Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay

  • Recent studies demonstrate that the reproduction of prey is suppressed by the fear of predators. However, it will not respond immediately to fear, but rather reduce after a time lag. We propose a diffusive predator-prey model incorporating fear response delay into prey reproduction. Detailed bifurcation analysis reveals that there are three different cases for the effect of the fear response delay on the system: it might have no effect, both stabilizing and destabilizing effect, or destabilizing effect on the stability of the positive equilibrium, respectively, which are found by numerical simulations to correspond to low, intermediate or high level of fear. For the second case, through ordering the critical values of Hopf bifurcation, we prove the existence of stability switches for the system. Double Hopf bifurcation analysis is carried out to better understand how the fear level and delay jointly affect the system dynamics. Using the normal form method and center manifold theory, we derive the normal form of double Hopf bifurcation, and obtain bifurcation sets around double Hopf bifurcation points, from which all the dynamical behaviors can be explored, including periodic solutions, quasi-periodic solutions and even chaotic phenomenon.

    Citation: Mengting Sui, Yanfei Du. Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay[J]. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262

    Related Papers:

    [1] Ruizhi Yang, Dan Jin . Dynamics in a predator-prey model with memory effect in predator and fear effect in prey. Electronic Research Archive, 2022, 30(4): 1322-1339. doi: 10.3934/era.2022069
    [2] Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150
    [3] Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128
    [4] Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109
    [5] Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215
    [6] Jiani Jin, Haokun Qi, Bing Liu . Hopf bifurcation induced by fear: A Leslie-Gower reaction-diffusion predator-prey model. Electronic Research Archive, 2024, 32(12): 6503-6534. doi: 10.3934/era.2024304
    [7] San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045
    [8] Jiange Dong, Xianyi Li . Bifurcation of a discrete predator-prey model with increasing functional response and constant-yield prey harvesting. Electronic Research Archive, 2022, 30(10): 3930-3948. doi: 10.3934/era.2022200
    [9] Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309
    [10] Jie Xia, Xianyi Li . Bifurcation analysis in a discrete predator–prey model with herd behaviour and group defense. Electronic Research Archive, 2023, 31(8): 4484-4506. doi: 10.3934/era.2023229
  • Recent studies demonstrate that the reproduction of prey is suppressed by the fear of predators. However, it will not respond immediately to fear, but rather reduce after a time lag. We propose a diffusive predator-prey model incorporating fear response delay into prey reproduction. Detailed bifurcation analysis reveals that there are three different cases for the effect of the fear response delay on the system: it might have no effect, both stabilizing and destabilizing effect, or destabilizing effect on the stability of the positive equilibrium, respectively, which are found by numerical simulations to correspond to low, intermediate or high level of fear. For the second case, through ordering the critical values of Hopf bifurcation, we prove the existence of stability switches for the system. Double Hopf bifurcation analysis is carried out to better understand how the fear level and delay jointly affect the system dynamics. Using the normal form method and center manifold theory, we derive the normal form of double Hopf bifurcation, and obtain bifurcation sets around double Hopf bifurcation points, from which all the dynamical behaviors can be explored, including periodic solutions, quasi-periodic solutions and even chaotic phenomenon.



    The interaction between predator and prey is one of the most important topics in mathematical biology and theoretical ecology [1]. The direct interaction, which is reflected by predation, has been extensively studied [2,3,4,5]. Recently, a field experimental study [6] provided evidence that show that indirect effect (e.g., fear effect) cannot be ignored. Even though there is no direct killing between predators and prey, the presence of predators cause a reduction in prey population [6,7]. To explore the impact that fear can have on population dynamics, Wang et al. [8] formulated the following model incorporating the cost of fear

    {dudt=f(k,v(t))r0u(t)du(t)au2(t)g(u(t))v(t),dvdt=cg(u(t))v(t)mv(t), (1.1)

    where u and v are the population densities of the prey and predator, respectively. Function f(k,v(t)) reflects the cost of anti-predation response due to fear, where k measures the level of fear. r0 is the reproduction rate of prey in the absence of predator, d and m represent the natural death rate, a reflects the death rate due to intra-specific competition, the positive constant c is the efficiency in biomass transfer, g(u(t)) is the functional response. After their work, other biological phenomena, such as the Allee effect, harvesting, cooperation hunting, prey refuge, group defense and so on, were introduced into predator-prey models with the fear effect [1,9,10,11,12,13,14,15].

    In fact, besides the cost in the reproduction of prey due to fear, there are also some benefits for an anti-predation response. Wang and Zou [16] described such benefits with g(u(t),k) instead of g(u(t)), and considered both linear functional response and Holling type II functional response. In the model with the Holling type II functional response, they choose g(u(t),k)=11+c1kpu(t)1+qu(t), f(k,v)=11+c2kv(t) and get the following model

    {dudt=r0u(t)1+c2kv(t)du(t)au2(t)pu(t)v(t)1+qu(t)11+c1k,dvdt=pu(t)v(t)1+qu(t)c1+c1kmv(t), (1.2)

    where c1,c2 represent the decreasing rate of reproduction and predation, respectively.

    To obtain more resources, species tend to migrate from a high population density area to a low population density area. Therefore, spatial diffusion should be considered when we model the interaction between predator and prey. Many diffusive models have been proposed to investigate the influence of the cost of fear on the spatial distribution of species [17,18,19,20]. Wang and Zou [21] proposed a reaction-diffusion-advection predator prey model, in which conditions of spatial pattern formation are obtained.

    In reality, there are time delays in almost every process of predator and prey interaction. Many kinds of delays have been incorporated into predator-prey models with the fear effect [16,17,18,22,23,24,25]. Since the biomass transfer is not instantaneous after the predation of prey, the biomass transfer delay is considered in [16]. A generation time delay in prey has been considered in [23]. In fact, the reproduction of prey will not respond immediately to fear, but will rather reduce after a time lag. Such a fear response delay has been considered in [18,24,25].

    Motivated by [16] and the previous work, we propose the following model

    {tu(x,t)=d1Δu(x,t)+r0u(x,t)1+c2kv(x,tτ)du(x,t)au2(x,t)pu(x,t)v(x,t)1+qu(x,t)11+c1k,x(0,lπ),t>0,tv(x,t)=d2Δv(x,t)+pu(x,t)v(x,t)1+qu(x,t)c1+c1kmv(x,t),x(0,lπ),t>0,ux(0,t)=ux(lπ,t)=vx(0,t)=vx(lπ,t)=0,t>0,u(x,t)=u0(x,t)0,v(x,t)=v0(x,t)0,x(0,lπ),t[τ,0], (1.3)

    where d1,d2>0 represent the diffusion coefficients of prey and predator, respectively, and τ is the fear response delay in prey reproduction.

    Although there have been some literature on a predator-prey model with fear effect and delay, no work has been done to explore the joint impact of fear level and fear response delay on the population dynamics. In this paper, we aim to reveal how these two parameters k and τ jointly affect the dynamics of system (1.3) from the view of bifurcation analysis. The characteristic equation may have no, one or two pairs of purely imaginary roots under different conditions. Correspondingly, the fear response delay may have no effect, both stabilizing and destabilizing effect, or destabilizing effect on the stability of the positive equilibrium. By the discussion of the monotonicity of the critical values of Hopf bifurcation, we get the order of all the Hopf bifurcation values. We prove the existence of stability switches for system (1.3) under suitable condition. Through numerical simulations, we find that the level of fear k has a crucial role in the stability of positive equilibrium and the occurrence of Hopf bifurcation induced by fear response delay. To reveal the complex phenomena induced by k and τ, double Hopf bifurcation analysis is carried out, which can induce complex spatio-temporal dynamics [26,27]. It provides a qualitative classification of dynamical behaviors on the (k,τ) plane, which can help us to explicitly observe the different dynamics corresponding to different values of k and τ, including quasi-periodic oscillations on two or three dimensional torus, and even chaos.

    The rest of this paper is organized as follows. In Section 2, we analyze the local stability of equilibria and obtain the condition of Hopf bifurcation in three different cases. In Section 3, we give the condition for double Hopf bifurcation and derive the normal form of the double Hopf bifurcation. In Section 4, numerical simulations are presented to verify our theoretical results. Finally, conclusions and discussions are given in Section 5.

    In this section, we discuss the local stability of equilibria. System (1.3) has three possible equilibria. The trivial equilibrium E0(0,0) always exists. When r0d>0, system (1.3) has a predator-free equilibrium Eu(r0da,0). Moreover, if

    (H1)0<m(1+c1k)pcmq(1+c1k)<r0da

    holds, there is a unique positive equilibrium E(u,v), where

    u=m(1+c1k)pcmq(1+c1k),v=a1+a214a0a22a0, (2.1)

    with

    {a0=c2pk,a1=p+(d+au)(1+qu)(1+c1k)c1k,a2=(d+aur0)(1+qu)(1+c1k). (2.2)

    It is well known that the laplacian operator Δ has eigenvalues n2/l2(n=0,1,2,).

    Now, we study the local stability for each of equilibria. The linearization of system (1.3) at an equilibrium (ˉu,ˉv) can be written as

    t(u(x,t)v(x,t))=D(Δu(x,t)Δv(x,t))+G0(u(x,t)v(x,t))+G1(u(x,tτ)v(x,tτ)), (2.3)

    where

    D=(d100d2),G0=(J11J12J21J22),G1=(0K1200),

    with

    J11=r01+c2kˉvd2aˉupˉv(1+c1k)(1+qˉu)2,J12=pˉu(1+c1k)(1+qˉu),J21=cpˉv(1+c1k)(1+qˉu)2,J22=cpˉu(1+qˉu)(1+c1k)m,K12=r0c2kˉu(1+c2kˉv)2.

    The characteristic equation is given by

    det(λ+d1n2l2J11 J12K12eλτJ21 λ+d2n2l2J22)=0,n=0,1,2,. (2.4)

    From (2.4), we get the corresponding characteristic equation at E0(0,0) as

    (λ+d1n2l2r0+d)(λ+d2n2l2+m)=0,n=0,1,2,.

    Thus, we have

    λ1,n=d1n2l2+r0d,λ2,n=d2n2l2m,n=0,1,2,.

    Obviously, λ2,n<0. If r0<d, λ1,n<0 for all nN0, and if r0>d, λ1,0>0. Thus, if r0<d, E0 is locally asymptotically stable. If r0>d, E0 is unstable.

    If r0>d, the predator-free equilibrium Eu(r0da,0) exists. The corresponding characteristic equation is given by

    (λ+d1n2l2+r0d)(λ+d2n2l2+mcp(r0d)(1+c1k)[a+q(r0d)])=0,n=0,1,2,.

    Since r0d>0, λ1,n=d1n2l2(r0d)<0. Now we consider the sign of λ2,n=[d2n2l2+mcp(r0d)(1+c1k)[a+q(r0d)]]. If (H1) holds, λ2,0=m+cp(r0d)(1+c1k)[a+q(r0d)]>0, which indicates that Eu is unstable. If m(1+c1k)pcmq(1+c1k)>r0da>0 holds, λ2,n<0 for all nN0, and Eu is locally asymptotically stable.

    Theorem 1. (i) When r0<d, there is only the trivial equilibrium E0 for system (1.3), which is locally asymptotically stable; when r0>d, E0 is unstable and there is a predator-free equilibrium Eu.

    (ii) When m(1+c1k)pcmq(1+c1k)>r0da>0, Eu is locally asymptotically stable; When (H1) holds, Eu is unstable and there is a positive equilibrium E.

    For the positive equilibrium E(u,v) of system (1.3), we have

    J11=au+pquv(1+c1k)(1+qu)2,J12=pu(1+c1k)(1+qu)<0,J21=cpv(1+c1k)(1+qu)2>0,J22=0,K12=r0c2ku(1+c2kv)2<0. (2.5)

    The characteristic equation becomes

    λ2+Tnλ+Dn+Meλτ=0,n=0,1,2,, (2.6)

    where

    Tn=(d1+d2)n2l2J11,Dn=d1d2n4l4J11d2n2l2J21J12,M=J21K12>0.

    When τ=0, the characteristic equation (2.6) becomes

    λ2+Tnλ+Dn+M=0,n=0,1,2,. (2.7)

    Obviously, D0+M=J21J12J21K12>0. Hence, if

    (H2)J11=au+pquv(1+c1k)(1+qu)2<0

    holds, we have  Tn>0, Dn+M>0 for all nN0, thus all roots of Eq (2.7) have negative real parts. If J11>0, we have T0<0 and D0+M>0, the roots of (2.7) have positive real parts when n=0, which means that E is unstable in the absence of diffusion. To sum up, Turing instability induced by diffusion will not occur.

    In the following, we assume that (H2) always holds, which ensures  Tn>0, Dn>0 and Dn+M>0. Thus, a change of stability at E(u,v) can only happen when there is at least one root of Eq (2.6) across the imaginary axis on the complex plane. After excluding Turing instability induced by diffusion, now we seek the critical values of τ where the roots of Eq (2.6) will cross the imaginary axis from the left half plane to the right half plane. Plugging λ=iωn(ωn>0) into Eq (2.6), we obtain

    ω2n+Tniωn+Dn+M(cosωnτisinωnτ)=0.

    Separating the real and imaginary parts, we obtain

    {cos(ωnτ)=ω2nDnM=Cn(ωn),sin(ωnτ)=TnωnM=Sn(ωn). (2.8)

    Squaring and adding both equations of (2.8), we have

    ω4n+(T2n2Dn)ω2n+D2nM2=0. (2.9)

    The number of positive roots of Eq (2.9) is relevant to the signs of T2n2Dn and D2nM2. For the convenience of discussion, we make the following assumptions.

    (H3)T202D00 and D0M0.(H4)T202D0<0 and D0M>0.(H5)T202D0<0 and D0M=0.(H6)T202D00 and D0M<0.(H7)T202D0<0 and D0M<0.

    Before the discussion of these cases, we need to investigate some properties of T2n2Dn and D2nM2.

    Lemma 1. Suppose that (H1) and (H2) hold. T2n2Dn and D2nM2 are both monotonically increasing with respect to n.

    Proof. When (H2) holds, we have

    d(T2n2Dn)dn=2(TndTndndDndn)=2(d212n3l4+d222n3l4J11d12nl2)>0,

    which means that T2n2Dn is monotonically increasing with respect to n. Similarly, when J11<0, we have Dn>0, and

    d(D2nM2)dn=2DndDndn=2Dn(d1d24n3l4J11d22nl2)>0.

    We first consider the case when (H3) holds.

    Lemma 2. If (H1), (H2) and (H3) hold, then all roots of Eq (2.6) have negative real parts for all τ0.

    Proof. Noticing that Dn+M>0 for all nN0, thus D0M0 leads to D20M20. From Lemma 1 and (H3), we can deduce that T2n2Dn0 and D2nM20 for all nN0. Therefore, Eq (2.9) has no positive roots, and Eq (2.6) has no imaginary roots.

    Theorem 2. If (H1), (H2) and (H3) hold, the positive equilibrium E is locally asymptotically stable for all τ0.

    Lemma 3. If (H1), (H2) and (H4) hold, there exists n1N0 such that the following conclusions hold.

    (i) If either n>n1 or 0nn1 and Δ=(T2n2Dn)24(D2nM2)<0, then all roots of Eq (2.6) have negative real parts for all τ0.

    (ii) If Δ>0 for 0nn1, then Eq (2.6) has a pair of imaginary roots ±iω+n(±iωn, respectively) for 0nn1.

    Proof. Similar as the proof in Lemma 2, we have D2nM2>0 for all nN0. When T202D0<0, from Lemma 1, there exists n1N0 such that

    {T2n2Dn0,n>n1,T2n2Dn<0,0nn1. (2.10)

    It means that Eq (2.9) has no positive roots for n>n1. For 0nn1, if Δ=(T2n2Dn)24(D2nM2)<0, Eq (2.9) has no positive roots, and if Δ>0, Eq (2.9) has two positive roots ω±n, where

    ω±n=[12(2DnT2n±Δ)]12. (2.11)

    Correspondingly, Eq (2.6) has a pair of purely imaginary roots ±iω+n(±iωn,respectively) for 0nn1.

    Since (H2) holds, we have Tn>0. Noticing that M=J21K12>0, from (2.8), we have Sn(ω±n)>0, and we get

    τj±n=1ω±n(arccos[(ω±n)2DnM]+2jπ),0nn1,j=0,1,2,. (2.12)

    Lemma 4. Assume that (H1), (H2) and (H4) hold. If Δ>0, then Re[dλdτ]τ=τj+n>0, Re[dλdτ]τ=τjn<0, for 0nn1, j=0,1,2,, where n1 is defined in (2.10).

    Proof. Differentiating two sides of (2.6) with respect to τ, we obtain

    Re[dλdτ]1τ=τj±n=2ω±ncos(ω±nτ)+Tnsin(ω±nτ)Mω±n=2ω±n(ω±2nDn)+T2nω±nM2ω±n=±ΔM2.

    The sign of Re[dλdτ]τ=τj±n is the same as that of Re[dλdτ]1τ=τj±n, thus we have Re[dλdτ]τ=τj+n>0, Re[dλdτ]τ=τjn<0.

    Now we consider the order of the critical values Hopf bifurcation.

    Lemma 5. Assume that (H1), (H2) and (H4) hold, and Δ>0 for 0nn1. τj+n is monotonically increasing and τjn is monotonically decreasing with respect to n (0<nn1), where n1 and τj±n are defined in (2.10) and (2.12), respectively.

    Proof. See Appendix A.

    Theorem 3. Assume that (H1), (H2) and (H4) hold, and n1 is defined in (2.10).

    (i) If Δ<0 for 0nn1, E is locally asymptotically stable for all τ>0.

    (ii) If Δ>0 for 0nn1, system (1.3) undergoes a Hopf bifurcation at E when τ=τj±n(0nn1,j=0,1,2,). Moreover, E is locally asymptotically stable when τ[0,τ0+0)(τ00,τ1+0)(τ(s1)0,τs+0), and it is unstable when τ(τ0+0,τ00)(τ1+0,τ10)(τs+0,+).

    Proof. (ⅰ) From Lemma 3, when Δ<0, Eq (2.6) has no purely imaginary roots, thus the stability of E can not be changed by delay.

    (ⅱ) When Δ>0, from Lemma 5, we have τj+0<τj+1<<τj+n1, and τjn1<<τj0,j=0,1,2,. From ω+n1>ωn1, we have arccos[(ω+n1)2Dn1M]<arccos[(ωn1)2Dn1M], and τj+n1<τjn1. Therefore, τj+0<<τj+n1<τjn1<<τj0.

    We claim that there exists a positive integer s such that τ0+0<τ00<τ1+0<τ10<<τs+0<τ(s+1)+0<τs0. Since ω0<ω+0, τj+0τ(j1)+0=2πω+0<τj0τ(j1)0=2πω0, which means that the alternation for τj+0 and τj0 cannot persist for the entire sequence, and there exists a positive integer s such that τ(s1)+0<τ(s1)0<τs+0<τ(s+1)+0<τs0. From Lemma 4, we can get the stable and unstable interval for τ.

    First, we consider the case when (H5) holds. If (H5) holds, Eq (2.9) has a positive root ω+0=2D0T20 when n=0. Similar as Lemma 3, there exists n1N0 such that (2.10) holds. If Δ>0 for 0<nn1, Eq (2.6) has a pair of imaginary roots ±iω+n(±iωn,respectively) for 0<nn1.

    Lemma 6. If (H1), (H2) and (H5) hold, there exists n1N0 defined in (2.10).

    (i) If either n>n1 or 0<nn1 and Δ<0, then all roots of Eq (2.6) have negative real parts for all τ0.

    (ii) If n=0, then Eq (2.6) has a pair of imaginary roots ±iω+0.

    (iii) If Δ>0 for 0<nn1, then Eq (2.6) has a pair of imaginary roots ±iω+n(±iωn, respectively) for 0<nn1.

    Remark 1. Similar as the proof of Lemma 5 and Theorem 3, we can get τj+0<τj+1<<τj+n1<τjn1<<τj1, where τj±n is defined in (2.12). Similar as Lemma 4, we can verify that Re[dλdτ]τ=τj+n>0 for 0nn1 and Re[dλdτ]τ=τjn<0 for 0<nn1. Being different from case II, stability switches can not happen.

    Now we consider the case when (H6) holds. From Lemma 1, and T202D00, we have T2n2Dn0 for all nN0. From Lemma 1, combining D0M<0 with Dn+M>0, we get

    D2nM2{0,n>n2,<0,0nn2. (2.13)

    Equation (2.9) has only one positive root ω+n for 0nn2, where

    ω+n=[12(2DnT2n+Δ)]12.

    Lemma 7. Assume that (H1), (H2) and (H6) hold, and n2N0 is defined in (2.13).

    (i) If n>n2, then all roots of Eq (2.6) have negative real parts for all τ0.

    (ii) If 0nn2, then Eq (2.6) has a pair of imaginary roots ±iω+n for 0nn2.

    It is easy to get Sn(ω+n)>0, from (2.8), we have

    τj+n=1ω+n(arccos[(ω+n)2DnM]+2jπ),0nn2,j=0,1,2,. (2.14)

    When (H7) holds, from Lemma 1, we can deduce that there exists n1 and n2 such that

    T2n2Dn{0,n>n1,<0,0nn1,D2nM2{0,n>n2,<0,0nn2. (2.15)

    If n1n2, Eq (2.9) has only one positive root ω+n for 0nn2, where

    τj+n=1ω+n(arccosCn(ω+n)+2jπ),0nn2,j=0,1,2,,

    and Eq (2.9) has no positive roots for n>n2.

    If n2n1, Eq (2.9) has one positive root ω+n for 0nn2, and has two positive roots ω±n for n2+1<nn1, and has no positive roots for n>n1. When n=n2+1, if D2n2+1M2>0, Eq (2.9) has two positive roots ω±n2+1, and if D2n2+1M2=0, Eq (2.9) has one positive root ω+n2+1. Similar as the proof in Lemma 5, we have

    {τj+0<<τj+n2<τj+n2+1<<τj+n1<τjn1<<τjn2+2<τjn2+1,if D2n2+1M2>0,τj+0<<τj+n2<τj+n2+1<<τj+n1<τjn1<<τjn2+2,if D2n2+1M2=0,

    where

    {τj+n=1ω+n(arccosCn(ω+n)+2jπ),0nn2,j=0,1,2,,τj±n=1ω±n(arccosCn(ω±n)+2jπ),n2+1<nn1,j=0,1,2,,τj+n2+1={1ω±n2+1(arccosCn2+1(ω±n2+1)+2jπ),ifD2n2+1M2>0,j=0,1,2,,1ω+n2+1(arccosCn2+1(ω+n2+1)+2jπ),ifD2n2+1M2=0,j=0,1,2,.

    For either n1n2 or n2n1, we have

    τ0+0=min{τj±n or τj+n0nmax{n1,n2},j=0,1,2,}. (2.16)

    From the previous discussion, we can get the following conclusions.

    Theorem 4. Assume that (H1) and (H2) hold, and n1, n2 are defined in (2.15). If (H5), (H6) or (H7) holds, E is locally asymptotically stable when τ[0,τ0+0), and E is unstable when τ>τ0+0.

    Proof. If (H5), (H6) or (H7) holds, we easily know that τ0+0 is the smallest critical value of Hopf bifurcation values. Similar as Lemma 4, we can get the transversality condition, which can lead to the results.

    In order to figure out the joint effect of fear level k and fear response delay τ on the dynamical behavior of system (1.3), we carry out double Hopf bifurcation analysis. We first verify the condition for the occurrence of double Hopf bifurcation.

    Remark 2. Assume that (H1), (H2) and (H4) hold. If there exists k=k such that τj1+n1=τj2n2=τ. Then system (1.3) may undergo a double Hopf bifurcation at the positive equilibrium E(u,v) when (k,τ)=(k,τ).

    Define the real-valued Hilbert space X:={(u,v)H2(0,lπ)×H2(0,lπ):(ux,vx)|x=0,lπ=0}, and the corresponding complex space XC:={x1+ix2,x1,x2X}. Let ˜u=u(x,τt)u, ˜v=v(x,τt)v, and drop the hats for the convenience of notation. Denote U(t)=(u(x,t),v(x,t))T, and Ut(θ)=U(t+θ),1θ0. System (1.3) can be written as

    dUdt=τDΔU(t)+τ[G0(k)Ut(0)+G1(k)Ut(1)]+τ[12!(Fuuu2t(0)+2Fuvut(0)vt(0)+Fvtvtv2t(1)+2Fuvtut(0)vt(1))+13!(Fuuuu3t(0)+3Fuuvu2t(0)vt(0)+3Fuvtvtut(0)v2t(1)+Fvtvtvtv3t(1))], (3.1)

    where the coefficients of nonlinear terms Fuu,Fuv, etc., are in Appendix B, and

    G0(k)=(J11(k)J12(k)J21(k)0),G1(k)=(0K12(k)00).

    Here J11(k),J12(k),J21(k), and K12(k) are defined in (2.5), in which u,v are functions of k defined in (2.1).

    Let k=k+α1, τ=τ+α2, then (α1,α2)=(0,0) is a double Hopf bifurcation point. System (3.1) can be written as

    dUdt=D0ΔU(t)+L0Ut+˜F(α1,α2,Ut), (3.2)

    where

    D0=τD,L0Ut=τ(G0(k)Ut(0)+G1(k)Ut(1)),˜F(α1,α2,Ut)=12!F2(α1,α2,Ut)+13!F3(α1,α2,Ut)+,

    with

    F2(α1,α2,Ut)=2((α1D(1,0)1+α2D(0,1)1)ΔU+α1L(1,0)1Ut+α2L(0,1)1Ut)+2Fuvut(0)vt(0)+Fuuu2t(0)+2Fuvtut(0)vt(1)+Fvtvtv2t(1),F3(0,0,Ut)=Fvtvtvtv3t(1)+3Fuuvu2t(0)vt(0)+3Fuvtvtut(0)v2t(1)+Fuuuu3t(0),

    and

    D(1,0)1=0,D(0,1)1=D,L(0,1)1Ut=G0(k)Ut(0)+G1(k)Ut(1),L(1,0)1Ut=τ(J11(k)J12(k)J21(k)0)Ut(0)+τ(0K12(k)00)Ut(1).

    We leave the detailed calculation and some expressions in Appendix C.

    Define an enlarged phase space BC by

    BC:={ψ:[1,0]XC:ψ is continuous on[1,0),limθ0ψ(θ)XC}.

    Equation (3.2) can be written as an abstract ordinary differential equation in BC

    dUtdt=AUt+X0˜F(α1,α2,Ut), (3.3)

    where Aφ=˙φ+X0[D0Δφ(0)+L0(φ)˙φ(0)], with

    X0(θ)={0,1θ<0,I2,θ=0.

    It is obvious that {±iω+n1τ,±iωn2τ} are the pure imaginary eigenvalues of the operator A and the corresponding eigenfunctions in BC are {ϕ1(θ)γn1,ϕ2(θ)γn1,ϕ3(θ)γn2,ϕ4(θ)γn2}, respectively, where

    γn(x)=cosnlxcosnlxL2={1lπ,n=0,2lπcosnlx,n1,

    and ϕ1(θ)=(1,p12)Teiω+n1τθ, ϕ3(θ)=(1,p32)Teiωn2τθ, ϕ2(θ)=¯ϕ1(θ), ϕ4(θ)=¯ϕ3(θ), with

    p12=J11(k)+n21l2d1+iω+n1J12(k)+K12(k)eiω+n1τ,p32=J11(k)+n22l2d1+iωn2J12(k)+K12(k)eiωn2τ.

    A direct calculation derived that ψ1(s)=D1(1,q12)eiω+n1τs, ψ3(s)=D3(1,q32)eiωn2τs are the formal adjoint eigenvectors of ϕ1(θ), ϕ3(θ) and satisfy (ϕ1,ψ1)1=1, (ϕ3,ψ3)2=1. Here

    q12=J11(k)+n21l2d1+iω+n1J21(k),q32=J11(k)+n22l2d1+iωn2J21(k),D1=(1+p12q12+τK12(k)p12eiω+n1τ)1,D3=(1+p32q32+τK12(k)p32eiωn2τ)1,

    and (,)k are the adjoint bilinear form

    (ψ,ϕ)k=ψ(0)ϕ(0)01θξ=0ψ(ξθ)dηk(θ,τ)ϕ(ξ)dξ,k=1,2,

    with a function of bounded variation ηk(θ,τ),θ[1,0] are given by

    D0n2kl2ϕ(0)+L0(ϕ)=01dηk(θ,τ)ϕ(θ),ϕC,k=1,2.

    Denote

    Φ1(θ)=(ϕ1(θ),ϕ2(θ)),Φ2(θ)=(ϕ3(θ),ϕ4(θ)),Ψ1(s)=(ψ1(s),ψ2(s))T, Ψ2(s)=(ψ3(s),ψ4(s))T,

    where ψ2(s)=¯ψ1(s), ψ4(s)=¯ψ3(s).

    Now we decompose BC into the direct sum of the central subspace P and its complementary space Kerπ

    BC=PKerπ, (3.4)

    where π:BCP is the projection defined by

    π(φ)=2k=1Φk(Ψk,φ(),βnk)kβnk.

    According to (3.4), Ut can be decomposed as

    Ut(θ)=ϕ1(θ)z1(t)γn1+ϕ2(θ)z2(t)γn1+ϕ3(θ)z3(t)γn2+ϕ4(θ)z4(t)γn2+yt(θ),

    where (z1,z2,z3,z4)T=(˜z1,˜z2)=(Ψk,Ut,βnk)k, and yt(θ)Kerπ.

    Then system (3.3) in BC is equivalent to the system

    {˙z1=iω+n1τz1+ψ1(0)12!˜F2(z,y,α)+13!˜F3(z,y,α),βn1+,˙z2=iω+n1τz2+ψ2(0)12!˜F2(z,y,α)+13!˜F3(z,y,α),βn1+,˙z3=iωn2τz3+ψ3(0)12!˜F2(z,y,α)+13!˜F3(z,y,α),βn2+,˙z4=iωn2τz4+ψ4(0)12!˜F2(z,y,α)+13!˜F3(z,y,α),βn2+,dydt=A1y+(Iπ)X0[12!˜F2(z,y,α)+13!˜F3(z,y,α)],

    where A1 is the restriction of A on Q1KerπKerπ, A1φ=Aφ for φQ1, and ˜F2(z,y,α) and ˜F3(z,0,0) are in Appendix D.

    According to [28], we can obtain the normal form of the double Hopf bifurcation up to the third order as follows

    ˙z1=iω+n1τz1+B11α1z1+B21α2z1+B2100z21z2+B1011z1z3z4,˙z2=iω+n1τz2+¯B11α1z2+¯B21α2z2+¯B2100z1z22+¯B1011z2z3z4,˙z3=iωn2τz3+B13α1z3+B23α2z3+B0021z23z4+B1110z1z2z3,˙z4=iωn2τz4+¯B13α1z4+¯B23α2z4+¯B0021z3z24+¯B1110z1z2z4. (3.5)

    Here

    B11=ψ1(0)(n21l2D(1,0)1ϕ1(0)+L(1,0)1ϕ1), B21=ψ1(0)(n21l2D(0,1)1ϕ1(0)+L(0,1)1ϕ1),B13=ψ3(0)(n22l2D(1,0)1ϕ3(0)+L(1,0)1ϕ3), B23=ψ3(0)(n22l2D(0,1)1ϕ3(0)+L(0,1)1ϕ3),B2100=C2100+32(D2100+E2100), B1011=C1011+32(D1011+E1011),B0021=C0021+32(D0021+E0021),  B1110=C1110+32(D1110+E1110),

    where

    C2100=16lπψ1(0)F2100,C1011=16lπψ1(0)F1011,C0021=16lπψ3(0)F0021,C1110=16lπψ3(0)F1110,
    D2100=16τ(2iω+n1f1(1)2000f1(1)1100+1iω+n1f1(1)1100f1(1)2000+1iω+n1f1(1)1100f1(2)1100+213iω+n1f1(1)0200f1(2)2000+1iωn2f1(1)1010f1(3)1100+12iω+n1iωn2f1(1)0110f1(3)2000+1iωn2f1(1)1001f1(4)1100+12iω+n1+iωn2f1(1)0101f1(4)2000),D1011=16τ(2iω+n1f1(1)2000f1(1)0011+1iωn2f1(1)1010f1(1)1001+1iωn2f1(1)1001f1(1)1010+1iω+n1f1(1)1100f1(2)0011+12iω+n1iωn2f1(1)0110f1(2)1001+12iω+n1+iωn2f1(1)0101f1(2)1010+1iωn2f1(1)1010f1(3)0011+2iω+n12iωn2f1(1)0020f1(3)1001+1iω+n1f1(1)0011f1(3)1010+1iωn2f1(1)1001f1(4)0011+1iω+n1f1(1)0011f1(4)1001+2iω+n1+2iωn2f1(1)0002f1(4)1010),D0021=16τ(2iω+n1f1(3)1010f1(1)0011+12iωn2iω+n1f1(3)1001f1(1)0020+1iω+n1f1(3)0110f1(2)0011+12iωn2+iω+n1f1(3)0101f1(2)0020+21iωn2f1(3)0020f1(3)0011+1iωn2f1(3)0011f1(3)0020+1iωn2f1(3)0011f1(4)0011+213iωn2f1(3)0002f1(4)0020),
    D1110=16τ(22iω+n1+iωn2f1(3)2000f1(1)0110+1iωn2f1(3)1100f1(1)1010+1iω+n1f1(3)1010f1(1)1100+1iωn2f1(3)1100f1(2)0110+212iω+n1+iωn2f1(3)0200f1(2)1010+1iω+n1f1(3)0110f1(2)1100+1iω+n1f1(3)1010f1(3)0110+1iω+n1f1(3)0110f1(3)1010+2iωn2f1(3)0020f1(3)1100+1iω+n1+2iωn2f1(3)1001f1(4)0110+1iω+n1+2iωn2f1(3)0101f1(4)1010+1iωn2f1(3)0011f1(4)1100),E2100=16lπψ1(0)[Syz1(w01100)+Syz2(w02000)],E1011=16lπψ1(0)[Syz1(w00011)+Syz3(w01001)+Syz4(w01010)],E0021=16lπψ3(0)[Syz3(w00011)+Syz4(w00020)],E1110=16lπψ3(0)[Syz1(w00110)+Syz2(w01010)+Syz3(w01100)],

    with

    w02000(θ)=1iω+n1τϕ1(θ)f1(1)2000+13iω+n1τϕ2(θ)f1(2)2000+12iω+n1τ+iωn2τϕ3(θ)f1(3)2000+12iω+n1τiωn2τϕ4(θ)f1(4)20001lπe2iω+n1τθ[2iω+n1τ+L0(e2iω+n1τI)]1F2000,w01100(θ)=1iω+n1τϕ1(θ)f1(1)1100+1iω+n1τϕ2(θ)f1(2)1100+1iωn2τϕ3(θ)f1(3)1100+1iωn2τϕ4(θ)f1(4)11001lπ[L0(I)]1F1100,w00011(θ)=1iω+n1τϕ1(θ)f1(1)0011+1iω+n1τϕ2(θ)f1(2)0011+1iωn2τϕ3(θ)f1(3)0011+1iωn2τϕ4(θ)f1(4)00111lπ[L0(I)]1F0011,w01001(θ)=1iωn2τϕ1(θ)f1(1)1001+12iω+n1τ+iωn2τϕ2(θ)f1(2)1001+1iω+n1τ+2iωn2τϕ3(θ)f1(3)1001+1iω+n1τϕ4(θ)f1(4)10011lπe(iω+n1iωn2)τθ[(iω+n1iωn2)τ+L0(e(iω+n1iωn2)τI)]1F1001,w01010(θ)=1iωn2τϕ1(θ)f1(1)1010+12iω+n1τiωn2τϕ2(θ)f1(2)1010+1iω+n1τϕ3(θ)f1(3)1010+1iω+n1τ2iωn2τϕ4(θ)f1(4)1010
    1lπe(iω+n1+iωn2)τθ[(iω+n1+iωn2)τ+L0(e(iω+n1+iωn2)τI)]1F1010,w00110(θ)=12iω+n1τiωn2τϕ1(θ)f1(1)0110+1iωn2τϕ2(θ)f1(2)0110+1iω+n1τϕ3(θ)f1(3)0110+1iω+n1τ2iωn2τϕ4(θ)f1(4)01101lπe(iω+n1+iωn2)τθ[(iω+n1+iωn2)τ+L0(e(iω+n1+iωn2)τI)]1F0110,w00020(θ)=1iω+n1τ2iωn2τϕ1(θ)f1(1)0020+1iω+n1τ2iωn2τϕ2(θ)f1(2)0020+1iωn2τϕ3(θ)f1(3)0020+13iωn2τϕ4(θ)f1(4)00201lπe2iωn2τθ[2iωn2τ+L0(e2iωn2τI)]1F0020.

    In this section, we carry out some numerical simulations to explain the theoretical results obtained in the previous sections. Symbolic mathematical software Matlab is used to plot numerical graphs. The delayed reaction-diffusion system (1.3) is numerically solved by transforming the continuous system to discrete system using discretization of time and space. In the discrete system, the Laplacian describing diffusion is calculated using finite differences, and the time evolution is solved using the Euler method. Fix the parameters as

    r0=0.5,d=0.1,a=0.2,c=0.9,m=0.05,c1=1,c2=1,p=0.9,q=0.4,d1=0.3,d2=0.6,l=8. (4.1)

    From Theorems 2, 3, 4, we have found that the fear response delay may have no effect, both stabilizing and destabilizing effect, or destabilizing effect on the stability of E, respectively. In this section, we explore that the impact level of fear k has on the stability of the positive equilibrium E and the occurrence of Hopf bifurcation induced by fear response delay.

    Choose k=0.04, such that the level of fear is low. We can verify that (H1) holds, thus E(0.0659,0.4483) exists. Moreover, (H2) and (H3) hold, which indicates that E is locally asymptotically stable for all τ>0 from Theorem 2. This indicates that the fear response delay does not change the stability of the positive equilibrium, which means that the fear response delay cannot support periodic oscillations in prey and predator populations with a low level of fear.

    When k is increased to k=1, the positive equilibrium E(0.1299,0.49) still exists, and (H1), (H2) and (H4) hold. By calculating the critical value of Hopf bifurcation in Table 1, we can get τ0+0<τ0+1<τ01<τ00<τ1+0<τ1+1<τ2+0<τ2+1. From Theorem 3, E is locally asymptotically stable when τ[0,τ0+0)(τ00,τ1+0) (see Figures 1 and 3), and it is unstable when τ(τ0+0,τ00)(τ1+0,+) (see Figures 2 and 4). It means that the fear response delay may lead to stability switches with an intermediate level of fear, which is not observed in the model with biomass transfer delay in [16].

    Table 1.  Critical values of Hopf bifurcation τj±n.
    j=0 j=1 j=2
    n=0 τ0+0=3.048 τ1+0=54.609 τ2+0=106.169
    n=0 τ00=42.19 τ10=132.59 τ20=222.997
    n=1 τ0+1=6.203 τ1+1=59.224 τ2+1=112.244
    n=1 τ01=36.98 τ11=122.64 τ21=208.296

     | Show Table
    DownLoad: CSV
    Figure 1.  The positive equilibrium E is asymptotically stable when τ=1[0,τ0+0).
    Figure 2.  E is unstable and there is a bifurcating periodic solution when τ=15(τ0+0,τ00).
    Figure 3.  The positive equilibrium E is asymptotically stable when τ=52(τ00,τ1+0).
    Figure 4.  E is unstable and there is a bifurcating periodic solution when τ=70(τ1+0,+).

    When k is further increased to k=2.5, we can verify that (H1), (H2) and (H7) hold, and we get τ0+0=7.3868. From Theorem 4, E is locally asymptotically stable when τ[0,τ0+0), and it is unstable when τ>τ0+0.

    In this section, we explore the joint effect of the fear level and fear response delay on the dynamics of system (1.3). We choose k and τ as bifurcation parameters, and other parameters are the same as in (4.1). From (2.12), we can draw Hopf bifurcation curves with varying k (see Figure 5). Hopf bifurcation curve τ1+0 intersects with τ00 at HH (k,τ)=(1.3677,58.6677). On HH, Eq (2.6) has two pairs of purely imaginary roots ±iω0=±0.0503 i, ±iω+0=±0.1146 i. Therefore, HH is a possible double Hopf bifurcation point. By the calculation of normal form of double Hopf bifurcation, we can get the coefficients in (3.5) with

    B110.66820.4239 i,    B210.0108+0.0082 i,B130.44820.2903 i,  B230.03817+0.0424 i,B21000.33852.1955 i,  B10110.5068+0.6607 i,B00210.73640.774 i, B11104.2835+2.426 i.
    Figure 5.  The bifurcation diagram on (k,τ) plane.

    Make the polar coordinate transformation

    z1=r1cosθ1+ir1sinθ1,z2=r1cosθ1ir1sinθ1,z3=r2cosθ2+ir2sinθ2,z4=r2cosθ2ir2sinθ2,

    and denote

    ϵ1=Sign(ReB2100),ϵ2=Sign(ReB0021),ˉr1=r1|B2100|,ˉr2=r2|B0021|,ˉt=tε1.

    Removing the bars, the four-dimensional system (3.5) is transformed into the following two-dimensional amplitude system

    ˙r1=r1(0.6682 α10.0108 α2+r21+0.6885 r22),˙r2=r2(0.4482 α1+0.0382 α212.6574 r21r22). (4.2)

    By a simple calculation, we find that the system (4.2) admits the following equilibria

    A0=(0,0),A1=(0.0108α20.6682α1,0), for α1<0.0162α2,A2=(0,0.0382α20.4482α1), for α1<0.0852α2,A3=(0.0466α1+0.002α2,1.0382α1+0.0128α2), for 0.0429α2<α1<0.0123α2.

    According to [29], the unfolding of system (4.2) is of type VIa. There are rays L1L8 near the bifurcation point, which divide the phase space into eight parts (see Figure 6(a)), where

    L1:τ=τ+11.7406 (kk)(k>k);L2:τ=τ+61.7604 (kk)(k>k);L3:τ=τ+81.0926 (kk)(k>k);L4:τ=τ+100.462 (kk)+o(kk)(k>k);L5:τ=τ+100.462 (kk)(k>k);L6:τ=τ23.3449 (kk)(k<k);L7:τ=τ+11.7406 (kk)(k<k);L8:τ=τ+61.7604 (kk)(k<k).
    Figure 6.  (a) The bifurcation diagram on (k,τ) plane. (b) The complete bifurcation set around HH.

    The detailed dynamics in D1D8 near (k,τ) have shown in Figure 6(b).

    Notice that A0 of (4.2) corresponds to the positive equilibrium E of the original system (1.3). A1 and A2 correspond to the periodic solution of (1.3). A3 corresponds to quasi-periodic solution on a 2-torus. Periodic orbit of (4.2) corresponds to quasi-periodic solution on a 3-torus. Due to the fact that double Hopf bifurcation point HH is the intersection of 0-mode Hopf bifurcation curves τ1+0 and τ00, all periodic or quasi-periodic solutions near HH are spatially homogeneous.

    In region D1, system (4.2) has only one equilibrium A0, which is a saddle. It means that the positive equilibrium E(u,v) of system (1.3) is unstable.

    In region D2, the trivial equilibrium A0 of (4.2) is a sink, and A1 is unstable. This means that E(u,v) of system (1.3) is asymptotically stable as shown in Figure 7, when k=0.9 and τ=50 are chosen in D2. In addition, the spatially homogeneous periodic solution is born, which is unstable.

    Figure 7.  When k=0.9,τ=50 in D2, the positive equilibrium E(u,v) is asymptotically stable.

    In region D3, system (4.2) has three equilibria: A0, A1 and A2. A2 is stable while other equilibria are unstable. When k=0.8 and τ=55 are chosen in D3, system (1.3) has spatially homogeneous periodic solution, which is stable (see Figure 8).

    Figure 8.  When k=0.8,τ=55 in D3, the periodic solutions are stable.

    In region D4, there are four equilibria of (4.2): A0, A1, A2 and A3. A3 is stable while other equilibria are unstable. Since the double Hopf bifurcation occurs at the intersection of two Hopf bifurcation curves with n=0, we choose x=0 to demonstrate the rich dynamical phenomena of system (1.3). When parameters are chosen as k=1.345 and τ=59.5 in D4, system (1.3) has a quasi-periodic solution on 2-torus, which is shown in Figure 9.

    Figure 9.  There is a stable quasi-periodic solution on a 2-torus of system (1.3) when (k,τ)=(1.345,59.5)D4.

    In region D5, there are four equilibria and a periodic orbit of (4.2). A0, A1, A2 and A3 are all unstable, and the periodic orbit is stable. This means that system (1.3) has a stable quasi-periodic solution on a 3-torus. Figure 10 illustrates the existence of a stable quasi-periodic solution on a 3-torus when k=1.364 and τ=59.4 are chosen in D5.

    Figure 10.  There is a stable quasi-periodic solution on a 3-torus of system (1.3) when (k,τ)=(1.364,59.4)D5.

    In region D6, system (4.2) has four equilibria: A0, A1, A2 and A3, which are all unstable. When the parameter enters region D6, the 3-torus disappears through heteroclinic orbits, and the system (1.3) generates strange attractor according to the "the Ruelle-Takens-Newhouse" scenario. Figure 11 illustrates system (1.3), which demonstrates chaotic phenomena when k=1.3801 and τ=59.978 are chosen in D6. The right figures of Figure 9, 10 and 11 are the results of Poincaré map on a Poincaré section.

    Figure 11.  When (k,τ)=(1.3801,59.978)D6, the system has chaotic behavior and there is a strange attractor of system (1.3).

    In region D7, system (4.2) has three equilibria: A0, A1 and A2, which are all unstable. This means that the positive equilibrium E(u,v) and the spatially homogeneous periodic solutions of system (1.3) are all unstable.

    In region D8, system (4.2) has two equilibria: A0 and A2, which are both unstable. This means that the positive equilibrium E(u,v) and the spatially homogeneous periodic solution of system (1.3) are all unstable.

    There have been more and more evidences showing that fear from the predator may affect the behavior and physiology of prey, which could reduce the reproduction of prey [1,6,7]. Moreover, anti-predation behavior of scared prey can also reduce the chance of prey being caught by predators. Therefore, we consider a predator-prey model with both costs and benefits due to fear based on [16]. Since the cost of fear in prey reproduction is not instantaneous, we incorporate a fear response delay into the model. Our aim is to explore how the fear level and fear response delay jointly affect the dynamics of the predator-prey system from the point of view of both codimension-1 and codimension-2 bifurcation analysis.

    First, we find that Turing instability does not occur in our model. Since the pioneering work of Turing [30], quite a number of literature has revealed that diffusion might destabilize an equilibrium and generate spatial pattern formation. However, not all diffusive systems exhibit such diffusion-driven instability. In some chemical and biological models, Turing instability induced by diffusion are not observed [31,32,33].

    Next, we explore the role of fear response delay on the dynamics. Through the analysis of the characteristic equation, it turns out that there are three different cases: the fear response delay may have no effect, both stabilizing and destabilizing effect, or destabilizing effect on the stability of E, which are shown in Theorems 2, 3, 4, respectively. Combining with numerical simulations, it is found that the the occurrence of Hopf bifurcation induced by fear response delay has closely connection with the fear level. When the fear level is low, the delay could not change the stability of the positive equilibrium, which is asymptotically stable for all τ. From the point of view of biology, the fear response delay cannot support periodic oscillations in prey and predator populations with a low level of fear. When the fear level is in an intermediate range, the fear response delay can induce stability switches, which indicates that the delay has effects on both destabilizing and stabilizing the dynamics. It is observed that such stability switches cannot be generated by the biomass transfer delay [16]. When the fear level is high, the fear response delay can destabilize the equilibrium and induce Hopf bifurcation.

    Finally, to better reveal the joint effect of fear level and fear response delay on the dynamics of system (1.3), we carry out codimension-2 bifurcation analysis. On the (k,τ) plane, by finding the interactions of Hopf bifurcation curves, we can find the possible critical point of double Hopf bifurcation. After the calculation of the normal form for double Hopf bifurcation, we can obtain the bifurcation set, and figure out all the dynamical behaviors around the critical point. There are periodic solutions, quasi-periodic solutions and even chaos observed near the double Hopf bifurcation point.

    In this paper, we assume that the interaction between individuals of a species is local. The individuals at different locations may compete for common resource or communicate visually or by chemical means. Regarding fear effect, the fear of predator depends on both the local and nearby appearance of the predator. Taking these into account, the influence of the nonlocal effect on the spatiotemporal dynamics of the diffusive predator-prey model with fear would be worth further study. In addition, we consider diffusion induced by random movement, however, cognition and memory also have an important influence on the animal movement. Shi et al. [34] proposed a memory-based diffusion equation of a single species. Memory-based diffusion can be modified and applied to the model with fear effect, which will provide much more realistic results and bring mathematical challenges.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The work is supported by National Natural Science Foundation of China (Nos. 11901369, 11971281 and 12071268) and PhD research startup foundation of Shaanxi University of Science and Technology (2023BJ-13).

    The authors declare there is no conflicts of interest.

    By direct calculation, we have

    d[(ω+n)2Dn]dn=2ω+ndω+ndndDndn=12{d(T2n2Dn)dn+1Δ[(T2n2Dn)d(T2n2Dn)dn2d(D2nM2)dn]}dDndn.

    From Lemma 1, we have d(T2n2Dn)dn>0, d(D2nM2)dn>0. Moreover, from (2.10), we know that T2n2Dn<0 for 0nn1. Combining dDndn>0 and M>0, we have d[(ω+n)2Dn]dn<0, and d[arccos((ω+n)2DnM)+2jπ]dn>0. Similarly, we can also prove that dω+ndn<0. Thus, τj+n is monotonically increasing with respect to n.

    Now we consider the monotonicity for τjn.

    d[(ωn)2Dn]dn=12{d(T2n2Dn)dn1Δ[(T2n2Dn)d(T2n2Dn)dn2d(D2nM2)dn]}dDndn=12d(T2n2Dn)dn[Δ+(T2n2Dn)Δ]+1Δd(D2nM2)dndDndn.

    Since D2nM2>0 for all nN0, and T2n2Dn<0 for 0nn1, then Δ+T2n2Dn<0 for 0nn1. Combining with Lemma 1, we have d(T2n2Dn)dn[Δ+(T2n2Dn)Δ]>0. From T2n2Dn<0, and D2n>M2, we have 4D2nΔ=4(D2nM2)+T2n(4DnT2n)>0, and thus 2DnΔ>0. From dDndn>0, we get 1Δd(D2nM2)dndDndn=dDndn[2DnΔΔ]>0. Therefore, d[(ωn)2Dn]dn>0, and d[arccos((ωn)2DnM)+2jπ]dn<0. d[(ωn)2Dn]dn>0 can also lead to dωndn>0. Thus, τjn is monotonically decreasing with respect to n.

    Fuu=2τ(a+pqv(1+qu)3(1+c1k), pqcv(1+qu)3(1+c1k))T,Fuv=τ(p(1+qu)2(1+c1k), cp(1+qu)2(1+c1k))T,Fuvt=τ(r0c2k(1+c2kv)2, 0)T,Fvtvt=2τ(r0uc22k2(1+c2kv)3, 0)T,Fuuu=6τ(pq2v(1+qu)4(1+c1k), pq2cv(1+qu)4(1+c1k))T,Fuuv=2τ(pq(1+qu)3(1+c1k), pqc(1+qu)3(1+c1k))T,Fuvtvt=2τ(r0c22k2(1+c2kv)3, 0)T,Fvtvtvt=6τ(r0uc32k3(1+c2kv)4, 0)T.

    From the expression of F2(α1,α2,Ut),

    L(1,0)1Ut=τ[G0(k+α1)Ut(0)+G1(k+α1)Ut(1)]L0Ut=τ[G0(k)Ut(0)+G1(k)Ut(1)],

    where

    J11(k)=pq[u(k)v(k)+u(k)v(k)]xypqu(k)v(k)hx2y3au(k),J12(k)=pu(k)xypu(k)[c1y+qu(k)x]x2y2,J21(k)=cpv(k)xy2cpv(k)hx2y3,K12(k)=r0c2[u(k)+ku(k)]z2r0kc22u(k)[v(k)+kv(k)]z3,

    with x=1+c1k, y=1+qu(k), z=1+c2kv(k), h=c1y+2qu(k)x,

    a0(k)=c2p,a1(k)=ac1ku(k)xy+(d+au(k))[qc1ku(k)x+yc1(c1k+x)],a2(k)=au(k)xy+[d+au(k)r0][qu(k)x+c1y],u(k)=cc1mp[cpmq(1+c1k)]2,v(k)=a1(k)+[a21(k)4a0(k)a2(k)]12[a1(k)a1(k)2(a0(k)a2(k)+a0(k)a2(k))]2a0(k)+a0(k)[a1(k)(a21(k)4a0(k)a2(k))12]2a20(k).
    ˜F2(z,y,α)=q1+q2+q3+q4=2Fq1q2q3q4γq1+q2n1(x)γq3+q4n2(x)zq11zq22zq33zq44+Syz1(y)z1γn1+Syz2(y)z2γn1+Syz3(y)z3γn2+Syz4(y)z4γn2+o(z2,y2),˜F3(z,0,0)=q1+q2+q3+q4=3Fq1q2q3q4γq1+q2n1(x)γq3+q4n2(x)zq11zq22zq33zq44,

    Here Syzi(i=1,2,3,4) are linear operators and

    Syzi(y)=(Fy1(0)zi,Fy2(0)zi)y(0)+(Fy1(1)zi,Fy2(1)zi)y(1),

    where

    Fy1(0)z1=2(Fuu+p12Fuv+p12eiω+n1τFuvt),Fy2(0)z1=2Fuv,Fy1(1)z1=0,Fy2(1)z1=2(Fuvt+p12eiω+n1τFvtvt),Fy1(0)z3=2(Fuu+p32Fuv+p32eiωn2τFuvt),Fy2(0)z3=2Fuv,Fy1(1)z3=0,Fy2(1)z3=2(Fuvt+p32eiωn2τFvtvt),Fy1(0)z2=¯Fy1(0)z1,Fy2(0)z2=¯Fy2(0)z1,Fy1(1)z2=¯Fy1(1)z1,Fy2(1)z2=¯Fy2(1)z1,Fy1(0)z4=¯Fy1(0)z3,Fy2(0)z4=¯Fy2(0)z3,Fy1(1)z4=¯Fy1(1)z3,Fy2(1)z4=¯Fy2(1)z3,

    and

    F2000=Fuu+2p12Fuv+2p12eiω+n1τFuvt+p212e2iω+n1τFvtvt,F1100=2(Fuu+(p12+ˉp12)Fuv+(p12eiω+n1τ+ˉp12eiω+n1τ)Fuvt+p12ˉp12Fvtvt),F1010=2(Fuu+(p12+p32)Fuv+(p12eiω+n1τ+p32eiωn2τ)Fuvt+p12p32e(iω+n1+iωn2)τFvtvt),F1001=2(Fuu+(p12+ˉp32)Fuv+(p12eiω+n1τ+ˉp32eiωn2τ)Fuvt+p12ˉp32e(iω+n1+iωn2)τFvtvt),F0020=Fuu+2p32Fuv+2p32eiωn2τFuvt+p232e2iωn2τFvtvt,F0011=2(Fuu+(p32+ˉp32)Fuv+(p32eiωn2τ+ˉp32eiωn2τ)Fuvt+p32ˉp32Fvtvt),F2100=3(Fuuu+(2p12+ˉp12)Fuuv+(2p12ˉp12+p212e2iω+n1τ)Fuvtvt+p212ˉp12eiω+n1τFvtvtvt),F1011=6(Fuuu+(p12+p32+ˉp32)Fuuv+(p32ˉp32+p12ˉp32e(iω+n1+iωn2)τ+p12p32e(iω+n1+iωn2)τ)Fuvtvt+p12p32ˉp32eiω+n1τFvtvtvt),F0021=3(Fuuu+(2p32+ˉp32)Fuuv+(2p32ˉp32+p232e2iωn2τ)Fuvtvt+p232ˉp32eiωn2τFvtvtvt),F1110=6(Fuuu+(p12+ˉp12+p32)Fuuv+(p12ˉp12+p32ˉp12e(iω+n1iωn2)τ+p12p32e(iω+n1+iωn2)τ)Fuvtvt+p12p32ˉp12eiωn2τFvtvtvt),F0200=¯F2000,F0110=¯F1001,F0101=¯F1010,F0002=¯F0020.

    Denote f1(k)q1q2q3q4=1lπψk(0)Fq1q2q3q4,(q1+q2+q3+q4=2).



    [1] É. Diz-Pita, M. V. Otero-Espinar, Predator–prey models: A review of some recent advances, Mathematics, 9 (2021), 1783. https://https://doi.org/10.3390/math9151783 doi: 10.3390/math9151783
    [2] Q. J. A. Khan, E. Balakrishnan, G. C. Wake, Analysis of a predator-prey system with predator switching, B. Math. Biol., 66 (2004), 109–123. https://doi.org/10.1016/j.bulm.2003.08.005 doi: 10.1016/j.bulm.2003.08.005
    [3] S. Liu, E. Beretta, A stage-structured predator-prey model of Beddington-DeAngelis type, Siam. J. Appl. Math., 66 (2006), 1101–1129. https://doi.org/10.1137/050630003 doi: 10.1137/050630003
    [4] J. M. Jeschke, M. Kopp, R. Tollrian, Predator functional responses: Discriminating between handling and digesting prey, Ecol. Monogr., 72 (2002), 95–112. https://doi.org/10.1890/0012-9615(2002)072[0095:PFRDBH]2.0.CO;2 doi: 10.1890/0012-9615(2002)072[0095:PFRDBH]2.0.CO;2
    [5] C. S. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, Mem. Entomol. Soc. Can., 97 (1965), 5–60. https://doi.org/10.4039/entm9745fv doi: 10.4039/entm9745fv
    [6] L. Y. Zanette, A. F. White, M. C. Allen, M. Clinchy, Perceived predation risk reduces the number of offspring songbirds produce per year, Science, 334 (2011), 1398–140. https://doi.org/10.1126/science.1210908 doi: 10.1126/science.1210908
    [7] W. Cresswell, Predation in bird populations, J. Ornithol., 152 (2011), 251–263. https://doi.org/10.1007/s10336-010-0638-1 doi: 10.1007/s10336-010-0638-1
    [8] X. Wang, L. Zanette, X. Zou, Modelling the fear effect in predator-prey interactions, J. Math. Biol., 73 (2016), 1179–1204. https://doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
    [9] Y. Shi, J. Wu, Q. Cao, Analysis on a diffusive multiple Allee effects predator-prey model induced by fear factors, Nonlinear Anal. Real, 59 (2021), 103249. https://doi.org/10.1016/j.nonrwa.2020.103249 doi: 10.1016/j.nonrwa.2020.103249
    [10] X. Zhang, H. Zhao, Y. Yuan, Impact of discontinuous harvesting on a diffusive predator-prey model with fear and Allee effect, Z. Angew. Math. Phys., 73 (2022), 1–29. https://doi.org/10.1007/s00033-022-01807-8 doi: 10.1007/s00033-022-01807-8
    [11] S. Pal, N. Pal, S. Samanta, J. Chattopadhyay, Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model, Math. Biosci. Eng., 16 (2019), 5146–5179. https://doi.org/10.3934/mbe.2019258 doi: 10.3934/mbe.2019258
    [12] P. Panday, N. Pal, S. Samanta, P. Tryjanowski, Dynamics of a stage-structured predator-prey model: cost and benefit of fear-induced group defense, J. Theor. Biol., 528 (2021), 110846. https://doi.org/10.1016/j.jtbi.2021.110846 doi: 10.1016/j.jtbi.2021.110846
    [13] S. K. Sasmal, Y. Takeuchi, Dynamics of a predator-prey system with fear and group defense, J. Math. Anal. Appl., 481 (2020), 123471. https://doi.org/10.1016/j.jmaa.2019.123471 doi: 10.1016/j.jmaa.2019.123471
    [14] N. Zhang, Y. Kao, B. Xie, Impact of fear effect and prey refuge on a fractional order prey-predator system with Beddington-DeAngelis functional response, Chaos, 32 (2022), 043125. https://doi.org/10.1063/5.0082733 doi: 10.1063/5.0082733
    [15] H. Chen, C. Zhang, Dynamic analysis of a Leslie-Gower-type predator-prey system with the fear effect and ratio-dependent Holling III functional response, Nonlinear Anal-Model., 27 (2022), 904–926. https://doi.org/10.15388/namc.2022.27.27932 doi: 10.15388/namc.2022.27.27932
    [16] Y. Wang, X. Zou, On a predator-prey system with digestion delay and anti-predation strategy, J. Nonlinear Sci., 30 (2020), 1579–1605. https://doi.org/10.1007/s00332-020-09618-9 doi: 10.1007/s00332-020-09618-9
    [17] B. Dai, G. Sun, Turing-Hopf bifurcation of a delayed diffusive predator-prey system with chemotaxis and fear effect, Appl. Math. Lett., 111 (2021), 106644. https://doi.org/10.1016/j.aml.2020.106644 doi: 10.1016/j.aml.2020.106644
    [18] X. Zhang, Q. An, L. Wang, Spatiotemporal dynamics of a delayed diffusive ratio-dependent predator-prey model with fear effect, Nonlinear Dyn., 105 (2021), 3775–3790. https://doi.org/10.1007/s11071-021-06780-x doi: 10.1007/s11071-021-06780-x
    [19] C. Wang, S. Yuan, H. Wang, Spatiotemporal patterns of a diffusive prey-predator model with spatial memory and pregnancy period in an intimidatory environment, J. Math. Biol., 84 (2022), 12. https://doi.org/10.1007/s00285-022-01716-4 doi: 10.1007/s00285-022-01716-4
    [20] J. Liu, Y. Kang, Spatiotemporal dynamics of a diffusive predator-prey model with fear effect, Nonlinear Anal-Model., 27 (2022), 841–862. https://doi.org/10.15388/namc.2022.27.27535 doi: 10.15388/namc.2022.27.27535
    [21] X. Wang, X. Zou, Pattern formation of a predator-prey model with the cost of anti-predator behaviors, Math. Biosci. Eng., 15 (2018), 775–805. https://doi.org/10.3934/mbe.2018035 doi: 10.3934/mbe.2018035
    [22] J. P. Tripathi, S. Bugalia, D. Jana, N. Gupta, V. Tiwari, J. Li, et al., Modeling the cost of anti-predator strategy in a predator-prey system: The roles of indirect effect, Math. Methods Appl. Sci., 45 (2022), 4365-4396. https://doi.org/10.1002/mma.8044 doi: 10.1002/mma.8044
    [23] D. Duan, B. Niu, J. Wei, Hopf-Hopf bifurcation and chaotic attractors in a delayed diffusive predator-prey model with fear effect, Chaos Solitons Fractals, 123 (2019), 206–216. https://doi.org/10.1016/j.chaos.2019.04.012 doi: 10.1016/j.chaos.2019.04.012
    [24] P. Panday, S. Samanta, N. Pal, J. Chattopadhyay, Delay induced multiple stability switch and chaos in a predator-prey model with fear effect, Math. Comput. Simul., 172 (2020), 134–158. https://doi.org/10.1016/j.matcom.2019.12.015 doi: 10.1016/j.matcom.2019.12.015
    [25] B. Dubey, A. Kumar, Stability switching and chaos in a multiple delayed prey-predator model with fear effect and anti-predator behavior, Math. Comput. Simulat., 188 (2021), 164–192. https://doi.org/10.1016/j.matcom.2021.03.037 doi: 10.1016/j.matcom.2021.03.037
    [26] Y. Song, Q. Shi, Stability and bifurcation analysis in a diffusive predator-prey model with delay and spatial average, Math. Method. Appl. Sci., 5 (2023), 5561–5584. https://doi.org/10.1002/mma.8853 doi: 10.1002/mma.8853
    [27] D. Geng, W. Jiang, Y. Lou, H. Wang, Spatiotemporal patterns in a diffusive predator-prey system with nonlocal intraspecific prey competition, Stud. Appl. Math., 148 (2022), 396–432. https://doi.org/10.1111/sapm.12444 doi: 10.1111/sapm.12444
    [28] Y. Du, B. Niu, Y. Guo, J. Wei, Double Hopf bifurcation in delayed reaction-diffusion systems, J. Dyn. Differ. Equ., 32 (2020), 313–358. https://doi.org/10.1007/s10884-018-9725-4 doi: 10.1007/s10884-018-9725-4
    [29] J. Guckenheimer, P. Holmes, Local codimension two bifurcations of flows in Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, Springer, (1983), 397–411. https://doi.org/10.1007/978-1-4612-1140-2
    [30] A. M. Turing, The chemical basis of morphogenesis, B. Math. Biol., 52 (1990), 153–197. https://doi.org/10.1016/S0092-8240(05)80008-4 doi: 10.1016/S0092-8240(05)80008-4
    [31] Y. Almirantis, S. Papageorgiou, Cross-diffusion effects on chemical and biological pattern formation, J. Theor. Biol., 151 (1991), 289–311. https://doi.org/10.1016/S0022-5193(05)80379-0 doi: 10.1016/S0022-5193(05)80379-0
    [32] J. Chattopadhyay, P. K. Tapaswi, Effect of cross-diffusion on pattern formation-a nonlinear analysis, Acta Appl. Math., 48 (1997), 1–12. https://doi.org/10.1023/A:1005764514684 doi: 10.1023/A:1005764514684
    [33] J. Zhao, J. Wei, Dynamics in a diffusive plankton system with delay and toxic substances effect, Nonlinear Anal. Real, 22 (2015), 66–83. https://doi.org/10.1016/j.nonrwa.2014.07.010 doi: 10.1016/j.nonrwa.2014.07.010
    [34] J. Shi, C. Wang, H. Wang, X. Yan, Diffusive spatial movement with memory, J. Dyn. Differ., 32 (2020), 979–1002. https://doi.org/10.1007/s10884-019-09757-y doi: 10.1007/s10884-019-09757-y
  • This article has been cited by:

    1. Xin Du, Quansheng Liu, Yuanhong Bi, Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay, 2023, 32, 2688-1594, 293, 10.3934/era.2024014
    2. Huazhou Mo, Yuanfu Shao, Stability and bifurcation analysis of a delayed stage-structured predator–prey model with fear, additional food, and cooperative behavior in both species, 2025, 2025, 2731-4235, 10.1186/s13662-025-03879-y
  • Reader Comments
  • © 2023 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(1393) PDF downloads(130) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog