)
(6)
The next step is to find the corresponding first-order system of equations with respect to time partitioning. Consider the auxiliary variables
and
. A first-order system of equations corresponding to (6) is as follows:
(7)
where
. Let
, and
be numerical approximations of
, and
, respectively. In the corresponding weak formulation of (7), we seek numerical approximation
such that for all test functions
and all
, it holds that
(8)
where
and
are numerical flux and numerical trace, respectively. The numerical flux
is defined as
(9)
where the stabilization parameter
is responsible for the stability of the numerical scheme. Although definition of the numerical fluxes is not unique, their appropriate definition is needed to guarantee stability of the method. In the HDG scheme, boundary conditions are imposed to the definition of the numerical traces. So the numerical trace
at time
is defined as
(10)
where
is Dirichlet boundary data (2) and
for
are global unknowns at time level
which are added to the local unknowns
and
. We note that the boundary conditions are imposed on the boundary faces and numerical trace
acts on the interior faces as the global unknown
. Thus, it is needed to have
extra global equations for adding to weak formulation (8), which is obtained by applying conservation of the numerical fluxes. Indeed, we achieve the following global equations by imposing the zero jump to the numerical fluxes (9) on all interior faces,
(11)
Summing Equations (8) and (11) over all elements and using (9) and (10), we get the following system of equations:
(12)
where
and
Remark 1.If system (1) is equipped with the Neumann boundary conditions (3), because there is not any boundary value for
, then definition of the numerical traces (10) changes to
(13)
According to (13), it is needed to have
global equations which act on all faces with imposing the Neumann boundary conditions, that is,
(14)
With summing Equations (8) and (14) over all elements and using (9) and (13), we get the following system for
,
where
and
3.1 Stability Analysis
In this section, we focus on investigation of stability of the proposed HDG method. In studying reaction–diffusion system (1), some specific assumptions have to be considered over
. Without loss of generality, we assume that
is bounded over the spatial domain
, that is, there are positive constants
such that
Theorem 1.If-component reaction–diffusion system (1) is equipped with the homogeneous Dirichlet boundary conditions, then the proposed HDG method is stable provided that
.
Because
and
are single-value functions and
, for
, we conclude that
(20)
By substituting (20) in the sum of (19) over all elements, we get
(21)
Using the definition of numerical flux
from (9) in (21), we obtain
According to the assumption of the theorem, that is,
, for
, it can be concluded that
(22)
Summing (16) over all elements, and then using
and (22), we get
or equivalently
(23)
Using the Cauchy–Schwarz inequality, we have
(24)
which by assuming
to be sufficiently small, and by using the boundedness of
's, it leads to
Corollary 1.If-component reaction–diffusion system (1) is equipped with the homogeneous Neumann boundary conditions then the proposed HDG method is stable provided that
.
Proof.Based on the definition of global Equation (14) and
, we get
The fully discrete form of (12) has been presented based on the backward Euler method and HDG method, respectively, for the time and space discretization. The aim of this section is to solve the nonlinear system (12). So it is needed to linearize the nonlinear terms or exploit an appropriate numerical method for solving nonlinear systems, for instance, Newton–Raphson method. Using the Taylor series expansion to each
, for
, leads to the following second-order linearization formula:
(25)
Although it is enough to use a first-order linearization technique, we apply a second-order technique just to be sure that this linearization does not affect the temporal order of accuracy of the method. Using (25) into (12), for
, we have
(26)
where
The corresponding matrix–vector equation of (26) has not to be solved globally because of its high computational costs. Therefore, it is suggested to use the Schur complement technique that is raised in solving a system of linear equations. At first, the matrix–vector equation of (26) should be rewritten in the form of a
block matrix, and then, Schur complement technique leads to solve two smaller systems instead of solving a large system (26); see [16]. The idea of applying the Schur complement is not novel, and it is actually the basic HDG static condensation/local elimination but extended towards a system of equations. This utilized technique has been used successfully in previous works [37-39] and some more performance-oriented works [12, 40]. Substructuring techniques used in [40] can speed up the inversion of the block matrices which is relevant for variable-coefficient equations and can be applied to nonlinear system (1).
4 Numerical Simulations
In this section, we aim to evaluate the performance of the proposed HDG method in solving various systems of nonlinear reaction–diffusion equations, specifically the Brusselator and glycolysis systems. In Section 4.1, we investigate the Brusselator system, analyzing its properties and presenting numerical results through three examples. In the first example, the HDG method demonstrates the optimal order of convergence for both the approximate solutions and their first derivatives with respect to the spatial variables. Example 2 examines cases where analytical solutions are unavailable, showcasing the method's convergence towards equilibrium points. Example 3 further investigates two Brusselator systems, highlighting stable and unstable equilibrium points. Section 4.2 introduces the glycolysis system, followed by a performance assessment of the proposed HDG method in three examples. First, we confirm the optimal order of convergence for both the approximate solutions and their first derivatives. Then, in the fifth example, we explore the convergence of numerical results towards equilibrium points for systems with varying stability. Finally, we use the proposed HDG method to solve a three-component system. Note that, in all tests, we employ triangular meshes generated using the Delaunay triangulation algorithm.
4.1 Brusselator System
Brusselator system, a two-component reaction–diffusion system, is derived by considering system (1) with
, and
. See [32, 33]. In this form, the Brusselator system is expressed as
(27)
where
and
are real value constants, and
. For the small value of
, the dissipative terms in system (27) become negligible, which leads to the following system of ordinary differential equations:
(28)
System (28) has a single equilibrium (critical) point at
. If
and
are chosen such that
, it can be shown that the corresponding equilibrium point is stable, and the steady-state solutions
will converge to this equilibrium point. Otherwise, the equilibrium point is unstable and no convergence is observed. This topic is addressed in Examples 2 and 3. However, before exploring those, we will first examine the order of convergence for the approximate solutions and their first derivatives.
Example 1.In this example, we will employ the proposed HDG method to solve the Brusselator equation (27) with Dirichlet boundary conditions and coefficients
, and
, that is, [34]
(29)
where
. The exact solutions are given by
and
. Therefore, the right-hand side functions
and
, as well as the initial and boundary conditions, can be derived easily using the exact solutions. The results are reported in Table 1 for approximate polynomials of degree
at final time
. As expected, the
-error norms are satisfactory, demonstrating that the approximate solutions accurately estimate the analytical solutions. Also, in Table 1, we can observe the optimal order of convergence
in the spatial domain
for
and their first-order derivatives. To achieve the desired order of convergence, when halving the spatial mesh size
at each step, the time step size should be adjusted by dividing it by
. To this end, the first time step size is considered as
. In Table 2, we present another test to evaluate the time convergence order of the method. As expected, we observe first-order convergence for both the approximate solutions and their first derivatives, which is attributed to the use of the Euler method for time discretization.
Additionally, we conducted another test to evaluate the accuracy of the proposed method at a longer final
. In Figure 2, the contour plot of the absolute value of pointwise errors between the numerical solutions and analytical solutions are illustrated for approximation polynomials of degree one,
, and
. It is shown that the approximate solutions estimate the analytical solutions with high accuracy.
The contour plot of the absolute value of pointwise errors of numerical solutions and analytical solutions for Example 1 at final time
with
. [Colour figure can be viewed at wileyonlinelibrary.com]
TABLE 1.
error norms and corresponding spatial accuracy of the associated approximate solutions for Example 1 with polynomials of degree
and
at final time
.
Spatial order
Spatial order
0
0.2
2.8482 E-3
4.4793 E-2
0.1
1.5502 E-3
0.87
2.4915 E-2
0.85
0.05
8.3760 E-4
0.89
1.2562 E-2
0.99
1
0.4
5.5333 E-4
1.0097 E-2
0.2
1.2440 E-4
2.15
2.7227 E-3
1.89
0.1
3.2677 E-5
1.93
7.1223 E-4
1.94
2
0.4
6.4595 E-5
4.9854 E-4
0.2
7.3372 E-6
3.14
5.6062 E-5
3.15
0.1
8.8214 E-7
3.06
6.7099 E-6
3.06
Spatial order
Spatial order
0
0.2
2.4321 E-2
3.7904 E-1
0.2
1.3360 E-2
0.86
3.6709 E-1
0.87
0.05
6.5237 E-3
1.03
1.0216 E-1
0.98
1
0.4
5.1940 E-3
5.4826 E-2
0.2
1.5904 E-3
1.71
1.4923 E-2
1.88
0.1
4.5656 E-4
1.80
3.7385 E-3
2.00
2
0.4
5.4681 E-4
4.3203 E-3
0.2
5.9730 E-5
3.19
4.8290 E-4
3.16
0.1
6.9986 E-6
3.09
5.8120 E-5
3.05
TABLE 2.
error norms and corresponding temporal accuracy of the associated approximate solutions for Example 1 with polynomials of degree one and
at final time
.
Temporal order
Temporal order
0.8
8.7476 E-4
4.0123 E-3
0.4
4.1218 E-4
1.09
1.8911 E-3
1.09
0.2
1.18081 E-4
1.19
8.3003 E-4
1.19
0.1
8.9022 E-5
1.02
4.0922 E-4
1.02
0.05
4.4423 E-5
1.01
2.0423 E-4
1.01
0.025
2.2114 E-5
1.00
1.0378 E-4
0.98
Temporal order
Temporal order
0.8
4.1741 E-1
1.9506
0.4
2.2220 E-1
0.91
1.03777
0.91
0.2
1.2684 E-1
0.81
5.9328 E-1
0.81
0.1
6.4439 E-2
0.98
3.0220 E-1
0.97
0.05
3.2529 E-2
0.99
1.5393 E-1
0.97
0.025
1.6405 E-2
0.99
8.0188 E-2
0.95
Example 2.Consider the Brusselator system (27) with imposing homogeneous Neumann boundary conditions and setting
, and
; see [33, 34]. Also, consider the following initial conditions:
(30)
where
and
. The approximations of
and
are reported in Table 3 at points
, and
for
with approximate polynomial of degree one and
. Due to
, the approximate solutions
converge to critical point
. Moreover, in Figure 3, by showing
and
at times
, the convergence can be seen in the spatial domain by concentrating on the range of color data.
For further investigation, again consider (27) with homogeneous Neumann boundary conditions and initial data
(31)
where
and
. Moreover, the coefficients are chosen as
, and
where
. The reported results in Table 4 show the convergence of
to equilibrium point
at different points
, and
, for
. Due to the linearity of the initial conditions (31), the faster convergence is observed in Table 4 in comparison to Table 3.
Next, we investigate the order of convergence using the following formula:
(32)
where
, and
represent the approximate solutions with spatial mesh sizes
, and
, respectively. As shown in Table 5, we achieve the optimal convergence order for polynomial degrees
, with
, and
.
TABLE 3.
The values of the associated approximate solutions for Example 2, with initial conditions (30) at selected different points. The approximate polynomials are degree one,
, and
.
Obtained approximate solutions
(left) and
(right) for Example 2 at times
(from top to bottom) with polynomial of degree one,
, and
. [Colour figure can be viewed at wileyonlinelibrary.com]
TABLE 4.
The values of the associated approximate solutions for Example 2, with initial conditions (31) at selected different points. The approximate polynomials are degree one,
, and
.
(0.2, 0.2)
(0.4, 0.6)
(0.5, 0.5)
(0.8, 0.9)
1
2.3430
0.4167
2.4466
0.3954
2.4738
0.3906
2.6070
0.3684
2
2.0950
0.4686
2.1258
0.4600
2.1343
0.4577
2.1762
0.4471
3
2.0219
0.4915
2.0298
0.4887
2.0320
0.4880
2.0433
0.4842
4
2.0044
0.4981
2.0061
0.4975
2.0066
0.4973
2.0091
0.4963
5
2.0008
0.4996
2.0011
0.4995
2.0012
0.4995
2.0017
0.4993
6
2.0001
0.4999
2.0002
0.4999
2.0002
0.4999
2.0003
0.4999
7
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
8
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
9
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
10
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
2.0000
0.5000
2
0.5
2
0.5
2
0.5
2
0.5
TABLE 5.
The evaluated order of convergence in the spatial domain using the relation (32) for Example 2. The results are presented for polynomial degrees
and stabilization parameters
at final time
.
0
1
0.26
0.18
0.19
0.09
2
0.25
0.19
0.20
0.10
3
0.54
0.49
0.60
0.46
4
0.85
0.86
0.83
0.72
5
0.95
0.88
0.91
0.83
Expected order
1
1
1
1
1
1
1.32
0.44
1.16
0.39
2
1.88
1.65
1.79
1.72
3
1.90
1.71
1.85
1.89
4
1.92
1.89
1.93
1.90
Expected order
2
2
2
2
Example 3.In this numerical example, we intend to examine Brusselator system with stable and unstable equilibrium points. Consider Brusselator system (27) with homogeneous Neumann boundary conditions and initial conditions
and
where
; see [32, 33]. In Figure 4, the results are reported for
, and
. Indeed
and the approximate solutions converge to the stable equilibrium point
. In contrast, by choosing
, and
, we get
, and theoretically, we expect that the solution does not exhibit any convergence behavior. As seen, the HDG method produces divergence approximation solutions according to the theoretical achievements for unstable equilibrium point; see Figure 5.
Obtained approximate solutions
(the left side) and
(the right side) for Example 3 with
, and
at times
. The approximate polynomials are degree one,
, and
. [Colour figure can be viewed at wileyonlinelibrary.com]
Obtained approximate solutions
(the left side) and
(the right side) for Example 3 with
, and
at times
. The approximate polynomials are degree one,
, and
. [Colour figure can be viewed at wileyonlinelibrary.com]
4.2 Glycolysis System
By considering (1) with
, and
, the following glycolysis reaction–diffusion system [32] is achieved
(33)
where
and
are real value constants. Similar to the Brusselator system, for small values of
, the dissipative terms in (33) become negligible, reducing system (33) to a system of ODEs. Under the condition
(34)
the equilibrium point
is stable and approximate solutions
converge to this point as
goes to infinity [35, 41]. Otherwise, the equilibrium point is unstable. In the following, we demonstrate that this stability property is preserved by the proposed HDG method.
Example 4.Consider the following glycolysis system which is subjected to Dirichlet boundary conditions:
where
, and the right-hand side functions
and
are obtained by exact solutions
and
. The
error norms and corresponding order of accuracy are given in Table 6, for approximate polynomials of degree
at final time
. In this test, the initial time step size is set to
, and with each halving of the spatial mesh size
is updated by dividing it by
. As seen, the results are satisfactory, and the optimal order of convergence is gained in the spatial domain
for
and their first-order partial derivatives.
TABLE 6.
error norms and corresponding spatial accuracy of the associated approximate solutions for Example 4 with approximate polynomials of degree
and
at final time
.
Order
Order
0
0.2
2.4452 E-3
3.7635 E-2
0.1
1.3375 E-3
0.88
1.9306 E-2
0.96
0.05
7.3079 E-4
0.88
9.8531 E-3
0.97
1
0.2
1.9501 E-4
6.6192 E-3
0.1
3.8494 E-5
2.34
1.7486 E-3
1.92
0.05
9.0851 E-6
2.08
4.4117 E-4
1.99
2
0.2
1.2422 E-4
9.8089 E-4
0.1
1.3110 E-5
3.24
1.0141 E-4
3.27
0.05
1.4484 E-6
3.18
1.1180 E-5
3.18
Order
Order
0
0.2
2.1956 E-3
3.6199 E-2
0.1
1.1486 E-3
0.93
1.9115 E-2
0.92
0.05
5.9749 E-4
0.94
9.6999 E-3
0.98
1
0.2
1.8493 E-4
4.9198 E-3
0.1
4.5424 E-5
2.03
1.3165 E-3
1.90
0.05
1.1069 E-5
2.04
3.3108 E-4
1.99
2
0.2
1.5340 E-4
1.1535 E-3
0.1
1.6412 E-5
3.22
1.3508 E-4
3.09
0.05
1.8499 E-6
3.15
1.6044 E-5
3.07
Example 5.Consider system (33) with Neumann boundary conditions and initial conditions
and
, as presented in [32]. The values of approximate solution
are reported in Table 7 at the points
, and
for
, and
. As observed, the approximate solution converges to equilibrium point
, as the condition in (34) is satisfied, making
a stable equilibrium point. In contrast, by setting
, and
, the system yields an unstable equilibrium point. Therefore, as we expected, no convergence is observed in Figure 6.
Obtained approximate solutions
(the left side) and
(the right side) for Example 5 with
, and
at times
. The approximate polynomials are degree one,
, and
. [Colour figure can be viewed at wileyonlinelibrary.com]
TABLE 7.
The values of the associated approximate solutions for Example 5, with
, and
. The approximate polynomials are degree one,
, and
.
(0.2, 0.2)
(0.4, 0.6)
(0.5, 0.5)
(0.8, 0.9)
2
0.4426
0.0662
0.4979
0.0647
0.5049
0.0645
0.5759
0.0624
4
0.2763
0.0698
0.2837
0.0696
0.2846
0.0696
0.2940
0.0695
6
0.2536
0.0701
0.2546
0.0701
0.2548
0.0701
0.2560
0.0701
8
0.2505
0.0702
0.2506
0.0702
0.2507
0.0702
0.2508
0.0702
10
0.2501
0.0702
0.2501
0.0702
0.2501
0.0702
0.2501
0.0702
12
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
14
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
16
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
18
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
20
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.2500
0.0702
0.25
0.25
0.25
0.25
Example 6.By setting
, and
in system (1), the following three-component system is obtained.
(35)
which is equipped with Neumann boundary conditions; see [29]. In Figure 7, the approximate solutions are plotted for
, and the following initial conditions
The results are obtained using approximate polynomials of degree one, with
, and
. Compared to [29] and by focusing on the color data range, it is evident that system (35) reaches the true constant steady state, demonstrating the capability of the proposed HDG method to provide accurate and acceptable approximate solutions.
Obtained approximate solutions
(the left side),
(in the middle), and
(the right side) for Example 6 at times
. The approximate polynomials are degree one,
, and
. [Colour figure can be viewed at wileyonlinelibrary.com. [Colour figure can be viewed at wileyonlinelibrary.com]
5 Conclusion
In this paper, a stable HDG method has been proposed for solving the 2D nonlinear
-component reaction–diffusion system. The semidiscrete scheme is obtained by using the backward Euler method for time discretization. The fully discrete scheme has been achieved by applying an HDG method for the spatial domain. A suitable definition of numerical fluxes and traces is the success key of the HDG method which is led it to the numerically stable method. In Theorem 1, we have proved that by considering homogeneous Dirichlet boundary conditions and under some mild conditions on the stabilization parameters, the method is numerically stable in the sense of the energy method. In Corollary 1, the stability theorem has been extended for problems with homogeneous Neumann boundary conditions. Due to the existence of nonlinear terms in system (1), a section is devoted to presenting a linearization method. Therefore, a linear system has been presented by introducing some bilinear and functional terms. To get lower computational costs, we offer Schur's complement idea for solving the linear system. Sufficient numerical results are provided for confirming the stability of the proposed method by applying the HDG method to a three-component system, Brusselator system, and glycolysis system. For both systems, the HDG method demonstrated the optimal rate of convergence. Indeed, for a generated mesh with approximation polynomials of order,
in each element, approximate solutions, and their first partial derivatives converge to their corresponding exact values with order
. Selected physical properties of Brusselator and glycolysis systems are discussed in numerical examples. Approximate solutions have shown convergence behavior to the stable equilibrium points of Brusselator and glycolysis systems, and conversely, no convergence is observed to the unstable equilibrium points although in both cases the solution converge to a certain solution. The HDG method is one of the successful methods that can be constructed for such systems of differential equations, especially in higher dimensions.
Author Contributions
Shima Baharlouei: conceptualization, writing – original draft, writing – review and editing, investigation, visualization, validation, methodology, software, formal analysis. Reza Mokhtari: writing – review and editing, investigation, supervision, project administration. Nabi Chegini: writing – review and editing; supervision.
Conflicts of Interest
The authors declare no conflicts of interest.
References
1A. S. Mikhailov and K. Showalter, “Control of Waves, Patterns and Turbulence in Chemical Systems,” Physics Reports425, no. 2-3 (2006): 79–194.
2R. E. Baker, E. A. Gaffney, and P. K. Maini, “Partial Differential Equations for Self-Organization in Cellular and Developmental Biology,” Nonlinearity21, no. 11 (2008): R251.
3L. Marcon and J. Sharpe, “Turing Patterns in Development: What About the Horse Part?” Current Opinion in Genetics and Development22, no. 6 (2012): 578–584.
4M. Otsuji, S. Ishihara, C. Co, K. Kaibuchi, A. Mochizuki, and S. Kuroda, “A Mass Conserved Reaction-Diffusion System Captures Properties of Cell Polarity,” PLoS Computational Biology3, no. 6 (2007): e108.
5A. Gierer and H. Meinhardt, “ Biological Pattern Formation Involving Lateral Inhibition,” in Lectures on Mathematics in the Life Sciences, vol. 7 (American Mathematical Society, 1974), 163–182.
6M. Mimura, M. Tabata, and Y. Hosono, “Multiple Solutions of Two-Point Boundary Value Problems of Neumann Type With a Small Parameter,” SIAM Journal on Mathematical Analysis11, no. 4 (1980): 613–631.
7Y. Nishiura and H. Fujii, “Stability of Singularly Perturbed Solutions to Systems of Reaction-Diffusion Equations,” SIAM Journal on Mathematical Analysis18, no. 6 (1987): 1726–1770.
11J. Morgan and B. Q. Tang, “Boundedness for Reaction-Diffusion Systems With Lyapunov Functions and Intermediate Sum Conditions,” Nonlinearity33, no. 7 (2020): 3105.
13M. Kronbichler and W. A. Wall, “A Performance Comparison of Continuous and Discontinuous Galerkin Methods With Fast Multigrid Solvers,” SIAM Journal on Scientific Computing40, no. 5 (2018): A3423–A3448.
14B. Cockburn, B. Dong, and J. Guzmán, “A Superconvergent LDG-Hybridizable Galerkin Method for Second-Order Elliptic Problems,” Mathematics of Computation77, no. 264 (2008): 1887–1916.
15B. Cockburn, J. Gopalakrishnan, and R. Lazarov, “Unified Hybridization of Discontinuous Galerkin, Mixed, and Continuous Galerkin Methods for Second Order Elliptic Problems,” SIAM Journal on Numerical Analysis47, no. 2 (2009): 1319–1365.
16S. Baharloui, R. Mokhtari, and N. Chegini, “A Stable Numerical Scheme Based on the Hybridized Discontinuous Galerkin Method for the Ito-Type Coupled KdV System,” Communications on Applied Mathematics and Computation4 (2022): 1351–1373.
17S. Baharlouei, R. Mokhtari, and N. Chegini, “Solving Two-Dimensional Coupled Burgers Equations via a Stable Hybridized Discontinuous Galerkin Method,” Iranian Journal of Numerical Analysis and Optimization13, no. 3 (2023): 397–425.
18H. T. Leng and Y. P. Chen, “A Hybridizable Discontinuous Galerkin Method for Second Order Elliptic Equations With Dirac Delta Source,” ESAIM: M2AN56 (2022): 385–406.
19M. Li and X. B. Luo, “An EMC-HDG Scheme for the Convection-Diffusion Equation With Random Diffusivity,” Numerical Algorithms90, no. 4 (2022): 1755–1776.
22M. Fatih Karaaslan, “Solvability of the Fractional Volterra-Fredholm Integro Differential Equation by Hybridizable Discontinuous Galerkin Method,” Mathematical Methods in the Applied Sciences42, no. 16 (2019): 5626–5634.
23M. F. Karaaslan, “The Accuracy of an HDG Method for Conservative Fractional Diffusion Equations,” Mathematical Methods in the Applied Sciences41, no. 17 (2018): 8201–8211.
24S. Baharlouei, R. Mokhtari, and F. Mostajeran, “DNN-HDG: A Deep Learning Hybridized Discontinuous Galerkin Method for Solving Some Elliptic Problems,” Engineering Analysis with Boundary Elements151 (2023): 656–669.
25B. Cockburn, “Hybridizable Discontinuous Galerkin Methods for Second-Order Elliptic Problems: Overview, a New Result and Open Problems,” Japan Journal of Industrial and Applied Mathematics40, no. 3 (2023): 1637–1676.
26P. Castillo and S. Gomez, “Conservative Super-Convergent and Hybrid Discontinuous Galerkin Methods Applied to Nonlinear Schrödinger Equations,” Applied Mathematics and Computation371 (2020): 124950.
27R. Jiwari and A. Gerisch, “A Local Radial Basis Function Differential Quadrature Semi-Discretisation Technique for the Simulation of Time-Dependent Reaction-Diffusion Problems,” Engineering Computations38, no. 6 (2021): 2666–2691.
28R. Jiwari, S. Tomasiello, and F. Tornabene, “A Numerical Algorithm for Computational Modelling of Coupled Advection-Diffusion-Reaction Systems,” Engineering Computations35, no. 3 (2018): 1383–1401.
29C. Liu, C. Wang, Y. Wang, and S. M. Wise, “Convergence Analysis of the Variational Operator Splitting Scheme for a Reaction-Diffusion System With Detailed Balance,” SIAM Journal on Numerical Analysis60, no. 2 (2022): 781–803.
30Z. Lin, R. Ruiz-Baier, and C. Tian, “Finite Volume Element Approximation of an Inhomogeneous Brusselator Model With Cross-Diffusion,” Journal of Computational Physics256 (2014): 806–823.
31M. Dehghan and M. Abbaszadeh, “Variational Multiscale Element Free Galerkin (VMEFG) and Local Discontinuous Galerkin (LDG) Methods for Solving Two-Dimensional Brusselator Reaction-Diffusion System With and Without Cross-Diffusion,” Computer Methods in Applied Mechanics and Engineering300 (2016): 770–797.
32N. Ahmed, A. Korkmaz, M. Rafiq, et al., “A Novel Time Efficient Structure-Preserving Splitting Method for the Solution of Two-Dimensional Reaction-Diffusion Systems,” Advances in Difference Equations2020, no. 1 (2020): 1–26.
33R. C. Mittal and R. Jiwari, “Numerical Solution of Two-Dimensional Reaction-Diffusion Brusselator System,” Applied Mathematics and Computation217, no. 12 (2011): 5404–5415.
34S. O. K. Uysal and Y. Cicek, “Numerical Solution of the Brusselator Model by Time Splitting Method,” Cumhuriyet Science Journal42, no. 1 (2021): 75–87.
35R. E. Mickens, “ Positivity Preserving Discrete Model for the Coupled ODE's Modeling Glycolysis,” In Conference Publications (Vol. 2003, No. (Special), p. 623-629). American Institute of Mathematical Sciences, (2003).
38N. C. Nguyen, J. Peraire, and B. Cockburn, “High-Order Implicit Hybridizable Discontinuous Galerkin Methods for Acoustics and Elastodynamics,” Journal of Computational Physics230, no. 10 (2011): 3695–3718.
39N. C. Nguyen, J. Peraire, and B. Cockburn, “An Implicit High-Order Hybridizable Discontinuous Galerkin Method for the Incompressible Navier-Stokes Equations,” Journal of Computational Physics230, no. 4 (2011): 1147–1170.
40M. Kronbichler, S. Schoeder, C. Müller, and W. A. Wall, “Comparison of Implicit and Explicit Hybridizable Discontinuous Galerkin Methods for the Acoustic Wave Equation,” International Journal for Numerical Methods in Engineering106, no. 9 (2016): 712–739.
Please check your email for instructions on resetting your password.
If you do not receive an email within 10 minutes, your email address may not be registered,
and you may need to create a new Wiley Online Library account.
Request Username
Can't sign in? Forgot your username?
Enter your email address below and we will send you your username
If the address matches an existing account you will receive an email with instructions to retrieve your username
The full text of this article hosted at iucr.org is unavailable due to technical difficulties.