Processing math: 32%

Delay induced spatiotemporal patterns in a diffusive intraguild predation model with Beddington-DeAngelis functional response

  • Received: 16 February 2017 Revised: 03 August 2017 Published: 01 June 2018
  • MSC : Primary: 35B32, 35B36, 35K57; Secondary: 92D25, 92D40

  • A diffusive intraguild predation model with delay and Beddington-DeAngelis functional response is considered. Dynamics including stability and Hopf bifurcation near the spatially homogeneous steady states are investigated in detail. Further, it is numerically demonstrated that delay can trigger the emergence of irregular spatial patterns including chaos. The impacts of diffusion and functional response on the model's dynamics are also numerically explored.

    Citation: Renji Han, Binxiang Dai, Lin Wang. Delay induced spatiotemporal patterns in a diffusive intraguild predation model with Beddington-DeAngelis functional response[J]. Mathematical Biosciences and Engineering, 2018, 15(3): 595-627. doi: 10.3934/mbe.2018027

    Related Papers:

    [1] Xin-You Meng, Yu-Qian Wu . Bifurcation analysis in a singular Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting. Mathematical Biosciences and Engineering, 2019, 16(4): 2668-2696. doi: 10.3934/mbe.2019133
    [2] Fang Liu, Yanfei Du . Spatiotemporal dynamics of a diffusive predator-prey model with delay and Allee effect in predator. Mathematical Biosciences and Engineering, 2023, 20(11): 19372-19400. doi: 10.3934/mbe.2023857
    [3] Xiaoying Wang, Xingfu Zou . Pattern formation of a predator-prey model with the cost of anti-predator behaviors. Mathematical Biosciences and Engineering, 2018, 15(3): 775-805. doi: 10.3934/mbe.2018035
    [4] Yan Wang, Minmin Lu, Daqing Jiang . Viral dynamics of a latent HIV infection model with Beddington-DeAngelis incidence function, B-cell immune response and multiple delays. Mathematical Biosciences and Engineering, 2021, 18(1): 274-299. doi: 10.3934/mbe.2021014
    [5] Kalyan Manna, Malay Banerjee . Stability of Hopf-bifurcating limit cycles in a diffusion-driven prey-predator system with Allee effect and time delay. Mathematical Biosciences and Engineering, 2019, 16(4): 2411-2446. doi: 10.3934/mbe.2019121
    [6] Yue Xing, Weihua Jiang, Xun Cao . Multi-stable and spatiotemporal staggered patterns in a predator-prey model with predator-taxis and delay. Mathematical Biosciences and Engineering, 2023, 20(10): 18413-18444. doi: 10.3934/mbe.2023818
    [7] G.A.K. van Voorn, D. Stiefs, T. Gross, B. W. Kooi, Ulrike Feudel, S.A.L.M. Kooijman . Stabilization due to predator interference: comparison of different analysis approaches. Mathematical Biosciences and Engineering, 2008, 5(3): 567-583. doi: 10.3934/mbe.2008.5.567
    [8] Zhihong Zhao, Yan Li, Zhaosheng Feng . Traveling wave phenomena in a nonlocal dispersal predator-prey system with the Beddington-DeAngelis functional response and harvesting. Mathematical Biosciences and Engineering, 2021, 18(2): 1629-1652. doi: 10.3934/mbe.2021084
    [9] Yuhong Huo, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Renji Han . Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor. Mathematical Biosciences and Engineering, 2023, 20(10): 18820-18860. doi: 10.3934/mbe.2023834
    [10] Ming Chen, Menglin Gong, Jimin Zhang, Lale Asik . Comparison of dynamic behavior between continuous- and discrete-time models of intraguild predation. Mathematical Biosciences and Engineering, 2023, 20(7): 12750-12771. doi: 10.3934/mbe.2023569
  • A diffusive intraguild predation model with delay and Beddington-DeAngelis functional response is considered. Dynamics including stability and Hopf bifurcation near the spatially homogeneous steady states are investigated in detail. Further, it is numerically demonstrated that delay can trigger the emergence of irregular spatial patterns including chaos. The impacts of diffusion and functional response on the model's dynamics are also numerically explored.


    1. Introduction

    Competition and predation are two fundamental ecological relationships among species and have been widely studied [1]. Recently, it is been recognized that intraguild predation (IGP), which is a combination of competition and predation, has significant impacts on the distribution, abundance, persistence and evolution of the species involved [2]. As a result, growing attention has been paid to IGP models [3,4,5,6,7,8,9].

    The general framework of IGP described below was established by Holt and Polis [5]

    {˙R(t)=R(φ(R)ρ1(R,N,P)Nρ2(R,N,P)P),˙N(t)=N(e1ρ1(R,N,P)Rρ3(R,N,P)Pm1),˙P(t)=P(e2ρ2(R,N,P)R+e3ρ3(R,N,P)Nm2), (1)

    where R(t),N(t),P(t) represent the densities of basal resource, IG prey and IG predator, respectively. The quantities ρ2(R,N,P)R and ρ3(R,N,P)N are functional responses of the IG predator to the resource and IG prey, respectively; ρ1(R,N,P)R is the functional response of the IG prey to the basal resource; and m1 and m2 are density-independent morality rates. The parameters e1 and e2 are the conversion rates of resource consumption into reproduction for the IG prey and IG predator, respectively; the parameter e3 denotes the conversion rate of the IG predator from its consumption of IG prey; Rφ(R) is recruitment of the basal resource.

    Functional response describes how the consumption rate of individual consumers varies with respect to resource density and is often used to model predator-prey interactions. For IGP models, several functional response functions have been studied. For instance, Velazquez et al. [10] and Hsu et al. [11] investigated the case with a linear functional response. Abrams and Fung [12] considered Holling type-II functional response. Verdy and Amarasekare [13] and Freeze et al. [14] investigated Holling type-II and ratio-dependent functional responses, respectively. Kang and Wedekin [15] considered Holling-III functional response.

    Note that the reproduction of predator following the consumption of prey is not instantaneous, but rather is mediated by some reaction-time lag required for gestation. Time delay plays an important role in ecology and it can induce very complex dynamical behaviors [16,17,18,19,20,21,22]. For IGP models, it has been shown that a time delay greatly impacts their dynamics [23,24]. In [24], Shu et al. investigated the complex dynamics of the following IGP model

    {˙R(t)=rR(t)(1R(t)K)c1R(t)N(t)c2R(t)P(t),˙N(t)=e1c1R(tτ)N(tτ)c3N(t)P(t)m1N(t),˙P(t)=e2c2R(t)P(t)+e3c3N(t)P(t)m2P(t), (2)

    where r is the growth rate of R in the absence of N and P, K is the carrying capacity of resource. c1 is the predation rate of IG prey for resource, c2 is the predation rate of IG predator for resource, c3 is the consumption rate of IG predator to IG prey and all other parameters have the same meanings as those in (1).

    Note that for each species, individuals tend to migrate towards regions with lower population densities. Hence the species are distributed over space and interact with each other within their spatial domains. To take spatial effects into consideration, reaction diffusion equations become a natural choice [25,26,27,28,29,30,31,32,33,34,35,36]. In this work, we consider a reaction diffusion IGP model with delay and Beddington-DeAngelis functional response.

    Suppose ΩRn is a bounded domain with smooth boundary Ω. Let R(t,x), N(t,x), P(t,x) represent the densities of basal resource, IG prey and IG predator at time t and location x, respectively. The basal resource is assumed to grow logistically. We assume the basal resource is consumed by the IG prey at a rate c1R(t,x)N(t,x), and the IG prey is consumed by the IG predator is c3N(t,x)P(t,x) at time t and location x. In this paper, we will assume the functional response takes the Beddington-DeAngelis (B-D) form, i.e., the consumption of the resource by the IG predator is characterized by c2P(t,x)R(t,x)1+a1R(t,x)+a2P(t,x). The reproduction of IG prey from consuming the basal resource is e1c1R(tτ,x)N(tτ,x), where the time-lag parameter is introduced in a manner analogous to the treatment in [24]. We further assume the populations cannot cross the boundary of Ω. Our model then reads as

    {R(t,x)t=˜d1ΔR+R(r(1RK)c1Nc2P1+a1R+a2P),t>0,xΩN(t,x)t=˜d2ΔN+e1c1N(tτ,x)R(tτ,x)c3NPm1N,t>0,xΩ,P(t,x)t=˜d3ΔP+P(e2c2R1+a1R+a2P+e3c3Nm2),t>0,xΩ,Rν=Nν=Pν=0,t>0,xΩ,R(t,x)=˜ϕ1(t,x)0,(t,x)[τ,0]×Ω,N(t,x)=˜ϕ2(t,x)0,(t,x)[τ,0]×Ω,P(t,x)=˜ϕ3(t,x)0,(t,x)[τ,0]×Ω, (3)

    where ˜d1,˜d2,˜d3 denote the diffusion coefficients of the three species, respectively; Δ is the Laplacian operator in the n dimensional space, ν is the outward unit normal vector on Ω, and the homogeneous Neumann boundary conditions reflect the situation where the population cannot across the boundary of Ω. The meanings and units of the parameters of model (3) are summarized in Table 1.

    Table 1. Parameters definitions in model (3) and their units, where [resource] indicates basal resource density, [IG prey] indicates IG prey density, and [IG predator] indicates IG predator density.
    Symbol Parameter Definition Units
    r Basal resource intrinsic growth rate [time]1
    K Basal resource carrying capacity [Basal resource density]
    c1 Predation rate of IG prey on resource [IG prey]1 [time]1
    c2 Predation rate of IG predator on resource [IG predator]1[time]1
    c3 Predation rate of IG predaotr on IG prey [IG preys][IG predator]1
    [time]1
    e1 Conversion rate from resource to IG prey [IG preys][resource]1
    e2 Conversion rate from resource to IG predator [IG predators][resource]1
    e3 Conversion rate from IG prey to IG predator [IG predators][IG prey]1
    a1 [Half saturation constant]1 [resource]1
    a2 [Half saturation constant]1 [IG predator]1
    m1 Mortality rate of IG prey [time]1
    m2 Mortality rate of IG predator [time]1
    ˜d1 Diffusion coefficient of resource [length]2[time]1
    ˜d2 Diffusion coefficient of IG prey [length]2[time]1
    ˜d3 Diffusion coefficient of IG predatior [length]2[time]1
    L The size of spatial domain Ω [length]
     | Show Table
    DownLoad: CSV

    For rescalling, we let

    u1(t,x)=R(t,x)K,u2(t,x)=c1N(t,x)r,u3(t,x)=c2P(t,x)r,γ1=m1r,γ2=m2r,β1=e1c1Kr,β2=e2c2Kr,α=c3c2,β=e3c3c1,b=a1K,c=a2rc2,d1=˜d1rL2,d2=˜d2rL2,d3=˜d3rL2,ˆx=xL,ˆt=tr,ˆτ=τr,ϕ1(t,x)=˜ϕ1(t,x)K,ϕ2(t,x)=c1˜ϕ2(t,x)r,ϕ3(t,x)=c2˜ϕ3(t,x)r.

    Then model (1.3) becomes

    {u1(t,x)t=d1Δu1(t,x)+f1(u,v),t>0,xΩ,u2(t,x)t=d2Δu2(t,x)+f2(u,v),t>0,xΩ,u3(t,x)t=d3Δu3(t,x)+f3(u,v),t>0,xΩ,u1(t,x)ν=u2(t,x)ν=u3(t,x)ν=0,t>0,xΩ,u1(t,x)=ϕ1(t,x)0,(t,x)[τ,0]×Ω,u2(t,x)=ϕ2(t,x)0,(t,x)[τ,0]×Ω,u3(t,x)=ϕ3(t,x)0,(t,x)[τ,0]×Ω, (4)

    where u=(u1(t,x),u2(t,x),u3(t,x)),v=(u1(tτ,x),u2(tτ,x)) and

    f1(u,v)=u1(t,x)(1u1(t,x)u2(t,x)u3(t,x)1+bu1(t,x)+cu3(t,x)),
    f2(u,v)=β1u1(tτ,x)u2(tτ,x)αu2(t,x)u3(t,x)γ1u2(t,x),
    f3(u,v)=u3(t,x)(β2u1(t,x)1+bu1(t,x)+cu3(t,x)+βu2(t,x)γ2).

    Throughout the paper, we denote ˉΩ=ΩΩ, DT=(0,T]×Ω,ˉDT=[0,T]×ˉΩ, Q0=[τ,0]×Ω,ˉQ0=[τ,0]×ˉΩ, QT=[τ,T]×Ω,ˉQT=[τ,T]×ˉΩ. Denote by Cγ(DT) the space of Hölder continuous functions in DT with exponent γ(0,1). The space of continuous functions in ˉDT is denoted by C(ˉDT). For vector-value functions we use the product spaces

    C(ˉDT)C(ˉDT)×C(ˉDT)×C(ˉDT),Cγ(ˉDT)Cγ(ˉDT)×Cγ(ˉDT)×Cγ(ˉDT).

    Denote

    X={(u1,u2,u3)TH2(Ω)×H2(Ω)×H2(Ω):u1ν=u2ν=u3ν=0onΩ}

    with the usual inner product ,.

    The rest of the paper is organized as follows. In Section 2, we study the existence and uniqueness of solution of (4) and estimate the solution's priori bounds. In Section 3, we discuss the existence of nonnegative spatially homogeneous steady states. In Section 4, we carry out stability analysis and Hopf bifurcation analysis about the unique positive spatially homogeneous steady state of System (4). Numerical simulations are presented in Section 5 to illustrate the impacts of delay, diffusion and the functional response on the dynamics of our IGP model. We conclude this paper with a brief summary and discussion in Section 6.


    2. Existence of solution of System (4) and priori bound estimation

    Theorem 2.1. Consider System (4), we have the following conclusions.

    (ⅰ) Given any initial condition (ϕ1(t,x),ϕ2(t,x),ϕ3(t,x))Cγ(Q0) with

    0ϕi(t,x)Li,(t,x)Q0,i=1,2,3, (5)

    where Li,i=1,2,3 are positive constants satisfying

    1L1γ1β1,L2<γ2β,L31c[β2L1γ2βL2bL11], (6)

    System (4) admits a unique solution (u1(t,x),u2(t,x),u3(t,x)) satisfying

    0ui(t,x)Li,fort>0,xΩ,i=1,2,3.

    (ⅱ) For any solution (u1(t,x),u2(t,x),u3(t,x)) of System (4), it holds true that

    lim suptu1(t,x)1,lim suptΩu2(t,x)dxJ1,lim suptΩu3(t,x)dxJ2

    where J1=(β14γ1+β1)|Ω|, J2=(β24κ+β1βακJ1+β2)|Ω| with κ=min{γ1,γ2}.

    Furthermore, if d1=d2=d3 and τ=0, then for any xˉΩ,

    lim suptu2(t,x)β2α4κβ+β1κ(β14γ1+β1)+β2αβ,lim suptu3(t,x)β24κ+β1βακ(β14γ1+β1)+β2.

    Proof. Note that fi(u,v) is mixed quasi-monotone in a subset Λ×Λ of R3×R2 for each i=1,2,3, we can apply [27,Therorem 2.2] to establish the existence and uniqueness of solutions to System (4). To this end, we first need to construct a pair of coupled upper and lower solutions of System (4), which we denote by (˜u1,˜u2,˜u3) and (ˆu1,ˆu2,ˆu3), respectively. In view of [27,Definition 2.1], the required upper and lower solutions should satisfy the boundary-initial inequalities and the following differential inequalities

    {˜u1td1Δ˜u1+˜u1(1˜u1ˆu2ˆu31+b˜u1+cˆu3)~u2td2Δ˜u2+β1˜u1˜u2αˆu3˜u2γ1˜u2,˜u3td3Δ˜u3+β2˜u1˜u31+b˜u1+c˜u3+β˜u2˜u3γ2˜u3 (7)

    and

    {ˆu1td1Δˆu1+ˆu1(1ˆu1˜u2˜u31+bˆu1+c˜u3)ˆu2td2Δˆu2+β1ˆu1ˆu2α˜u3ˆu2γ1ˆu2ˆu3td3Δˆu3+β2ˆu1ˆu31+bˆu1+cˆu3+βˆu2ˆu3γ2ˆu3. (8)

    Take (ˆu1,ˆu2,ˆu3)=(0,0,0) and (˜u1,˜u2,˜u3)=(L1,L2,L3). Clearly, ˜uiν=000=ˆuiν. It follows from (5) and (6) that the pair (ˆu1,ˆu2,ˆu3)=(0,0,0) and (˜u1,˜u2,˜u3)=(L1,L2,L3) are coupled upper and lower solutions of System (4). It is easy to check that fi(u,v)(i=1,2,3) satisfy the Lipschitz condition for 0uiMi,0viLi,i=1,2,3, and we denote the Lipschitz constants by Ki,i=1,2,3. By [27,Theorem 2.2], System (4) admits a unique global solution (u1(t,x),u2(t,x),u3(t,x)), which satisfies

    (0,0,0)(u1(t,x),u2(t,x),u3(t,x))(L1,L2,L3),t0,xΩ.

    This completes the proof of ().

    Next we establish the priori bound of solutions to System (4). To estimate u1(t,x), we observe that u1(t,x) satisfies

    {u1(t,x)td1Δu1(t,x)+u1(t,x)(1u1(t,x)),t>0,xΩ,u1(t,x)n=0,t>0,xΩ.

    It follows from the standard comparison principle [37,Lemma 3.4.2] of parabolic equations that lim suptu1(t,x)1. Thus for any ε>0, there exists a T1>0 such that u1(t,x)1+ε for tT1.

    To estimate the priori bounds of u2(t,x) and u3(t,x), we set

    U1(t)=Ωu1(t,x)dx,U2(t)=Ωu2(t,x)dx,U3(t)=Ωu3(t,x)dx.

    Then

    dU1(t)dt=Ωu1tdx=Ωd1Δu1dx+Ω[u1(t,x)(1u1(t,x)u2(t,x)u3(t,x)1+bu1(t,x)+cu3(t,x))]dx,
    dU2(t)dt=Ωu2tdx=Ωd2Δu2dx+Ω[β1u1(tτ,x)u2(tτ,x)αu2(t,x)u3(t,x)γ1u2(t,x)]dx,
    dU3(t)dt=Ωu3tdx=Ωd3Δu3dx+Ω[u3(t,x)(β2u1(t,x)1+bu1(t,x)+cu3(t,x)+βu2(t,x)γ2)]dx.

    From the Neumann boundary conditions, we further obtain

    d(β1U1(t)+U2(t+τ))dt=β1Ωu1tdx+Ωu2(t+τ,x)tdx=β1Ω(u1u21)dxΩβ1u1u31+bu1+cu3dxΩαu2(t+τ,x)u3(t+τ,x)dxΩγ1u2(t+τ,x)dxβ14|Ω|+γ1β1(1+ε)|Ω|γ1(β1U1(t)+U2(t+τ)),t>T1.

    By the comparison principle, we have

    lim supt(β1U1(t)+U2(t+τ))β14γ1|Ω|+β1|Ω|J1.

    Similarly, there exists T2>T1 such that Ωu2(t,x)dx=U2(t)J1+ε for tT2. Thus, for tT2+τ, we have

    d(β2U1(t)+βαU2(t)+U3(t))dt=β2Ω(u1u21)dx+Ωβ1βαu1(tτ,x)u2(tτ,x)dxβγ1αΩu2dxγ2Ωu3dxβ24|Ω|+β1βα(1+ε)Ωu2(tτ,x)dxβγ1αU2γ2U3β24|Ω|+β1βα(1+ε)(J1+ε)|Ω|+β2κ(1+ε)|Ω|κ(β2U1+βαU2+U3),

    where κ=min{γ1,γ2}. This implies that

    lim supt(β2U1+βαU2+U3)β24κ|Ω|+β1βακJ1|Ω|+β2|Ω|J2.

    and hence lim suptΩu3(t,x)dx=lim suptU3(t)J2.

    For the case with d1=d2=d3=d and τ=0, we can similarly show that for any ε>0, there exists T3>T1 such that 0u1(t,x)1+ε and 0u2(t,x)β14γ1+β1 for t>T3. Moreover, let S(t,x)=β2u1(t,x)+βαu2(t,x)+u3(t,x). Then

    {St=dΔS+β2(u1u21)β2u1u2+ββ1αu1u2βγ1αu2γ2u3,t>T3,xΩ,Sν=0,t>T3,xΩ,S(T3,x)=β2u1(T3,x)+βαu2(T3,x)+u3(T3,x),xΩ.

    Thus for t>T3, we have

    β2(u1u21)β2u1u2+ββ1αu1u2βγ1αu2γ2u3β24+β1βα(1+ε)(β14γ1+β1+ε)+β2κ(1+ε)κS.

    Consider the system

    {Wt=dΔW+β24+β1βα(1+ε)(β14γ1+β1+ε)+β2κ(1+ε)κW,t>T3,xΩWν=0,t>T3,xΩ,W(T3,x)=β2u1(T3,x)+βαu2(T3,x)+u3(T3,x),xΩ.

    It follows from [37,Theorem 2.4.6] that the solution W(t,x) satisfies

    limtW(t,x)=β24κ+β1βακ(1+ε)(β14γ1+β1+ε)+β2(1+ε).

    The comparison argument implies that

    lim suptu2(t,x)lim suptαβS(t,x)β2α4κβ+β1κ(β14γ1+β1)+β2αβ

    and

    lim suptu3(t,x)lim suptS(t,x)β24κ+β1βακ(β14γ1+β1)+β2.

    This completes the proof.


    3. Spatially homogeneous steady states of System (4)

    Same as in [24], we denote by Ri=βiγi=eiciKmi(i=1,2) the reproduction numbers for the IG prey and IG predator, respectively. Consider (4), we easily have the following existence results on trivial and semi-trivial spatially homogeneous steady states.

    Proposition 1. (ⅰ) The trivial steady state E0=(0,0,0) always exists.

    (ⅱ) There is a weakly semi-trivial steady state in the absence of IG Prey and IG Predator E1=(1,0,0).

    (ⅲ) The IG Prey-only strong semi-trivial steady state E10:=(1R1,11R1,0) exists if and only if R1>1.

    (ⅳ) The IG Predator-only strong semi-trivial steady state E01:=(ˆu1,0,R2(1ˆu1)ˆu1) exists if and only if R2>1+b, where

    ˆu1=(R2cR2b)+(R2cR2b)2+4cR22cR2.

    System (4) admits a positive steady state E:=(u1,u2,u3) if E is a positive solution to the following three equations:

    1u1u2u31+bu1+cu3=0, (9)
    β1u1αu3γ1=0, (10)
    β2u11+bu1+cu3+βu2γ2=0. (11)

    It follows from (10) that

    u1=αβ1u3+γ1β1. (12)

    Combining (9), (10) and (12), we obtain

    u32+pu3+q=0, (13)

    where

    p=cγ2β21+bαβ1(γ2β)+αβ1(ββ2)+cββ1(γ1β1)+ββ21+2bαβγ1αβ(bα+cβ1),q=β21(γ2β)+bβ1γ1(γ2β)+β1γ1(ββ2)+β(bγ21β21)αβ(bα+cβ1).

    For the distribution of roots of Eq. (13), we have the following results.

    Lemma 3.1. (ⅰ) If p<0,p24q>0 and q>0, then Eq. (13) has two positive roots u3±=p±p24q2.

    (ⅱ) If p<0 and p24q=0, then Eq. (13) has a unique positive root u30=p2.

    (ⅲ) If q<0, then Eq. (13) has a unique positive root u3=u3+.

    Remark 1. Set R3=βγ2, then q<0 provided that R1>b and R2>R3>1.

    According to Lemma 3.1, the following proposition is valid.

    Proposition 2. (ⅰ) When p24q>0,p<0,q>0 and

    R2<min{(bα+cβ1)(p+p24q)+2(bγ1+β1)α(p+p24q)+2γ1,(bα+cβ1)(pp24q)+2(bγ1+β1)α(pp24q)+2γ1}

    System (4) has two positive constant steady states E+=(u1+,u2+,u3+) and

    E=(u1,u2,u3),whereu1+=αu3++γ1β1,u2+=γ2ββ2β×αu3++γ1(bα+cβ1)u3++bγ1+β1,u3+=p+p24q2andu1=αu3+γ1β1,u2=γ2ββ2β×αu3+γ1(bα+cβ1)u3+bγ1+β1,u3=pp24q2.

    (ⅱ) When p24q=0,p<0 and R2<(bα+cβ1)p+2(bγ1+β1)2γ1αp, E+ and E merge, denoted by E0=(u10,u20,u30), where u10=pα2β1+γ1β1,u20=γ2ββ2β2γ1pα2(bγ1+β1)(bα+cβ1)p,u30=p2.

    (ⅲ) When q<0 and R2<(bα+cβ1)(p+p24q)+2(bγ1+β1)α(p+p24q)+2γ1, System (4) has only one positive constant steady state E=(u1,u2,u3), where u1=u1+,u2=u2+,u3=u3+.

    Remark 2. There exist parameter values such that Proposition 3.4 holds. For example, choosing α=0.68,β=0.9,β1=1.9,β2=1.8,γ1=0.2,γ2=0.76,,b=2.5,c=10. A direct calculation yields only one positive steady state

    E=(0.1891,0.7562,0.2344).

    4. Dynamics of System (4)

    Let 0=μ1<μ2μ3 be the eigenvalues of Δ on Ω under no-flux boundary conditions, and E(μi) be the eigenspace corresponding to μi with multiplicity mi1,iN{1,2,}. Set Xij:={eϕij:eR3}, where {ϕij} is an orthonormal basis of E(μi) for j=1,2,,dimE(μi). For X:={wC1(ˉΩ):w1ν=w2ν=w3ν=0onΩ}, we have the following lemma from [37].

    Lemma 4.1.

    X=i=1Xi,whereXi=dimE(μi)j=1Xij.

    4.1. Stability of E0 and E1

    In the following, we consider the stability of E0 and E1.

    Theorem 4.2.

    (ⅰ) The trivial steady state E0 is always unstable.

    (ⅱ) The semi-trivial steady state E1 is locally asymptotically stable if R1<1 and R2<1+b and unstable if either R1>1 or R2>1+b.

    Proof. Linearizing System (4) at a constant steady state u=(u1,u2,u3) gives

    ut=(DΔ+ˆJ1)u+ˆJ2uτ, (14)

    where D=diag{d1,d2,d3},u=(u1(t,x),u2(t,x),u3(t,x)),uτ=(u1(tτ,x),u2(tτ,x),u3(tτ,x) and

    ˆJ1=[12u1u2u3+cu32(1+bu1+cu3)2u1(1+bu1)u1(1+bu1+cu3)20αu3γ1αu2β2(1+cu3)u3(1+bu1+cu3)2βu3β2u1(1+bu1)(1+bu1+cu3)2+βu2γ2],ˆJ2=[000β1u2β1u10000].

    From Lemma 4.1, we know that the eigenvalues of the System (4) is confined on the subspace Xi, and λ is an eigenvalue of (14) on Xi if and only if it is an eigenvalue of the matrix μiD+J, where J=ˆJ1+ˆJ2eλτ. Then the characteristic equation of (14) is

    det(λI3+μiDJ)=0, (15)

    where I3 stands for the 3×3 identity matrix.

    () If u=E0, then we obtain the following characteristic equation

    (λ+d1μi1)(λ+d2μi+γ1)(λ+d3μi+γ2)=0,

    which gives three sets of eigenvalues, namely, λ1i=d1μi+1,λ2i=d2μiγ1,λ3i=d3μiγ2,i=1,2,. Clearly, for i=1, λ11=1>0. From [40,Corollary 1.11], the trivial steady state E0 is unstable.

    () If u=E1, then we obtain the following characteristic equation

    (λ+d1μi+1)(λ+d2μiβ1eλτ+γ1)(λ+d3μiβ21+b+γ2)=0,

    which gives the eigenvalues λ1i=d1μi1,λ3i=d3μi+β21+bγ2 and λ2i is determined by λ2i+d2μi+γ1β1eλ2iτ=0. Clearly, λ1i<0 for all μi. If R2<1+b, we get λ3i<0 for all μi. If R1<1, then we have β1<γ1+d2μi for all μi. Thus it follows from [24,Lemma 6] that the eigenvalues λ2i have negative real parts for all μi. It follows from [40,Corollary 1.11] that the equilibrium E1 is locally asymptotically stable for R1<1 and R2<1+b. If R1>1 then there exists μ1=0 such that d2μ1+γ1<β1, so it follows from the [24,Lemma 6] that at least one of the eigenvalues λ2i has a positive real part. If R2>1+b then there exists μ1=0 such that λ31>0. From [40,Corollary 1.11], the steady state E1 is unstable in either case.

    Theorem 4.3. Suppose that R1<1,R2<1+b and 1γ2β>1c. Then E1 is globally asymptotically stable.

    Proof. Let (ˆu1,ˆu2,ˆu3)=(ε,0,0) and (˜u1,˜u2,˜u3)=(M1,M2,M3), where M1=1+ε<γ1β1,M2=γ2βε>0, M3=1c(β2M1γ2βM2bM11)=1c(β2(1+ε)βεb(1+ε)1) >0 and ε is arbitrary small positive constant.

    We claim that (ˆu1,ˆu2,ˆu3)=(ε,0,0) and (˜u1,˜u2,˜u3)=(M1,M2,M3) are also coupled upper and lower solutions of System (4). In fact, it follows from 1γ2β>1c that 1γ2β>M31+cM3, and thus we get ε(1ε(γ2βε)M31+bε+cM3)0. Hence, we obtain ˆu1(1ˆu1˜u2˜u31+bˆu1+c˜u3)0. It is easy to check that (ˆu1,ˆu2,ˆu3)=(ε,0,0) and (˜u1,˜u2,˜u3)=(M1,M2,M3) satisfy the differential inequalities in (7) and (8). Thus the claim holds. Define the iterated sequences (ˉu(m)1,ˉu(m)2,ˉu(m)3) and (u_(m)1,u_(m)2,u_(m)3) as follows:

    ˉu(m)1=ˉu(m1)1+1K1[ˉu(m1)1(1ˉu(m1)1u_(m1)2u_(m1)31+bˉu(m1)1+cu_(m1)3)],u_(m)1=u_(m1)1+1K1[u_(m1)1(1u_(m1)1ˉu(m1)2ˉu(m1)31+bu_(m1)1+cˉu(m1)3)],ˉu(m)2=ˉu(m1)2+1K2[ˉu(m1)2(β1ˉu(m1)1αu_(m1)3γ1)],u_(m)2=u_(m1)2+1K2[u_(m1)2(β1u_(m1)1αˉu(m1)3γ1)],ˉu(m)3=ˉu(m1)3+1K3[ˉu(m1)3(β2ˉu(m1)11+bˉu(m1)1+cˉu(m1)3+βˉu(m1)2γ2)],u_(m)3=u_(m1)3+1K3[u_(m1)3(β2u_(m1)11+bu_(m1)1+cu_(m1)3+βu_(m1)2γ2)],

    where m=1,2,,(ˉu(0)1,ˉu(0)2,ˉu(0)3)=(M1,M2,M3), (u_(0)1,u_(0)2,u_(0)3)=(ε,0,0) and Ki,i=1,2,3 are the Lipschitz constants in Theorem 2.1. It is easy to see that f(u,v)(f1(u,v),f2(u,v),f3(u,v)) is a C1 function of u,v and is mixed quasi-monotone in a subset Λ×Λ of R3×R2. We can deduce from the induction method that

    ˆuu_(m)u_(m+1)ˉu(m+1)ˉu(m)˜u. (16)

    It follows from (16) that the limits

    limmˉu(m)1=ˉu1,limmˉu(m)2=ˉu2,limmˉu(m)3=ˉu3,limmu_(m)1=u_1,limmu_(m)2=u_2,limmu_(m)3=u_3 (17)

    exist and satisfy the following equations

    f1(ˉu1,u_2,u_3)=0,f2(ˉu1,ˉu2,u_3)=0,f3(ˉu1,ˉu2,ˉu3)=0, (18)
    f1(u_1,ˉu2,ˉu3)=0,f2(u_1,u_2,ˉu3)=0,f3(u_1,u_2,u_3)=0, (19)

    where

    ˆu=(ˆu1,ˆu2,ˆu3),u_(m)=(u_(m)1,u_(m)2,u_(m)3),u_(m+1)=(u_(m+1)1,u_(m+1)2,u_(m+1)3),
    ˉu(m+1)=(ˉu(m+1)1,ˉu(m+1)2,ˉu(m+1)3),ˉu(m)=(ˉu(m)1,ˉu(m)2,ˉu(m)3),˜u=(˜u1,˜u2,˜u3).

    Since u_(0)2=0 and u_(0)3=0, we get u_(m)2=0 and u_(m)3=0. It follows from (17) that u_2=u_3=0. Thus, it follows from the first equality of (18) that \bar u_1 = 1. Substituting \bar u_1 = 1 into the second equality of (18) and noting that \mathcal{R}_1<1 yields \bar u_2 = 0. In view of the third equality of (18), we have \bar u_3\left(\frac{\beta_2}{1+b+c\bar u_3}-\gamma_2\right) = 0. Since \mathcal{R}_2<1+b, we obtain \frac{\beta_2}{1+b+c\bar u_3}-\gamma_2<0. This implies that \bar u_3 = 0. Hence, it follows from the first equality of (19) that \underline u_1 = 1. Therefore, we have (\bar u_1, \bar u_2, \bar u_3) = (1, 0, 0) = (\underline u_1, \underline u_2, \underline u_3). It follows from [27,Theorem 3.2] that for any initial function {\bf\phi} = (\phi_1(t,x), \phi_2(t,x), \phi_3(t,x)) satisfying {\bf{\widehat u}} \leq {\bf{\phi}} \leq {\bf{\widetilde u}} in Q_0, the solution {\bf{u}}\equiv(u_1(t,x), u_2(t,x), u_3(t,x)) of System (4) satisfies \lim\limits_{t\rightarrow\infty}{\bf{u}} = (1,0,0). This completes the proof.


    4.2. Stability of E_{10} and E_{01}

    Next, we consider the stability of the two strong semi-trivial steady states: E_{10} and E_{01}. For the IG prey-only strong semi-trivial steady state E_{10}, we have the following result.

    Theorem 4.4. Consider System (4) with \mathcal{R}_1>1.

    (ⅰ) If -\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2<0, then E_{10} is unstable.

    (ⅱ) If -\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2>0 and 1<\mathcal{R}_1\leq 3, then E_{10} is locally asymptotically stable for all \tau\geq 0.

    (ⅲ) If -\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2>0 and \mathcal{R}_1>3, then there exists \tau_{00}>0 such that E_{10} is locally asymptotically stable for \tau\in[0,\tau_{00}) and is unstable for \tau>\tau_{00}. Further there exists a sequence of delays, \{\tau_i^j\}_{j = 0}^{+\infty} for i = 1,2,\cdots, N_1, at which E_{10} undergoes Hopf bifurcations. Here, \tau_{00} and \tau_i^j are given in the proof of this theorem.

    Proof. For E_{10}, the characteristic equation is

    m_1(\lambda)[g_1(\lambda)+h_1(\lambda)e^{-\lambda\tau}] = 0, \label{Eq4.7} (20)

    with m_1(\lambda) = \lambda+d_3\mu_i-\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2, h_1(\lambda) = -(\lambda+d_1\mu_i+\frac{1}{\mathcal{R}_1})\frac{\beta_1}{\mathcal{R}_1}+\frac{\beta_1}{\mathcal{R}_1}(1-\frac{1}{\mathcal{R}_1}), and g_1(\lambda) = (\lambda+d_1\mu_i+\frac{1}{\mathcal{R}_1})(\lambda+d_2\mu_i+\gamma_1). Note that -\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2<0, there exists \mu_1 = 0 such that -d_3\mu_1+\frac{\beta_2}{\mathcal{R}_1+b}+\beta (1-\frac{1}{\mathcal{R}_1})-\gamma_2>0 holds. Therefore, m_1(\lambda) has at least one zero with a positive real part and the characteristic Eq. (20) has at least one positive root with a positive real part. The proof of (ⅰ) is complete.

    Denote

    \bar{E}_1 = d_1\mu_i+\frac{1}{\mathcal{R}_1}>0, \bar L_1 = d_2\mu_italic>0, \quad \quad \bar J_1 = \gamma_1(1-\frac{1}{\mathcal{R}_1})>0.

    Then we have

    \label{Eq4.8} g_1(\lambda)+h_1(\lambda) = \lambda^2+(\bar E_1+\bar L_1)\lambda+\bar E_1\bar L_1+\bar J_1. (21)

    Since \bar E_1+\bar L_1>0 and \bar E_1\bar L_1+\bar J_1>0, it is easy to see that Eq. (21) has no positive zeros for all \mu_i.

    Define

    \begin{split} \label{Eq4.9} G(u)\equiv&\mid g_1(i\sqrt{u})\mid^2-\mid h_1(i\sqrt u)\mid^2\\ = &u^2+(\bar L_1^2+2\bar L_1\gamma_1+\bar E_1^2)u+\bar E_1^2\bar L_1^2+2\bar E_1^2\bar L_1\gamma_1-\bar J_1^2+2\bar E_1\gamma_1\bar J_1. \end{split} (22)

    Since \bar L_1^2+2\bar L_1\gamma_1+\bar E_1^2 is positive for all \mu_i, if \mathcal{F}(\mu_i)\equiv \bar E_1^2\bar L_1^2+2\bar E_1^2\bar L_1\gamma_1-\bar J_1^2+2\bar E_1\gamma_1\bar J_1>0 then G(u) has no positive zeros.

    Next, we discuss the distribution of positive zeros of G(u). Clearly \mathcal{F}(0) = \frac{\beta_1^2}{\mathcal{R}_1^2}(1-\frac{1}{\mathcal{R}_1})(\frac{3}{\mathcal{R}_1}-1). If 1<\mathcal{R}_1\leq 3, then \mathcal{F}(0)\geq0. Since \mathcal{F}(\mu_i) = (d_1\mu_i+\frac{1}{\mathcal{R}_1})^2d_2^2\mu_i^2+2(d_1\mu_i+\frac{1}{\mathcal{R}_1})^2d_2\mu_i\gamma_1+\frac{\beta_1}{\mathcal{R}_1}(1-\frac{1}{\mathcal{R}_1})[2(d_1\mu_i+\frac{1}{\mathcal{R}_1})\frac{\beta_1}{\mathcal{R}_1}-\frac{\beta_1}{\mathcal{R}_1}(1-\frac{1}{\mathcal{R}_1})] is increasing in \mu_i, we obtain \mathcal{F}(\mu_i)\geq 0 for all \mu_i. Thus, G(u) has no positive zeros for all \mu_i. It follows from [24,Lemma 11] that E_{10} is locally asymptotically stable for all \tau\geq 0. This completes the proof of ().

    If \mathcal{R}_1>3, then \mathcal{F}(0)<0. Since \mathcal{F}(\mu_i) is increasing in \mu_i and \lim\limits_{i\rightarrow\infty}\mathcal{F}(\mu_i) = \infty, there exists a constant n\in\mathbb{N} such that

    \mathcal{F}(\mu_i)\geq 0 \mbox{ for }\quad italic>N_1, \mbox{ and } \mathcal{F}(\mu_i)<0 \mbox{ for }\quad i\leq N_1.

    This implies that (22) has no positive root for italic>N_1, has N_1 positive roots for i\leq N_1, denoted by u_1, u_2, \cdots,u_{N_1}. From (22), we get u_i = \omega_i^2 = \frac{-Tr_i+\sqrt{Tr_i^2-4\delta_i}}{2}, where Tr_i = \bar L_1^2+2\bar L_1\gamma_1+\bar E_1^2, \delta_i = \bar E_1^2\bar L_1^2+2\bar E_1^2\bar L_1\gamma_1-\bar J_1^2+2\bar E_1\bar J_1\gamma_1. Similar to the argument of [24], we obtain

    \tau_{i}^j = \frac{1}{\omega_{i}}\left\{\arccos\left(\frac{\mathcal{B}_{1i}} {\sqrt{\mathcal{B}_{1i}^2+\mathcal{C}_{1i}^2}}\right)+2j\pi\right\}, \quad i = 1,2,\cdots,N_1, \quad j = 0,1,2,\cdots,

    where

    \begin{split} \mathcal{B}_{1i} = &(\frac{\beta_1}{\mathcal{R}_1})d_1\mu_i+\frac{\beta_1}{\mathcal{R}_1^2}-\frac{\beta_1}{\mathcal{R}_1} (1-\frac{1}{\mathcal{R}_1})((d_1\mu_i+\frac{1}{\mathcal{R}_1})(d_2\mu_i+\gamma_1)-\omega_i^2)+\\ &\frac{\beta_1}{\mathcal{R}_1}\omega_i^2(d_1\mu_i+d_2\mu_i+\frac{1}{\mathcal{R}_1}+\gamma_1), \end{split}
    \begin{split} \mathcal{C}_{1i} = &-(\frac{\beta_1}{\mathcal{R}_1} d_1\mu_i+\frac{\beta_1}{\mathcal{R}_1^2}-\frac{\beta_1}{\mathcal{R}_1} (1-\frac{1}{\mathcal{R}_1}))(d_1\mu_i+d_2\mu_i+\frac{1}{\mathcal{R}_1}+\gamma_1)\omega_i+\\ &\frac{\beta_1}{\mathcal{R}_1}\omega_i((d_1\mu_i+\frac{1}{\mathcal{R}_1})(d_2\mu_i+\gamma_1)-\omega_i^2). \end{split}

    Denote

    \tau_{00} = \min\limits_{i = 1,2,\cdots,N_1}\{\tau_i^0\}.

    It follows easily from -\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2>0 that d_3\mu_i-\frac{\beta_2}{\mathcal{R}_1+b}-\beta (1-\frac{1}{\mathcal{R}_1})+\gamma_2>0 for all \mu_i. Then all zeros of m_1(\lambda) have negative real parts. Since all zeros of g_1(\lambda)+h_1(\lambda) have negative real parts, we conclude that all zeros of Eq. (20) have negative real parts for \tau = 0. Since \tau_{00} is the minimum value of \tau so that Eq. (20) has purely imaginary roots, applying Lemma 1.1 of [39], we get E_{10} is locally asymptotically stable for \tau\in[0,\tau_{00}) and is unstable for \tau>\tau_{00}. From (22), we obtain G^\prime(u_i)>0 for i = 1,2,\cdots, N_1. Thus, E_{10} undergoes a sequence of Hopf bifurcations as \tau increases through \tau_i^j for i = 1,2,\cdots,N_1, j = 0,1,2,\cdots. This completes the proof of ().

    For the IG predator-only strong semi-trivial steady state E_{01}, we have the following result.

    Theorem 4.5. Consider System (4) with \mathcal{R}_2>1+b.

    (ⅰ) If \alpha\mathcal{R}_2\hat u_1(1-\hat u_1)+\gamma_1<\beta_1 \hat u_1, then E_{01} is unstable.

    (ⅱ) If \alpha\mathcal{R}_2\hat u_1(1-\hat u_1)+\gamma_1>\beta_1 \hat u_1 and b<\hat u_1(\mathcal{R}_2+b), then E_{01} is locally asymptotically stable for all \tau\geq 0. Here \hat u_1 is defined in Proposition 3.1.

    Proof. For {\bf{u}}^\diamond = (\hat{u}_1, 0, \mathcal{R}_2(1-\hat{u}_1)\hat{u}_1), we get the characteristic equation

    \label{Eq4.10} m_2(\lambda)g_2(\lambda) = 0, (23)

    with

    m_2(\lambda) = (\lambda+d_2\mu_i+\alpha\hat u_3+\gamma_1-\beta_1\hat u_1e^{-\lambda\tau}),

    and

    \begin{array}{l} {g_2}(\lambda ) = (\lambda + {d_1}{\mu _i} + {{\hat u}_1} - \frac{{b{{\hat u}_1}{{\hat u}_3}}}{{{{(1 + b{{\hat u}_1} + c{{\hat u}_3})}^2}}})(\lambda + {d_3}{\mu _i} + {\gamma _2} - \frac{{{\beta _2}{{\hat u}_1}(1 + b{{\hat u}_1})}}{{{{(1 + b{{\hat u}_1} + c{{\hat u}_3})}^2}}}) + \\ \quad \quad \quad \quad \frac{{{\beta _2}{{\hat u}_1}{{\hat u}_3}(1 + b{{\hat u}_1})(1 + c{{\hat u}_3})}}{{{{(1 + b{{\hat u}_1} + c{{\hat u}_3})}^4}}}, \end{array}

    where

    \hat{u}_1 = \frac{-(\mathcal{R}_2-c\mathcal{R}_2-b)+\sqrt{(\mathcal{R}_2-c\mathcal{R}_2-b)^2+4c\mathcal{R}_2}}{2c\mathcal{R}_2}, \hat{u}_3 = \mathcal{R}_2(1-\hat{u}_1)\hat{u}_1.

    Since \alpha\mathcal{R}_2\hat u_1(1-\hat u_1)+\gamma_1<\beta_1 \hat u_1 and \hat u_3 = \mathcal{R}_2\hat u_1(1-\hat u_1) then there exists \mu_1 = 0 such that d_2\mu_1+\alpha\hat u_3+\gamma_1<\beta_1\hat u_1 holds. Hence, it follows from [24,Lemma 6] that m_2(\lambda) has at least one zero with a positive real part. The proof of (ⅰ) is complete.

    Denote

    \begin{align*} &\bar E_2 = d_1\mu_i+\hat u_1-\frac{b\hat u_1\hat u_3}{(1+b\hat u_1+c\hat u_3)^2},\, \bar L_2 = d_3\mu_i+\gamma_2-\frac{\beta_2\hat u_1(1+b\hat u_1)}{(1+b\hat u_1+c\hat u_3)^2},\\ &\bar J_2 = \frac{\beta_2\hat u_1\hat u_3(1+b\hat u_1)(1+c\hat u_3)}{(1+b\hat u_1+c\hat u_3)^4}. \end{align*}

    Then we have

    g_2(\lambda) = \lambda^2+(\bar E_2+\bar L_2)\lambda+\bar E_2\bar L_2+\bar J_2.

    Since \bar E_2+\bar L_2 = d_1\mu_i+d_3\mu_i+\hat u_1-\frac{b\hat u_1\hat u_3}{(1+b\hat u_1+c\hat u_3)^2}+\gamma_2-\frac{\beta_2\hat u_1(1+b\hat u_1)}{(1+b\hat u_1+c\hat u_3)^2} and \bar E_2\bar L_2+\bar J_2 = (d_1\mu_i+\hat u_1-\frac{b\hat u_1\hat u_3}{(1+b\hat u_1+c\hat u_3)^2})(d_3\mu_i+\gamma_2-\frac{\beta_2\hat u_1(1+b\hat u_1)}{(1+b\hat u_1+c\hat u_3)^2})+\frac{\beta_2\hat u_1\hat u_3(1+b\hat u_1)(1+c\hat u_3)}{(1+b\hat u_1+c\hat u_3)^2}. Obviously, \gamma_2 = \frac{\beta_2\hat u_1}{1+b\hat u_1+c\hat u_3}>\frac{\beta_2\hat u_1(1+b\hat u_1)}{(1+b\hat u_1+c\hat u_3)^2}. Since b<\hat u_1(\mathcal{R}_2+b), we have b\hat u_3<\mathcal{R}_2^2\hat u_1^2. Noting that \mathcal{R}_2\hat u_1 = 1+b\hat u_1+c\hat u_3, we obtain \frac{b\hat u_3}{(1+b\hat u_1+c\hat u_3)^2}<1. This implies \hat u_1>\frac{b\hat u_1\hat u_3}{(1+b\hat u_1+c\hat u_3)^2}. Thus g_2(\lambda) has no nonnegative zeros for all \tau\geq 0. It follows from \alpha\mathcal{R}_2\hat u_1(1-\hat u_1)+\gamma_1>\beta_1 \hat u_1 that all roots of m_2(\lambda) have negative real parts. Thus all zeros of Eq. (23) have negative real parts for all \mu_i, and hence E_{01} is locally asymptotically stable. This completes the proof of ().


    4.3. Stability of the unique positive spatially homogenous steady state E^\ast


    4.3.1. Stability and Hopf bifurcation

    In this subsection, by taking \tau as the bifurcation parameter, we investigate the stability and the Hopf bifurcation near the unique positive spatially homogeneous steady state E^\ast. For this purpose, we always assume

    ({H_1})\;q < 0\;{\rm{and}}\;{{\cal R}_2} < \frac{{(b\alpha + c{\beta _1})( - p + \sqrt {{p^2} - 4q} ) + 2(b{\gamma _1} + {\beta _1})}}{{\alpha ( - p + \sqrt {{p^2} - 4q} ) + 2{\gamma _1}}}.

    The above assumption guarantees the uniqueness of the positive spatially homogeneous steady state E^*. For the spatially homogeneous positive steady state E^\ast = (u_1^\ast, u_2^\ast, u_3^\ast), the characteristic equation is given below:

    \label{Eq4.11} \lambda^3+b_{2i}\lambda^2+b_{1i}\lambda+b_{0i}+e^{-\lambda\tau}(c_{2}\lambda^2+c_{1i}\lambda+c_{0i}) = 0, (24)

    where

    \begin{align*} b_{2i} = d_1\mu_i+d_2\mu_i+d_3\mu_i+u_1^\ast-bA_3+\beta_1u_1^\ast+c\beta_2A_3,\\ b_{1i} = (d_1\mu_i+u_1^\ast-bA_3)(d_2\mu_i+\beta_1u_1^\ast+d_3\mu_i+c\beta_2A_3)+\\ \quad \quad (d_2\mu_i+\beta_1u_1^\ast)(d_3\mu_i+c\beta_2A_3)+\alpha\beta u_2^\ast u_3^\ast+A_1A_2\beta_2u_1^\ast u_3^\ast, \end{align*}
    \begin{array}{l} b_{0i}=(d_1\mu_i+u_1^\ast-bA_3)(d_2\mu_i+\beta_1u_1^\ast)(d_3\mu_i+c\beta_2A_3)+(d_1\mu_i+u_1^\ast-bA_3)\alpha\beta u_2^\ast u_3^\ast\\ \quad \quad -\alpha\beta_2A_2u_1^\ast u_2^\ast u_3^\ast+(d_2\mu_i+\beta_1u_1^\ast)A_1A_2\beta_2u_1^\ast u_3^\ast,\\ c_2=-\beta_1u_1^\ast,\\ c_{1i}=\beta_1u_1^\ast u_2^\ast-\beta_1u_1^\ast(d_1\mu_i+d_3\mu_i+u_1^\ast-bA_3+c\beta_2A_3),\\ c_{0i}=-d_1d_3\beta_1u_1^\ast\mu_i^2+d_3\mu_i\beta_1u_1^\ast u_2^\ast-d_1\mu_ic\beta_2A_3\beta_1u_1^\ast-d_3\mu_i(u_1^\ast-bA_3)\beta_1u_1^\ast+\\ \quad \quad (c\beta_2A_3+A_1\beta u_3^\ast)\beta_1u_1^\ast u_2^\ast-(u_1^\ast-bA_3)c\beta_2A_3\beta_1u_1^\ast-A_1A_2\beta_2 u_1^\ast u_3^\ast\beta_1 u_1^\ast,\\ A_1=\frac{1+bu_1^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}, A_2=\frac{1+cu_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}, A_3=\frac{u_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}. \end{array}

    Denote M_1 = u_1^\ast-bA_3, M_2 = \beta_1u_1^\ast, M_3 = c\beta_2A_3, M_4 = \alpha\beta u_2^\ast u_3^\ast, M_5 = A_1A_2\beta_2u_1^\ast u_3^\ast, M_6 = \alpha\beta_2A_2u_1^\ast u_2^\ast u_3^\ast, M_7 = A_1\beta u_2^\ast u_3^\ast. Clearly, M_italic>0 for i = 2,3,4,5,6,7.

    Furthermore, we assume that

    ({H_2}){\mkern 1mu} \;{M_1} > 0.
    ({H_3}){\mkern 1mu} \;{M_6} - {M_2}{M_7} > 0.
    ({H_4})\;{M_1}{M_4} + {M_2}{M_3}u_2^ * + {M_2}{M_7} - {M_6} > 0.
    (H_3^\prime )\;{M_6} - {M_2}{M_7} \le 0.
    (H_4^\prime )\;{M_1}({M_1}{M_3} + {M_5} + u_2^ * {M_2}) + {M_3}({M_1}{M_3} + {M_4} + {M_5}) + {M_6} - {M_2}{M_7} > 0.

    Theorem 4.6. (ⅰ) Assume that (H_1)-(H_4) hold. Then the spatially homogeneous positive steady state E^\ast of System (4) with \tau = 0 is locally asymptotically stable.

    (ⅱ) Assume that (H_1)-(H_2) and (H_3^\prime)-(H_4^\prime) hold. Then the spatially homogeneous positive steady state E^\ast of System (4) with \tau = 0 is locally asymptotically stable.

    Proof. When \tau = 0, (24) reduces to the following equation

    \label{Eq4.12} \lambda^3+(b_{2i}+c_2)\lambda^2+(b_{1i}+c_{1i})\lambda+b_{0i}+c_{0i} = 0. (25)

    Since M_1>0, a direct calculation yields

    b_{2i}+c_2 = (d_1+d_2+d_3)\mu_i+M_1+M_3>0.
    \begin{array}{l} {b_{1i}} + {c_{1i}} = ({d_1}{d_2} + {d_1}{d_3} + {d_2}{d_3})\mu _i^2 + ({M_3}{d_1} + {M_3}{d_2} + {M_1}{d_2} + {M_1}{d_3}){\mu _i} + \\ {M_1}{M_3} + {M_4} + {M_5} + u_2^*{M_2} > 0. \end{array}
    \begin{split} b_{0i}+c_{0i} = &d_1d_2d_3\mu_i^3+(M_3 d_1d_2+M_1 d_2d_3)\mu_i^2+\\ &(M_2u_2^\ast d_3+M_1M_3d_2+M_4d_1+M_5d_2)\mu_i+M_1M_4+M_2M_3u_2^\ast+\\ &M_2M_7-M_6. \end{split}
    \begin{equation*}\begin{split} (b_{2i}+c_2)(b_{1i}+c_{1i})-(b_{0i}+c_{0i})=&d_1\mu_i(b_{1i}+c_{1i}-d_2d_3\mu_i^2-M_3d_2\mu_i-M_4)+\\ &d_3\mu_i(b_{1i}+c_{1i}-u_2^\ast M_2)+\\ &d_{2}\mu_i(b_{1i}+c_{1i}-M_1d_3\mu_i-M_1M_3-M_5)+\\ &M_1(b_{1i}+c_{1i}-M_4)+M_3(b_{1i}+c_{1i}-u_2^\ast M_2)+\\ &M_6-M_2M_7. \end{split}\end{equation*}

    Since {b_{1i}} + {c_{1i}} - {d_2}{d_3}\mu _i^2 - {M_3}{d_2}{\mu _i} - {M_4} > 0,{b_{1i}} + {c_{1i}} - u_2^*{M_2} > 0,{b_{1i}} + {c_{1i}} - {M_1}{d_3}{\mu _i} - {M_1}{M_3} - {M_5} > 0,{b_{1i}} + {c_{1i}} - {M_4} > 0,{b_{1i}} + {c_{1i}} - u_2^ * {M_2} > 0, it follows from (H_3) that (b_{2i}+c_2)(b_{1i}+c_{1i})-(b_{0i}+c_{0i})>0 for all \mu_i.

    It follows from (H_2) and (H_4) that b_{0i}+c_{0i}>0 for all \mu_i. Thus, by the Routh-Hurwitz stability criterion, all the roots of (25) have negative real parts. This completes the proof of part (). Similarly, noting that (b_{2i}+c_2)(b_{1i}+c_{1i})-(b_{0i}+c_{0i}) is increasing in \mu_i under (H_2), we can prove part ().

    Next, we discuss the effect of the delay \tau\neq 0 on the stability of the positive steady state E^\ast. We first determine critical values of \tau at which a pair of simple purely imaginary eigenvalues appears.

    Let \lambda = \mathrm{i}\omega\,(\omega>0) be a root of Eq. (24). Substituting \lambda = \mathrm{i}\omega into Eq. (24) yields

    -\mathrm{i}\omega^3-b_{2i}\omega^2+\mathrm{i}b_{1i}\omega+b_{0i}+(-c_2\omega^2+\mathrm{i}\omega c_{1i}+c_{0i})e^{-\mathrm{i}\omega\tau} = 0,

    which implies that

    \label{Eq4.13} b_{2i}\omega^2-b_{0i} = (c_{0i}-c_2\omega^2)\cos\omega\tau+c_{1i}\omega\sin\omega\tau, (26)
    \label{Eq4.14} -\omega^3+b_{1i}\omega = (c_{0i}-c_2\omega^2)\sin\omega\tau-c_{1i}\omega\cos\omega\tau. (27)

    It follows from (26) and (27) that

    \label{Eq4.15} \omega^6+(b^2_{2i}-2b_{1i}-c_2^2)\omega^4+(b_{1i}^2-2b_{0i}b_{2i}+2c_{0i}c_2-c_{1i}^2)\omega^2+b_{0i}^2-c_{0i}^2 = 0. (28)

    Let \omega^2 = s, and denote p_i = b^2_{2i}-2b_{1i}-c_2^2, q_i = b_{1i}^2-2b_{0i}b_{2i}+2c_{0i}c_2-c_{1i}^2, r_i = b_{0i}^2-c_{0i}^2. Then (28) is reduced to

    \label{Eq4.16} h(s)\equiv s^3+p_is^2+q_is+r_i = 0. (29)

    From (24), we get

    \begin{array}{l} \quad \quad {p_i} = (d_1^2 + d_2^2 + d_3^2)\mu _i^2 + 2({M_1}{d_1} + {M_2}{d_2} + {M_3}{d_3}){\mu _i} + \\ \quad \quad \quad \quad M_1^2 + M_3^2 - 2{M_4} - 2{M_5},\\ {b_{0i}} - {c_{0i}} = {d_1}{d_2}{d_3}\mu _i^3 + ({d_1}{d_2}{M_3} + {d_2}{d_3}{M_1} + 2{d_1}{d_3}{M_2})\mu _i^2 + \\ \quad \quad \quad \quad (2{d_1}{M_2}{M_3} + {d_2}{M_1}{M_3} + 2{d_3}{M_1}{M_2} + {d_1}{M_4} + {d_2}{M_5} - {d_3}u_2^ * {M_2}){\mu _i}\\ \quad \quad \quad \quad + 2{M_1}{M_2}{M_3} + 2{M_2}{M_5} + {M_1}{M_4} - u_2^ * {M_2}{M_3} - {M_2}{M_7} - {M_6}. \end{array}

    In the following, we need to seek conditions required for Eq. (29) to have at least one positive root. For this purpose, we further make the following hypotheses: (H_5)2M_1M_2M_3+2M_2M_5+M_1M_4-u_2^\ast M_2M_3-M_2M_7-M_6<0.

    \begin{array}{l} ({H_6})\;2{d_1}{M_2}{M_3} + {d_2}{M_1}{M_3} + 2{d_3}{M_1}{M_2} + {d_1}{M_4} + {d_2}{M_5} - {d_3}u_2^ * {M_2} \ge 0. \end{array}

    Since b_{0i}-c_{0i} is increasing in \mu_i under (H_2) and (H_6), it follows from (H_5) that there exists N_2\in\mathbb{N} such that

    b_{0i}-c_{0i}<0 \quad \mathrm{for}\quad 1\leq i\leq N_2 \quad \mathrm{and} \quad b_{0i}-c_{0i}\geq 0\quad \mathrm{for}\quad i\geq N_2+1,\quad i\in\mathbb{N}.

    According to the above analysis, we have the following lemma.

    Lemma 4.7.(ⅰ) Assume that (H_2), (H_4)-(H_6) hold. Then Eq. (29) has at least one positive root for each i\in\{1,2,\cdots,N_2\}.

    (ⅱ) Assume that and (H_2), (H_3^\prime) and (H_5)-(H_6) hold. Then Eq. (29) has at least one positive root for each i\in\{1,2,\cdots,N_2\}.

    Proof. It follows from (H_2) and (H_4) that b_{0i}+c_{0i}>0 for all i\in\mathbb{N}. Thus r_i<0 if and only if b_{0i}-c_{0i}<0. It follows from (H_2), (H_5) and (H_6) that there exists \mu_i\,(i = 1,2,\cdots,N_2) such that r_i<0. Since \lim\limits_{s\rightarrow\infty}h(s) = \infty for fixed \mu_i, then (29) has at least one positive root for each i\in\{1,2,\cdots,N_2\}. This completes proof of part (). Similarly, we can prove part ().

    Remark 3. From Lemma 4.2, without loss of generality, for each i, 1\leq i\leq N_2, we may assume that it has three positive roots, which are denoted by s_{1,i}, s_{2,i}, s_{3,i}, respectively. Then for each i, 1\leq i\leq N_2, (18) has three positive roots \omega_{k,i} = \sqrt{s_{k,i}}, k = 1,2,3. By (26) and (27), we get

    \cos\omega_{k,i}\tau_{k,i} = \frac{(c_{1i}-b_{2i}c_2)\omega_{k,i}^4+(b_{2i}c_{0i}+b_{0i}c_2-c_{1i}b_{1i})\omega_{k,i}^2-b_{0i}c_{0i}}{(c_{0i}-c_2\omega_{k,i}^2)^2+c_{1i}^2\omega_{k,i}^2}.

    Let

    \label{Eq4.17} \tau_{k,j}^i = \frac{\left(\arccos\left(\frac{(c_{1i}-b_{2i}c_2)\omega_{k,i}^4+(b_{2i}c_{0i}+b_{0i}c_2-c_{1i}b_{1i})\omega_{k,i}^2-b_{0i}c_{0i}}{(c_{0i}-c_2\omega_{k,i}^2)^2+c_{1i}^2\omega_{k,i}^2}\right)+2j\pi\right)}{\omega_{k,i}} (30)

    for 1\leq i\leq N_2, k = 1,2,3, j = 0,1,2,\cdots. Then \pm\mathrm{i}\omega_{k,i} is a pair of purely imaginary roots of (28) with \tau = \tau_{k,j}^i. Define

    \label{Eq4.18} \tau^\ast = \tau_{k_0,0}^{i_0} = \min\limits_{k = {1,2,3},\,i = {1,2,\cdots,N_2}}\tau_{k,0}^i, \quad \quad \omega^\ast = \omega_{k_0,i_0}. (31)

    Lemma 4.8. Let \lambda(\tau) = \alpha(\tau)\pm\mathrm{i}\beta(\tau) be the roots of Eq. (14) near \tau = \tau^\ast satisfying \alpha(\tau^\ast) = 0, \beta(\tau^\ast) = \omega^\ast. Suppose that h^\prime((\omega^{\ast})^2)>0. Then \pm\mathrm{i}\omega^\ast is a pair of simple purely imaginary roots of Eq. (24). Moreover, the following transversality condition holds:

    \mathrm{sign}\left \{\frac{d(\mathrm{Re}\lambda(\tau))}{d\tau}\right\}_{\tau = \tau^\ast,\lambda = \mathrm{i}\omega^\ast}>0.

    Proof. We denote P(\lambda) = \lambda^3+b_{2i}\lambda^2+b_{1i}\lambda+b_{0i}, Q(\lambda) = c_{2}\lambda^2+c_{1i}\lambda+c_{0i}. Then (24) can be rewritten as

    \label{Eq4.19} P(\lambda)+Q(\lambda)e^{-\lambda\tau} = 0. (32)

    It is easy to know (26) and (27) are equivalent to the following equations

    \begin{split} &\mathrm{Re}P(\mathrm{i}\omega) = -\mathrm{Re}Q(\mathrm{i}\omega)\cos\omega\tau-\mathrm{Im} Q(\mathrm{i}\omega)\sin\omega\tau,\\ &\mathrm{Im}P(\mathrm{i}\omega) = \mathrm{Re}Q(\mathrm{i}\omega)\sin\omega\tau-\mathrm{Im} Q(\mathrm{i}\omega)\cos\omega\tau. \end{split}

    Thus

    \label{Eq4.20} h(\omega^2) = (\mathrm{Re}P(\mathrm{i}\omega))^2+(\mathrm{Im}P(\mathrm{i}\omega))^2- ((\mathrm{Re}Q(\mathrm{i}\omega))^2+(\mathrm{Im}Q(\mathrm{i}\omega))^2). (33)

    Differentiating both sides of (33) with respect to \omega yields

    \label{Eq4.21} 2\omega h^\prime(\omega^2) = \mathrm{i}[P^\prime(\mathrm{i}\omega)\bar{P}(\mathrm{i}\omega)-\bar P^\prime(\mathrm{i}\omega)P(\mathrm{i}\omega)-Q^\prime(\mathrm{i}\omega)\bar{Q}(\mathrm{i}\omega)+\bar Q^\prime(\mathrm{i}\omega)Q(\mathrm{i}\omega)]. (34)

    Substituting \mathrm{i}\omega^\ast into (32) yields

    \label{Eq4.22}|P(\mathrm{i}\omega^\ast)| = |Q(\mathrm{i}\omega^\ast)|. (35)

    If \mathrm{i}\omega^\ast is not simple, then \mathrm{i}\omega^\ast must satisfy P^\prime(\mathrm{i}\omega^\ast)+[Q^\prime(\mathrm{i}\omega^\ast)-\tau^\ast Q(\mathrm{i}\omega^\ast)]e^{-\mathrm{i}\omega^\ast\tau^\ast} = 0. Note that e^{-\mathrm{i}\omega^\ast\tau^\ast} = -P(\mathrm{i}\omega^\ast)/Q(\mathrm{i}\omega^\ast), we obtain \tau^\ast = \frac{Q^\prime(\mathrm{i}\omega^\ast)}{Q(\mathrm{i}\omega^\ast)}-\frac{P^\prime(\mathrm{i}\omega^\ast)}{P(\omega^\ast)}. Using (34) and (35), we have

    \begin{split} \mathrm{Im}\tau^\ast = &\mathrm{Im}\left[\frac{Q^\prime(\mathrm{i}\omega^\ast)}{Q(\mathrm{i}\omega^\ast)}-\frac{P^\prime(\mathrm{i}\omega^\ast)}{P(\mathrm{i}\omega^\ast)}\right] = \mathrm{Im}\left[\frac{Q^\prime(\mathrm{i}\omega^\ast)\bar Q(\mathrm{i}\omega^\ast)}{Q(\mathrm{i}\omega^\ast)\bar Q(\mathrm{i}\omega^\ast)}-\frac{P^\prime(\mathrm{i}\omega^\ast)\bar P(\mathrm{i}\omega^\ast)}{P(\mathrm{i}\omega^\ast)\bar P(\mathrm{i}\omega^\ast)}\right]\\ & = \frac{1}{Q(\mathrm{i}\omega^\ast)\bar Q(\mathrm{i}\omega^\ast)}\mathrm{Im}\left[Q^\prime(\mathrm{i}\omega^\ast)\bar Q(\mathrm{i}\omega^\ast)-P^\prime(\mathrm{i}\omega^\ast)\bar P(\mathrm{i}\omega^\ast)\right]\\ & = \frac{Q^\prime(\mathrm{i}\omega^\ast)\bar Q(\mathrm{i}\omega^\ast)-P^\prime(\mathrm{i}\omega^\ast)\bar P(\mathrm{i}\omega^\ast)-\bar{Q}^\prime(\mathrm{i}\omega^\ast)Q(\mathrm{i}\omega^\ast)+\bar{P}^\prime(\mathrm{i}\omega^\ast) P(\mathrm{i}\omega^\ast)}{2\mathrm{i}|Q(\mathrm{i}\omega^\ast)|^2}\\ & = \frac{\omega^\ast h^\prime((\omega^{\ast})^2)}{|Q(\mathrm{i}\omega^\ast)|^2}. \end{split}

    This is a contradiction. Thus \pm\mathrm{i}\omega^\ast is a pair of simple purely imaginary roots of Eq. (24).

    Since \pm\mathrm{i}\omega^\ast are simple purely imaginary roots and

    P^\prime(\mathrm{i}\omega^\ast)+[Q^\prime(\mathrm{i}\omega^\ast)-\tau^\ast Q(\mathrm{i}\omega^\ast)]e^{-\mathrm{i}\omega^\ast\tau^\ast}\neq 0,

    we may consider \lambda = \lambda(\tau) to be a differentiable function. Differentiating (22) with respect to \tau yields

    \frac{d\lambda(\tau)}{d\tau} = \frac{\lambda Q(\lambda)}{P^\prime(\lambda)e^{\lambda\tau}+Q^\prime(\lambda)-\tau Q(\lambda)}.

    Using (32) again, we obtain

    \label{Eq4.23} \left(\frac{d\lambda(\tau)}{d\tau}\right)^{-1} = -\frac{P^\prime(\lambda)}{\lambda P(\lambda)}+\frac{Q^\prime(\lambda)}{\lambda Q(\lambda)}-\frac{\tau}{\lambda}. (36)

    Thus, from (36), we have

    \begin{split} \mathrm{sign}\left \{\frac{d(\mathrm{Re}\lambda(\tau))}{d\tau}\right\}_{\tau = \tau^\ast,\lambda = \mathrm{i}\omega^\ast} = &\mathrm{sign}\left\{\mathrm{Re}\left[-\frac{P^\prime(\lambda)}{\lambda P(\lambda)}+\frac{Q^\prime(\lambda)}{\lambda Q(\lambda)}-\frac{\tau}{\lambda}\right]\right\}_{\tau = \tau^\ast,\lambda = \mathrm{i}\omega^\ast}\\ = &\mathrm{sign}\left\{\mathrm{Re}\bar\lambda\left[Q^\prime(\lambda)\bar{Q(\lambda)}-P^\prime(\lambda)\bar{P(\lambda)}\right]\right\}_{\lambda = \mathrm{i}\omega^\ast}\\ = &\mathrm{sign}\left\{(\omega^\ast)^2h^\prime((\omega^\ast)^2)\right\}>0. \end{split}

    This completes the proof.

    When \tau\in[0,\tau^\ast), we know that (24) has no roots on the imaginary axis. By the eigenvalue theory of [39], the sum of orders of the zeros of Eq. (24) for \tau\in[0,\tau^\ast) is equal to Eq. (25). Then Eq. (24) only has negative real part roots for \tau\in[0,\tau^\ast), which implies that (u_1^\ast, u_2^\ast, u_3^\ast) is locally asymptotically stable for \tau\in[0,\tau^\ast). Combining Theorem 4.6, Lemma 4.7 and 4.8, we arrive the following theorem.

    Theorem 4.9. Assume that either (H_1)-(H_6) or (H_1),(H_2), (H_3^\prime),(H_4^\prime),(H_5), (H_6) hold. We have the following results

    (ⅰ) The spatially homogeneous positive steady state E^\ast = (u_1^\ast, u_2^\ast, u_3^\ast) of System (4) is locally asymptotically stable for \tau\in [0,\tau^\ast) and unstable when \tau>\tau^\ast, where \tau^\ast is difined in (31).

    (ⅱ) Furthermore, suppose that h^\prime((\omega^\ast)^2)>0. Then System (4) undergoes Hopf bifurcation at the positive steady state E^\ast = (u_1^\ast, u_2^\ast, u_3^\ast) when \tau = \tau_{k,j}^i, i.e., a family of spatially periodic solutions bifurcate from E^\ast = (u_1^\ast, u_2^\ast, u_3^\ast) when \tau crosses through the critical values \tau_{k,j}^i, where \tau_{k,j}^i is defined in (30).

    Remark 4. There exist parameter values such that hypotheses (H_1)-(H_6) hold. For example, we choose parameter values \alpha = 0.7,\beta = 0.9,{\beta _1} = 1.95,{\beta _2} = 1.8,{\gamma _1} = 0.2,{\gamma _2} = 0.8,b = 2.5,c = 12,{d_1} = 0.25,{d_2} = 0.28,{d_3} = 0.2. Using any CAS, it is easy to check that hypotheses (H_1)-(H_6) are satisfied.


    4.3.2. Properties of Hopf bifurcations

    In this section, we investigate the direction of Hopf bifurcation and the stability of bifurcated periodic solutions by using the normal form theory and center manifold reduction. For convenience, for fixed k\in\{1,2,3\}, j = 0,1,2,\cdots, we denote \tau_{k,j}^i (1\leq i\leq N_2) by \hat\tau, and denote \omega_{k,i} (1\leq i\leq N_2) by \hat\omega.

    Let \tau = \hat\tau+\mu, \mu\in \mathbb{R} and \widetilde u_1(t,\cdot) = u_1(\tau t,\cdot)-u_1^\ast, \widetilde u_2(t,\cdot) = u_2(\tau t,\cdot)-u_2^\ast, \widetilde u_3(t,\cdot) = u_3(\tau t, \cdot)-u_3^\ast and \widetilde U(t) = (\widetilde u_1(t,\cdot), \widetilde u_2(t,\cdot),\widetilde u_3(t,\cdot)) as u_1-u_1^\ast, u_2-u_2^\ast, u_3-u_3^\ast, then drop the tildes for simplicity. System (1.4) can be rewritten in the phase space \mathcal{C} = C([-1,0], X) as

    \dot U(t) = \hat\tau D\triangle U(t)+L(\hat\tau)(U_t)+F(U_t,\mu), \label{Eq4.24} (37)

    where D = \mathrm{diag}\{d_1, d_2, d_3\}, L(\delta)(\cdot):\mathcal{C}\rightarrow X and F: \mathcal{C}\times R\rightarrow X are given by

    L(\delta)\phi = \delta\left[ \begin{array}{c} (\frac{bu_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}-u_1^\ast)\phi_1(0)-u_1^\ast\phi_2(0)-\frac{(1+bu_1^\ast)u_1^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_3(0)\\ \beta_1u_2^\ast\phi_1(-1)+\beta_1u_1^\ast\phi_2(-1)-\beta_1u_1^\ast\phi_2(0)-\alpha u_2^\ast\phi_3(0)\\ \frac{\beta_2(1+cu_3^\ast)u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_1(0)+\beta u_3^\ast\phi_2(0)-\frac{c\beta_2u_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_3(0) \end{array}\right]

    and

    F(\phi,\mu) = \mu D\Delta\phi(0)+L(\mu)\phi+f(\phi,\mu),

    where

    \begin{split} f(\phi,\mu) = &(\hat\tau+\mu)\times\\ &\left[\begin{array}{l} (\phi_1(0)+u_1^\ast)(-\phi_1(0)-\phi_2(0)-\frac{\phi_3(0)+u_3^\ast}{1+b(\phi_1(0)+u_1^\ast)+c(\phi_3(0)+u_3^\ast)}+\\ \frac{u_3^\ast}{1+bu_1^\ast+cu_3^\ast})-((\frac{bu_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}-u_1^\ast)\phi_1(0)-u_1^\ast\phi_2(0)-\\ \frac{(1+bu_1^\ast)u_1^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_3(0))\\\\ \beta_1\phi_1(-1)\phi_2(-1)-\alpha \phi_2(0)\phi_3(0)\\\\ (\phi_3(0)+u_3^\ast)(\frac{\beta_2(\phi_1(0)+u_1^\ast)}{1+b(\phi_1(0)+u_1^\ast)+c(\phi_3(0)+u_3^\ast)}+\beta \phi_2(0)-\frac{\beta_2 u_1^\ast}{1+bu_1^\ast+cu_3^\ast})\\ -(\frac{\beta_2(1+cu_3^\ast)u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_1(0)+\beta u_3^\ast\phi_2(0)-\frac{c\beta_2u_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\phi_3(0)) \end{array}\right] \end{split}

    for \phi = (\phi_1, \phi_2, \phi_3)\in\mathcal{C}. Let \mathcal{A} to be the infinitesimal generator of the semigroup induced by the solution of the linearized equation of (37)

    \dot U(t) = \hat\tau D\triangle U(t)+L(\hat\tau)(U_t).

    Thus Eq. (37) can be written in the following abstract form

    \frac{dU_t}{dt} = \mathcal{A}U_t+X_0F(U_t,\mu),

    where X_0(\theta) = \left\{ \begin{array}{ll} 0,\theta\in[-1,0),\\ I, \theta = 0. \end{array}\right. Recall that the Banach space decomposition {\bf{X}} = \bigoplus_{i = 0}^\infty{\bf{X}}_i. In view of the Resiz representation theorem, there exists a 3\times 3 matrix function \eta(\theta, \hat\tau)\,(-1\leq\theta\leq 0), whose entries are bounded variation such that

    -\hat\tau D\mu_i\phi(0)+L(\hat\tau)(\phi) = \int^0_{-1}d[\eta(\theta, \hat\tau)]\phi(\theta),

    for \phi\in C([-1,0], R^3). We can choose

    \begin{equation*}\begin{split} \eta(\theta, \hat\tau)=& \hat\tau\left[ \begin{array}{ccc} -d_1\mu_i+\frac{bu_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}-u_1^\ast&-u_1^\ast&-\frac{(1+bu_1^\ast)u_1^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}\\ 0&-d_2\mu_i-\beta_1u_1^\ast&-\alpha u_2^\ast\\ \frac{\beta_2(1+cu_3^\ast)u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2}&\beta u_3^\ast&-d_3\mu_i-\frac{c\beta_2u_1^\ast u_3^\ast}{(1+bu_1^\ast+cu_3^\ast)^2} \end{array}\right]\\ &\times\delta(\theta)+\hat\tau\left[ \begin{array}{ccc} 0&0&0\\ -\beta_1u_2^\ast&-\beta_1 u_1^\ast&0\\ 0&0&0 \end{array}\right]\delta(\theta+1), \end{split}\end{equation*}

    where \delta is a Dirac delta function.

    Let us define C^\ast = C([0,1], \mathbb{R}^{3\ast}), where \mathbb{R}^{3\ast} is the 3-dimensional vector space of row vectors, \mathcal{A}^\ast with domain dense in C^\ast and range in C^\ast. Let \mathcal {P} and \mathcal {P}^\ast be the center subspace, that is, the generalized eigenspace of \mathcal {A} and \mathcal {A}^\ast associated with \Lambda^\ast = \{\mathrm{i}\hat\omega\hat\tau, -\mathrm{i}\hat\omega\hat\tau\}.

    For \Phi\in C([-1,0], R^3), \Psi\in C([0,1], R^{3\ast}), we define

    \mathcal{A}(\Phi(\theta)) = \left\{\begin{array}{ll} \frac{d\Phi(\theta)}{d\theta}, \quad \quad \quad \quad \ \theta\in[-1,0),\\ \int_{-1}^0[d\eta(\theta, \hat\tau)]\Phi(\theta), \quad \theta = 0, \end{array}\right.

    and

    \mathcal{A}^\ast(\Psi(s)) = \left\{\begin{array}{ll} -\frac{d\Psi(s)}{ds},&s\in(0,1],\\ \int_{-1}^0\Psi(-\theta)[d\eta(\theta, \hat\tau)],&s = 1. \end{array}\right.

    Then \mathcal{A}^\ast is the formal adjoint of \mathcal{A} under the bilinear pairing

    \begin{align*}\begin{split} \langle\psi, \phi\rangle = &\psi(0)\phi(0)-\int^0_{-1}\int^\theta_0\psi(\xi-\theta)d[\eta(\theta, \hat\tau)]\phi(\xi)d\xi\\ = &\bar\psi(0)\phi(0)+\hat\tau\int^0_{-1}\bar\psi(\xi+1) \left[ \begin{array}{ccc} 0&0&0\\ \beta_1u_2^\ast&\beta_1u_1^\ast&0\\ 0&0&0 \end{array}\right]\phi(\xi)d\xi, \end{split}\end{align*}

    for \phi\in C([-1,0], R^3), \psi\in C([0,1], R^{3\ast}).

    In view of the definition of the two infinitesimal generators \mathcal{A} and \mathcal{A}^\ast, we have the following conclusions.

    Lemma 4.10. Let

    \begin{split} \eta_1 = &\frac{(d_3\mu_i+M_3)(G_2^2+H_2^2)-\beta u_3^\ast(G_1G_2+H_1H_2)}{u_3^\ast\beta_2A_2(G_2^2+H_2^2)}+\\ &\mathrm{i}\frac{(\hat\omega(G_2^2+H_2^2)-\beta u_3^\ast(G_2H_1-G_1H_2))}{u_3^\ast\beta_2A_2(G_2^2+H_2^2)}, \end{split}
    \begin{split} \eta_1^\ast = &\frac{-\alpha u_2^\ast(G_3G_4+H_3H_4)-(d_3\mu_i+M_3)(G_4^2+H_4^2)}{u_1^\ast A_1(G_4^2+H_4^2)}-\\ &\mathrm{i}\frac{\alpha u_2^\ast(G_4H_3-G_3H_4)+\hat\omega(G_4^2+H_4^2)}{u_1^\ast A_1(G_4^2+H_4^2)}, \end{split}
    \eta_2 = \frac{G_1G_2+H_1H_2+\mathrm{i}(G_2H_1-G_1H_2)}{G_2^2+H_2^2},\eta_2^\ast = \frac{G_3G_4+H_3H_4+\mathrm{i}(G_4H_3-G_3H_4)}{G_4^2+H_4^2},

    where G_i, H_i\,(i = 1,2,3,4) are defined as follows

    G_1 = M_5-\hat\omega+(d_1\mu_i+M_1)(d_3\mu_i+M_3), \quad \quad \quad \ H_1 = \hat\omega(d_1\mu_i+M_1+d_3\mu_i+M_3),

    G_2 = \beta u_3^\ast d_1\mu_i+\beta M_1u_3^\ast-\beta_2 u_1^\ast u_3^\ast A_2, \quad \quad \quad \quad uad \quad H_2 = \beta u_3^\ast\hat\omega,

    G_3 = -M_5-(d_3\mu_i+M_3)(d_1\mu_i+M_1)+(\hat\omega)^2,\quad \quad H_3 = -\hat\omega(d_3\mu_i+M_3)-\hat\omega(d_1\mu_i+M_1),

    G_4 = u_2^\ast A_1M_2\cos\hat\omega\hat\tau+\alpha u_2^\ast(d_1\mu_i+M_1), \quad \quad uad \quad H_4 = -u_2^\ast A_1M_2\sin\hat\omega\hat\tau+\alpha u_2^\ast\hat\omega.

    Then

    p(\theta) = e^{\mathrm{i}\hat\omega\hat\tau\theta}(\eta_1, \eta_2, 1)^T is the eigenfunction of \mathcal{A} with respect to \mathrm{i}\hat\omega\hat\tau;

    p^\ast (s) = e^{-\mathrm{i}\hat\omega\hat\tau s}(\eta_1^\ast, \eta_2^\ast, 1) is the eigenfunction of \mathcal{A}^\ast with respect to \mathrm{i}\hat\omega\hat\tau.

    Proof. The proof is standard and we omit it here.

    Clearly, from Lemma 4.10, we know the center subspace of Eq. (37) is

    \mathcal{P} = \mathrm{span}\{ p(\theta), \bar {p(\theta)}\}, \mathcal{P}^\ast = \mathrm{span}\{p^\ast(s), \bar {p^\ast (s)}\}.

    Then \mathcal{C} can be decomposed as \mathcal{C} = \mathcal{P}\oplus\mathcal{Q}, where

    \mathcal{Q} = \{\psi\in\mathcal{C}: \langle\widehat\psi, \psi\rangle = 0 \quad \mathrm{for}\,\mathrm{all}\, \widehat\psi\in \mathcal{P}^\ast \}.

    It follows from Lemma 4.12 that

    \begin{split} \langle p^\ast(s), p(\theta)\rangle = &\bar{ p^\ast(0)}p(0)+\\ &\hat\tau\int_{-1}^0e^{-\mathrm{i}\hat\omega\hat\tau(\xi+1)}(\bar{\eta^\ast_1}, \bar{\eta^\ast_2}, 1)\left[ \begin{array}{ccc} 0&0&0\\ \beta_1u_2^\ast&\beta_1u_1^\ast&0\\ 0&0&0 \end{array}\right](\eta_1, \eta_2, 1)^Te^{\mathrm{i}\hat\omega\hat\tau\xi}d\xi\\ = &\bar{\eta^\ast_1}\eta_1+\bar{\eta^\ast_2}\eta_2+1+\beta_1u_2^\ast\hat\tau e^{-\mathrm{i}\hat\omega\hat\tau}\eta_1\bar{\eta_2^\ast}+\beta_1u_1^\ast\hat\tau e^{-\mathrm{i}\hat\omega\hat\tau}\eta_2\bar{\eta_2^\ast}. \end{split}

    Thus we choose \mathfrak{D} = 1/(\bar{\eta^\ast_1}\eta_1+\bar{\eta^\ast_2}\eta_2+1+\beta_1u_2^\ast\hat\tau e^{-\mathrm{i}\hat\omega\hat\tau}\eta_1\bar{\eta_2^\ast}+\beta_1u_1^\ast\hat\tau e^{-\mathrm{i}\hat\omega\hat\tau}\eta_2\bar{\eta_2^\ast}). Let \Phi = (p(\theta), \bar{p(\theta)}), \Psi = \mathfrak{D}(p^\ast(s), \bar {p^\ast(s)})^T\equiv (q(s), \bar {q(s)})^T, then \langle \Psi, \Phi\rangle = I, where I is the identity matrix in \mathbb{R}^{2\times2}.

    In what follows, in order to determine the bifurcation direction and stability, we compute the coordinates to describe the center manifold \mathcal{C}_0. As the formulas to be developed for the bifurcation direction and stability are all relative to \mu = 0 only, we set \mu = 0 in Eq. (4.24) and obtain a center manifold

    W(z, \bar z)(\theta) = W_{20}(\theta)\frac{z^2}{2}+W_{11}(\theta)z\bar z+W_{02}(\theta)\frac{\bar z^2}{2}+\cdots \label{Eq4.25} (38)

    with the range in \mathcal{Q}. The flow of Eq. (4.24) on the center manifold can be written as

    U_t = \Phi\cdot(z(t), \bar {z(t)})^T+W(z(t),\bar {z(t)}). \label{Eq4.26} (39)

    Moreover, from (39), z satisfies

    \begin{split} \dot z(t) = &\frac{d}{dt}\langle q(s), U_t\rangle = \langle q(s), \mathcal{A}U_t\rangle+\langle q(s), X_0F(U_t,0)\rangle\\ = &\mathrm{i}\hat\omega\hat\tau z+\bar {q(0)}f(W(z,\bar z)+2\mathrm{Re}\{zp(\theta)\},0)\\ = &\mathrm{i}\hat\omega\hat\tau z+g(z,\bar z), \label{Eq4.27} \end{split} (40)

    where

    \begin{split} g(z, \bar z) = g_{20}\frac{z^2}{2}+g_{11}z\bar z+g_{02}\frac{\bar z^2}{2}+g_{21}\frac{z^2\bar z}{2}+\cdot\cdot\cdot. \end{split}

    By the Taylor expansion

    \begin{array}{l} \frac{{v + {v^ * }}}{{1 + b(u + {u^ * }) + c(v + {v^ * })}} = \frac{{{v^ * }}}{{1 + b{u^ * } + c{v^ * }}} - \frac{{b{v^ * }u}}{{{{(1 + b{u^ * } + c{v^ * })}^2}}} + \frac{{(1 + b{u^ * })v}}{{{{(1 + b{u^ * } + c{v^ * })}^2}}} + \\ {C_{11}}{u^2} + {C_{12}}uv + {C_{13}}{v^2} + {C_{14}}{u^2}v + {C_{15}}u{v^2} + \\ {C_{16}}{u^3} + {C_{17}}{v^3} + O(4),\\ \frac{{u + {u^ * }}}{{1 + b(u + {u^ * }) + c(v + {v^ * })}} = \frac{{{u^ * }}}{{1 + b{u^ * } + c{v^ * }}} + \frac{{(1 + c{v^ * })u}}{{{{(1 + b{u^ * } + c{v^ * })}^2}}} - \frac{{c{u^ * }v}}{{{{(1 + b{u^ * } + c{v^ * })}^2}}}\\ + {C_{21}}{u^2} + {C_{22}}uv + {C_{23}}{v^2} + {C_{24}}{u^2}v + {C_{25}}u{v^2} + \\ {C_{26}}{u^3} + {C_{27}}{v^3} + O(4), \end{array} (41)

    where

    \begin{array}{l} {C_{11}} = \frac{{{b^2}{v^ * }}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},{C_{12}} = \frac{{bc{v^ * } - b(1 + b{u^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},{C_{13}} = - \frac{{c(1 + b{u^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},\\ {C_{14}} = \frac{{{b^2}(1 + b{u^ * } - 2c{v^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},{C_{15}} = \frac{{bc(2(1 + b{u^ * }) - c{v^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},{C_{16}} = \frac{{ - {b^3}{v^ * }}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}}, \end{array}
    \begin{array}{l} {C_{17}} = \frac{{{c^3}(1 + b{u^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},{C_{21}} = - \frac{{b(1 + c{v^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},{C_{22}} = \frac{{bc{u^ * } - c(1 + c{v^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},\\ {C_{23}} = \frac{{{c^2}{u^ * }}}{{{{(1 + b{u^ * } + c{v^ * })}^3}}},{C_{24}} = \frac{{bc(2(1 + c{v^ * }) - b{u^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},{C_{25}} = \frac{{{c^2}(1 + c{v^ * } - 2b{u^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},\\ {C_{26}} = \frac{{{b^3}(1 + c{v^ * })}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}},{C_{27}} = - \frac{{{c^3}{u^ * }}}{{{{(1 + b{u^ * } + c{v^ * })}^4}}}. \end{array}

    Noting that (40), we get

    g(z,\bar z) = \bar {\mathfrak{D}}(\bar{\eta_1^\ast}, \bar{\eta_2^\ast}, 1)f(W(z,\bar z)+zp(\theta)+\bar z\bar {p(\theta)},0). \label{Eq4.29} (42)

    Substituting (38) into (42) and combining (41) yield

    \begin{split} g_{20} = &2\bar{\mathfrak{D}}\hat\tau[\bar{\eta_1^\ast}\left(({bA_3}/{u_1^\ast}-C_{11}u_1^\ast-1)\eta_1^2-\eta_1\eta_2-C_{13}u_1^\ast-(A_1+C_{12}u_1^\ast)\eta_1\right) +\\ &\bar{{\eta_2^\ast}}(\beta_1\eta_1\eta_2e^{-2\mathrm{i}\hat\omega\hat\tau}-\alpha\eta_2) +\bar{{\eta_3^\ast}}(u_3^\ast\beta_2C_{21}\eta_1^2+C_{23}u_3^\ast\beta_2-M_3/u_3^\ast+\\ &(\beta_2A_2+\beta_2C_{22}u_3^\ast)\eta_1+\beta\eta_2)], \end{split}
    \begin{split} g_{11} = &\bar{\mathfrak{D}}\hat\tau[\bar{\eta_1^\ast}(2({bA_3}/{u_1^\ast}-C_{11}u_1^\ast-1)\eta_1\bar{\eta_1}-(\eta_1\bar{\eta_2}+\eta_2\bar{\eta_1})-2C_{13}u_1^\ast-\\ &(A_1+C_{12}u_1^\ast)(\eta_1+\bar{\eta_1}))+\bar{\eta_2^\ast}(\beta_1(\eta_1\bar{\eta_2}+\bar{\eta_1}\eta_2)-\alpha(\eta_2+\bar{\eta_2}))+\\ &\bar{\eta_3^\ast}(2u_3^\ast\beta_2C_{21}\eta_1\bar{\eta_1}+2(C_{23}u_3^\ast\beta_2-M_3/u_3^\ast)+\\ &(\beta_2A_2+\beta_2C_{22}u_3^\ast)(\eta_1+\bar{\eta_1})+\beta(\eta_2+\bar{\eta_2}))], \end{split}
    \begin{split} g_{21} = &2\bar{\mathfrak{D}}\hat\tau[\bar{\eta_1^\ast}(2({bA_3}/{u_1^\ast}-C_{11}u_1^\ast-1)(\eta_1W_{11}^{(1)}(0)+\frac{\bar{\eta_1}}{2}W_{20}^{(1)}(0)) -(\frac{W_{20}^{(1)}(0)}{2}\bar{\eta_2}+\\ &W_{11}^{(2)}(0)\eta_1+W_{11}^{(1)}(0)\eta_2+\frac{W_{20}^{(2)}(0)}{2}\bar{\eta_1})-2C_{13}u_1^\ast(W_{11}^{(3)}(0)+\frac{W_{20}^{(3)}(0)}{2})-\\ &(A_1+C_{12}u_1^\ast)(\frac{W_{20}^{(1)}(0)}{2}+W_{11}^{(1)}(0)+W_{11}^{(3)}(0)\eta_1+\frac{W_{20}^{(3)}(0)}{2}\bar{\eta_1})-\\ &3(C_{11}+C_{16}u_1^\ast)\eta_1\eta_1\bar{\eta_1}-(C_{12}+C_{14}u_1^\ast)(\eta_1\eta_1+\eta_1\bar{\eta_1}+\bar{\eta_1}\eta_1)-3C_{17}u_1^\ast-\\ &(C_{13}+C_{15}u_1^\ast)(2\eta_1+\bar{\eta_1}))+\bar{\eta_2^\ast}(\beta_1\eta_1W_{11}^{2}(-1)e^{-\mathrm{i}\hat\omega\hat\tau}+ \beta_1\bar{\eta_1}e^{\mathrm{i}\hat\omega\hat\tau}\frac{W_{20}^{2}(-1)}{2}+\\ &\beta_1\bar{\eta_2}e^{\mathrm{i}\hat\omega\hat\tau}\frac{W_{20}^{1}(-1)}{2}+\beta_1\eta_2W_{11}^{1}(-1)e^{-\mathrm{i}\hat\omega\hat\tau}-\alpha\eta_2W_{11}^{(3)}(0) -\alpha\bar{\eta_2}\frac{W_{20}^{(3)}(0)}{2}-\\ &\alpha\frac{W_{20}^{(2)}(0)}{2}-\alpha W_{11}^{(2)}(0))+\bar{\eta_3^\ast}(2u_3^\ast\beta_2C_{21}(W_{11}^{(1)}(0)\eta_1+\frac{\bar {\eta_1}}{2}W_{20}^{(1)}(0))+\\ &2(C_{23}u_3^\ast\beta_2-M_3/u_3^\ast)(W_{11}^{(3)}(0)+\frac{W_{20}^{(3)}(0)}{2})+(\beta_2A_2+\beta_2C_{22}u_3^\ast)(\frac{W_{20}^{(1)}(0)}{2}+\\ &\eta_1W_{11}^{(3)}(0)+\bar{\eta_1}\frac{W_{20}^{(3)}(0)}{2}+W_{11}^{(1)}(0))+\beta(\eta_2W_{11}^{(3)}(0)+\bar{\eta_2}\frac{W_{20}^{(3)}(0)}{2} + \end{split}
    \begin{split}&\frac{W_{20}^{(2)}(0)}{2}+W_{11}^{(2)}(0))+3C_{26}u_3^\ast\beta_2\eta_1\eta_1\bar{\eta_1}+3(C_{23}\beta_2+u_3^\ast\beta_2C_{27})+\\ &(C_{21}\beta_2+u_3^\ast\beta_2C_{24})(\eta_1^2+2\eta_1\bar{\eta_1})+(C_{22}\beta_2+u_3^\ast\beta_2C_{25})(2\eta_1+\bar{\eta_1}))], \end{split}
    g_{02} = 2\bar{g_{20}}\bar{\mathfrak{D}}/\mathfrak{D}.

    Since W(z(t),\bar{z(t)}) satisfies the following equation

    \begin{split} \dot W = &\mathcal{A}W+X_0f(\Phi\cdot(z, \bar z)^T+W(z,\bar z),0)-\Phi\Psi(0)f(\Phi\cdot(z,\bar z)^T+W(z,\bar z),0)\\ = &\mathcal{A}W+H_{20}\frac{z^2}{2}+H_{11}z\bar z+H_{02}\frac{\bar z^2}{2}+\cdot\cdot\cdot \label{Eq4.30} \end{split} (43)

    we obtain

    \left\{\begin{array}{ll} (2\mathrm{i}\hat\omega\hat\tau-\mathcal{A})W_{20} = H_{20},\\ -\mathcal{A}W_{11} = H_{11},\\ (-2\mathrm{i}\hat\omega\hat\tau-\mathcal{A})W_{02} = H_{02}. \end{array}\right. \label{Eq4.31} (44)

    Since \mathcal{A} has only two eigenvalues \pm \mathrm{i}\hat\omega\hat\tau, then Eq. (44) has only a unique solution W_{ij}.

    We first compute H_{ij}(\theta), \theta\in[-1, 0]. From (43), we know that for -1\leq\theta<0,

    H(z, \bar z) = -\Phi\Psi(0)f(\Phi\cdot(z,\bar z)^T+W(z,\bar z), 0).

    Therefore, by comparing the coefficients, and notice that

    H(z, \bar z)(0) = f(\Phi\cdot(z, \bar z)^T+W(z,\bar z), 0)-\Phi\Psi(0)f(\Phi\cdot(z,\bar z)^T+W(z,\bar z), 0),

    we obtain

    \begin{split} &H_{20}(\theta) = \\ &\left\{\begin{array}{ll} -(g_{20}p(\theta)+\bar{g_{02}}\bar{p(\theta)}),\quad \quad \theta\in[-1,0),\\ 2\hat\tau\left[\begin{array}{c} ({bA_3}/{u_1^\ast}-C_{11}u_1^\ast-1)\eta_1^2-\eta_1\eta_2-C_{13}u_1^\ast-(A_1+C_{12}u_1^\ast)\eta_1\\ \beta_1\eta_1\eta_2e^{-2\mathrm{i}\hat\omega\hat\tau}-\alpha\eta_2\\ u_3^\ast\beta_2C_{21}\eta_1^2+C_{23}u_3^\ast\beta_2-M_3/u_3^\ast+(\beta_2A_2+\beta_2C_{22}u_3^\ast)\eta_1+\beta\eta_2 \end{array}\right]\\ -[g_{20}p(0)+\bar{g_{02}}\bar {p(0)}], \quad \quad \theta = 0. \end{array}\right. \end{split}
    \begin{split} &H_{11}(\theta) = \\ &\left\{\begin{array}{ll} -(g_{11}p(\theta)+\bar{g_{11}}\bar{p(\theta)}),\quad \quad \theta\in[-1,0),\\ 2\hat\tau\left[\begin{array}{c} (\frac{bA_3}{u_1^\ast}-C_{11}u_1^\ast-1)\eta_1\bar{\eta_1}-\mathrm{Re}\{\eta_1\bar{\eta_2}\}-C_{13}u_1^\ast -(A_1+C_{12}u_1^\ast)\mathrm{Re}\{\eta_1\}\\ \beta_1\mathrm{Re}\{\eta_1\bar{\eta_2}\}-\alpha\mathrm{Re}\{\eta_2\}\\ u_3^\ast\beta_2C_{21}\eta_1\bar{\eta_1}+C_{23}u_3^\ast\beta_2-\frac{M_3}{u_3^\ast}+(\beta_2A_2+\beta_2C_{22}u_3^\ast)\mathrm{Re}\{\eta_1\}+\beta\mathrm{Re}\{\eta_2\} \end{array}\right]\\ -[g_{11}p(0)+\bar{g_{11}}\bar{p(0)}], \quad \quad \theta = 0. \end{array}\right. \end{split}

    It follows from (44) and the definition of \mathcal{A} that

    \begin{split} &\dot W_{20}(\theta) = 2\mathrm{i}\omega_n\hat\tau W_{20}(\theta)+[g_{20}p(\theta)+\bar{g_{02}}\bar p(\theta)], -1\leq\theta\leq 0,\\ &-\dot W_{11}(\theta) = -[g_{11}p(\theta)+\bar{g_{11}}\bar{p(\theta)}], -1\leq \theta\leq 0. \end{split}

    Noting that p(\theta) = p(0)e^{\mathrm{i}\hat\omega\hat\tau\theta}, -1\leq\theta\leq 0, we have

    \begin{split} &W_{20}(\theta) = [\frac{\mathrm{i}g_{20}}{\hat\omega\hat\tau}p(\theta)+\frac{\mathrm{i}\bar{g_{02}}}{3\hat\omega\hat\tau}\bar{p(\theta)}] +E_1e^{2\mathrm{i}\hat\omega\hat\tau\theta},\\ &W_{11}(\theta) = [\frac{-\mathrm{i}g_{11}}{\hat\omega\hat\tau}p(\theta)+\frac{\mathrm{i}\bar{g_{11}}}{\hat\omega\hat\tau}\bar{p(\theta)}] +E_2. \label{Eq4.32} \end{split} (45)

    Utilizing the definition of \mathcal{A}, (44) and (45) yields

    \begin{split} &E_1 = 2\left[\begin{array}{c} ({bA_3}/{u_1^\ast}-C_{11}u_1^\ast-1)\eta_1^2-\eta_1\eta_2-C_{13}u_1^\ast-(A_1+C_{12}u_1^\ast)\eta_1\\ \beta_1\eta_1\eta_2e^{-2\mathrm{i}\hat\omega\hat\tau}-\alpha\eta_2\\ u_3^\ast\beta_2C_{21}\eta_1^2+C_{23}u_3^\ast\beta_2-M_3/u_3^\ast+(\beta_2A_2+\beta_2C_{22}u_3^\ast)\eta_1+\beta\eta_2 \end{array}\right]\times\\ &\left[ \begin{array}{ccc} 2\mathrm{i}\hat\omega+d_1\mu_i+M_1&u_1^\ast&u_1^\ast A_1\\ -\beta_1u_2^\ast e^{-2\mathrm{i}\hat\omega\hat\tau}&2\mathrm{i}\hat\omega+d_2\mu_i+M_2(1-e^{-2\mathrm{i}\hat\omega\hat\tau})&\alpha u_2^\ast\\ -\beta_2A_2u_3^\ast &-\beta u_3^\ast&2\mathrm{i}\hat\omega+d_3\mu_i+M_3 \end{array}\right ]^{-1}\\ \end{split}

    and

    \begin{array}{l} {E_2} = 2{\left[ {\begin{array}{*{20}{c}} {{d_1}{\mu _i} + u_1^ * - b{A_3}}&{u_1^ * }&{u_1^ * {A_1}}\\ { - {\beta _1}u_2^ * }&{{d_2}{\mu _i}}&{\alpha u_2^ * }\\ { - {\beta _2}{A_2}u_3^ * }&{ - \beta u_3^ * }&{{d_3}{\mu _i} + c{\beta _2}{A_3}} \end{array}} \right]^{ - 1}} \times \\ \quad \left[ {\begin{array}{*{20}{c}} {(\frac{{b{A_3}}}{{u_1^ * }} - {C_{11}}u_1^ * - 1){\eta _1}\overline {{\eta _1}} - {\rm{Re}}\{ {\eta _1}\overline {{\eta _2}} \} - {C_{13}}u_1^ * - ({A_1} + {C_{12}}u_1^ * ){\rm{Re}}\{ {\eta _1}\} }\\ {{\beta _1}{\rm{Re}}\{ {\eta _1}\overline {{\eta _2}} \} - \alpha {\rm{Re}}\{ {\eta _2}\} }\\ {u_3^ * {\beta _2}{C_{21}}{\eta _1}\overline {{\eta _1}} + {C_{23}}u_3^ * {\beta _2} - \frac{{{M_3}}}{{u_3^ * }} + ({\beta _2}{A_2} + {\beta _2}{C_{22}}u_3^ * ){\rm{Re}}\{ {\eta _1}\} + \beta {\rm{Re}}\{ {\eta _2}\} } \end{array}} \right]. \end{array}

    Now, we can compute the following values

    \begin{split} &c_1(0) = \frac{\mathrm{i}}{2\hat\omega\hat\tau}(g_{20}g_{11}-2|g_{11}|^2-\frac{|g_{02}|^2}{3})+\frac{ g_{21}}{2}, \nu_2 = -\frac{\mathrm{Re}(c_1(0))}{\mathrm{Re}(\lambda^\prime(\hat\tau))},\\ &\beta_2 = 2\mathrm{Re}(c_1(0)), T_2 = -\frac{\mathrm{Im}(c_1(0))+\nu_2\mathrm{Im}(\lambda^\prime(\hat\tau))}{\hat\omega\hat\tau}, \end{split}

    which determine the properties of bifurcating periodic solutions at critical value \hat\tau, that is, \nu_2 determines the directions of the Hopf bifurcation; \beta_2 determines the stability of the bifurcating periodic solutions; T_2 determines the period of bifurcating periodic solutions. Moreover, by Hassard [41], we have the following result.

    Theorem 4.11. Assume that the conditions of Theorem 4.5 are satisfied, we have

    (ⅰ) If \nu_2>0\,(<0), then the direction of the Hopf bifurcation is forward (backward).

    (ⅱ)If \beta_2<0\,(>0), then the bifurcating periodic solutions are orbitally stable (unstable).

    (ⅲ) If T_2>0\,(<0), then the period of the bifurcating periodic solutions increases (decreases).


    5. Numerical simulations


    5.1. The spatiotemporal dynamics in one-dimensional space

    In this subsection, we numerically explore the dynamic behavior of System (4) with one-dimensional space, namely n=1, \Delta=\frac{\partial^2}{\partial x^2} and we take \Omega=(0,\pi).

    For the choice of the parameter values in System (4), we refer to [10,11,12,15] and choose parameter values as follows

    ({P_1}):\alpha = 0.7,\beta = 0.9,{\beta _1} = 1.95,{\beta _2} = 1.8,{\gamma _1} = 0.2,{\gamma _2} = 0.8,b = 2.5,c = 5.5,{d_1} = 0.4,{d_2} = 0.3,{d_3} = 0.2

    and the initial conditions as

    (I{C_1}):{\phi _1}(t,x) = 0.1717 + 0.001\cos x,{\phi _2}(t,x) = 0.7509 + 0.001\cos x,{\phi _3}(t,x) = 0.1926 + 0.001\cos x.

    With these parameter values, System (4) admits a unique positive spatially homogeneous steady state E^\ast = (u_1^\ast, u_2^\ast, u_3^\ast)\approx (0.1717, 0.7509, 0.1926). It is easy to check that the hypotheses (H_1)-(H_6) hold and h^\prime((\omega^\ast)^2) = 0.0938>0. By Theorem 4.9, local Hopf bifurcation occurs at \tau^\ast\approx0.7859. We use the forward Euler method to find numerical solutions to System (4) with \tau = 0.7<\tau^* and \tau = 1.2>\tau^*, respectively. As illustrated in Figures 1 and 2, when \tau = 0.7<\tau^*, solutions of (4) approach the steady state E^*, while when \tau>\tau^*, sustained oscillations are observed. Calculations give c_1(0) =-4.2942-30.9399\mathrm{i}, \nu_2 = 83.06, \beta_2 = -8.5884, T_2 = 185.7969. Thus the Hopf bifurcation is forward and the bifurcated periodic solutions from E^\ast are stable and the period of bifurcated periodic solution increases in \tau for \tau>\tau^\ast.

    Figure 1. Numerical solutions of (4) with \tau = 0.7<\tau^\ast\approx0.7895 (only the u_1 component is plotted here): the positive spatially homogeneous steady state is locally stable.
    Figure 2. Numerical solutions of (4) with \tau = 1.2>\tau^\ast\approx0.7895: a periodic solution bifurcates from the positive spatially homogeneous steady state E^\ast.

    5.2. The spatiotemporal dynamics in two-dimensional space

    Consider System (4) with u_1=u_1(t,x,y), u_2=u_2(t,x,y), u_3=u_3(t,x,y) and \Delta=\frac{\partial^2}{\partial x^2}+\frac{\partial^2}{\partial y^2}. For this purpose, the domain of System (4) is confined to a fixed spatial domain \Omega=[0, L]\times [0,L]\subset R^2 with L=400. we solve System (4) on a grid with 400\times 400 sites by a simple Euler method with a time step size of \delta t=0.01 and a space step size of \delta x=\delta y=1. The discretization of the Laplacian term takes the form

    \begin{split} \Delta u|_{(i,j)} = &\frac{1}{s^2}[A_l(i,j)u(i-1,j)+A_r(i,j)u(i+1,j)+\\ &A_d(i,j)u(i,j-1)+A_u(i,j)u(i,j+1)-4u(i,j)], \end{split}

    where (i,j) denote the lattice sites and s = 1 is the lattice constant. The matrix elements of A_l, A_r, A_d, A_u are unity except at the boundary. When (i,j) lies on the left boundary, that is i = 0, we define A_l(i,j)u(i-1,j)\equiv u(i+1,j), which guarantees zero-flux of individuals in the left boundary. Similarly we define A_r(i,j), A_d(i,j), A_u(i,j) such that the zero-flux boundary condition is satisfied.

    With the given Neumann boundary conditions, the eigenvalues of -\Delta on \Omega are \mu_i = \frac{\pi^2}{L^2}(n^2+m^2), n, m\in \mathbb{Z}, where \mathbb{Z} represents the integer set. In order to discuss the impacts of delay and diffusion on the dynamics of System (4), we will compare the temporal model (that is, System (4) without diffusion) with System (4). We take parameters as

    ({P_2}):{\mkern 1mu} \alpha = 0.7,\beta = 0.9,{\beta _1} = 1.95,{\beta _2} = 1.85,{\gamma _1} = 0.2,{\gamma _2} = 0.8,b = 2.5,c = 5,{d_1} = 1,{d_2} = 2,{d_3} = 4.

    We consider two types of different initial conditions:

    (IC_2): \phi_1(t,x,y) = u_1^{\ast}+0.001\sin x\sin y, \phi_2(t,x,y) = u_2^{\ast}+0.001\sin x\sin y,

    \phi_3(t,x,y) = u_3^{\ast}+0.001\sin x \sin y,

    and

    (IC_2^\prime):\left\{\begin{array}{lll} u_1(t,x,y) = 0.1754-\varepsilon_1(x-0.1y-225)(x-0.1y-675),\\ u_2(t,x,y) = 0.7419-\varepsilon_2(x-450)-\varepsilon_3(y-150),\\ u_3(t,x,y) = 0.2029-\varepsilon_4(x-350)-\varepsilon_5(y-200)\\ \end{array}\right.

    for (t,x,y)\in[-\tau,0]\times\Omega. Here \varepsilon_1 = 2\times 10^{-7}, \varepsilon_2 = 3\times 10^{-5}, \varepsilon_3 = 1.2\times 10^{-3}, \varepsilon_4 = 6\times 10^{-5}, \varepsilon_5 = 3\times 10^{-5}.

    Under (P_2) and (IC_2), it is easy to check that all conditions of Theorem 4.9 are satisfied and there is a unique positive steady state E^{\ast}\approx(0.1754, 0.7419, 0.2029). The corresponding Hopf bifurcation value is computed as \tau^1_{1,0}\approx 0.8028.

    Figure 3 depicts the population dynamics of the temporal model and the spatiotemporal model at \tau = 1. Both the temporal model and the spatiotemporal model undergo periodic oscillations. However, the temporal model exhibits irregular transient oscillations initially.

    Figure 3. Numerical solutions of the temporal model (left) and numerical solutions of the spatiotemporal model (right) with \tau = 1, (P_2) and (IC_2). Here, for the spatiotemporal model (4), average population density for each species is plotted.

    If we increase \tau to \tau = 1.5 and use initial condition (IC_2'), as shown in Figure 4, the temporal model still exhibits regular oscillations, while the spatiotemporal model exhibits irregular oscillations and the calculated Lyapunov exponent is 0.0011>0 (By the method proposed in [42]), which indicates the occurrence of chaos.

    Figure 4. Numerical solutions of the temporal model (left) and numerical solutions of the spatiotemporal model (right) with \tau = 1.5, (P_2) and (IC_2'). Here, periodic oscillations are observed for the temporal model and chaotic behavior is observed for the spatiotemporal model.

    Figure Figure 5 depicts the snapshots of the contour maps of specie u_1 for the temporal model and the spatiotemporal model at time t = 5000. The temporal model exhibits the spiral wave pattern. However, the spatiotemporal model presents the chaotic wave pattern. To have a better understanding on the evolution process of the spatiotemporal pattern, in Figure Figure 6, we present the snapshots of contour maps of the basal resource u_1 at t = 200,500,1000,1200,1500,2500, respectively. As pointed out in [43], the spirals are usually observed under suitable parametric conditions near Turing-Hopf bifurcation threshold. In addition, in the spatially extended system the existence of a stable limit cycle normally results in the formation of chaotic spatiotemporal patterns [44]. As can be seen from Figure Figure 6, as time t increases, an chaotic wave spatial pattern is gradually formed starting from a regular spiral wave pattern.

    Figure 5. Snapshots of contour maps of the basal resource u_1 for the temporal model (left) and spatiotemporal model (right) at t = 2000 with \tau = 1.5, (P_2) and (IC_2^\prime).
    Figure 6. Snapshots of contour maps of the time evolution of the specie u_1 at t = 200, 500, 1000, 1200, 1500, 2500 with \tau = 1.5 under (P_2) and (IC_2^\prime).

    5.2.1. The effect of delay

    To explore the impact of delay, in Figure Figure 7, we take the snapshots of the contour maps of specie u_1 at time t = 1500 for several different values of \tau. As can be seen in Figure Figure 7, the time delay can lead to the formation of an irregular spatial pattern from a regular spiral pattern in the whole domain as the time delay increases and surpasses some critical value.

    Figure 7. Snapshots of contour maps of the time evolution of the basal resource u_1 with different values of \tau at time t = 1500 under (P_2) and (IC_2^\prime). (\mathrm{ⅰ})\,\tau = 0.86; (\mathrm{ⅱ})\,\tau = 1; (\mathrm{ⅲ})\,\tau = 1.2; (\mathrm{ⅳ})\,\tau = 1.4; (\mathrm{ⅴ})\,\tau = 1.6; (\mathrm{ⅵ})\,\tau = 1.9.

    5.2.2. The effect of diffusion

    As seen from Figure Figure 6, System (4) has a regular spiral wave pattern when \tau = 1.5, d_1 = 1, d_2 = 2 and d_3 = 4 at t = 1500. To numerically examine how the diffusion affects the pattern, we take the snapshots of contour maps of u_1 at t = 1500 with several different choices of the diffusion coefficients. As shown in Figure Figure 8, spiral wave pattern emerges firstly when d_1 = d_2 = d_3 = 0, then as the three diffusion coefficients change to d_1 = d_2 = 0.01,d_3 = 0.04, the spiral wave structure disappears around the center of the spirals wave, with the increase in these diffusion coefficients, it grows steadily, and eventually the chaotic wave pattern dominates the whole domain. Differing from the instability mechanism in Figure Figure 7, the spirals wave loses its stability due to the Doppler effect [45].

    Figure 8. Snapshots of contour maps of the basal resource u_1 at time t = 1500 with different diffusion coefficients, \tau = 1.5, under (P_2) and (IC_2^\prime).

    5.2.3. The impact of the prey saturation constant b

    Figure 9 demonstrates how the prey saturation constant constant b affects the pattern formation of the basal resource u_1 at time t = 1500 with \tau = 1.5: when b is small, we observe a pattern with stripes firstly; as the constant b increases, the spiral wave pattern emerges, then it grows steadily, as b goes beyond a certain value, chaotic wave pattern appears.

    Figure 9. Snapshots of contour maps of the time evolution of the basal resource u_1 with different values of b and parameter values \alpha = 0.7, \beta = 0.9, \beta_1 = 1.95, \beta_2 = 1.85, \gamma_1 = 0.2, \gamma_2 = 0.8, c = 5 at times t = 1500 and \tau = 1.5 under(IC_2^\prime).

    5.2.4. The impact of the predator interference constant c

    To see how the predator interference constant c influences the spatiotemporal pattern, we numerically simulation (4) with different values of c and plot the snapshots of the contour maps of the basal resource u_1 at t = 1500 in Figure 10. It is illustrated in Figure Figure 10 that the predator interference constatn c can also lead to the formation of chaotic wave spatial pattern, which can be preceded from the evolution of a regular spiral patterns as the predator interference constant c decreases.

    Figure 10. 10. Snapshots of contour maps of the time evolution of the basal resource u_1 with different values of c and parameter values \alpha = 0.7, \beta = 0.9, \beta_1 = 1.95, \beta_2 = 1.85, \gamma_1 = 0.2, \gamma_2 = 0.8, b = 0.25 at times t = 1500 and \tau = 1.5 under(IC_2^\prime).

    6. A summary and discussion

    In this work, we have investigated the spatiotemporal dynamics of a diffusive IGP model with delay and the Beddington-DeAngelis functional response. we have established locally asymptotically stability results of the trivial, semi-trivial and strong semi-trivial steady states. In the case that there is a unique positive spatially homogeneous steady state E^\ast, we have carried out the Hopf bifurcation analysis. Unlike competition models with monotone response functions ([17]) where delay does not induce sustained oscillations, in our IGP models, delay promotes complex dynamics including bistability, and the emergence of spiral wave pattern and chaotic wave pattern.

    Compared with the temporal model in [24], we also observe bistability is possible in System (4). In addition, the diffusion also has impacts on the formation of spatiotemporal patterns as it can change the distribution of characteristic roots of the corresponding characteristic equations, and hence has an important effect on the dynamics for the constant steady state of System (4). This has been illustrated via numerical simulations as well (See Figure 8). Moreover, we have observed that the functional response can also influence the formation of complex patterns. As demonstrated in Figures 9 and 10, the functional responses can also trigger the emergence of spiral wave pattern and chaotic wave spatial pattern.


    Acknowledgments

    The authors were very grateful to two anonymous reviewers' very helpful comments and suggestions. The project was partially supported by the National Natural Science Foundation of China (No.51479215, No. 11401577) and the Natural Sciences and Engineering Research Council of Canada (RGPIN-2015-05686).


    [1] [ J. D. Murray, Mathematical Biology II: Spatial Models and Biomedical Applications, Spring-Verlag, New York, 2003.
    [2] [ G. A. Polis,C. A. Myers,R. D. Holt, The ecology and evolution of intraguild predation: Potential competitors that each other, Ann. Rev. Ecol. Sys., 20 (1989): 297-330.
    [3] [ M. H. Posey,A. H. Hines, Complex predator-prey interactions within an estuarine benthic community, Ecol., 72 (1991): 2155-2169.
    [4] [ G. A. Polis,R. D. Holt, Intraguild predation: The dynamics of complex trophic interactions, Trends Ecol. Evol., 7 (1992): 151-154.
    [5] [ R. D. Holt,G. A. Polis, A theoretical framework for intraguild predation, Am. Nat., 149 (1997): 745-764.
    [6] [ M. Arim,P. A. Marquet, Intraguild predation: A widespread interaction related to species biology, Ecol. Let., 7 (2004): 557-564.
    [7] [ P. Amarasekare, Trade-offs, temporal, variation, and species coexistence in communities with intraguild predation, Ecol., 88 (2007): 2720-2728.
    [8] [ R. Hall, Intraguild predation in the presence of a shared natural enemy, Ecol., 92 (2011): 352-361.
    [9] [ Y. S. Wang,D. L. DeAngelis, Stability of an intraguild predation system with mutual predation, Commun. Nonlinear Sci. Numer. Simulat., 33 (2016): 141-159.
    [10] [ I. Velazquez,D. Kaplan,J. X. Velasco-Hernandez,S. A. Navarrete, Multistability in an open recruitment food web model, Appl. Math. Comp., 163 (2005): 275-294.
    [11] [ S. B. Hsu,S. Ruan,T. H. Yang, Analysis of three species Lotka-Volterra food web models with omnivory, J. Math. Anal. Appl., 426 (2015): 659-687.
    [12] [ P. A. Abrams,S. R. Fung, Prey persistence and abundance in systems with intraguild predation and type-2 functional response, J. Theor. Biol., 264 (2010): 1033-1042.
    [13] [ A. Verdy,P. Amarasekare, Alternative stable states in communities with intraguild predatiion, J. Theor. Biol., 262 (2010): 116-128.
    [14] [ M. Freeze,Y. Chang,W. Feng, Analysis of dynamics in a complex food chain with ratio-dependent functional response, J. Appl. Anal. Comput., 4 (2014): 69-87.
    [15] [ Y. Kang,L. Wedekin, Dynamics of a intraguild predation model with generalist or specialist predator, J. Math. Biol., 67 (2013): 1227-1259.
    [16] [ H. I. Freedman,V. S. H. Rao, Stability criteria for a system involving two time delays, SIAM J. Appl. Math., 46 (1986): 552-560.
    [17] [ G. S. K. Wolkowicz,H. X. Xia, Global asymptotic behavior of chemostat model with discrete delays, SIAM J. Appl. Math., 57 (1997): 1019-1043.
    [18] [ Y. L. Song,M. A. Han,J. J. Wei, Stability and Hopf bifurcation analysis on a simplified BAM neural network with delays, Physica D, 200 (2005): 185-204.
    [19] [ S. Ruan, On nonlinear dynamics of predator-prey models with discrete delay, Math. Mod. Nat. Phen., 4 (2009): 140-188.
    [20] [ X. Y. Meng,H. F. Huo,X. B. Zhao,H. Xiang, Stability and Hopf bifurcation in a three-species system with feedback delays, Nonlinear Dyn., 64 (2011): 349-364.
    [21] [ M. Y. Li,H. Shu, Multiple stable periodic oscillations in a mathematical model of CTL-response to HTLV-I infection, Bull. Math. Biol., 73 (2011): 1774-1793.
    [22] [ H. Shu,L. Wang,J. Watmough, Sustained and transient oscillations and chaos induced by delayed antiviral inmune response in an immunosuppressive infective model, J. Math. Biol., 68 (2014): 477-503.
    [23] [ M. Yamaguchi,Y. Takeuchi,W. Ma, Dynamical properties of a stage structured three-species model with intra-guild predation, J. Comput. Appl. Math., 201 (2007): 327-338.
    [24] [ H. Shu,X. Hu,L. Wang,J. Watmough, Delayed induced stability switch, multitype bistability and chaos in an intraguild predation model, J. Math. Biol., 71 (2015): 1269-1298.
    [25] [ A. Okubo and S. A. Levin, Diffusion and Ecological Problems: Modern perspectives, Springer-Verlag, New York, 2001.
    [26] [ T. Faria, Stability and bifurcation for a delayed predator-prey model and the effect of diffusion, J. Math. Anal. Appl., 254 (2001): 433-463.
    [27] [ C. V. Pao, Systems of parabolic equations with continuous and discrete delays, J. Math. Anal. Appl., 205 (1997): 157-185.
    [28] [ C. V. Pao, Convergence of solutions of reaction-diffusion systems with time delays, Nonlinear Anal., 48 (2002): 349-362.
    [29] [ J. Wang,J. P. Shi,J. J. Wei, Dyanmics and pattern formation in a diffusive predator-prey system with strong Allee effect in prey, J. Diff. Equat., 251 (2011): 1276-1304.
    [30] [ C. Tian, Delay-driven spatial patterns in a plankton allelopathic system, Chaos, 22(2012), 013129, 7 pp.
    [31] [ C. Tian,L. Zhang, Hopf bifurcation analysis in a diffusive food-chain model with time delay, Comput. Math. Appl., 66 (2013): 2139-2153.
    [32] [ W. Zuo,J. Wei, Global stability and Hopf bifurcations of a Beddington-DeAngelis type predator-prey system with diffusion and delay, Appl. Math. Comput., 223 (2013): 423-435.
    [33] [ J. Zhao,J. Wei, Dynamics in a diffusive plankton system with delay and toxic substances effect, Nonlinear Anal., 22 (2015): 66-83.
    [34] [ L. Zhu,H. Zhao,X. M. Wang, Bifurcation analysis of a delay reaction-diffusion malware propagation model with feedback control, Commun. Nonlinear Sci. Numer. Simulat., 22 (2015): 747-768.
    [35] [ Y. Li,M. X. Wang, Hopf bifurcation and global stability of a delayed predator-prey model with prey harvesting, Comput. Math. Appl., 69 (2015): 398-410.
    [36] [ H. Y. Zhao,X. Zhang,X. Huang, Hopf bifurcation and spatial patterns of a delayed biological economic system with diffusion, Appl. Math. Comput., 266 (2015): 462-480.
    [37] [ Q. X. Ye, Z. Y. Li, M. X. Wang and Y. P. Wu, Introduction to Reaction-diffusion Equations (Second Edition), Science Press, Bei Jing, 2011.
    [38] [ D. Henry, Geometric Theory of Semilinear Parabolic Equations, Springer-Verlag, Berlin/New York, 1981.
    [39] [ S. Ruan,J. Wei, On the zeros of a third degree exponential polynomial with applications to a delayed model for the control of testoterone secretion, Math. Med. Biol., 18 (2001): 41-52.
    [40] [ J. Wu, Theory and Applications of Partial Functional Differential Equations, Springer-Verlag, New York, 1996.
    [41] [ B. Hassard, N. Kazarinoff and Y. Wan, Theory and Applications of Hopf bifurcation, Cambridge University Press, Cambridge, 1981.
    [42] [ J. Y. Wakano,C. Hauert, Pattern formation and chaos in spatial ecological public goods games, J. Theor. Biol., 268 (2011): 30-38.
    [43] [ M. Banerjee, S. Ghoral and N. Mukherjee, Approximated spiral and target patterns in Bazykin's prey-predator model: Multiscale perturbation analysis, Int. J. Bifurcat. Chaos, 27 (2017), 1750038, 14 pp.
    [44] [ H. Malchow, S. V. Petrovskii and E. Venturino, Spatiotemporal Patterns in Ecology and Epidemiology: Theory, Models, Simulations, Chapman & Hall / CRC Press, 2008.
    [45] [ Q. Ouyang, Pattern Formation in Reaction-Diffusion Systems Shanghai Scientific and Technological Education Publishing House, SHANGHAI, 2000.
  • This article has been cited by:

    1. Jingen Yang, Sanling Yuan, Tonghua Zhang, Complex dynamics of a predator–prey system with herd and schooling behavior: with or without delay and diffusion, 2021, 0924-090X, 10.1007/s11071-021-06343-0
    2. Dawei Zhang, Binxiang Dai, Spreading and vanishing in a diffusive intraguild predation model with intraspecific competition and free boundary, 2019, 42, 0170-4214, 6917, 10.1002/mma.5797
    3. Dawei Zhang, Binxiang Dai, A free boundary problem for the diffusive intraguild predation model with intraspecific competition, 2019, 474, 0022247X, 381, 10.1016/j.jmaa.2019.01.050
    4. Dawei Zhang, Binxiang Dai, The diffusive intraguild predation model with intraspecific competition and double free boundaries, 2020, 0003-6811, 1, 10.1080/00036811.2020.1716971
    5. Zhenzhen Li, Binxiang Dai, ANALYSIS OF DYNAMICS IN A GENERAL INTRAGUILD PREDATION MODEL WITH INTRASPECIFIC COMPETITION, 2019, 9, 2156-907X, 1493, 10.11948/2156-907X.20180296
    6. Renji Han, Binxiang Dai, Yuming Chen, Pattern formation in a diffusive intraguild predation model with nonlocal interaction effects, 2019, 9, 2158-3226, 035046, 10.1063/1.5084948
    7. Juping Ji, Russell Milne, Hao Wang, Stoichiometry and environmental change drive dynamical complexity and unpredictable switches in an intraguild predation model, 2023, 86, 0303-6812, 10.1007/s00285-023-01866-z
    8. Zhenzhen Li, Binxiang Dai, Renji Han, Hopf Bifurcation in a Reaction–Diffusion–Advection Population Model with Distributed Delay, 2022, 32, 0218-1274, 10.1142/S0218127422502479
    9. Juping Ji, Lin Wang, Competitive exclusion and coexistence in an intraguild predation model with Beddington–DeAngelis functional response, 2022, 107, 10075704, 106192, 10.1016/j.cnsns.2021.106192
    10. Renji Han, Global classical solvability and stabilization in a reaction–diffusion intraguild predation model with chemotaxis, 2022, 73, 0044-2275, 10.1007/s00033-022-01777-x
    11. 小宁 王, Stability and Bifurcation Analysis of a Spatiotemporal Intraguild PredationModel with Fear Effect and Beddington-DeAngelis Functional Response, 2022, 12, 2160-7583, 2081, 10.12677/PM.2022.1212225
    12. Heping Jiang, Stable spatially inhomogeneous periodic solutions for a diffusive Leslie–Gower predator–prey model, 2024, 1598-5865, 10.1007/s12190-024-02018-2
    13. Renji Han, Sanaa Moussa Salman, Nonlinear dynamics and pattern formation in a space-time discrete diffusive intraguild predation model, 2024, 01672789, 134295, 10.1016/j.physd.2024.134295
    14. Suparna Dash, Kankan Sarkar, Subhas Khajanchi, Spatiotemporal Dynamics of an Intraguild Predation Model with Intraspecies Competition, 2024, 34, 0218-1274, 10.1142/S0218127424300301
    15. Liqiong Pu, Haotian Tang, Jiashan Zheng, Smooth Solution in an N$$ N $$‐Dimensional Chemotaxis Model With Intraguild Predation, 2025, 0170-4214, 10.1002/mma.10694
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4496) PDF downloads(610) Cited by(15)

Article outline

Figures and Tables

Figures(10)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog