This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Hopf bifurcation and periodic solutions in a coupled Brusselator model of chemical reactions

Yihuan Sun1,  Shanshan Chen2111Corresponding Author, Email: [email protected]  
1 School of Mathematics, Harbin Institute of Technology,  
  Harbin, Heilongjiang, 150001, P.R.China.  
2 Department of Mathematics, Harbin Institute of Technology,  
  Weihai, Shandong, 264209, P.R.China.  
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:

AX,B+XY+C,  2X+Y3X,XD,A\rightarrow X,\;\;B+X\rightarrow Y+C,\;\;2X+Y\rightarrow 3X,\;\;X\rightarrow D, (1.1)

where AA and BB are the concentrations of the initial substances, XX and YY are the concentrations of the intermediate reactants, and CC and DD are the concentrations of the final products. If the concentrations AA and BB depend only on space, these chemical reactions can be modelled by the following reaction-diffusion model: (see [19, 30])

{Xt=d1ΔX+A(x)(B(x)+1)X+X2Y,xΩ,t>0,Yt=d2ΔY+B(x)XX2Y,xΩ,t>0.\begin{cases}X_{t}=d_{1}\Delta X+A(x)-(B(x)+1)X+X^{2}Y,&x\in\Omega,\;t>0,\\ Y_{t}=d_{2}\Delta Y+B(x)X-X^{2}Y,&x\in\Omega,\;t>0.\end{cases} (1.2)

Here d1,d2>0d_{1},d_{2}>0 are the dispersal rates. If AA and BB 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:

di2x2di2x2cixd_{i}\displaystyle\frac{\partial^{2}}{\partial x^{2}}\to d_{i}\displaystyle\frac{\partial^{2}}{\partial x^{2}}-c_{i}\displaystyle\frac{\partial}{\partial x}

for the case that Ω\Omega is one dimension.

If chemical reactions (1.1) take places in nn-boxes, then model (1.2) takes the following discrete form:

{dxjdt=d1k=1npjkxk+aj(βbj+1)xj+xj2yj,j=1,,n,t>0,dyjdt=d2k=1nqjkyk+βbjxjxj2yj,j=1,,n,t>0,𝒙(0)=𝒙0()𝟎,𝒚(0)=𝒚0()𝟎.\begin{cases}\displaystyle\frac{dx_{j}}{dt}=d_{1}\sum_{k=1}^{n}p_{jk}x_{k}+a_{j}-(\beta b_{j}+1)x_{j}+x_{j}^{2}y_{j},&j=1,\dots,n,\;\;t>0,\\ \displaystyle\frac{dy_{j}}{dt}=d_{2}\sum_{k=1}^{n}q_{jk}y_{k}+\beta b_{j}x_{j}-x_{j}^{2}y_{j},&j=1,\dots,n,\;\;t>0,\\ \bm{x}(0)=\bm{x}_{0}\geq(\not\equiv)\bm{0},\;\bm{y}(0)=\bm{y}_{0}\geq(\not\equiv)\bm{0}.\end{cases} (1.3)

Here n2n\geq 2 is the number of boxes in chemical reactions; 𝒙=(x1,,xn)T\bm{x}=(x_{1},\dots,x_{n})^{T} and 𝒚=(y1,,yn)T\bm{y}=(y_{1},\dots,y_{n})^{T}, where xjx_{j} and yjy_{j} denote the concentrations of XX and YY in box ii at time tt, respectively; the nonlinear term xj2yj{x_{j}^{2}}y_{j} describes the autocatalytic step in box jj; and aj>0{a_{j}}>0 and βbj>0\beta{b_{j}}>0 denote the input concentrations of initial substances AA and BB in box jj, respectively. We remark that β\beta introduced here is the scaling parameter, and we choose it as the bifurcation parameter. Moreover, (pjk)\left(p_{jk}\right) and (qjk)\left(q_{jk}\right) are the coupling matrices, where pjk,qjk(jk)p_{jk},q_{jk}(j\neq k) describe the rates of movement from box kk to box jj for the two reactants, respectively, and pjj(=kjpkj)p_{jj}(=-\sum_{k\neq j}p_{kj}) and qjj(=kjqkj)q_{jj}(=-\sum_{k\neq j}q_{kj}) denote the rates of leaving box jj for j=1,,nj=1,\dots,n.

Model (1.3) was first considered by Prigogine and Lefever in [30] for the case of two boxes (n=2n=2), where the coupling matrices (pjk)\left(p_{jk}\right) and (qjk)\left(q_{jk}\right) were assumed to be symmetric, and the two boxes were identical (that is, a1=a2=aa_{1}=a_{2}=a and b1=b2=1b_{1}=b_{2}=1). In this case, model (1.3) admits a unique space-independent steady state:

xi=a,yi=β/afori=1,2,x_{i}=a,\;\;y_{i}=\beta/a\;\;\text{for}\;\;i=1,2,

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 (pjk)\left(p_{jk}\right) and (qjk)\left(q_{jk}\right) 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:

  1. (𝐀𝟏)(\bf A_{1})

    The coupling matrices P:=(pjk)P:=(p_{jk}) and Q:=(qjk)Q:=(q_{jk}) are irreducible and essentially nonnegative;

  2. (𝐀𝟐)(\bf A_{2})

    d2d1:=θ>0\displaystyle\frac{d_{2}}{d_{1}}:=\theta>0.

Here we remark that real matrices with nonnegative off-diagonal elements are called essentially nonnegative. It follows from (𝐀𝟏)(\bf A_{1}) and Perron-Frobenius theorem that s(P)=s(Q)=0s(P)=s(Q)=0, where s(P)s(P) and s(Q)s(Q) are spectrum bounds of PP and QQ, respectively. Clearly, s(P)s(P) and s(Q)s(Q) are also simple eigenvalues of PP and QQ with strongly positive eigenvectors 𝝃{\bm{\xi}} and 𝜼{\bm{\eta}}, respectively, where

𝝃=(ξ1,,ξn)T𝟎,andj=1nξj=1,𝜼=(η1,,ηn)T𝟎,andj=1nηj=1.\begin{split}&\bm{\xi}=(\xi_{1},\cdots,\xi_{n})^{T}\gg\bm{0},\;\;\text{and}\;\;\sum_{j=1}^{n}\xi_{j}=1,\\ &\bm{\eta}=(\eta_{1},\cdots,\eta_{n})^{T}\gg\bm{0},\;\;\text{and}\;\;\sum_{j=1}^{n}\eta_{j}=1.\\ \end{split} (1.4)

Here (x1,,xn)T𝟎(x_{1},\cdots,x_{n})^{T}\gg\bm{0} represents that xj>0x_{j}>0 for j=1,,nj=1,\dots,n. Assumption (𝐀𝟐)(\bf A_{2}) is a mathematically technical condition, and it means that the dispersal rates of the two reactants are proportional. Then letting t~=d1t\tilde{t}=d_{1}t, denoting λ=1/d1\lambda=1/d_{1}, and dropping the tilde sign, model (1.3) can be transformed to the following equivalent model:

{dxjdt=k=1npjkxk+λ[aj(βbj+1)xj+xj2yj],j=1,,n,t>0,dyjdt=θk=1nqjkyk+λ(βbjxjxj2yj),j=1,,n,t>0,𝒙(0)=𝒙0()𝟎,𝒚(0)=𝒚0()𝟎,\begin{cases}\displaystyle\frac{dx_{j}}{dt}=\sum_{k=1}^{n}p_{jk}x_{k}+\lambda\left[a_{j}-(\beta b_{j}+1)x_{j}+x_{j}^{2}y_{j}\right],&j=1,\dots,n,\;\;t>0,\\ \displaystyle\frac{dy_{j}}{dt}=\theta\sum_{k=1}^{n}q_{jk}y_{k}+\lambda\left(\beta b_{j}x_{j}-x_{j}^{2}y_{j}\right),&j=1,\dots,n,\;\;t>0,\\ \bm{x}(0)=\bm{x}_{0}\geq(\not\equiv)\bm{0},\;\bm{y}(0)=\bm{y}_{0}\geq(\not\equiv)\bm{0},\end{cases} (1.5)

where θ\theta is defined in assumption (𝐀𝟐)(\bf A_{2}).

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 β\beta 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 𝒵\mathcal{Z} to be 𝒵:=𝒵i𝒵={x1+ix2|x1,x2𝒵}\mathcal{Z}_{\mathbb{C}}:=\mathcal{Z}\oplus{\rm i}\mathcal{Z}=\{x_{1}+{\rm i}x_{2}|x_{1},x_{2}\in\mathcal{Z}\}, and define the kernel of a linear operator TT by 𝒩(T)\mathcal{N}(T). For μ\mu\in\mathbb{C}, we define the real part by eμ\mathcal{R}e\mu. For the complex valued space n\mathbb{C}^{n}, we choose the standard inner product 𝒖,𝒗=j=1nu¯jvj\langle\bm{u},\bm{v}\rangle=\sum_{j=1}^{n}\overline{u}_{j}v_{j}, and consequently, the norm is defined by

𝒖2=(j=1n|uj|2)12for𝒖n.\|\bm{u}\|_{2}=\left(\sum_{j=1}^{n}|u_{j}|^{2}\right)^{\frac{1}{2}}\;\;\text{for}\;\;\bm{u}\in\mathbb{C}^{n}.

Moreover, for 𝒙=(x1,,xn)T,𝒚=(y1,,yn)Tn\bm{x}=(x_{1},\dots,x_{n})^{T},\bm{y}=(y_{1},\dots,y_{n})^{T}\in\mathbb{C}^{n}, we denote

(𝒙,𝒚)T=(𝒙𝒚)=(x1,,xn,y1,,yn)T.(\bm{x},\bm{y})^{T}=\left({\begin{array}[]{c}{\bm{x}}\\ {\bm{y}}\end{array}}\right)=(x_{1},\dots,x_{n},y_{1},\dots,y_{n})^{T}.

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 λ\lambda 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

{k=1npjkxk+λ[aj(βbj+1)xj+xj2yj]=0,j=1,,n,θk=1nqjkyk+λ(βbjxjxj2yj)=0,j=1,,n.\begin{cases}\sum_{k=1}^{n}p_{jk}x_{k}+\lambda\left[a_{j}-(\beta b_{j}+1)x_{j}+x_{j}^{2}y_{j}\right]=0,&j=1,\dots,n,\\ \theta\sum_{k=1}^{n}q_{jk}y_{k}+\lambda\left(\beta b_{j}x_{j}-x_{j}^{2}y_{j}\right)=0,&j=1,\dots,n.\end{cases} (2.1)

Note that (𝒙,𝒚)=(c𝝃,d𝜼)(\bm{x},\bm{y})=(c\bm{\xi},d\bm{\eta}) solves (2.1) for all c,dc,d\in\mathbb{R} when λ=0\lambda=0. 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

n=span{𝝃}=span{𝜼},\mathbb{R}^{n}={\rm span}\{\bm{\xi}\}\oplus\mathcal{M}={\rm span}\{\bm{\eta}\}\oplus\mathcal{M},

where 𝝃\bm{\xi} and 𝜼\bm{\eta} are defined in (1.4), and

:={𝒙=(x1,,xn)Tn:j=1nxj=0}.\mathcal{M}:=\left\{\bm{x}=(x_{1},\dots,x_{n})^{T}\in\mathbb{R}^{n}:\sum_{j=1}^{n}x_{j}=0\right\}.

Letting

𝒙=c𝝃+𝒖,c,𝒖,𝒚=r𝜼+𝒗,r,𝒗,\begin{split}&\bm{x}=c\bm{\xi}+\bm{u},\;\;c\in\mathbb{R},\;\bm{u}\in\mathcal{M},\\ &\bm{y}=r\bm{\eta}+\bm{v},\;\;r\in\mathbb{R},\;\bm{v}\in\mathcal{M},\\ \end{split} (2.2)

and plugging (2.2) into (2.1), we see that (𝒙,𝒚)(\bm{x},\bm{y}) (defined in (2.2)) is a solution of (2.1), if and only if (c,r,𝒖,𝒗)2×2(c,r,\bm{u},\bm{v})\in\mathbb{R}^{2}\times\mathcal{M}^{2} solves

𝑭(c,r,𝒖,𝒗,β,λ)=(f1,f21,,f2n,f3,f41,,f4n)T=𝟎,\bm{F}(c,r,\bm{u},\bm{v},\beta,\lambda)=(f_{1},f_{21},\dots,f_{2n},f_{3},f_{41},\dots,f_{4n})^{T}=\bm{0}, (2.3)

where 𝑭(c,r,𝒖,𝒗,β,λ):2×2×2(×)2\bm{F}(c,r,\bm{u},\bm{v},\beta,\lambda):\mathbb{R}^{2}\times\mathcal{M}^{2}\times\mathbb{R}^{2}\to\left(\mathbb{R}\times\mathcal{M}\right)^{2}, and

{f1(c,r,𝒖,𝒗,β,λ):=j=1n[aj(βbj+1)(cξj+uj)+(cξj+uj)2(rηj+vj)],f2j(c,r,𝒖,𝒗,β,λ):=λ[aj(βbj+1)(cξj+uj)+(cξj+uj)2(rηj+vj)]+k=1npjkukλnf1,j=1,,n,f3(c,r,𝒖,𝒗,β,λ):=j=1n[βbj(cξj+uj)(cξj+uj)2(rηj+vj)],f4j(c,r,𝒖,𝒗,β,λ):=λ[βbj(cξj+uj)(cξj+uj)2(rηj+vj)]+θk=1nqjkvkλnf3,j=1,,n.\begin{cases}f_{1}(c,r,\bm{u},\bm{v},\beta,\lambda):=\displaystyle\sum_{j=1}^{n}\left[a_{j}-(\beta b_{j}+1)(c\xi_{j}+u_{j})+(c\xi_{j}+u_{j})^{2}(r\eta_{j}+v_{j})\right],\\ f_{2j}(c,r,\bm{u},\bm{v},\beta,\lambda):=\lambda\left[a_{j}-(\beta b_{j}+1)(c\xi_{j}+u_{j})+(c\xi_{j}+u_{j})^{2}(r\eta_{j}+v_{j})\right]\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\displaystyle+\sum_{k=1}^{n}p_{jk}u_{k}-\displaystyle\frac{\lambda}{n}f_{1},~{}~{}~{}~{}~{}~{}j=1,\dots,n,\\ f_{3}(c,r,\bm{u},\bm{v},\beta,\lambda):=\displaystyle\sum_{j=1}^{n}\left[\beta b_{j}(c\xi_{j}+u_{j})-(c\xi_{j}+u_{j})^{2}(r\eta_{j}+v_{j})\right],\\ f_{4j}(c,r,\bm{u},\bm{v},\beta,\lambda):=\lambda\left[\beta b_{j}(c\xi_{j}+u_{j})-(c\xi_{j}+u_{j})^{2}(r\eta_{j}+v_{j})\right]\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\displaystyle+\theta\sum_{k=1}^{n}q_{jk}v_{k}-\displaystyle\frac{\lambda}{n}f_{3},~{}~{}~{}~{}~{}~{}j=1,\dots,n.\\ \end{cases} (2.4)

We first solve 𝑭(c,r,𝒖,𝒗,β,λ)=𝟎\bm{F}(c,r,\bm{u},\bm{v},\beta,\lambda)=\bm{0} for λ=0\lambda=0.

Lemma 2.1.

Assume that λ=0\lambda=0. For fixed β>0\beta>0, 𝐅(c,r,𝐮,𝐯,β,λ)=𝟎\bm{F}(c,r,\bm{u},\bm{v},\beta,\lambda)=\bm{0} has a unique solution (c0,r0β,𝐮0,𝐯0)2×2(c_{0},r_{0\beta},\bm{u}_{0},\bm{v}_{0})\in\mathbb{R}^{2}\times\mathcal{M}^{2}, where

c0=j=1naj,r0β=βj=1nbjξj(j=1naj)(j=1nξj2ηj),𝒖0=𝟎,𝒗0=𝟎.c_{0}=\sum_{j=1}^{n}a_{j},\;\;r_{0\beta}=\frac{\beta\sum_{j=1}^{n}b_{j}\xi_{j}}{\left(\sum_{j=1}^{n}a_{j}\right)\left(\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}\right)},\;\;\bm{u}_{0}=\bm{0},\;\;\bm{v}_{0}=\bm{0}. (2.5)
Proof.

Plugging λ=0\lambda=0 into f2j=f4j=0f_{2j}=f_{4j}=0 for j=1,,nj=1,\dots,n, we have 𝒖=𝒖0=𝟎\bm{u}=\bm{u}_{0}=\bm{0} and 𝒗=𝒗0=𝟎\bm{v}=\bm{v}_{0}=\bm{0}. Then plugging 𝒖=𝒗=𝟎\bm{u}=\bm{v}=\bm{0} into f1=f3=0f_{1}=f_{3}=0, we have

j=1n[ajc(βbj+1)ξj+c2rξj2ηj]=0,j=1n[cβbjξjc2rξj2ηj]=0,\sum_{j=1}^{n}\left[a_{j}-c(\beta b_{j}+1)\xi_{j}+c^{2}r\xi_{j}^{2}\eta_{j}\right]=0,\;\;\sum_{j=1}^{n}\left[c\beta b_{j}\xi_{j}-c^{2}r\xi_{j}^{2}\eta_{j}\right]=0,

which implies that

c=c0=j=1naj,r=r0β=βj=1nbjξj(j=1naj)(j=1nξj2ηj).c=c_{0}=\sum_{j=1}^{n}a_{j},\;\;r=r_{0\beta}=\frac{\beta\sum_{j=1}^{n}b_{j}\xi_{j}}{\left(\sum_{j=1}^{n}a_{j}\right)\left(\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}\right)}.

This completes the proof. ∎

Now we solve (2.1) (or equivalently, (2.3)) for λ>0\lambda>0.

Theorem 2.2.

For any fixed β1>0\beta_{1}>0, there exists δβ1(0,β1)\delta_{\beta_{1}}\in(0,\beta_{1}) and a continuously differentiable mapping (𝐱(λ,β),𝐲(λ,β)):[0,δβ1]×[β1δβ1,β1+δβ1]n×n\left(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}\right):[0,\delta_{\beta_{1}}]\times[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}]\to\mathbb{R}^{n}\times\mathbb{R}^{n} such that (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) is the unique positive solution of (2.1) for (λ,β)(0,δβ1]×[β1δβ1,β1+δβ1](\lambda,\beta)\in(0,\delta_{\beta_{1}}]\times[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}]. Moreover,

𝒙(λ,β)=c(λ,β)𝝃+𝒖(λ,β),𝒚(λ,β)=r(λ,β)𝜼+𝒗(λ,β),\bm{x}^{(\lambda,\beta)}=c^{(\lambda,\beta)}\bm{\xi}+\bm{u}^{(\lambda,\beta)},\;\;\bm{y}^{(\lambda,\beta)}=r^{(\lambda,\beta)}\bm{\eta}+\bm{v}^{(\lambda,\beta)}, (2.6)

where (c(λ,β),r(λ,β),𝐮(λ,β),𝐯(λ,β))2×2\left(c^{(\lambda,\beta)},r^{(\lambda,\beta)},\bm{u}^{(\lambda,\beta)},\bm{v}^{(\lambda,\beta)}\right)\in\mathbb{R}^{2}\times\mathcal{M}^{2} solves Eq. (2.3) for (λ,β)[0,δβ1]×[β1δβ1,β1+δβ1](\lambda,\beta)\in[0,\delta_{\beta_{1}}]\times[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}], and

(c(0,β),r(0,β),𝒖(0,β),𝒗(0,β))=(c0,r0β,𝟎,𝟎)forβ[β1δβ1,β1+δβ1],\left(c^{(0,\beta)},r^{(0,\beta)},\bm{u}^{(0,\beta)},\bm{v}^{(0,\beta)}\right)=(c_{0},r_{0\beta},\bm{0},\bm{0})\;\;\text{for}\;\;\beta\in[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}], (2.7)

with c0c_{0} and r0βr_{0\beta} defined in Lemma 2.1.

Proof.

We first show the existence. It follows from Lemma 2.1 that

𝑭(c0,r0β1,𝟎,𝟎,β1,0)=𝟎,\bm{F}(c_{0},r_{0\beta_{1}},\bm{0},\bm{0},\beta_{1},0)=\bm{0},

where 𝑭\bm{F} is defined in (2.3). A direct computation implies that the Fréchet derivative of 𝑭\bm{F} with respect to (c,r,𝒖,𝒗)(c,r,\bm{u},\bm{v}) at (c0,r0β1,𝟎,𝟎,β1,0)(c_{0},r_{0\beta_{1}},\bm{0},\bm{0},\beta_{1},0) is as follows:

𝑮(c~,r~,𝒖~,𝒗~)=(g1,g21,,g2n,g3,g41,,g4n)T,\bm{G}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}})=(g_{1},g_{21},\dots,g_{2n},g_{3},g_{41},\dots,g_{4n})^{T},

where c~,r~\tilde{c},\tilde{r}\in\mathbb{R}, 𝒖~,𝒗~\bm{\tilde{u}},\bm{\tilde{v}}\in\mathcal{M}, and

{g1(c~,r~,𝒖~,𝒗~):=j=1n[(2c0r0β1ξjηjβ1bj1)(c~ξj+u~j)+(c0ξj)2(r~ηj+v~j)],g2j(c~,r~,𝒖~,𝒗~):=k=1npjku~k,j=1,,n,g3(c~,r~,𝒖~,𝒗~):=j=1n[(β1bj2c0r0β1ξjηj)(c~ξj+u~j)(c0ξj)2(r~ηj+v~j)],g4j(c~,r~,𝒖~,𝒗~):=θk=1nqjkv~k,j=1,,n.\begin{cases}\displaystyle g_{1}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}}):=\sum_{j=1}^{n}\left[{\left(2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}-\beta_{1}b_{j}-1\right){(\tilde{c}\xi_{j}+\tilde{u}_{j})}+{(c_{0}\xi_{j})}^{2}}{(\tilde{r}\eta_{j}+\tilde{v}_{j})}\right],\\ \displaystyle g_{2j}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}}):=\sum_{k=1}^{n}{p_{jk}\tilde{u}_{k}},~{}~{}~{}~{}~{}~{}j=1,\dots,n,\\ \displaystyle g_{3}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}}):=\sum_{j=1}^{n}\left[{\left(\beta_{1}b_{j}-2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}\right){(\tilde{c}\xi_{j}+\tilde{u}_{j})}-{(c_{0}\xi_{j})}^{2}}{(\tilde{r}\eta_{j}+\tilde{v}_{j})}\right],\\ \displaystyle g_{4j}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}}):=\theta\sum_{k=1}^{n}{q_{jk}\tilde{v}_{k}},~{}~{}~{}~{}~{}~{}j=1,\dots,n.\\ \end{cases}

If 𝑮(c~,r~,𝒖~,𝒗~)=𝟎\bm{G}(\tilde{c},\tilde{r},\bm{\tilde{u}},\bm{\tilde{v}})=\bm{0}, then 𝒖~=𝟎\bm{\tilde{u}}=\bm{0} and 𝒗~=𝟎\bm{\tilde{v}}=\bm{0}. Plugging 𝒖~=𝒗~=𝟎\bm{\tilde{u}}=\bm{\tilde{v}}=\bm{0} into g1=g3=0g_{1}=g_{3}=0, we have

(j=1n(2c0r0β1ξjηjβ1bj1)ξjj=1n(c0ξj)2ηjj=1n(β1bj2c0r0β1ξjηj)ξjj=1n(c0ξj)2ηj)(c~r~)=(00).\left(\begin{array}[]{cc}\sum_{j=1}^{n}\left(2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}-\beta_{1}b_{j}-1\right)\xi_{j}&\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}\\ \sum_{j=1}^{n}\left(\beta_{1}b_{j}-2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}\right)\xi_{j}&-\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}\end{array}\right)\left({\begin{array}[]{*{20}{c}}\tilde{c}\\ \tilde{r}\end{array}}\right)=\left({\begin{array}[]{*{20}{c}}0\\ 0\end{array}}\right).

Noticing that

|j=1n(2c0r0β1ξjηjβ1bj1)ξjj=1n(c0ξj)2ηjj=1n(β1bj2c0r0β1ξjηj)ξjj=1n(c0ξj)2ηj|0,\left|\begin{array}[]{cc}\sum_{j=1}^{n}\left(2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}-\beta_{1}b_{j}-1\right)\xi_{j}&\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}\\ \sum_{j=1}^{n}\left(\beta_{1}b_{j}-2c_{0}r_{0\beta_{1}}\xi_{j}\eta_{j}\right)\xi_{j}&-\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}\end{array}\right|\neq 0,

we obtain that c~=0\tilde{c}=0 and r~=0\tilde{r}=0. Therefore, 𝑮\bm{G} is bijective.

It follows from the implicit function theorem that there exist δβ1(0,β1)\delta_{\beta_{1}}\in(0,\beta_{1}) and a continuously differentiable mapping

(λ,β)[0,δβ1]×[β1δβ1,β1+δβ1](c(λ,β),r(λ,β),𝒖(λ,β),𝒗(λ,β))2×2(\lambda,\beta)\in[0,\delta_{\beta_{1}}]\times[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}]\mapsto\left(c^{(\lambda,\beta)},r^{(\lambda,\beta)},\bm{u}^{(\lambda,\beta)},\bm{v}^{(\lambda,\beta)}\right)\in\mathbb{R}^{2}\times\mathcal{M}^{2}

such that 𝑭(c(λ,β),r(λ,β),𝒖(λ,β),𝒗(λ,β),β,λ)=𝟎\bm{F}\left(c^{(\lambda,\beta)},r^{(\lambda,\beta)},\bm{u}^{(\lambda,\beta)},\bm{v}^{(\lambda,\beta)},\beta,\lambda\right)=\bm{0} and

(c(0,β1),r(0,β1),𝒖(0,β1),𝒗(0,β1))=(c0,r0β1,𝟎,𝟎).\left(c^{(0,\beta_{1})},r^{(0,\beta_{1})},\bm{u}^{(0,\beta_{1})},\bm{v}^{(0,\beta_{1})}\right)=(c_{0},r_{0\beta_{1}},\bm{0},\bm{0}).

Then we can choose δβ1>0\delta_{\beta_{1}}>0 (sufficiently small) such that (𝒙(λ,β),𝒚(λ,β))\left(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}\right) (defined in (2.6)) is a positive solution of (2.1) for (λ,β)(0,δβ1]×[β1δβ1,β1+δβ1](\lambda,\beta)\in(0,\delta_{\beta_{1}}]\times[\beta_{1}-\delta_{\beta_{1}},\beta_{1}+\delta_{\beta_{1}}].

Now, we show the uniqueness. From the implicit function theorem, we only need to verify that if (𝒙~(λ,β),𝒚~(λ,β))\left(\bm{\tilde{x}}^{(\lambda,\beta)},\bm{\tilde{y}}^{(\lambda,\beta)}\right) is a positive solution of (2.1), where

𝒙~(λ,β)=c~(λ,β)𝝃+𝒖~(λ,β),𝒚~(λ,β)=r~(λ,β)𝜼+𝒗~(λ,β),c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β),\bm{\tilde{x}}^{(\lambda,\beta)}=\tilde{c}^{(\lambda,\beta)}\bm{\xi}+\bm{\tilde{u}}^{(\lambda,\beta)},\;\bm{\tilde{y}}^{(\lambda,\beta)}=\tilde{r}^{(\lambda,\beta)}\bm{\eta}+\bm{\tilde{v}}^{(\lambda,\beta)},\;\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)}\in\mathbb{R},\;\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)}\in\mathcal{M},

then (c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β))(c0,r0β1,𝟎,𝟎)\left(\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)},\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)}\right)\to(c_{0},r_{0\beta_{1}},\bm{0},\bm{0}) as (λ,β)(0,β1)(\lambda,\beta)\to(0,\beta_{1}). Substituting (c,r,𝒖,𝒗)=(c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β))(c,r,\bm{u},\bm{v})=\left(\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)},\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)}\right) into (2.3), we see from the first and third equation of (2.4) that

c~(λ,β)=j=1nx~j(λ,β)=j=1naj,\tilde{c}^{(\lambda,\beta)}=\sum_{j=1}^{n}{\tilde{x}_{j}^{(\lambda,\beta)}}=\sum_{j=1}^{n}a_{j},

which implies that 𝒖~(λ,β)\bm{\tilde{u}}^{(\lambda,\beta)} is bounded in n\mathbb{R}^{n}. Then, up to a subsequence, we assume that

lim(λ,β)(0,β1)𝒖~(λ,β)=𝒖.\lim_{(\lambda,\beta)\to(0,\beta_{1})}\bm{\tilde{u}}^{(\lambda,\beta)}=\bm{u}^{*}\in\mathcal{M}.

From the third equation of (2.4), we have

j=1nβbjx~j(λ,β)=j=1n(x~j(λ,β))2y~j(λ,β),\sum_{j=1}^{n}\beta b_{j}\tilde{x}_{j}^{(\lambda,\beta)}=\sum_{j=1}^{n}\left(\tilde{x}_{j}^{(\lambda,\beta)}\right)^{2}\tilde{y}_{j}^{(\lambda,\beta)}, (2.8)

which implies that {(x~j(λ,β))2y~j(λ,β)}j=1n\left\{\left(\tilde{x}_{j}^{(\lambda,\beta)}\right)^{2}\tilde{y}_{j}^{(\lambda,\beta)}\right\}_{j=1}^{n} are bounded. Taking the limit of

f2j(c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β),β,λ)=0forj=1,,n,f_{2j}\left(\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)},\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)},\beta,\lambda\right)=0\;\;\text{for}\;\;j=1,\dots,n,

as (λ,β)(0,β1)(\lambda,\beta)\to(0,\beta_{1}), we see that P𝒖=𝟎P\bm{u}^{*}=\bm{0}, which yields 𝒖=𝟎\bm{u}^{*}=\bm{0}. Therefore,

lim(λ,β)(0,β1)𝒖~(λ,β)=𝟎andlim(λ,β)(0,β1)c~(λ,β)=c0.\lim_{(\lambda,\beta)\to(0,\beta_{1})}\bm{\tilde{u}}^{(\lambda,\beta)}=\bm{0}\;\;\;\;\text{and}\;\;\lim_{(\lambda,\beta)\to(0,\beta_{1})}\tilde{c}^{(\lambda,\beta)}=c_{0}. (2.9)

It follows from (2.8) and (2.9) that 𝒚~(λ,β)\bm{\tilde{y}}^{(\lambda,\beta)} is also bounded in n\mathbb{R}^{n}. Then, up to a subsequence, we assume that

lim(λ,β)(0,β1)r~(λ,β)=r0andlim(λ,β)(0,β1)𝒗~(λ,β)=𝒗.\lim_{(\lambda,\beta)\to(0,\beta_{1})}\tilde{r}^{(\lambda,\beta)}=r^{*}\geq 0\;\;\;\;\text{and}\;\lim_{(\lambda,\beta)\to(0,\beta_{1})}\bm{\tilde{v}}^{(\lambda,\beta)}=\bm{v}^{*}\in\mathcal{M}.

Taking the limit of

f4j(c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β),β,λ)=0forj=1,,n,f_{4j}\left(\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)},\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)},\beta,\lambda\right)=0\;\;\text{for}\;\;j=1,\dots,n,

as (λ,β)(0,β1)(\lambda,\beta)\to(0,\beta_{1}), we see that Q𝒗=𝟎Q\bm{v}^{*}=\bm{0}, which yields 𝒗=𝟎\bm{v}^{*}=\bm{0}. Consequently, taking the limit of

f3(c~(λ,β),r~(λ,β),𝒖~(λ,β),𝒗~(λ,β),β,λ)=0f_{3}\left(\tilde{c}^{(\lambda,\beta)},\tilde{r}^{(\lambda,\beta)},\bm{\tilde{u}}^{(\lambda,\beta)},\bm{\tilde{v}}^{(\lambda,\beta)},\beta,\lambda\right)=0

as (λ,β)(0,β1)(\lambda,\beta)\to(0,\beta_{1}), we have r=r0β1r^{*}=r_{0\beta_{1}}. Therefore,

lim(λ,β)(0,β1)𝒗~(λ,β)=𝟎andlim(λ,β)(0,β1)r~(λ,β)=r0β1.\lim_{(\lambda,\beta)\to(0,\beta_{1})}\bm{\tilde{v}}^{(\lambda,\beta)}=\bm{0}\;\;\;\;\text{and}\;\;\lim_{(\lambda,\beta)\to(0,\beta_{1})}\tilde{r}^{(\lambda,\beta)}=r_{0\beta_{1}}.

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 β\beta is in a small neighborhood of a given positive constant β1\beta_{1}. In the following, we will consider the solution of (2.1) for a wider range of β\beta.

Theorem 2.3.

Let :=[ϵ,1/ϵ]\mathcal{B}:=[\epsilon,1/\epsilon], where ϵ>0\epsilon>0 is sufficiently small. Then there exists δϵ>0\delta_{\epsilon}>0 and a continuously differentiable mapping (𝐱(λ,β),𝐲(λ,β)):[0,δϵ]×n×n\left(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}\right):[0,\delta_{\epsilon}]\times\mathcal{B}\to\mathbb{R}^{n}\times\mathbb{R}^{n} such that (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) is the unique positive solution of (2.1) for (λ,β)(0,δϵ]×(\lambda,\beta)\in(0,\delta_{\epsilon}]\times\mathcal{B}. Moreover,

𝒙(λ,β)=c(λ,β)𝝃+𝒖(λ,β),𝒚(λ,β)=r(λ,β)𝜼+𝒗(λ,β),\bm{x}^{(\lambda,\beta)}=c^{(\lambda,\beta)}\bm{\xi}+\bm{u}^{(\lambda,\beta)},\;\;\bm{y}^{(\lambda,\beta)}=r^{(\lambda,\beta)}\bm{\eta}+\bm{v}^{(\lambda,\beta)}, (2.10)

where (c(λ,β),r(λ,β),𝐮(λ,β),𝐯(λ,β))2×2\left(c^{(\lambda,\beta)},r^{(\lambda,\beta)},\bm{u}^{(\lambda,\beta)},\bm{v}^{(\lambda,\beta)}\right)\in\mathbb{R}^{2}\times\mathcal{M}^{2} solves Eq. (2.3) for (λ,β)[0,δϵ]×(\lambda,\beta)\in[0,\delta_{\epsilon}]\times\mathcal{B}, and

(c(0,β),r(0,β),𝒖(0,β),𝒗(0,β))=(c0,r0β,𝟎,𝟎)forβ,\left(c^{(0,\beta)},r^{(0,\beta)},\bm{u}^{(0,\beta)},\bm{v}^{(0,\beta)}\right)=(c_{0},r_{0\beta},\bm{0},\bm{0})\;\;\text{for}\;\;\beta\in\mathcal{B}, (2.11)

with c0c_{0} and r0βr_{0\beta} defined in Lemma 2.1.

Proof.

It follows from Theorem 2.2, for any β~\tilde{\beta}\in\mathcal{B}, there exists δβ~(0,β~)\delta_{\tilde{\beta}}\in(0,\tilde{\beta}) such that, for (λ,β)(0,δβ~]×[β~δβ~,β~+δβ~](\lambda,\beta)\in(0,\delta_{\tilde{\beta}}]\times[\tilde{\beta}-\delta_{\tilde{\beta}},\tilde{\beta}+\delta_{\tilde{\beta}}], (2.1) admits a unique positive solution (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}), where 𝒙(λ,β)\bm{x}^{(\lambda,\beta)} and 𝒚(λ,β)\bm{y}^{(\lambda,\beta)} are defined in (2.6) and continuously differentiable for (λ,β)[0,δβ~]×[β~δβ~,β~+δβ~](\lambda,\beta)\in[0,\delta_{\tilde{\beta}}]\times[\tilde{\beta}-\delta_{\tilde{\beta}},\tilde{\beta}+\delta_{\tilde{\beta}}]. Clearly,

β~(β~δβ~,β~+δβ~).\mathcal{B}\subseteq\bigcup_{\tilde{\beta}\in\mathcal{B}}{\left({\tilde{\beta}-\delta_{\tilde{\beta}},\tilde{\beta}+\delta_{\tilde{\beta}}}\right)}.

Noticing that \mathcal{B} is compact, we see that there exist finite open intervals (see Fig. 1), denoted by (β~lδβ~l,β~l+δβ~l)\left({\tilde{\beta}_{l}}-\delta_{\tilde{\beta}_{l}},{\tilde{\beta}_{l}}+\delta_{\tilde{\beta}_{l}}\right) for l=1,,sl=1,\dots,s, such that

l=1s(β~lδβ~l,β~l+δβ~l).\mathcal{B}\subseteq\bigcup_{l=1}^{s}{\left({\tilde{\beta}_{l}}-\delta_{\tilde{\beta}_{l}},{\tilde{\beta}_{l}}+\delta_{\tilde{\beta}_{l}}\right)}.

Choose δϵ=min1lsδβ~l\delta_{\epsilon}=\min_{1\leq l\leq s}\delta_{\tilde{\beta}_{l}}. Then, for (λ,β)(0,δϵ]×(\lambda,\beta)\in(0,\delta_{\epsilon}]\times\mathcal{B}, (2.1) admits a unique positive solution (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}), which is defined in (2.10) and continuously differentiable for (λ,β)[0,δϵ]×(\lambda,\beta)\in[0,\delta_{\epsilon}]\times\mathcal{B}. Eq. (2.11) can be proved by using the similar arguments as in the proof of Theorem 2.2, and here we omit the proof. ∎

Refer to caption
Figure 1: Finite open intervals covering \mathcal{B}.

3 Stability and Hopf bifurcation

Throughout this section, we assume that β\beta\in\mathcal{B}, where \mathcal{B} is defined in Theorem 2.3. It follows from Theorem 2.3 that there exists δϵ>0\delta_{\epsilon}>0 such that, for (λ,β)(0,δϵ]×(\lambda,\beta)\in(0,\delta_{\epsilon}]\times\mathcal{B}, (1.5) admits a unique positive equilibrium (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}), where 𝒙(λ,β)\bm{x}^{(\lambda,\beta)} and 𝒚(λ,β)\bm{y}^{(\lambda,\beta)} are defined in (2.10). Here we use β\beta as the bifurcation parameter, and we will show that there exists a Hopf bifurcation curve β=βλ\beta=\beta_{\lambda} when λ\lambda is small, see Fig. 2.

Linearizing model (1.5) at (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}), we obtain

{dx~jdt=k=1npjkx~k+λ[M1j(λ,β)x~j+M2j(λ,β)y~j],j=1,,n,t>0,dy~jdt=θk=1nqjky~k+λ[M3j(λ,β)x~jM2j(λ,β)y~j],j=1,,n,t>0,\begin{cases}\displaystyle\frac{d\tilde{x}_{j}}{dt}=\sum_{k=1}^{n}p_{jk}\tilde{x}_{k}+\lambda\left[M_{1j}^{(\lambda,\beta)}\tilde{x}_{j}+M_{2j}^{(\lambda,\beta)}\tilde{y}_{j}\right],~{}~{}~{}~{}j=1,\dots,n,\;t>0,\\ \displaystyle\frac{d\tilde{y}_{j}}{dt}=\theta\sum_{k=1}^{n}q_{jk}\tilde{y}_{k}+\lambda\left[M_{3j}^{(\lambda,\beta)}\tilde{x}_{j}-M_{2j}^{(\lambda,\beta)}\tilde{y}_{j}\right],~{}~{}~{}~{}j=1,\dots,n,\;t>0,\\ \end{cases}

where

M1j(λ,β)=2xj(λ,β)yj(λ,β)βbj1,M2j(λ,β)=(xj(λ,β))2,M3j(λ,β)=βbj2xj(λ,β)yj(λ,β).M_{1j}^{(\lambda,\beta)}=2x_{j}^{(\lambda,\beta)}y_{j}^{(\lambda,\beta)}-\beta b_{j}-1,\;\;M_{2j}^{(\lambda,\beta)}={\left(x_{j}^{(\lambda,\beta)}\right)}^{2},\;\;M_{3j}^{(\lambda,\beta)}=\beta b_{j}-2x_{j}^{(\lambda,\beta)}y_{j}^{(\lambda,\beta)}. (3.1)

Let

Aβ(λ):=(P00θQ)+λ(M1M2M3M2),A_{\beta}(\lambda):=\left({\begin{array}[]{*{20}{c}}P&0\\ 0&\theta Q\end{array}}\right)+\lambda\left({\begin{array}[]{*{20}{c}}M_{1}&M_{2}\\ M_{3}&-M_{2}\end{array}}\right), (3.2)

where M1=diag(M1j(λ,β)),M2=diag(M2j(λ,β))M_{1}={\rm diag}\left(M_{1j}^{(\lambda,\beta)}\right),M_{2}={\rm diag}\left(M_{2j}^{(\lambda,\beta)}\right) and M3=diag(M3j(λ,β))M_{3}={\rm diag}\left(M_{3j}^{(\lambda,\beta)}\right). Then, μ\mu\in\mathbb{C} is an eigenvalue of Aβ(λ)A_{\beta}(\lambda) if there exists (𝝋,𝝍)T(𝟎)2n(\bm{\varphi},\bm{\psi})^{T}(\neq\bm{0})\in\mathbb{C}^{2n} such that

{μφj=k=1npjkφk+λ[M1j(λ,β)φj+M2j(λ,β)ψj],j=1,,n,μψj=θk=1nqjkψk+λ[M3j(λ,β)φjM2j(λ,β)ψj],j=1,,n.\begin{cases}\displaystyle\mu\varphi_{j}=\sum_{k=1}^{n}p_{jk}\varphi_{k}+\lambda\left[M_{1j}^{(\lambda,\beta)}\varphi_{j}+M_{2j}^{(\lambda,\beta)}\psi_{j}\right],&j=1,\dots,n,\\ \displaystyle\mu\psi_{j}=\theta\sum_{k=1}^{n}q_{jk}\psi_{k}+\lambda\left[M_{3j}^{(\lambda,\beta)}\varphi_{j}-M_{2j}^{(\lambda,\beta)}\psi_{j}\right],&j=1,\dots,n.\\ \end{cases} (3.3)

For further application, we first give a priori estimates for solutions of eigenvalue problem (3.3).

Lemma 3.1.

Let \mathcal{B} and δϵ\delta_{\epsilon} are defined in Theorem 2.3. Assume that, for λ(0,δϵ]\lambda\in(0,\delta_{\epsilon}], (μλ,βλ,𝛗λ,𝛙λ)(\mu_{\lambda},\beta_{\lambda},\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda}) solves (3.3) with eμλ0\mathcal{R}e\mu_{\lambda}\geq 0, (𝛗λ,𝛙λ)T(𝟎)2n(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}(\neq\bm{0})\in\mathbb{C}^{2n} and βλ\beta_{\lambda}\in\mathcal{B}. Then there exists λ1(0,δϵ)\lambda_{1}\in(0,\delta_{\epsilon}) such that |μλ/λ|\left|\mu_{\lambda}/\lambda\right| is bounded for λ(0,λ1]\lambda\in(0,{\lambda}_{1}].

Proof.

Ignoring a scalar factor, we assume that 𝝋λ22+𝝍λ22=𝝃22+𝜼22\|\bm{\varphi}_{\lambda}\|_{2}^{2}+\|\bm{\psi}_{\lambda}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}, where 𝝃\bm{\xi} and 𝜼\bm{\eta} are defined in (1.4). Substituting (μ,β,𝝋,𝝍)=(μλ,βλ,𝝋λ,𝝍λ)(\mu,\beta,\bm{\varphi},\bm{\psi})=(\mu_{\lambda},\beta_{\lambda},\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda}) into (3.3), we have

{μλφλj=k=1npjkφλk+λ[M1j(λ,βλ)φλj+M2j(λ,βλ)ψλj],j=1,,n,μλψλj=θk=1nqjkψλk+λ[M3j(λ,βλ)φλjM2j(λ,βλ)ψλj],j=1,,n,\begin{cases}\displaystyle\mu_{\lambda}\varphi_{\lambda j}=\sum_{k=1}^{n}p_{jk}\varphi_{\lambda k}+\lambda\left[M_{1j}^{(\lambda,\beta_{\lambda})}\varphi_{\lambda j}+M_{2j}^{(\lambda,\beta_{\lambda})}\psi_{\lambda j}\right],&j=1,\dots,n,\\ \displaystyle\mu_{\lambda}\psi_{\lambda j}=\theta\sum_{k=1}^{n}q_{jk}\psi_{\lambda k}+\lambda\left[M_{3j}^{(\lambda,\beta_{\lambda})}\varphi_{\lambda j}-M_{2j}^{(\lambda,\beta_{\lambda})}\psi_{\lambda j}\right],&j=1,\dots,n,\\ \end{cases} (3.4)

where Mlj(λ,β)M_{lj}^{(\lambda,\beta)} are defined in (3.1) for l=1,2,3l=1,2,3. Multiplying the first and second equation of (3.4) by φ¯λj\overline{\varphi}_{\lambda j} and ψ¯λj\overline{\psi}_{\lambda j}, respectively, and summing these equations over all jj yield

μλ(𝝋λ22+𝝍λ22)=j=1nk=1npjkφ¯λjφλk+j=1nk=1nqjkψ¯λjψλk+λj=1n(M1j(λ,βλ)φλj+M2j(λ,βλ)ψλj)φ¯λj+λj=1n(M3j(λ,βλ)φλjM2j(λ,βλ)ψλj)ψ¯λj.\begin{split}&\mu_{\lambda}\left(\|\bm{\varphi}_{\lambda}\|_{2}^{2}+\|\bm{\psi}_{\lambda}\|_{2}^{2}\right)\\ =&\sum_{j=1}^{n}\sum_{k=1}^{n}p_{jk}\overline{\varphi}_{\lambda j}\varphi_{\lambda k}+\sum_{j=1}^{n}\sum_{k=1}^{n}q_{jk}\overline{\psi}_{\lambda j}\psi_{\lambda k}\\ &+\lambda\sum_{j=1}^{n}\left(M_{1j}^{(\lambda,\beta_{\lambda})}\varphi_{\lambda j}+M_{2j}^{(\lambda,\beta_{\lambda})}\psi_{\lambda j}\right)\overline{\varphi}_{\lambda j}+\lambda\sum_{j=1}^{n}\left(M_{3j}^{(\lambda,\beta_{\lambda})}\varphi_{\lambda j}-M_{2j}^{(\lambda,\beta_{\lambda})}\psi_{\lambda j}\right)\overline{\psi}_{\lambda j}.\end{split} (3.5)

Note that there exists a positive constant MM^{*} such that

|M1j(λ,βλ)|,|M2j(λ,βλ)|,|M3j(λ,βλ)|Mforλ[0,δϵ]andj=1,,n.\left|M_{1j}^{(\lambda,\beta_{\lambda})}\right|,\;\left|M_{2j}^{(\lambda,\beta_{\lambda})}\right|,\;\left|M_{3j}^{(\lambda,\beta_{\lambda})}\right|\leq M^{*}\;\;\text{for}\;\;\lambda\in[0,\delta_{\epsilon}]\;\;\text{and}\;\;j=1,\dots,n. (3.6)

This, combined with (3.5), implies that μλ\mu_{\lambda} is bounded for λ[0,δϵ]\lambda\in[0,\delta_{\epsilon}].

Now we claim that there exists λ1(0,δϵ)\lambda_{1}\in(0,\delta_{\epsilon}) such that |μλ/λ|\left|\mu_{\lambda}/\lambda\right| is bounded for λ(0,λ1]\lambda\in(0,{\lambda}_{1}], where δϵ\delta_{\epsilon} is defined in Theorem 2.3. Note that 𝝋λ22+𝝍λ22=𝝃22+𝜼22\|\bm{\varphi}_{\lambda}\|_{2}^{2}+\|\bm{\psi}_{\lambda}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}, eμλ0\mathcal{R}e\mu_{\lambda}\geq 0, and μλ\mu_{\lambda} is bounded. If the claim is not true, then there exists a sequence {λl}l=1\{\lambda_{l}\}_{l=1}^{\infty} such that limlλl=0\lim_{l\to\infty}\lambda_{l}=0, liml|μλl/λl|=\lim_{l\to\infty}|\mu_{\lambda_{l}}/\lambda_{l}|=\infty, limlμλl=γ\lim_{l\to\infty}\mu_{\lambda_{l}}=\gamma with eγ0\mathcal{R}e\gamma\geq 0, and liml𝝋λl=𝝋\lim_{l\to\infty}\bm{\varphi}_{\lambda_{l}}=\bm{\varphi}_{*} and liml𝝍λl=𝝍\lim_{l\to\infty}\bm{\psi}_{\lambda_{l}}=\bm{\psi}_{*} with 𝝋22+𝝍22=𝝃22+𝜼22\|\bm{\varphi}_{*}\|_{2}^{2}+\|\bm{\psi}_{*}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}. Substituting λ=λl\lambda=\lambda_{l} into (3.4) and taking the limits of the two equations as ll\to\infty, we have

P𝝋γ𝝋=𝟎andQ𝝍γ𝝍=𝟎,P\bm{\varphi}_{*}-\gamma\bm{\varphi}_{*}=\bm{0}\;\;\text{and}\;\;Q\bm{\psi}_{*}-\gamma\bm{\psi}_{*}=\bm{0}, (3.7)

where PP and QQ are defined in assumption (𝐀𝟏)(\bf A_{1}). Without loss of generality, we assume that 𝝋𝟎\bm{\varphi}_{*}\neq\bm{0}. Then it follows from (3.7) that γ\gamma is an eigenvalue of PP, which implies that γ=s(P)\gamma=s(P) or eγ<s(P)\mathcal{R}e\gamma<s(P) from [33, Corollary 4.3.2]. Since s(P)=0s(P)=0 and eγ0\mathcal{R}e\gamma\geq 0, we have γ=0\gamma=0. Consequently, 𝝋=κ1𝝃\bm{\varphi}_{*}=\kappa_{1}\bm{\xi} and 𝝍=κ2𝜼\bm{\psi}_{*}=\kappa_{2}\bm{\eta}, where κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{C}, and κ10\kappa_{1}\neq 0. Summing the first equation of (3.4) over all jj, we have

μλj=1nφλj=j=1nλ[M1j(λ,βλ)φλj+M2j(λ,βλ)ψλj].\mu_{\lambda}\sum_{j=1}^{n}\varphi_{\lambda j}=\sum_{j=1}^{n}\lambda\left[M_{1j}^{(\lambda,\beta_{\lambda})}\varphi_{\lambda j}+M_{2j}^{(\lambda,\beta_{\lambda})}\psi_{\lambda j}\right].

This, combined with (3.6), implies that

lim supl|μλlλl|M|κ1|+|κ2||κ1|,\limsup_{l\to\infty}\left|\frac{\mu_{\lambda_{l}}}{\lambda_{l}}\right|\leq M^{*}\displaystyle\frac{|\kappa_{1}|+|\kappa_{2}|}{|\kappa_{1}|},

which contradicts liml|μλl/λl|=\lim_{l\to\infty}|\mu_{\lambda_{l}}/\lambda_{l}|=\infty. Therefore, the claim is true. ∎

To analyze the stability of (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}), 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 μ=iλν\mu={\rm i}\lambda\nu is an eigenvalue of (3.3), then ν\nu is bounded for λ(0,λ1]\lambda\in(0,\lambda_{1}], where λ1\lambda_{1} is defined in Lemma 3.1. Substituting μ=iλν(ν0)\mu={\rm i}\lambda\nu\;(\nu\geq 0) into (3.3), we have

{iλνφj=k=1npjkφk+λ[M1j(λ,β)φj+M2j(λ,β)ψj],j=1,,n,iλνψj=θk=1nqjkψk+λ[M3j(λ,β)φjM2j(λ,β)ψj],j=1,,n.\begin{cases}\displaystyle{\rm i}\lambda\nu\varphi_{j}=\sum_{k=1}^{n}p_{jk}\varphi_{k}+\lambda\left[M_{1j}^{(\lambda,\beta)}\varphi_{j}+M_{2j}^{(\lambda,\beta)}\psi_{j}\right],&j=1,\dots,n,\\ \displaystyle{\rm i}\lambda\nu\psi_{j}=\theta\sum_{k=1}^{n}q_{jk}\psi_{k}+\lambda\left[M_{3j}^{(\lambda,\beta)}\varphi_{j}-M_{2j}^{(\lambda,\beta)}\psi_{j}\right],&j=1,\dots,n.\\ \end{cases} (3.8)

Ignoring a scalar factor, (𝝋,𝝍)T2n(\bm{\varphi},\bm{\psi})^{T}\in\mathbb{C}^{2n} in (3.8) can be represented as

{𝝋=(φ1,,φn)T=δ𝝃+𝒘,whereδ0and𝒘,𝝍=(ψ1,,ψn)T=(s1+is2)𝜼+𝒛,wheres1,s2and𝒛,𝝋22+𝝍22=𝝃22+𝜼22.\begin{cases}\bm{\varphi}={\left(\varphi_{1},\cdots,\varphi_{n}\right)}^{T}=\delta\bm{\xi}+\bm{w},\;\;\text{where}\;\delta\geq 0\;\;\text{and}\;\;\bm{w}\in\mathcal{M}_{\mathbb{C}},\\ \bm{\psi}={\left(\psi_{1},\cdots,\psi_{n}\right)}^{T}=(s_{1}+{\rm i}s_{2})\bm{\eta}+\bm{z},\;\;\text{where}\;s_{1},s_{2}\in\mathbb{R}\;\;\text{and}\;\;\bm{z}\in\mathcal{M}_{\mathbb{C}},\\ \|\bm{\varphi}\|_{2}^{2}+\|\bm{\psi}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}.\end{cases} (3.9)

Then we see that (𝝋,𝝍,ν,β)(\bm{\varphi},\bm{\psi},\nu,\beta) is a solution of (3.8), where ν0\nu\geq 0, β\beta\in\mathcal{B}, and (𝝋,𝝍)(\bm{\varphi},\bm{\psi}) is defined in (3.9), if and only if

{𝑯(δ,s1,s2,𝒘,𝒛,ν,β,λ)=𝟎δ0,s1,s2,β,ν0,𝒘,𝒛\begin{cases}\bm{H}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda)=\bm{0}\\ \delta\geq 0,\;s_{1},s_{2}\in\mathbb{R},\;\beta\in\mathcal{B},\;\nu\geq 0,\;\bm{w},\bm{z}\in\mathcal{M}_{\mathbb{C}}\end{cases} (3.10)

is solvable for some value of (δ,s1,s2,𝒘,𝒛,ν,β)(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta). Here \mathcal{B} is defined in Lemma 2.3, and 𝑯(δ,s1,s2,𝒘,𝒛,ν,β,λ):3×()2×××[0,λ1](×)2×\bm{H}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda):\mathbb{R}^{3}\times\left(\mathcal{M}_{\mathbb{C}}\right)^{2}\times\mathbb{R}\times\mathcal{B}\times[0,\lambda_{1}]\to\left(\mathbb{C}\times\mathcal{M}_{\mathbb{C}}\right)^{2}\times\mathbb{R} is defined by

𝑯(δ,s1,s2,𝒘,𝒛,ν,β,λ)=(h1,h21,h2n,h3,h41,,h4n,h5)T,\bm{H}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda)=(h_{1},h_{21},\dots h_{2n},h_{3},h_{41},\dots,h_{4n},h_{5})^{T},

where

h1(δ,s1,s2,𝒘,𝒛,ν,β,λ):=j=1n[M1j(λ,β)(δξj+wj)+M2j(λ,β)[(s1+is2)ηj+zj]]iνδ,h2j(δ,s1,s2,𝒘,𝒛,h,β,λ):=k=1npjkwk+λ[M1j(λ,β)(δξj+wj)+M2j(λ,β)[(s1+is2)ηj+zj]]iλν(δξj+wj)λnh1,j=1,,n,h3(δ,s1,s2,𝒘,𝒛,ν,β,λ):=j=1n[M3j(λ,β)(δξj+wj)M2j(λ,β)[(s1+is2)ηj+zj]]iν(s1+is2),h4j(δ,s1,s2,𝒘,𝒛,ν,β,λ):=θk=1nqjkzk+λ[M3j(λ,β)(δξj+wj)M2j(λ,β)[(s1+is2)ηj+zj]]iλν[(s1+is2)ηj+zj]λnh3,j=1,,n,h5(δ,s1,s2,𝒘,𝒛,ν,β,λ):=(δ21)𝝃22+δj=1nξj(wj+w¯j)+(s12+s221)𝜼22+𝒘22+j=1nηj[s1(zj+z¯j)+is2(z¯jzj)]+𝒛22,\begin{split}h_{1}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda):=&\sum_{j=1}^{n}\left[M_{1j}^{(\lambda,\beta)}(\delta\xi_{j}+w_{j})+M_{2j}^{(\lambda,\beta)}[(s_{1}+{\rm i}s_{2})\eta_{j}+z_{j}]\right]-{\rm i}\nu\delta,\\ h_{2j}(\delta,s_{1},s_{2},\bm{w},\bm{z},h,\beta,\lambda):=&\sum_{k=1}^{n}p_{jk}w_{k}+\lambda\left[M_{1j}^{(\lambda,\beta)}(\delta\xi_{j}+w_{j})+M_{2j}^{(\lambda,\beta)}[(s_{1}+{\rm i}s_{2})\eta_{j}+z_{j}]\right]\\ &-{\rm i}\lambda\nu(\delta\xi_{j}+w_{j})-\frac{\lambda}{n}h_{1},\;\;\;\;j=1,\dots,n,\\ h_{3}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda):=&\sum_{j=1}^{n}\left[M_{3j}^{(\lambda,\beta)}(\delta\xi_{j}+w_{j})-M_{2j}^{(\lambda,\beta)}[(s_{1}+{\rm i}s_{2})\eta_{j}+z_{j}]\right]\\ &-{\rm i}\nu(s_{1}+{\rm i}s_{2}),\\ h_{4j}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda):=&\theta\sum_{k=1}^{n}q_{jk}z_{k}+\lambda\left[M_{3j}^{(\lambda,\beta)}(\delta\xi_{j}+w_{j})-M_{2j}^{(\lambda,\beta)}[(s_{1}+{\rm i}s_{2})\eta_{j}+z_{j}]\right]\\ &-{\rm i}\lambda\nu[(s_{1}+{\rm i}s_{2})\eta_{j}+z_{j}]-\frac{\lambda}{n}h_{3},\;\;\;\;j=1,\dots,n,\\ h_{5}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda):=&\left(\delta^{2}-1\right)\|\bm{\xi}\|_{2}^{2}+\delta\sum_{j=1}^{n}\xi_{j}(w_{j}+\overline{w}_{j})+\left(s_{1}^{2}+s_{2}^{2}-1\right)\|\bm{\eta}\|_{2}^{2}\\ &+\|\bm{w}\|_{2}^{2}+\sum_{j=1}^{n}\eta_{j}\left[s_{1}(z_{j}+\overline{z}_{j})+{\rm i}s_{2}(\overline{z}_{j}-z_{j})\right]+\|\bm{z}\|_{2}^{2},\end{split}

and M1j(λ,β),M2j(λ,β),M3j(λ,β)M_{1j}^{(\lambda,\beta)},M_{2j}^{(\lambda,\beta)},M_{3j}^{(\lambda,\beta)} are defined in (3.1).

We first solve (3.10) for λ=0\lambda=0.

Lemma 3.2.

Assume that λ=0\lambda=0. Then (3.10) admits a unique solution

(δ,s1,s2,𝒘,𝒛,ν,β)=(δ0,s10,s20,𝒘0,𝒛0,ν0,β0),(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta)=(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0}),

where

ν0=(j=1naj)j=1nξj2ηj,β0=ν02+1j=1nbjξj,𝒘0=𝟎,𝒛0=𝟎,δ0=𝝃22+𝜼22𝝃22+(1+1ν02)𝜼22,s10=δ0,s20=δ0ν0.\begin{split}&\nu_{0}=\left(\sum_{j=1}^{n}{a_{j}}\right)\sqrt{\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}},\;\beta_{0}=\frac{\nu_{0}^{2}+1}{\sum_{j=1}^{n}b_{j}\xi_{j}},\;\bm{w}_{0}=\bm{0},\;\bm{z}_{0}=\bm{0},\\ &\delta_{0}=\sqrt{\frac{\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}}{\|\bm{\xi}\|_{2}^{2}+\left(1+\frac{1}{\nu^{2}_{0}}\right)\|\bm{\eta}\|_{2}^{2}}},\;s_{10}=-\delta_{0},\;s_{20}=\frac{\delta_{0}}{\nu_{0}}.\\ \end{split}
Proof.

Substituting λ=0\lambda=0 into h2j=h4j=0h_{2j}=h_{4j}=0 for j=1,,nj=1,\dots,n, we have 𝒘=𝒘0=𝟎\bm{w}=\bm{w}_{0}=\bm{0} and 𝒛=𝒛0=𝟎\bm{z}=\bm{z}_{0}=\bm{0}. Then plugging 𝒘=𝒛=𝟎\bm{w}=\bm{z}=\bm{0} and λ=0\lambda=0 into h1=h3=0h_{1}=h_{3}=0, we have

(j=1nM1j(0,β)ξjiνj=1nM2j(0,β)ηjj=1nM3j(0,β)ξjj=1nM2j(0,β)ηjiν)(δs1+is2)=(00),\left({\begin{array}[]{*{20}{c}}\sum_{j=1}^{n}M_{1j}^{(0,\beta)}\xi_{j}-{\rm i}\nu&\sum_{j=1}^{n}M_{2j}^{(0,\beta)}\eta_{j}\\ \sum_{j=1}^{n}M_{3j}^{(0,\beta)}\xi_{j}&-\sum_{j=1}^{n}M_{2j}^{(0,\beta)}\eta_{j}-{\rm i}\nu\end{array}}\right)\left({\begin{array}[]{*{20}{c}}\delta\\ s_{1}+{\rm i}s_{2}\end{array}}\right)=\left({\begin{array}[]{*{20}{c}}0\\ 0\end{array}}\right), (3.11)

where

M1j(0,β)=2c0r0βξjηjβbj1,M2j(0,β)=c02ξj2,M3j(0,β)=βbj2c0r0βξjηj,M_{1j}^{(0,\beta)}=2c_{0}r_{0\beta}\xi_{j}\eta_{j}-\beta b_{j}-1,\;\;M_{2j}^{(0,\beta)}=c_{0}^{2}\xi_{j}^{2},\;\;M_{3j}^{(0,\beta)}=\beta b_{j}-2c_{0}r_{0\beta}\xi_{j}\eta_{j}, (3.12)

and c0c_{0} and r0βr_{0\beta} are defined in (2.5). Then we see from (2.5) that

j=1nM1j(0,β)ξj=βj=1nbjξj1,j=1nM2j(0,β)ηj=(j=1naj)2j=1nξj2ηj,j=1nM3j(0,β)ξj=βj=1nbjξj.\begin{split}&\sum_{j=1}^{n}M_{1j}^{(0,\beta)}\xi_{j}=\beta\sum_{j=1}^{n}b_{j}\xi_{j}-1,\;\;\sum_{j=1}^{n}M_{2j}^{(0,\beta)}\eta_{j}=\left(\sum_{j=1}^{n}a_{j}\right)^{2}\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j},\\ &\sum_{j=1}^{n}M_{3j}^{(0,\beta)}\xi_{j}=-\beta\sum_{j=1}^{n}b_{j}\xi_{j}.\end{split} (3.13)

Then it follows from (3.11) that

β=β0=(j=1naj)2j=1nξj2ηj+1j=1nbjξj,ν=ν0=(j=1naj)j=1nξj2ηj,s1=s10=δ,s2=s20=δν.\begin{split}&\beta=\beta_{0}=\frac{{\left({\sum_{j=1}^{n}a_{j}}\right)}^{2}\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}+1}{\sum_{j=1}^{n}b_{j}\xi_{j}},\;\nu=\nu_{0}=\left(\sum_{j=1}^{n}a_{j}\right)\sqrt{\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}},\\ &s_{1}=s_{10}=-\delta,\;s_{2}=s_{20}=\frac{\delta}{\nu}.\end{split} (3.14)

Again, plugging 𝒘=𝒛=𝟎\bm{w}=\bm{z}=\bm{0} and λ=0\lambda=0 into h5=0h_{5}=0, we have

(δ21)𝝃22+(s12+s221)𝜼22=0.\left(\delta^{2}-1\right)\|\bm{\xi}\|_{2}^{2}+\left(s_{1}^{2}+s_{2}^{2}-1\right)\|\bm{\eta}\|_{2}^{2}=0.

This, combined with (3.14), implies that

δ=δ0=𝝃22+𝜼22𝝃22+(1+1ν02)𝜼22.\delta=\delta_{0}=\sqrt{\frac{\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2}}{\|\bm{\xi}\|_{2}^{2}+\left(1+\frac{1}{\nu^{2}_{0}}\right)\|\bm{\eta}\|_{2}^{2}}}.

Now, we solve (3.10) for λ>0\lambda>0.

Theorem 3.3.

There exists λ2(0,λ1)\lambda_{2}\in(0,\lambda_{1}), where λ1\lambda_{1} is obtained in Lemma 3.1, and a continuously differentiable mapping

λ(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ):[0,λ2](0,)×2×()2×(0,+)×\lambda\mapsto(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}):[0,\lambda_{2}]\to(0,\infty)\times\mathbb{R}^{2}\times\left(\mathcal{M}_{\mathbb{C}}\right)^{2}\times(0,+\infty)\times\mathcal{B}

such that (3.10) has a unique solution (δλ,s1λ,s2λ,𝐰λ,𝐳λ,νλ,βλ)(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}) for λ[0,λ2]\lambda\in[0,\lambda_{2}]. Moreover, for λ=0\lambda=0,

(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ)=(δ0,s10,s20,𝒘0,𝒛0,ν0,β0),\left(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}\right)=(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0}),

where (δ0,s10,s20,𝐰0,𝐳0,ν0,β0)(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0}) is defined in Lemma 3.2.

Proof.

We first show the existence. It follows from Lemma 3.2 that 𝑯(K0)=𝟎\bm{H}\left(K_{0}\right)=\bm{0}, where K0=(δ0,s10,s20,𝒘0,𝒛0,ν0,β0,0)K_{0}=(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0},0). A direct computation implies that the Fréchet derivative of 𝑯(δ,s1,s2,𝒘,𝒛,ν,β,λ)\bm{H}(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta,\lambda) with respect to (δ,s1,s2,𝒘,𝒛,ν,β)(\delta,s_{1},s_{2},\bm{w},\bm{z},\nu,\beta) at K0K_{0} is as follows:

𝑲(δ^,s^1,s^2,𝒘^,𝒛^,ν^,β^):3×()2×2(×)2×,𝑲(δ^,s^1,s^2,𝒘^,𝒛^,ν^,β^)=(k1,k21,,k2n,k3,k41,,k4n,k5)T(δ^,s^1,s^2,𝒘^,𝒛^,ν^,β^),\begin{split}&\bm{K}(\hat{\delta},\hat{s}_{1},\hat{s}_{2},\bm{\hat{w}},\bm{\hat{z}},\hat{\nu},\hat{\beta}):\mathbb{R}^{3}\times\left(\mathcal{M}_{\mathbb{C}}\right)^{2}\times\mathbb{R}^{2}\to\left(\mathbb{C}\times\mathcal{M}_{\mathbb{C}}\right)^{2}\times\mathbb{R},\\ &\bm{K}(\hat{\delta},\hat{s}_{1},\hat{s}_{2},\bm{\hat{w}},\bm{\hat{z}},\hat{\nu},\hat{\beta})=(k_{1},k_{21},\dots,k_{2n},k_{3},k_{41},\dots,k_{4n},k_{5})^{T}(\hat{\delta},\hat{s}_{1},\hat{s}_{2},\bm{\hat{w}},\bm{\hat{z}},\hat{\nu},\hat{\beta}),\end{split}

where

k1:=j=1n[(2c0r0β0ξjηjβ0bj1)(δ^ξj+w^j)+(c0ξj)2((s^1+is^2)ηj+z^j)]iν^δ0iν0δ^+β^δ0j=1nbjξj,k2j:=k=1npjkw^k,j=1,,n,k3:=j=1n[(β0bj2c0r0β0ξjηj)(δ^ξj+w^j)(c0ξj)2((s^1+is^2)ηj+z^j)]iν0(s^1+is^2)+ν^δ0(i+1ν0)β^δ0j=1nbjξj,\begin{split}k_{1}:=&\sum_{j=1}^{n}\left[\left(2c_{0}r_{0\beta_{0}}\xi_{j}\eta_{j}-\beta_{0}b_{j}-1\right)\left(\hat{\delta}\xi_{j}+\hat{w}_{j}\right)+{(c_{0}\xi_{j})}^{2}((\hat{s}_{1}+{\rm i}\hat{s}_{2})\eta_{j}+\hat{z}_{j})\right]\\ &-{\rm i}\hat{\nu}\delta_{0}-{\rm i}\nu_{0}\hat{\delta}+\hat{\beta}\delta_{0}\sum_{j=1}^{n}b_{j}\xi_{j},\\ k_{2j}:=&\sum_{k=1}^{n}{p_{jk}\hat{w}_{k}},~{}~{}~{}~{}~{}j=1,\dots,n,\\ k_{3}:=&\sum_{j=1}^{n}\left[(\beta_{0}b_{j}-2c_{0}r_{0\beta_{0}}\xi_{j}\eta_{j})\left(\hat{\delta}\xi_{j}+\hat{w}_{j}\right)-{(c_{0}\xi_{j})}^{2}((\hat{s}_{1}+{\rm i}\hat{s}_{2})\eta_{j}+\hat{z}_{j})\right]\\ &-{\rm i}\nu_{0}(\hat{s}_{1}+{\rm i}\hat{s}_{2})+\hat{\nu}\delta_{0}\left({\rm i}+\frac{1}{\nu_{0}}\right)-\hat{\beta}\delta_{0}\sum_{j=1}^{n}b_{j}\xi_{j},\\ \end{split}
k4j:=θk=1nqjkz^k,j=1,,n,k5:=2δ0𝝃22δ^2δ0𝜼22s^1+2δ0ν0𝜼22s^2+δ0j=1nξj(w^j+w^j¯)+δ0j=1nηj[iν0(z^j¯z^j)(z^j+z^j¯)].\begin{split}k_{4j}:=&\theta\sum_{k=1}^{n}{q_{jk}\hat{z}_{k}},~{}~{}~{}~{}~{}j=1,\dots,n,\\ k_{5}:=&2\delta_{0}\|\bm{\xi}\|_{2}^{2}\hat{\delta}-2\delta_{0}\|\bm{\eta}\|_{2}^{2}\hat{s}_{1}+\frac{2\delta_{0}}{\nu_{0}}\|\bm{\eta}\|_{2}^{2}\hat{s}_{2}+\delta_{0}\sum_{j=1}^{n}\xi_{j}(\hat{w}_{j}+\overline{\hat{w}_{j}})\\ &+\delta_{0}\sum_{j=1}^{n}\eta_{j}\left[\frac{\rm i}{\nu_{0}}(\overline{\hat{z}_{j}}-\hat{z}_{j})-(\hat{z}_{j}+\overline{\hat{z}_{j}})\right].\end{split}

If 𝑲(δ^λ,s^1,s^2,𝒘^,𝒛^,ν^,β^)=𝟎\bm{K}(\hat{\delta}_{\lambda},\hat{s}_{1},\hat{s}_{2},\bm{\hat{w}},\bm{\hat{z}},\hat{\nu},\hat{\beta})=\bm{0}, we see that 𝒘^=𝟎\bm{\hat{w}}=\bm{0} and 𝒛^=𝟎\bm{\hat{z}}=\bm{0}. Plugging 𝒘^=𝒛^=𝟎\bm{\hat{w}}=\bm{\hat{z}}=\bm{0} into k1=k3=k5=0k_{1}=k_{3}=k_{5}=0, we get

D(δ^,s^1,s^2,ν^,β^)T=𝟎,D(\hat{\delta},\hat{s}_{1},\hat{s}_{2},\hat{\nu},\hat{\beta})^{T}=\bm{0},

where

D=(d11j=1n(c0ξj)2ηj00δ0j=1nbjξjν00j=1n(c0ξj)2ηjδ00d31j=1n(c0ξj)2ηjν0δ0ν0δ0j=1nbjξj0ν0j=1n(c0ξj)2ηjδ00𝝃22𝜼221ν0𝜼2200),D=\left(\begin{array}[]{*{20}{c}}d_{11}&\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}&0&0&\delta_{0}\sum_{j=1}^{n}b_{j}\xi_{j}\\ -\nu_{0}&0&\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}&-\delta_{0}&0\\ d_{31}&-\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}&\nu_{0}&\frac{\delta_{0}}{\nu_{0}}&-\delta_{0}\sum_{j=1}^{n}b_{j}\xi_{j}\\ 0&-\nu_{0}&-\sum_{j=1}^{n}{(c_{0}\xi_{j})}^{2}\eta_{j}&\delta_{0}&0\\ \|\bm{\xi}\|_{2}^{2}&-\|\bm{\eta}\|_{2}^{2}&\frac{1}{\nu_{0}}\|\bm{\eta}\|_{2}^{2}&0&0\end{array}\right),\\

and

d11=j=1n(2c0r0β0ξjηjβ0bj1)ξj,d31=j=1n(β0bj2c0r0β0ξjηj)ξj.d_{11}=\sum_{j=1}^{n}\left(2c_{0}r_{0\beta_{0}}\xi_{j}\eta_{j}-\beta_{0}b_{j}-1\right)\xi_{j},\;\;d_{31}=\sum_{j=1}^{n}(\beta_{0}b_{j}-2c_{0}r_{0\beta_{0}}\xi_{j}\eta_{j})\xi_{j}.

A direct computation implies that

|D|=2δ02(j=1nbjξj)(𝜼22+ν02𝝃22+ν02𝜼22)0,|D|=2\delta^{2}_{0}\left(\sum_{j=1}^{n}b_{j}\xi_{j}\right)\left(\|\bm{\eta}\|_{2}^{2}+\nu_{0}^{2}\|\bm{\xi}\|_{2}^{2}+\nu_{0}^{2}\|\bm{\eta}\|_{2}^{2}\right)\neq 0,

which implies that δ^=0,s^1=0,s^2=0,ν^=0\hat{\delta}=0,\hat{s}_{1}=0,\hat{s}_{2}=0,\hat{\nu}=0 and β^=0\hat{\beta}=0. Therefore, 𝑲\bm{K} is bijection. It follows from the implicit function theorem that there exists λ2(0,λ1)\lambda_{2}\in(0,\lambda_{1}), where λ2\lambda_{2} is sufficiently small, and a continuously differentiable mapping

λ(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ):[0,λ2](0,)×2×()2×(0,+)×\lambda\mapsto(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}):[0,\lambda_{2}]\to(0,\infty)\times\mathbb{R}^{2}\times\left(\mathcal{M}_{\mathbb{C}}\right)^{2}\times(0,+\infty)\times\mathcal{B}

such that 𝑯(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ,λ)=𝟎\bm{H}\left(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda},\lambda\right)=\bm{0}, and for λ=0\lambda=0,

(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ)=(δ0,s10,s20,𝒘0,𝒛0,ν0,β0).\left(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}\right)=(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0}).

Then (δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ)\left(\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda},\nu_{\lambda},\beta_{\lambda}\right) is a positive solution of (3.10) for λ[0,λ2]\lambda\in[0,\lambda_{2}].

Now, we show the uniqueness. From the implicit function theorem, we only need to verify that if (δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ)\left(\delta^{\lambda},s_{1}^{\lambda},s_{2}^{\lambda},\bm{w}^{\lambda},\bm{z}^{\lambda},\nu^{\lambda},\beta^{\lambda}\right) is a solution of (3.10), then

(δλ,s1λ,s2λ,𝒘λ,𝒛λ,νλ,βλ)(δ0,s10,s20,𝒘0,𝒛0,ν0,β0)asλ0.\left(\delta^{\lambda},s_{1}^{\lambda},s_{2}^{\lambda},\bm{w}^{\lambda},\bm{z}^{\lambda},\nu^{\lambda},\beta^{\lambda}\right)\to(\delta_{0},s_{10},s_{20},\bm{w}_{0},\bm{z}_{0},\nu_{0},\beta_{0})\;\;\text{as}\;\;\lambda\to 0. (3.15)

Since

δλ𝝃+𝒘λ22+(s1λ+is2λ)𝜼+𝒛λ22=𝝃22+𝜼22,\|\delta^{\lambda}\bm{\xi}+\bm{w}^{\lambda}\|_{2}^{2}+\|(s^{\lambda}_{1}+{\rm i}s^{\lambda}_{2})\bm{\eta}+\bm{z}^{\lambda}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2},

we obtain that 𝒘λ\bm{w}^{\lambda}, 𝒛λ,δλ\bm{z}^{\lambda},\delta^{\lambda}, s1λs^{\lambda}_{1} and s2λs_{2}^{\lambda} are bounded. It follows from Lemma 3.1 that νλ\nu^{\lambda} is bounded. Then, up to a sequence, we can assume that

limλ0𝒘λ=𝒘,limλ0𝒛λ=𝒛,limλ0s1λ=s1,limλ0s2λ=s2,limλ0δλ=δ,limλ0βλ=β,limλ0νλ=ν.\begin{split}&\lim_{\lambda\to 0}\bm{w}^{\lambda}=\bm{w}^{*},\;\lim_{\lambda\to 0}\bm{z}^{\lambda}=\bm{z}^{*},\;\lim_{\lambda\to 0}s^{\lambda}_{1}=s_{1}^{*},\;\lim_{\lambda\to 0}s^{\lambda}_{2}=s_{2}^{*},\\ &\lim_{\lambda\to 0}\delta^{\lambda}=\delta^{*},\;\lim_{\lambda\to 0}\beta^{\lambda}=\beta^{*},\;\lim_{\lambda\to 0}\nu^{\lambda}=\nu^{*}.\end{split}

Taking the limit of (3.10) as λ0\lambda\to 0, we see that (δ,s1,s2,𝒘,𝒛,ν,β)(\delta^{*},s^{*}_{1},s^{*}_{2},\bm{w}^{*},\bm{z}^{*},\nu^{*},\beta^{*}) is a solution of (3.10) for λ=0\lambda=0. 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.

Assume that λ(0,λ2]\lambda\in(0,\lambda_{2}], where λ2\lambda_{2} defined in Theorem 3.3. Then (𝛗,𝛙,ν,β)(\bm{\varphi},\bm{\psi},\nu,\beta) solves

{(Aβ(λ)iλνI)(𝝋,𝝍)T=𝟎,ν0,β,(𝝋,𝝍)T(𝟎)2n,\begin{cases}\left(A_{\beta}(\lambda)-{\rm i}\lambda\nu I\right)(\bm{\varphi},\bm{\psi})^{T}=\bm{0},\\ \nu\geq 0,\;\beta\in\mathcal{B},\;(\bm{\varphi},\bm{\psi})^{T}(\neq\bm{0})\in\mathbb{C}^{2n},\\ \end{cases}

if and only if

ν=νλ,β=βλ,𝝋=κ𝝋λ=κ(δλ𝝃+𝒘λ),𝝍=κ𝝍λ=κ((s1λ+is2λ)𝜼+𝒛λ),\nu=\nu_{\lambda},\;\beta=\beta_{\lambda},\;\bm{\varphi}=\kappa\bm{\varphi}_{\lambda}=\kappa(\delta_{\lambda}\bm{\xi}+\bm{w}_{\lambda}),\;\bm{\psi}=\kappa\bm{\psi}_{\lambda}=\kappa((s_{1\lambda}+{\rm i}s_{2\lambda})\bm{\eta}+\bm{z}_{\lambda}),

where Aβ(λ)A_{\beta}(\lambda) is defined in (3.2), II is the identity matrix, κ\kappa is a nonzero constant, and (νλ,βλ,δλ,s1λ,s2λ,𝐰λ,𝐳λ)(\nu_{\lambda},\beta_{\lambda},\delta_{\lambda},s_{1\lambda},s_{2\lambda},\bm{w}_{\lambda},\bm{z}_{\lambda}) is defined in Theorem 3.3.

To show that iλνλ{\rm i}\lambda\nu_{\lambda} is a simple eigenvalue of Aβλ(λ)A_{\beta_{\lambda}}(\lambda), we need to consider the eigenvalues of AβλT(λ)A^{T}_{\beta_{\lambda}}(\lambda), where AβλT(λ)A^{T}_{\beta_{\lambda}}(\lambda) is the transpose of Aβλ(λ)A_{\beta_{\lambda}}(\lambda). Denote

~1={(x1,,xn)Tn:j=1nξjxj=0},~2={(x1,,xn)Tn:j=1nηjxj=0}.\begin{split}&\widetilde{\mathcal{M}}_{1}=\{(x_{1},\dots,x_{n})^{T}\in\mathbb{R}^{n}:\sum_{j=1}^{n}\xi_{j}x_{j}=0\},\\ &\widetilde{\mathcal{M}}_{2}=\{(x_{1},\dots,x_{n})^{T}\in\mathbb{R}^{n}:\sum_{j=1}^{n}\eta_{j}x_{j}=0\}.\end{split}

Then

n=span{(1,,1)T}~1=span{(1,,1)T}~2.\mathbb{R}^{n}={\rm span}\left\{\left(1,\cdots,1\right)^{T}\right\}\oplus\widetilde{\mathcal{M}}_{1}={\rm span}\left\{\left(1,\cdots,1\right)^{T}\right\}\oplus\widetilde{\mathcal{M}}_{2}.
Lemma 3.5.

Let AβλT(λ)A_{\beta_{\lambda}}^{T}(\lambda) be the transpose of Aβλ(λ)A_{\beta_{\lambda}}(\lambda), where λ(0,λ2]\lambda\in(0,\lambda_{2}] and λ2\lambda_{2} defined in Theorem 3.3. Then iλνλ-{\rm i}\lambda\nu_{\lambda} is an eigenvalue of AβλT(λ)A_{\beta_{\lambda}}^{T}(\lambda), and

𝒩[AβλT(λ)+iλνλI]=span[(𝝋~λ,𝝍~λ)T].\mathcal{N}[A_{\beta_{\lambda}}^{T}(\lambda)+{\rm i}\lambda\nu_{\lambda}I]={\rm span}[(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T}].

Moreover, ignoring a scalar factor, (𝛗~λ,𝛙~λ)(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda}) can be represented as

{𝝋~λ=δ~λ(1,,1)T+𝒘~λ,whereδ~λ0and𝒘~λ(~1),𝝍~λ=(s~1λ+is~2λ)(1,,1)T+𝒛~λ,wheres~1λ,s~2λand𝒛~λ(~2),𝝋~λ22+𝝍~λ22=2n,\begin{cases}\bm{\tilde{\varphi}}_{\lambda}=\tilde{\delta}_{\lambda}{\left(1,\cdots,1\right)}^{T}+\bm{\tilde{w}}_{\lambda},\;\;\text{where}\;\;\tilde{\delta}_{\lambda}\geq 0\;\;\text{and}\;\;\bm{\tilde{w}}_{\lambda}\in\left(\widetilde{\mathcal{M}}_{1}\right)_{\mathbb{C}},\\ \bm{\tilde{\psi}}_{\lambda}=(\tilde{s}_{1\lambda}+{\rm i}\tilde{s}_{2\lambda}){\left(1,\cdots,1\right)}^{T}+\bm{\tilde{z}}_{\lambda},\;\;\text{where}\;\;\tilde{s}_{1\lambda},\tilde{s}_{2\lambda}\in\mathbb{R}\;\;\text{and}\;\;\bm{\tilde{z}}_{\lambda}\in\left(\widetilde{\mathcal{M}}_{2}\right)_{\mathbb{C}},\\ \|\bm{\tilde{\varphi}}_{\lambda}\|_{2}^{2}+\|\bm{\tilde{\psi}}_{\lambda}\|_{2}^{2}=2n,\end{cases} (3.16)

and (δ~λ,s~1λ,s~2λ,𝐰~λ,𝐳~λ)=(δ~0,s~10,s~20,𝐰~0,𝐳~0)\left(\tilde{\delta}_{\lambda},\tilde{s}_{1\lambda},\tilde{s}_{2\lambda},\bm{\tilde{w}}_{\lambda},\bm{\tilde{z}}_{\lambda}\right)=(\tilde{\delta}_{0},\tilde{s}_{10},\tilde{s}_{20},\bm{\tilde{w}}_{0},\bm{\tilde{z}}_{0}) for λ=0\lambda=0, where

s~10=δ~0ν02ν02+1,s~20=δ~0ν0ν02+1,𝒘~0=𝟎,𝒛~0=𝟎,δ~0=2ν02+22ν02+1,\tilde{s}_{10}=\frac{\tilde{\delta}_{0}\nu_{0}^{2}}{\nu_{0}^{2}+1},\;\tilde{s}_{20}=\frac{\tilde{\delta}_{0}\nu_{0}}{\nu_{0}^{2}+1},\;\bm{\tilde{w}}_{0}=\bm{0},\;\bm{\tilde{z}}_{0}=\bm{0},\;\tilde{\delta}_{0}=\sqrt{\frac{2\nu_{0}^{2}+2}{2\nu_{0}^{2}+1}},

and ν0\nu_{0} is defined in Lemma 3.2.

Proof.

It follows from Theorem 3.4 that, for λ(0,λ2]\lambda\in(0,\lambda_{2}], ±iλνλ\pm{\rm i}\lambda\nu_{\lambda} is an eigenvalue of Aβλ(λ)A_{\beta_{\lambda}}(\lambda) and 𝒩[Aβλ(λ)iλνλI]\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I] is one-dimensional. Then iλνλ-{\rm i}\lambda\nu_{\lambda} is an eigenvalue of AβλT(λ)A^{T}_{\beta_{\lambda}}(\lambda), 𝒩[AβλT(λ)+iλνλI]=span[(𝝋~λ,𝝍~λ)T]\mathcal{N}[A_{\beta_{\lambda}}^{T}(\lambda)+{\rm i}\lambda\nu_{\lambda}I]={\rm span}[(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T}], and (𝝋~λ,𝝍~λ)(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda}) can be represented as (3.16). It follows from (3.16) that δ~λ\tilde{\delta}_{\lambda}, s~1λ\tilde{s}_{1\lambda}, s~2λ\tilde{s}_{2\lambda}, 𝒘~λ\bm{\tilde{w}}_{\lambda} and 𝒛~λ\bm{\tilde{z}}_{\lambda} are bounded. Then, up to a sequence, we assume that

limλ0𝝋~λ=𝝋~,limλ0𝝍~λ=𝝍~,limλ0𝒘~λ=𝒘~0,limλ0𝒛~λ=𝒛~0,limλ0s~1λ=s~10,limλ0s~2λ=s~20,limλ0δ~λ=δ~0,\begin{split}&\lim_{\lambda\to 0}\bm{\tilde{\varphi}}_{\lambda}=\bm{\tilde{\varphi}}_{*},\;\lim_{\lambda\to 0}\bm{\tilde{\psi}}_{\lambda}=\bm{\tilde{\psi}}_{*},\;\lim_{\lambda\to 0}\bm{\tilde{w}}_{\lambda}=\bm{\tilde{w}}_{0},\;\lim_{\lambda\to 0}\bm{\tilde{z}}_{\lambda}=\bm{\tilde{z}}_{0},\\ &\lim_{\lambda\to 0}\tilde{s}_{1\lambda}=\tilde{s}_{10},\;\lim_{\lambda\to 0}\tilde{s}_{2\lambda}=\tilde{s}_{20},\;\lim_{\lambda\to 0}\tilde{\delta}_{\lambda}=\tilde{\delta}_{0},\end{split}

where

𝝋~22+𝝍~22=2n.\|\bm{\tilde{\varphi}}_{*}\|_{2}^{2}+\|\bm{\tilde{\psi}}_{*}\|_{2}^{2}=2n. (3.17)

Note that limλ0βλ=β0\lim_{\lambda\to 0}\beta_{\lambda}=\beta_{0} and limλ0νλ=ν0\lim_{\lambda\to 0}\nu_{\lambda}=\nu_{0}, where ν0,β0\nu_{0},\beta_{0} are defined in Lemma 3.2. Taking the limit of

(AβλT(λ)+iλνλI)(𝝋~λ,𝝍~λ)T=𝟎,\left(A_{\beta_{\lambda}}^{T}({\lambda})+{\rm i}{\lambda}\nu_{\lambda}I\right)(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T}=\bm{0}, (3.18)

as λ0\lambda\to 0, we have PT𝝋=𝟎P^{T}\bm{\varphi}_{*}=\bm{0} and QT𝝍=𝟎Q^{T}\bm{\psi}_{*}=\bm{0}, which yield 𝒘~0=𝟎\bm{\tilde{w}}_{0}=\bm{0} and 𝒛~0=𝟎\bm{\tilde{z}}_{0}=\bm{0}. Noticing that P𝝃=Q𝜼=𝟎P\bm{\xi}=Q\bm{\eta}=\bm{0}, we see from (3.18) that

(j=1nM1j(0,β0)ξj+iν0j=1nM3j(0,β0)ξjj=1nM2j(0,β0)ηjj=1nM2j(0,β0)ηj+iν0)(δ~0s~10+is~20)=(00).\left({\begin{array}[]{*{20}{c}}\sum_{j=1}^{n}M_{1j}^{(0,\beta_{0})}\xi_{j}+{\rm i}\nu_{0}&\sum_{j=1}^{n}M_{3j}^{(0,\beta_{0})}\xi_{j}\\ \sum_{j=1}^{n}M_{2j}^{(0,\beta_{0})}\eta_{j}&-\sum_{j=1}^{n}M_{2j}^{(0,\beta_{0})}\eta_{j}+{\rm i}\nu_{0}\end{array}}\right)\left({\begin{array}[]{*{20}{c}}\tilde{\delta}_{0}\\ \tilde{s}_{10}+{\rm i}\tilde{s}_{20}\end{array}}\right)=\left({\begin{array}[]{*{20}{c}}0\\ 0\end{array}}\right).

This, combined with Eqs. (3.13) and (3.14), implies that

s~10=δ~0ν02ν02+1,s~20=δ~0ν0ν02+1.\tilde{s}_{10}=\frac{\tilde{\delta}_{0}\nu_{0}^{2}}{\nu_{0}^{2}+1},\;\tilde{s}_{20}=\frac{\tilde{\delta}_{0}\nu_{0}}{\nu_{0}^{2}+1}. (3.19)

Note from (3.17) that δ~02+s~102+s~202=2.\tilde{\delta}_{0}^{2}+\tilde{s}_{10}^{2}+\tilde{s}_{20}^{2}=2. This, combined with (3.19), implies that

δ~0=2ν02+22ν02+1.\tilde{\delta}_{0}=\sqrt{\frac{2\nu_{0}^{2}+2}{2\nu_{0}^{2}+1}}.

This completes the proof. ∎

Now, we show that iλνλ{\rm i}\lambda\nu_{\lambda} is simple.

Theorem 3.6.

Assume that λ(0,λ2]\lambda\in(0,\lambda_{2}], where λ2\lambda_{2} is sufficiently small. Then iλνλ{\rm i}\lambda\nu_{\lambda} is a simple eigenvalue of Aβλ(λ)A_{\beta_{\lambda}}(\lambda), where Aβ(λ)A_{\beta}(\lambda) is defined in (3.2).

Proof.

It follows from Theorem 3.4 that 𝒩[Aβλ(λ)iλνλI]=span[(𝝋λ,𝝍λ)T]\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]={\rm span}[(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}], where 𝝋λ,𝝍λ\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda} is defined in Theorem 3.3. Then we show that

𝒩[Aβλ(λ)iλνλI]2=𝒩[Aβλ(λ)iλνλI].\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]^{2}=\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I].

If 𝚿=(𝚿1,𝚿2)T𝒩[Aβλ(λ)iλνλI]2\bm{\Psi}=(\bm{\Psi}_{1},\bm{\Psi}_{2})^{T}\in\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]^{2}, where 𝚿1=(Ψ11,,Ψ1n)Tn\bm{\Psi}_{1}=(\Psi_{11},\dots,\Psi_{1n})^{T}\in{\mathbb{C}}^{n} and 𝚿2=(Ψ21,,Ψ2n)Tn\bm{\Psi}_{2}=(\Psi_{21},\dots,\Psi_{2n})^{T}\in{\mathbb{C}}^{n}, then

[Aβλ(λ)iλνλI]𝚿𝒩[Aβλ(λ)iλνλI]=span[(𝝋λ,𝝍λ)T],[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]\bm{\Psi}\in\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]={\rm span}[(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}],

which implies that there exists a constant ss such that

[Aβλ(λ)iλνλI]𝚿=s(𝝋λ,𝝍λ)T.[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]\bm{\Psi}=s(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}.

That is,

{sφλj=j=1npijΨ1j+λ[M1j(λ,βλ)Ψ1j+M2j(λ,βλ)Ψ2j]iλνλΨ1j,j=1,,n,sψλj=θj=1nqijΨ~2j+λ[M3j(λ,βλ)Ψ1jM2j(λ,βλ)Ψ2j]iλνλΨ2j,j=1,,n.\begin{cases}\displaystyle s\varphi_{\lambda j}=\sum_{j=1}^{n}p_{ij}\Psi_{1j}+\lambda\left[M_{1j}^{(\lambda,{\beta_{\lambda}})}\Psi_{1j}+M_{2j}^{(\lambda,{\beta_{\lambda}})}\Psi_{2j}\right]-{\rm i}\lambda\nu_{\lambda}\Psi_{1j},\;j=1,\dots,n,\\ \displaystyle s\psi_{\lambda j}=\theta\sum_{j=1}^{n}q_{ij}\tilde{\Psi}_{2j}+\lambda\left[M_{3j}^{(\lambda,{\beta_{\lambda}})}\Psi_{1j}-M_{2j}^{(\lambda,{\beta_{\lambda}})}\Psi_{2j}\right]-{\rm i}\lambda\nu_{\lambda}\Psi_{2j},\;j=1,\dots,n.\end{cases} (3.20)

It follows from Lemma 3.5 that

(AβλT(λ)+iλνλI)(𝝋~λ,𝝍~λ)T=𝟎.\left(A_{\beta_{\lambda}}^{T}(\lambda)+{\rm i}\lambda\nu_{\lambda}I\right)(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T}=\bm{0}.

Then, multiplying the first and second equation of (3.20) by φ~λj¯\overline{\tilde{\varphi}_{\lambda j}} and ψ~λj¯\overline{\tilde{\psi}_{\lambda j}}, respectively, and summing the results over all jj, we have

0=(AβλT(λ)+iλνλI)(𝝋~λ,𝝍~λ)T,𝚿=(𝝋~λ,𝝍~λ)T,(Aβλ(λ)iλνλI)𝚿=sj=1n(φ~λj¯φλj+ψ~λj¯ψλj).\begin{split}0&=\left\langle\left(A_{\beta_{\lambda}}^{T}(\lambda)+{\rm i}\lambda\nu_{\lambda}I\right)(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T},\bm{\Psi}\right\rangle=\left\langle(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T},\left(A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I\right)\bm{\Psi}\right\rangle\\ &=s\sum_{j=1}^{n}\left(\overline{\tilde{\varphi}_{\lambda j}}\varphi_{\lambda j}+\overline{\tilde{\psi}_{\lambda j}}\psi_{\lambda j}\right).\\ \end{split}

It follows from Theorem 3.3 and Lemma 3.5 that

limλ0j=1n(φ~λj¯φλj+ψ~λj¯ψλj)=δ0δ~0j=1n(ξj+ν0ν02+1(ν0i)(iν01)ηj)=2δ0δ~0iν0+1ν02+10,\begin{split}\lim_{\lambda\to 0}\sum_{j=1}^{n}\left(\overline{\tilde{\varphi}_{\lambda j}}\varphi_{\lambda j}+\overline{\tilde{\psi}_{\lambda j}}\psi_{\lambda j}\right)&=\delta_{0}\tilde{\delta}_{0}\sum_{j=1}^{n}\left(\xi_{j}+\frac{\nu_{0}}{\nu_{0}^{2}+1}(\nu_{0}-{\rm i})(\frac{{\rm i}}{\nu_{0}}-1)\eta_{j}\right)\\ &=2\delta_{0}\tilde{\delta}_{0}\frac{{\rm i}\nu_{0}+1}{\nu_{0}^{2}+1}\neq 0,\end{split} (3.21)

which implies that s=0s=0 for λ(0,λ2]\lambda\in(0,\lambda_{2}], where λ2\lambda_{2} is sufficiently small. Therefore,

𝒩[Aβλ(λ)iλνλI]j=𝒩[Aβλ(λ)iλνλI],j=2,3,.\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I]^{j}=\mathcal{N}[A_{\beta_{\lambda}}(\lambda)-{\rm i}\lambda\nu_{\lambda}I],\;\;j=2,3,\cdots.

This completes the proof. ∎

It follows from Theorem 3.6 that μ=iλνλ\mu={\rm i}\lambda\nu_{\lambda} is a simple eigenvalue of Aβλ(λ)A_{\beta_{\lambda}}(\lambda) for fixed λ(0,λ2]\lambda\in(0,\lambda_{2}]. Then, by using the implicit function theorem, we see that there exists a neighborhood V1×V2×OV_{1}\times V_{2}\times O of (𝝋λ,𝝍λ,iνλ,βλ)(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda},{\rm i}\nu_{\lambda},\beta_{\lambda}) ( V1V_{1} is defined as the neighborhood of (𝝋λ,𝝍λ)(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda}) ) and a continuously differentiable function (𝝋(β),𝝍(β),μ(β)):OV1×V2(\bm{\varphi}(\beta),\bm{\psi}(\beta),\mu(\beta)):O\to V_{1}\times V_{2} such that μ(βλ)=iνλ,(𝝋(βλ),𝝍(βλ))=(𝝋λ,𝝍λ)\mu(\beta_{\lambda})={\rm i}\nu_{\lambda},\;(\bm{\varphi}(\beta_{\lambda}),\bm{\psi}(\beta_{\lambda}))=(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda}). Moreover, for each βO\beta\in O, the only eigenvalue of Aβ(λ)A_{\beta}(\lambda) in V2V_{2} is μ(β)\mu(\beta), and

(Aβ(λ)μ(β)I)(𝝋(β),𝝍(β))T=𝟎.\left(A_{\beta}(\lambda)-\mu(\beta)I\right)(\bm{\varphi}(\beta),\bm{\psi}(\beta))^{T}=\bm{0}. (3.22)

Then, we show that the following transversality condition holds.

Theorem 3.7.

For λ(0,λ2]\lambda\in(0,\lambda_{2}], where λ2\lambda_{2} is sufficiently small,

de[μ(β)]dβ|β=βλ>0.\left.\frac{d\mathcal{R}e[\mu(\beta)]}{d\beta}\right|_{\beta=\beta_{\lambda}}>0.
Proof.

Differentiating (3.22) with respect to β\beta at β=βλ,\beta=\beta_{\lambda}, we have

dμdβ|β=βλ(𝝋λ,𝝍λ)T=(Aβλ(λ)iνλI)(d𝝋dβ,d𝝍dβ)T|β=βλ+dAβ(λ)dβ|β=βλ(𝝋λ,𝝍λ)T.\left.\frac{d\mu}{d\beta}\right|_{\beta=\beta_{\lambda}}(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}=\left(A_{\beta_{\lambda}}(\lambda)-{\rm i}\nu_{\lambda}I\right)\left.\left(\frac{d\bm{\varphi}}{d\beta},\frac{d\bm{\psi}}{d\beta}\right)^{T}\right|_{\beta=\beta_{\lambda}}+\left.\frac{dA_{\beta}(\lambda)}{d\beta}\right|_{\beta=\beta_{\lambda}}(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}. (3.23)

Note that

(𝝋~λ,𝝍~λ)T,(Aβλ(λ)iνλI)(d𝝋dβ,d𝝍dβ)T|β=βλ=(AβλT(λ)+iλνλI)(𝝋~λ,𝝍~λ)T,(d𝝋dβ,d𝝍dβ)T|β=βλ=0,\begin{split}&\left\langle(\bm{\tilde{\varphi}}_{\lambda},\bm{\tilde{\psi}}_{\lambda})^{T},\left(A_{\beta_{\lambda}}(\lambda)-{\rm i}\nu_{\lambda}I\right)\left.\left(\frac{d\bm{\varphi}}{d\beta},\frac{d\bm{\psi}}{d\beta}\right)^{T}\right|_{\beta=\beta_{\lambda}}\right\rangle\\ &=\left\langle\left(A_{\beta_{\lambda}}^{T}(\lambda)+{\rm i}\lambda\nu_{\lambda}I\right)({\bm{\tilde{\varphi}}_{\lambda}},{\bm{\tilde{\psi}}_{\lambda}})^{T},\left.\left(\frac{d\bm{\varphi}}{d\beta},\frac{d\bm{\psi}}{d\beta}\right)^{T}\right|_{\beta=\beta_{\lambda}}\right\rangle=0,\end{split}

where 𝝋~λ\bm{\tilde{\varphi}}_{\lambda} and 𝝍~λ\bm{\tilde{\psi}}_{\lambda} are defined in Lemma 3.5. Then we see from (3.23) that

dμdβ|β=βλ(𝝋~λ,𝝍~λ)T,(𝝋λ,𝝍λ)T=(𝝋~λ,𝝍~λ)T,dAβ(λ)dβ|β=βλ(𝝋λ,𝝍λ)T.\left.\frac{d\mu}{d\beta}\right|_{\beta=\beta_{\lambda}}\left\langle({\bm{\tilde{\varphi}}_{\lambda}},{\bm{\tilde{\psi}}_{\lambda}})^{T},(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}\right\rangle=\left\langle({\bm{\tilde{\varphi}}_{\lambda}},{\bm{\tilde{\psi}}_{\lambda}})^{T},\left.\frac{dA_{\beta}(\lambda)}{d\beta}\right|_{\beta=\beta_{\lambda}}(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}\right\rangle.\\ (3.24)

It follows from Theorem 2.3 that (𝒙(λ,β),𝒚(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) is continuously differentiable on [0,λ2]×[0,\lambda_{2}]\times\mathcal{B}. This, combined the fact that limλ0βλ=β0\lim_{\lambda\to 0}\beta_{\lambda}=\beta_{0}, implies that

limλ0dMlj(λ,β)dβ|β=βλ=dMlj(0,β)dβ|β=β0forl=1,2,3andj=1,,n,\lim_{\lambda\to 0}\left.\displaystyle\frac{dM^{(\lambda,\beta)}_{lj}}{d\beta}\right|_{\beta=\beta_{\lambda}}=\left.\displaystyle\frac{dM^{(0,\beta)}_{lj}}{d\beta}\right|_{\beta=\beta_{0}}\;\;\text{for}\;\;l=1,2,3\;\;\text{and}\;\;j=1,\dots,n, (3.25)

where Mlj(λ,β)M^{(\lambda,\beta)}_{lj} is defined in (3.1). Then we see from (2.5) and (3.25) that

limλ0dAβ(λ)dβ|β=βλ=λ(SOn×nSOn×n),\lim_{\lambda\to 0}\left.\frac{dA_{\beta}(\lambda)}{d\beta}\right|_{\beta=\beta_{\lambda}}=\lambda\left({\begin{array}[]{cc}S&O_{n\times n}\\ -S&O_{n\times n}\end{array}}\right),

where

S=diag(2j=1nbjξjj=1nξj2ηjξjηjbj),S={\rm diag}\left(\frac{2\sum_{j=1}^{n}b_{j}\xi_{j}}{\sum_{j=1}^{n}\xi_{j}^{2}\eta_{j}}\xi_{j}\eta_{j}-b_{j}\right),

and On×nO_{n\times n} is a zero matrix of n×n.n\times n. It follows from Theorem 3.3 and Lemma 3.5 that

limλ01λ(𝝋~λ,𝝍~λ)T,dAβ(λ)dβ|β=βλ(𝝋λ,𝝍λ)T=δ0δ~0j=1nbjξjν02+1(1+iν0).\lim_{\lambda\to 0}\displaystyle\frac{1}{\lambda}\left\langle({\bm{\tilde{\varphi}}_{\lambda}},{\bm{\tilde{\psi}}_{\lambda}})^{T},\left.\frac{dA_{\beta}(\lambda)}{d\beta}\right|_{\beta=\beta_{\lambda}}(\bm{\varphi}_{\lambda},\bm{\psi}_{\lambda})^{T}\right\rangle=\displaystyle\frac{\delta_{0}\tilde{\delta}_{0}\sum_{j=1}^{n}b_{j}\xi_{j}}{\nu_{0}^{2}+1}(1+{\rm i}\nu_{0}).

This, combined with (3.21) and (3.24), yields

limλ01λde[μ(β)]dβ|β=βλ=12j=1nbjξj>0.\lim_{\lambda\to 0}\frac{1}{\lambda}\left.\frac{d\mathcal{R}e[\mu(\beta)]}{d\beta}\right|_{\beta=\beta_{\lambda}}=\frac{1}{2}\sum_{j=1}^{n}b_{j}\xi_{j}>0.

Refer to caption
Figure 2: Bifurcation diagram of model (1.5) with respect to β\beta when λ\lambda is small.

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.

Assume that 0<ϵ10<\epsilon\ll 1. Then there exists λ2>0\lambda_{2}>0 (depending on ϵ\epsilon) and a continuously differentiable mapping

λβλ:[0,λ2]=[ϵ,1/ϵ]\lambda\mapsto\beta_{\lambda}:[0,\lambda_{2}]\to\mathcal{B}=[\epsilon,1/\epsilon]

such that, for each λ(0,λ2]\lambda\in(0,\lambda_{2}], the positive equilibrium (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) of model (1.5) is locally asymptotically stable when β[ϵ,βλ)\beta\in[\epsilon,\beta_{\lambda}) and unstable when β(βλ,1ϵ]\beta\in(\beta_{\lambda},\frac{1}{\epsilon}]. Moreover, system (1.5) undergoes a Hopf bifurcation at (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) when β=βλ\beta=\beta_{\lambda}.

Proof.

Note from [16, Chapter 2, Theorem 5.1] that the eigenvalues of Aβ(λ)A_{\beta}(\lambda) are continuous with respect to β.\beta. We only need to show that there exists λ2>0\lambda_{2}>0, which depends on ϵ\epsilon, such that

σ(Aβ(λ)){x+iy:x,y,x<0}forλ(0,λ2]andβ=ϵ.\sigma(A_{\beta}(\lambda))\subset\{x+{\rm i}y:x,y\in\mathbb{R},x<0\}\;\text{for}\;\lambda\in(0,\lambda_{2}]\;\text{and}\;\beta=\epsilon.

If it is not true, then there exists a sequence {λl}l=1\{\lambda_{l}\}_{l=1}^{\infty} such that limlλl=0\lim_{l\to\infty}\lambda_{l}=0, and

(Aϵ(λl)μI)(𝝋,𝝍)T=𝟎\left(A_{\epsilon}({\lambda_{l}})-\mu I\right)(\bm{\varphi},\bm{\psi})^{T}=\bm{0} (3.26)

is solvable for some value of (μλl,𝝋λl,𝝍λl)(\mu_{\lambda_{l}},\bm{\varphi}_{\lambda_{l}},\bm{\psi}_{\lambda_{l}}) with eμλl0\mathcal{R}e\mu_{\lambda_{l}}\geq 0 and (𝝋λl,𝝍λl)T(𝟎)2n(\bm{\varphi}_{\lambda_{l}},\bm{\psi}_{\lambda_{l}})^{T}(\neq\bm{0})\in\mathbb{C}^{2n}. Here Aϵ(λl)=Aβ(λl)|β=ϵA_{\epsilon}(\lambda_{l})=\left.A_{\beta}(\lambda_{l})\right|_{\beta=\epsilon}. Substituting (μ,𝝋,𝝍)=(μλl,𝝋λl,𝝍λl)(\mu,\bm{\varphi},\bm{\psi})=(\mu_{\lambda_{l}},\bm{\varphi}_{\lambda_{l}},\bm{\psi}_{\lambda_{l}}) into (3.26), we have

{μλlφλlj=k=1npjkφλlk+λl[M1j(λl,ϵ)φλlj+M2j(λl,ϵ)ψλlj],j=1,,n,μλlψλlj=θk=1nqjkψλlk+λl[M3j(λl,ϵ)φλljM2j(λl,ϵ)ψλlj],j=1,,n.\begin{cases}\displaystyle\mu_{\lambda_{l}}\varphi_{{\lambda_{l}}j}=\sum_{k=1}^{n}p_{jk}\varphi_{{\lambda_{l}}k}+\lambda_{l}\left[M_{1j}^{({\lambda_{l}},\epsilon)}\varphi_{{\lambda_{l}}j}+M_{2j}^{({\lambda_{l}},\epsilon)}\psi_{{\lambda_{l}}j}\right],\;\;j=1,\dots,n,\\ \displaystyle\mu_{\lambda_{l}}\psi_{{\lambda_{l}}j}=\theta\sum_{k=1}^{n}q_{jk}\psi_{{\lambda_{l}}k}+\lambda_{l}\left[M_{3j}^{({\lambda_{l}},\epsilon)}\varphi_{{\lambda_{l}}j}-M_{2j}^{({\lambda_{l}},\epsilon)}\psi_{{\lambda_{l}}j}\right],\;\;j=1,\dots,n.\\ \end{cases} (3.27)

Note from Lemma 3.1 that |μλl/λl|\left|\mu_{\lambda_{l}}/{\lambda_{l}}\right| is bounded. Then, up to a subsequence, we assume that limlμλl/λl=μ\lim_{l\to\infty}\mu_{\lambda_{l}}/{\lambda_{l}}=\mu^{*} with eμ0\mathcal{R}e\mu^{*}\geq 0. As in the proof of Lemma 3.5, we see that (𝝋λl,𝝍λl)T(𝟎)2n(\bm{\varphi}_{\lambda l},\bm{\psi}_{\lambda l})^{T}(\neq\bm{0})\in\mathbb{C}^{2n} can be represented as

{𝝋λl=δλl𝝃+𝒘λl,whereδλl0and𝒘λl,𝝍λl=(s1λl+is2λl)𝜼+𝒛λl,wheres1λl,s2λland𝒛λl,𝝋λl22+𝝍λl22=𝝃22+𝜼22,\begin{cases}\bm{\varphi}_{\lambda_{l}}=\delta_{\lambda_{l}}\bm{\xi}+\bm{w}_{\lambda_{l}},\;\;\text{where}\;\;\delta_{\lambda_{l}}\geq 0\;\;\text{and}\;\;\bm{w}_{\lambda_{l}}\in\mathcal{M}_{\mathbb{C}},\\ \bm{\psi}_{\lambda_{l}}=({s_{1}}_{\lambda_{l}}+{\rm i}{s_{2}}_{\lambda_{l}})\bm{\eta}+\bm{z}_{\lambda_{l}},\;\;\text{where}\;\;{s_{1}}_{\lambda_{l}},{s_{2}}_{\lambda_{l}}\in\mathbb{R}\;\;\text{and}\;\;\bm{z}_{\lambda_{l}}\in\mathcal{M}_{\mathbb{C}},\\ \|\bm{\varphi}_{\lambda_{l}}\|_{2}^{2}+\|\bm{\psi}_{\lambda_{l}}\|_{2}^{2}=\|\bm{\xi}\|_{2}^{2}+\|\bm{\eta}\|_{2}^{2},\end{cases} (3.28)

and 𝒘λl\bm{w}_{\lambda_{l}} and 𝒛λl\bm{z}_{\lambda_{l}} satisfy

liml𝒘λl=𝟎,liml𝒛λl=𝟎.\lim_{l\to\infty}\bm{w}_{\lambda_{l}}=\bm{0},\;\;\lim_{l\to\infty}\bm{z}_{\lambda_{l}}=\bm{0}.

Up to a subsequence, we also assume that limlδλl=δ\lim_{l\to\infty}\delta_{\lambda_{l}}=\delta^{*}, limls1λl=s1\lim_{l\to\infty}{s_{1}}_{\lambda_{l}}=s_{1}^{*} and limls2λl=s2\lim_{l\to\infty}{s_{2}}_{\lambda_{l}}=s_{2}^{*}. Dividing the first and second equation of (3.27) by λl\lambda_{l}, respectively, summing the results over all jj, and taking the limits as ll\to\infty, we have

{μδ=δj=1nM1j(0,ϵ)ξj+(s1+is2)j=1nM2j(0,ϵ)ηj,μ(s1+is2)=δj=1nM3j(0,ϵ)ξj(s1+is2)j=1nM2j(0,ϵ)ηj.\begin{cases}\displaystyle\mu^{*}\delta^{*}=\delta^{*}\sum_{j=1}^{n}M_{1j}^{(0,\epsilon)}\xi_{j}+\left(s_{1}^{*}+{\rm i}s_{2}^{*}\right)\sum_{j=1}^{n}M_{2j}^{(0,\epsilon)}\eta_{j},\\ \displaystyle\mu^{*}\left(s_{1}^{*}+{\rm i}s_{2}^{*}\right)=\delta^{*}\sum_{j=1}^{n}M_{3j}^{(0,\epsilon)}\xi_{j}-\left(s_{1}^{*}+{\rm i}s_{2}^{*}\right)\sum_{j=1}^{n}M_{2j}^{(0,\epsilon)}\eta_{j}.\\ \end{cases}

It follows from (3.28) that at least one of δ\delta^{*} and s1+is2s_{1}^{*}+{\rm i}s_{2}^{*} is not equal to zero. Consequently, μ\mu^{*} is an eigenvalue of matrix

(j=1nM1j(0,ϵ)ξjj=1nM2j(0,ϵ)ηjj=1nM3j(0,ϵ)ξjj=1nM2j(0,ϵ)ηj).\left({\begin{array}[]{cc}\sum_{j=1}^{n}M_{1j}^{(0,\epsilon)}\xi_{j}&\sum_{j=1}^{n}M_{2j}^{(0,\epsilon)}\eta_{j}\\ \sum_{j=1}^{n}M_{3j}^{(0,\epsilon)}\xi_{j}&-\sum_{j=1}^{n}M_{2j}^{(0,\epsilon)}\eta_{j}\end{array}}\right).

It follows from (3.12) and (3.13) that, for sufficiently small ϵ\epsilon,

j=1nM1j(0,ϵ)ξjj=1nM2j(0,ϵ)ηj<0,\sum_{j=1}^{n}M_{1j}^{(0,\epsilon)}\xi_{j}-\sum_{j=1}^{n}M_{2j}^{(0,\epsilon)}\eta_{j}<0,

which contradicts eμ0\mathcal{R}e\mu^{*}\geq 0. This completes the proof. ∎

Noticing that model (1.5) is equivalent to the original model (1.3), we have the following result.

Theorem 3.9.

Assume that 0<ϵ10<\epsilon\ll 1, and (𝐀𝟏)(𝐀𝟐)(\bf A_{1})-(\bf A_{2}) hold. Then there exists d>0d_{*}>0 (depending on ϵ\epsilon) and a continuously differentiable mapping

d1βd1:[d,)=[ϵ,1/ϵ]d_{1}\mapsto\beta^{d_{1}}:[d_{*},\infty)\to\mathcal{B}=[\epsilon,1/\epsilon]

such that, for each d1[d,)d_{1}\in[d_{*},\infty), the unique positive equilibrium of model (1.3) is locally asymptotically stable when β[ϵ,βd1)\beta\in[\epsilon,\beta^{d_{1}}) and unstable when β(βd1,1ϵ]\beta\in(\beta^{d_{1}},\frac{1}{\epsilon}]. Moreover, system (1.3) undergoes a Hopf bifurcation when β=βd1\beta=\beta^{d_{1}}.

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 P=QP=Q and the boxes are all identical (that is, ai=aa_{i}=a and bi=1b_{i}=1 for i=1,,ni=1,\dots,n). Then, model (1.5) is reduced to the following form:

{dxjdt=k=1npjkxk+λ[a(β+1)xj+xj2yj],j=1,,n,t>0,dyjdt=θk=1npjkyk+λ(βxjxj2yj),j=1,,n,t>0,𝒙(0)=𝒙0()𝟎,𝒚(0)=𝒚0()𝟎.\begin{cases}\displaystyle\frac{dx_{j}}{dt}=\sum_{k=1}^{n}p_{jk}x_{k}+\lambda\left[a-(\beta+1)x_{j}+x_{j}^{2}y_{j}\right],&j=1,\dots,n,\;\;t>0,\\ \displaystyle\frac{dy_{j}}{dt}=\theta\sum_{k=1}^{n}p_{jk}y_{k}+\lambda\left(\beta x_{j}-x_{j}^{2}y_{j}\right),&j=1,\dots,n,\;\;t>0,\\ \bm{x}(0)=\bm{x}_{0}\geq(\not\equiv)\bm{0},\;\bm{y}(0)=\bm{y}_{0}\geq(\not\equiv)\bm{0}.\end{cases} (4.1)

From Theorem 3.8, we have the following result on the dynamics of (4.1).

Proposition 4.1.

Assume that 0<ϵ10<\epsilon\ll 1. Then there exists λ2>0\lambda_{2}>0 (depending on ϵ\epsilon) and a Hopf bifurcation curve:

λβλ:[0,λ2]=[ϵ,1/ϵ]\lambda\mapsto\beta_{\lambda}:[0,\lambda_{2}]\to\mathcal{B}=[\epsilon,1/\epsilon]

such that, for each λ(0,λ2]\lambda\in(0,\lambda_{2}], the positive equilibrium (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) of system (4.1) is locally asymptotically stable when β[ϵ,βλ)\beta\in[\epsilon,\beta_{\lambda}) and unstable when β(βλ,1ϵ]\beta\in(\beta_{\lambda},\frac{1}{\epsilon}]. Moreover, system (4.1) undergoes a Hopf bifurcation at (𝐱(λ,β),𝐲(λ,β))(\bm{x}^{(\lambda,\beta)},\bm{y}^{(\lambda,\beta)}) when β=βλ\beta=\beta_{\lambda}, and βλ\beta_{\lambda} satisfies

limλ0βλ=β0=1+(na)2(j=1nξj3),\lim_{\lambda\to 0}\beta_{\lambda}=\beta_{0}=1+(na)^{2}\left(\sum_{j=1}^{n}\xi_{j}^{3}\right), (4.2)

where (ξ1,,ξn)(\xi_{1},\dots,\xi_{n}) satisfying (1.4) is the corresponding eigenfunction of PP with respect to eigenvalue 0.

Now, we show the effect of the coupling matrix on the Hopf bifurcation value. We recall the P=(pjk)P=(p_{jk}) is line-sum symmetric matrix if kjpjk=kjpkj\sum_{k\neq j}p_{jk}=\sum_{k\neq j}p_{kj} for all j=1,,nj=1,\dots,n.

Proposition 4.2.

Denote β0\beta_{0} by β0L\beta_{0}^{L} (respectively, β0NL\beta_{0}^{NL}) when P=(pjk)P=(p_{jk}) is line-sum symmetric (respectively, not line-sum symmetric). Then β0L<β0NL\beta_{0}^{L}<\beta_{0}^{NL}.

Proof.

It follows from (4.2) that β0=1+(na)2(j=1nξj3)\beta_{0}=1+(na)^{2}\left(\sum_{j=1}^{n}\xi_{j}^{3}\right). A direct computation implies that

𝝃=(ξ1,,ξn)T=(1n,,1n)T,\bm{\xi}=(\xi_{1},\dots,\xi_{n})^{T}=\left(\frac{1}{n},\dots,\frac{1}{n}\right)^{T},

if and only if PP is line-sum symmetric. Then we see that β0L=1+a2\beta_{0}^{L}=1+a^{2}. It follows from (1.4) and Hölder inequality that

1=(j=1nξj)3(j=1n132)2(j=1nξj3)=n2(j=1nξj3),1=\left(\sum_{j=1}^{n}\xi_{j}\right)^{3}\leq\left(\sum_{j=1}^{n}1^{\frac{3}{2}}\right)^{2}\left(\sum_{j=1}^{n}\xi_{j}^{3}\right)=n^{2}\left(\sum_{j=1}^{n}\xi_{j}^{3}\right), (4.3)

and the equality holds if and only if ξj=1/n\xi_{j}=1/n for j=1nj=1\cdots n. Then, we see from (4.2) and (4.3) that β0L<β0NL\beta_{0}^{L}<\beta_{0}^{NL}. ∎

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.

Refer to caption
Figure 3: The effect of coupling matrices on Hopf bifurcation value.

This phenomenon can also be illustrated numerically. We consider model (4.1), and choose the parameters and initial values as follows:

n=3,β=2.05,λ=0.1,θ=1,a=1,xj(0)=yj(0)=1,(j=1,2,3).n=3,\;\;\beta=2.05,\;\;\lambda=0.1,\;\;\theta=1,\;\;a=1,\;\;x_{j}(0)=y_{j}(0)=1,(j=1,2,3).

Here we choose two different coupling matrices:

PL:=(211132123),PNL:=(323232115),P^{L}:=\left({\begin{array}[]{ccc}-2&1&1\\ 1&-3&2\\ 1&2&-3\end{array}}\right),\;\;P^{NL}:=\left({\begin{array}[]{ccc}-3&2&3\\ 2&-3&2\\ 1&1&-5\end{array}}\right),

where PLP^{L} is line-sum symmetric, and PNLP^{NL} is not line-sum symmetric. Then when β(=2.05)\beta(=2.05) is fixed, model (4.1) admits a positive periodic solution for P=PLP=P^{L}, while the positive equilibrium is stable for P=PNLP=P^{NL}, see Fig. 4. This shows that the Hopf bifurcation value of model (4.1) for P=PLP=P^{L} is smaller than that for P=PNLP=P^{NL}.

Refer to caption
Refer to caption
Figure 4: The effect of coupling matrices. Here β=2.05\beta=2.05, and we only plot one box for simplicity. (Left): P=PLP=P^{L}; (Right): P=PNLP=P^{NL}.

Finally, we provide some numerical simulations to illustrate the theoretical results obtained in Section 3. Here we consider model (1.5), and choose n=5n=5, λ=0.1\lambda=0.1, θ=1\theta=1 and coupling matrices:

P:=(4111115211115111214211115),Q:=(5111116211115112116313136).P:=\left({\begin{array}[]{ccccc}-4&1&1&1&1\\ 1&-5&2&1&1\\ 1&1&-5&1&1\\ 1&2&1&-4&2\\ 1&1&1&1&-5\end{array}}\right),\;\;Q:=\left({\begin{array}[]{ccccc}-5&1&1&1&1\\ 1&-6&2&1&1\\ 1&1&-5&1&1\\ 2&1&1&-6&3\\ 1&3&1&3&-6\end{array}}\right).

Moreover, let β\beta be the variable parameter and choose the other parameters as Table 1.

Table 1: Initial concentrations of reactants in five boxes.
jj 1 2 3 4 5
aja_{j} 1 2 1 0.5 1
bjb_{j} 0.1 0.2 0.4 0.1 0.2
xj(0)x_{j}(0) 0.5 1 1 0.5 1
yj(0)y_{j}(0) 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.

Refer to caption
Refer to caption
Figure 5: The corresponding solution converges to the unique positive equilibrium for model (1.5). Here β=10\beta=10, and we only plot two boxes for simplicity.
Refer to caption
Refer to caption
Figure 6: The corresponding solution converges to a positive periodic solution for model (1.5). Here β=13\beta=13, and we only plot two boxes for simplicity.

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.