Hopf bifurcation and periodic solutions in a coupled Brusselator model of chemical reactions
Abstract
In this paper, we consider a coupled Brusselator model of chemical reactions, for which no symmetry for the coupling matrices is assumed. We show that the model can undergoes a Hopf bifurcation, and consequently periodic solutions can arise when the dispersal rates are large. Moreover, the effect of the coupling matrices on the Hopf bifurcation value is considered for a special case.
Keywords: Hopf bifurcation; Periodic solutions; Coupling
matrix; Line-sum symmetric matrix.
MSC 2010: 34C23, 37G15, 92C40
1 Introduction
Nonlinear oscillations often occur in many chemical reactions and physical processes, [9, 17, 19]. For example, sustained oscillations in Brusselator model were studied in [19]. Brusselator model, proposed by Prigogine and Lefever [30], describes a set of chemical reactions as follows:
(1.1) |
where and are the concentrations of the initial substances, and are the concentrations of the intermediate reactants, and and are the concentrations of the final products. If the concentrations and depend only on space, these chemical reactions can be modelled by the following reaction-diffusion model: (see [19, 30])
(1.2) |
Here are the dispersal rates. If and are spatially homogeneous, then model (1.2) admits a unique constant steady state for the homogeneous Neumann boundary condition. One can refer to [7, 10, 21, 23, 25, 36, 39] and references therein for the results on steady state and Hopf bifurcations near this constant steady state. The global bifurcation theory and some other methods were used to show the existence of non-constant steady states for a wide range of parameters, see, e.g., [3, 14, 26, 27, 28]. Moreover, the effect of advection was considered in [2, 18], and the diffusion term in (1.2) was replaced by the following form:
for the case that is one dimension.
If chemical reactions (1.1) take places in -boxes, then model (1.2) takes the following discrete form:
(1.3) |
Here is the number of boxes in chemical reactions; and , where and denote the concentrations of and in box at time , respectively; the nonlinear term describes the autocatalytic step in box ; and and denote the input concentrations of initial substances and in box , respectively. We remark that introduced here is the scaling parameter, and we choose it as the bifurcation parameter. Moreover, and are the coupling matrices, where describe the rates of movement from box to box for the two reactants, respectively, and and denote the rates of leaving box for .
Model (1.3) was first considered by Prigogine and Lefever in [30] for the case of two boxes (), where the coupling matrices and were assumed to be symmetric, and the two boxes were identical (that is, and ). In this case, model (1.3) admits a unique space-independent steady state:
and the existence of space-dependent steady state was showed in [30]. Model (1.3) with symmetric coupling matrices and identical boxes were also considered in [20, 22, 31, 32], where the steady state and Hopf bifurcation were studied in [20, 31]. One can also refer to [8, 29, 35, 40] and references therein for the steady state and Hopf bifurcations of other coupled models with symmetric coupling matrices.
It is well-known that the symmetric coupling matrices could mimic random diffusion, and the asymmetric case could mimic advective movements in the fluid. If the coupling matrices and are asymmetric, the steady states of model (1.3) are space-dependent even when the boxes are all identical. Consequently, we cannot obtain the explicit expression of the steady states, which brings difficulties in analyzing the Hopf bifurcation. In this paper, we aim to solve this problem and analyze the Hopf bifurcation of model (1.3).
Throughout the paper, we impose the following two assumptions:
-
The coupling matrices and are irreducible and essentially nonnegative;
-
.
Here we remark that real matrices with nonnegative off-diagonal elements are called essentially nonnegative. It follows from and Perron-Frobenius theorem that , where and are spectrum bounds of and , respectively. Clearly, and are also simple eigenvalues of and with strongly positive eigenvectors and , respectively, where
(1.4) |
Here represents that for . Assumption is a mathematically technical condition, and it means that the dispersal rates of the two reactants are proportional. Then letting , denoting , and dropping the tilde sign, model (1.3) can be transformed to the following equivalent model:
(1.5) |
where is defined in assumption .
We point out that the method in this paper is motivated by [4], where the steady state of the model is space-dependent. One can also refer to [1, 5, 6, 11, 12, 13, 15, 24, 34, 37, 38] on Hopf bifurcations near this type of steady state for delayed reaction-diffusion equations and delayed differential equations. We remark that, for the model in [4], the steady state does not depends on bifurcation parameter. But for patch models, the steady states always depend on the bifurcation parameter (see in model (1.5)), which brings some technical hurdles in analyzing Hopf bifurcations. Therefore, we need to improve the method in [4] here.
For simplicity, we use the following notations. We denote the complexification of a real linear space to be , and define the kernel of a linear operator by . For , we define the real part by . For the complex valued space , we choose the standard inner product , and consequently, the norm is defined by
Moreover, for , we denote
The rest of the paper is organized as follows. In Section 2, we study the existence and uniqueness of the positive equilibrium for model (1.5) (or respectively, (1.3)). In Section 3, we show the existence of Hopf bifurcation and the stability of the positive equilibrium for model (1.5) (or respectively, (1.3)) when is small. In Section 4, we show the effect of the coupling matrices on the Hopf bifurcation value for a special case. Finally, some numerical simulations are provided to illustrate the theoretical results.
2 Existence of positive equilibria
In this section, we consider the existence of positive equilibria of model (1.5), which satisfy
(2.1) |
Note that solves (2.1) for all when . Therefore, we cannot solve (2.1) by the direct application of the implicit function theorem. We need to split the phase space and find an equivalent system of (2.1). It is well-known that
where and are defined in (1.4), and
Letting
(2.2) |
and plugging (2.2) into (2.1), we see that (defined in (2.2)) is a solution of (2.1), if and only if solves
(2.3) |
where , and
(2.4) |
We first solve for .
Lemma 2.1.
Assume that . For fixed , has a unique solution , where
(2.5) |
Proof.
Plugging into for , we have and . Then plugging into , we have
which implies that
This completes the proof. ∎
Theorem 2.2.
Proof.
We first show the existence. It follows from Lemma 2.1 that
where is defined in (2.3). A direct computation implies that the Fréchet derivative of with respect to at is as follows:
where , , and
If , then and . Plugging into , we have
Noticing that
we obtain that and . Therefore, is bijective.
It follows from the implicit function theorem that there exist and a continuously differentiable mapping
such that and
Then we can choose (sufficiently small) such that (defined in (2.6)) is a positive solution of (2.1) for .
Now, we show the uniqueness. From the implicit function theorem, we only need to verify that if is a positive solution of (2.1), where
then as . Substituting into (2.3), we see from the first and third equation of (2.4) that
which implies that is bounded in . Then, up to a subsequence, we assume that
From the third equation of (2.4), we have
(2.8) |
which implies that are bounded. Taking the limit of
as , we see that , which yields . Therefore,
(2.9) |
It follows from (2.8) and (2.9) that is also bounded in . Then, up to a subsequence, we assume that
Taking the limit of
as , we see that , which yields . Consequently, taking the limit of
as , we have . Therefore,
Finally, we need to show that (2.7) holds. We can use the similar arguments as in the proof of the uniqueness, and here we omit the proof. ∎
In the above Theorem 2.2, we solve (2.1) when is in a small neighborhood of a given positive constant . In the following, we will consider the solution of (2.1) for a wider range of .
Theorem 2.3.
Proof.
It follows from Theorem 2.2, for any , there exists such that, for , (2.1) admits a unique positive solution , where and are defined in (2.6) and continuously differentiable for . Clearly,
Noticing that is compact, we see that there exist finite open intervals (see Fig. 1), denoted by for , such that

3 Stability and Hopf bifurcation
Throughout this section, we assume that , where is defined in Theorem 2.3. It follows from Theorem 2.3 that there exists such that, for , (1.5) admits a unique positive equilibrium , where and are defined in (2.10). Here we use as the bifurcation parameter, and we will show that there exists a Hopf bifurcation curve when is small, see Fig. 2.
Linearizing model (1.5) at , we obtain
where
(3.1) |
Let
(3.2) |
where and . Then, is an eigenvalue of if there exists such that
(3.3) |
For further application, we first give a priori estimates for solutions of eigenvalue problem (3.3).
Lemma 3.1.
Proof.
Ignoring a scalar factor, we assume that , where and are defined in (1.4). Substituting into (3.3), we have
(3.4) |
where are defined in (3.1) for . Multiplying the first and second equation of (3.4) by and , respectively, and summing these equations over all yield
(3.5) |
Note that there exists a positive constant such that
(3.6) |
This, combined with (3.5), implies that is bounded for .
Now we claim that there exists such that is bounded for , where is defined in Theorem 2.3. Note that , , and is bounded. If the claim is not true, then there exists a sequence such that , , with , and and with . Substituting into (3.4) and taking the limits of the two equations as , we have
(3.7) |
where and are defined in assumption . Without loss of generality, we assume that . Then it follows from (3.7) that is an eigenvalue of , which implies that or from [33, Corollary 4.3.2]. Since and , we have . Consequently, and , where , and . Summing the first equation of (3.4) over all , we have
This, combined with (3.6), implies that
which contradicts . Therefore, the claim is true. ∎
To analyze the stability of , we need to consider that whether the eigenvalues of (3.3) could pass through the imaginary axis. It follows from Lemma 3.1 that if is an eigenvalue of (3.3), then is bounded for , where is defined in Lemma 3.1. Substituting into (3.3), we have
(3.8) |
Ignoring a scalar factor, in (3.8) can be represented as
(3.9) |
Then we see that is a solution of (3.8), where , , and is defined in (3.9), if and only if
(3.10) |
is solvable for some value of . Here is defined in Lemma 2.3, and is defined by
where
and are defined in (3.1).
We first solve (3.10) for .
Lemma 3.2.
Proof.
Substituting into for , we have and . Then plugging and into , we have
Now, we solve (3.10) for .
Theorem 3.3.
Proof.
We first show the existence. It follows from Lemma 3.2 that , where . A direct computation implies that the Fréchet derivative of with respect to at is as follows:
where
If , we see that and . Plugging into , we get
where
and
A direct computation implies that
which implies that and . Therefore, is bijection. It follows from the implicit function theorem that there exists , where is sufficiently small, and a continuously differentiable mapping
such that , and for ,
Then is a positive solution of (3.10) for .
Now, we show the uniqueness. From the implicit function theorem, we only need to verify that if is a solution of (3.10), then
(3.15) |
Since
we obtain that , , and are bounded. It follows from Lemma 3.1 that is bounded. Then, up to a sequence, we can assume that
Taking the limit of (3.10) as , we see that is a solution of (3.10) for . This, combined with Lemma 3.2, implies that (3.15) holds. This completes the proof. ∎
Then from Theorem 3.3, we obtain the following result.
Theorem 3.4.
To show that is a simple eigenvalue of , we need to consider the eigenvalues of , where is the transpose of . Denote
Then
Lemma 3.5.
Proof.
It follows from Theorem 3.4 that, for , is an eigenvalue of and is one-dimensional. Then is an eigenvalue of , , and can be represented as (3.16). It follows from (3.16) that , , , and are bounded. Then, up to a sequence, we assume that
where
(3.17) |
Note that and , where are defined in Lemma 3.2. Taking the limit of
(3.18) |
as , we have and , which yield and . Noticing that , we see from (3.18) that
This, combined with Eqs. (3.13) and (3.14), implies that
(3.19) |
Note from (3.17) that This, combined with (3.19), implies that
This completes the proof. ∎
Now, we show that is simple.
Theorem 3.6.
Assume that , where is sufficiently small. Then is a simple eigenvalue of , where is defined in (3.2).
Proof.
It follows from Theorem 3.4 that , where is defined in Theorem 3.3. Then we show that
If , where and , then
which implies that there exists a constant such that
That is,
(3.20) |
It follows from Lemma 3.5 that
Then, multiplying the first and second equation of (3.20) by and , respectively, and summing the results over all , we have
It follows from Theorem 3.3 and Lemma 3.5 that
(3.21) |
which implies that for , where is sufficiently small. Therefore,
This completes the proof. ∎
It follows from Theorem 3.6 that is a simple eigenvalue of for fixed . Then, by using the implicit function theorem, we see that there exists a neighborhood of ( is defined as the neighborhood of ) and a continuously differentiable function such that . Moreover, for each , the only eigenvalue of in is , and
(3.22) |
Then, we show that the following transversality condition holds.
Theorem 3.7.
For , where is sufficiently small,
Proof.
Differentiating (3.22) with respect to at we have
(3.23) |
Note that
where and are defined in Lemma 3.5. Then we see from (3.23) that
(3.24) |
It follows from Theorem 2.3 that is continuously differentiable on . This, combined the fact that , implies that
(3.25) |
where is defined in (3.1). Then we see from (2.5) and (3.25) that
where
and is a zero matrix of It follows from Theorem 3.3 and Lemma 3.5 that
This, combined with (3.21) and (3.24), yields
∎

From Theorems 2.3, 3.3, 3.6 and 3.7, we can obtain the following results on the dynamics of model (1.5), see also Fig. 2.
Theorem 3.8.
Proof.
Note from [16, Chapter 2, Theorem 5.1] that the eigenvalues of are continuous with respect to We only need to show that there exists , which depends on , such that
If it is not true, then there exists a sequence such that , and
(3.26) |
is solvable for some value of with and . Here . Substituting into (3.26), we have
(3.27) |
Note from Lemma 3.1 that is bounded. Then, up to a subsequence, we assume that with . As in the proof of Lemma 3.5, we see that can be represented as
(3.28) |
and and satisfy
Up to a subsequence, we also assume that , and . Dividing the first and second equation of (3.27) by , respectively, summing the results over all , and taking the limits as , we have
It follows from (3.28) that at least one of and is not equal to zero. Consequently, is an eigenvalue of matrix
It follows from (3.12) and (3.13) that, for sufficiently small ,
which contradicts . This completes the proof. ∎
4 The effect of the coupling matrices
In this section, we show the effect of the coupling matrices on the Hopf bifurcation value. Moreover, some numerical simulations are given to illustrate the theoretical results. For simplicity, we consider a special case, where and the boxes are all identical (that is, and for ). Then, model (1.5) is reduced to the following form:
(4.1) |
Proposition 4.1.
Assume that . Then there exists (depending on ) and a Hopf bifurcation curve:
such that, for each , the positive equilibrium of system (4.1) is locally asymptotically stable when and unstable when . Moreover, system (4.1) undergoes a Hopf bifurcation at when , and satisfies
(4.2) |
where satisfying (1.4) is the corresponding eigenfunction of with respect to eigenvalue .
Now, we show the effect of the coupling matrix on the Hopf bifurcation value. We recall the is line-sum symmetric matrix if for all .
Proposition 4.2.
Denote by (respectively, ) when is line-sum symmetric (respectively, not line-sum symmetric). Then .
Proof.
From Propositions 4.1 and 4.2, we see that the Hopf bifurcation value with line-sum symmetric coupling matrix is smaller than that for the non-line-sum symmetric case, see Fig. 3.

This phenomenon can also be illustrated numerically. We consider model (4.1), and choose the parameters and initial values as follows:
Here we choose two different coupling matrices:
where is line-sum symmetric, and is not line-sum symmetric. Then when is fixed, model (4.1) admits a positive periodic solution for , while the positive equilibrium is stable for , see Fig. 4. This shows that the Hopf bifurcation value of model (4.1) for is smaller than that for .


Finally, we provide some numerical simulations to illustrate the theoretical results obtained in Section 3. Here we consider model (1.5), and choose , , and coupling matrices:
Moreover, let be the variable parameter and choose the other parameters as Table 1.
1 | 2 | 3 | 4 | 5 | |
---|---|---|---|---|---|
1 | 2 | 1 | 0.5 | 1 | |
0.1 | 0.2 | 0.4 | 0.1 | 0.2 | |
0.5 | 1 | 1 | 0.5 | 1 | |
2 | 1 | 2 | 1 | 1 |
We numerically show that model (1.5) can undergoes a Hopf bifurcation, and consequently periodic solutions can arise, see Figs. 5 and 6.




Statements and Declarations
Funding: This study was funded by National Natural Science Foundation of China (grant number 12171117) and Shandong Provincial Natural Science Foundation of China (grant number ZR2020YQ01).
Conflict of Interest: The authors declare that they have no conflict of interest.
Availability of date and materials: All data generated or analysed during this study are included in this published article.
References
- [1] Q. An, C. Wang, and H. Wang. Analysis of a spatial memory model with nonlocal maturation delay and hostile boundary condition. Discrete Contin. Dyn. Syst., 40(10):5845–5868, 2020.
- [2] P. Andresen, M. Bache, E. Mosekilde, G. Dewel, and P. Borckmans. Stationary space-periodic structures with equal diffusion coefficients. Phys. Rev. E, 60(1):297–301, 1999.
- [3] K. J. Brown and F. A. Davidson. Global bifurcation in the Brusselator system. Nonlinear Anal., 24(12):1713–1725, 1995.
- [4] S. Busenberg and W. Huang. Stability and Hopf bifurcation for a population delay model with diffusion effects. J. Differential Equations, 124(1):80–107, 1996.
- [5] S. Chen, Y. Lou, and J. Wei. Hopf bifurcation in a delayed reaction-diffusion-advection population model. J. Differential Equations, 264(8):5333–5359, 2018.
- [6] S. Chen and J. Shi. Stability and Hopf bifurcation in a diffusive logistic population model with nonlocal delay effect. J. Differential Equations, 253(12):3440–3470, 2012.
- [7] Y. Choi, T. Ha, J. Han, S. Kim, and D. S. Lee. Turing instability and dynamic phase transition for the Brusselator model with multiple critical eigenvalues. Discrete Contin. Dyn. Syst., 41(9):4255–4281, 2021.
- [8] M. Duan, L. Chang, and Z. Jin. Turing patterns of an SI epidemic model with cross-diffusion on complex networks. Phys. A, 533:122023, 2019.
- [9] R. J. Field and R. M. Noyes. Oscillations in chemical systems. IV. Limit cycle behavior in a model of a real chemical reaction. J. Chem. Phys., 60(5):1877–1884, 1974.
- [10] G. Guo, J. Wu, and X. Ren. Hopf bifurcation in general Brusselator system with diffusion. Appl. Math. Mech. (English Ed.), 32(9):1177–1186, 2011.
- [11] S. Guo. Stability and bifurcation in a reaction-diffusion model with nonlocal delay effect. J. Differential Equations, 259(4):1409–1448, 2015.
- [12] S. Guo and S. Yan. Hopf bifurcation in a diffusive Lotka-Volterra type system with nonlocal delay effect. J. Differential Equations, 260(1):781–817, 2016.
- [13] R. Hu and Y. Yuan. Spatially nonhomogeneous equilibrium in a reaction-diffusion system with distributed delay. J. Differential Equations, 250(6):2779–2806, 2011.
- [14] Y. Jia, Y. Li, and J. Wu. Coexistence of activator and inhibitor for Brusselator diffusion system in chemical or biochemical reactions. Appl. Math. Lett., 53:33–38, 2016.
- [15] Z. Jin and R. Yuan. Hopf bifurcation in a reaction-diffusion-advection equation with nonlocal delay effect. J. Differential Equations, 271:533–562, 2021.
- [16] T. Kato. Perturbation theory in a finite-dimensional space. Springer Berlin Heidelberg, 1995.
- [17] Y. Kuramoto. Chemical oscillations, waves, and turbulence, volume 19 of Springer Series in Synergetics. Springer Berlin Heidelberg, 1984.
- [18] S. P. Kuznetsov. Noise-induced absolute instability. Math. Comput. Simulation, 58(4-6):435–442, 2002.
- [19] R. Lefever and G. Nicolis. Chemical instabilities and sustained oscillations. J. Theor. Biol., 30(2):267–284, 1971.
- [20] I. Lengyel and I. R. Epstein. Diffusion-induced instability in chemically reacting systems: steady-state multiplicity, oscillation, and chaos. Chaos, 1(1):69–76, 1991.
- [21] B. Li and M. Wang. Diffusion-driven instability and Hopf bifurcation in Brusselator system. Appl. Math. Mech. (English Ed.), 29(6):825–832, 2008.
- [22] Q.-S. Li and J.-C. Shi. Cooperative effects of parameter heterogeneity and coupling on coherence resonance in unidirectional coupled brusselator system. Phys. Lett. A, 360(4):593–598, 2007.
- [23] Y. Li. Hopf bifurcations in general systems of Brusselator type. Nonlinear Anal. Real World Appl., 28:32–47, 2016.
- [24] Z. Li and B. Dai. Stability and Hopf bifurcation analysis in a Lotka-Volterra competition-diffusion-advection model with time delay effect. Nonlinearity, 34(5):3271–3313, 2021.
- [25] M. Liao and Q.-R. Wang. Stability and bifurcation analysis in a diffusive Brusselator-type system. Internat. J. Bifur. Chaos Appl. Sci. Engrg., 26(7):1650119, 11, 2016.
- [26] M. Ma and J. Hu. Bifurcation and stability analysis of steady states to a Brusselator model. Appl. Math. Comput., 236:580–592, 2014.
- [27] R. Peng and M. Wang. Pattern formation in the Brusselator system. J. Math. Anal. Appl., 309(1):151–166, 2005.
- [28] R. Peng and M. Yang. On steady-state solutions of the Brusselator-type system. Nonlinear Anal., 71(3-4):1389–1394, 2009.
- [29] J. Petit, M. Asllani, D. Fanelli, B. Lauwens, and T. Carletti. Pattern formation in a two-component reaction-diffusion system with delayed processes on a network. Phys. A, 462:230–249, 2016.
- [30] I. Prigogine and R. Lefever. Symmetry breaking instabilities in dissipative systems. II. J. Chem. Phys., 48(4):1695–1700, 1968.
- [31] I. Schreiber, M. Holodniok, M. Kubíček, and M. Marek. Periodic and aperiodic regimes in coupled dissipative chemical oscillators. J. Statist. Phys., 43(3-4):489–519, 1986.
- [32] J.-C. Shi and Q.-S. Li. The enhancement and sustainment of coherence resonance in a two-way coupled brusselator system. Phys. D, 225(2):229–234, 2007.
- [33] H. L. Smith. Monotone dynamical systems: an introduction to the theory of competitive and cooperative systems: an introduction to the theory of competitive and cooperative systems, volume 41. American Mathematical Society, 1995.
- [34] Y. Su, J. Wei, and J. Shi. Hopf bifurcations in a reaction-diffusion population model with delay effect. J. Differential Equations, 247(4):1156–1184, 2009.
- [35] C. Tian and S. Ruan. Pattern formation and synchronism in an allelopathic plankton model with delay in a network. SIAM J. Appl. Dyn. Syst., 18(1):531–557, 2019.
- [36] R. W. Wittenberg and P. Holmes. The limited effectiveness of normal forms: a critical review and extension of local bifurcation studies of the Brusselator PDE. Phys. D, 100(1-2):1–40, 1997.
- [37] X.-P. Yan and W.-T. Li. Stability of bifurcating periodic solutions in a delayed reaction-diffusion population model. Nonlinearity, 23(6):1413–1431, 2010.
- [38] X.-P. Yan and W.-T. Li. Stability and Hopf bifurcations for a delayed diffusion system in population dynamics. Discrete Contin. Dyn. Syst. Ser. B, 17(1):367–399, 2012.
- [39] X.-P. Yan, P. Zhang, and C.-H. Zhang. Turing instability and spatially homogeneous Hopf bifurcation in a diffusive Brusselator system. Nonlinear Anal. Model. Control, 25(4):638–657, 2020.
- [40] P. Yu and A. B. Gumel. Bifurcation and stability analyses for a coupled Brusselator model. J. Sound Vibration, 244(5):795–820, 2001.