Research article Special Issues

The effects of fear and delay on a predator-prey model with Crowley-Martin functional response and stage structure for predator

  • Taking into account the delayed fear induced by predators on the birth rate of prey, the counter-predation sensitiveness of prey, and the direct consumption by predators with stage structure and interference impacts, we proposed a prey-predator model with fear, Crowley-Martin functional response, stage structure and time delays. By use of the functional differential equation theory and Sotomayor's bifurcation theorem, we established some criteria of the local asymptotical stability and bifurcations of the system equilibrium points. Numerically, we validated the theoretical findings and explored the effects of fear, counter-predation sensitivity, direct predation rate and the transversion rate of the immature predator. We found that the functional response as well as the stage structure of predators affected the system stability. The fear and anti-predation sensitivity have positive and negative impacts to the system stability. Low fear level and high anti-predation sensitivity are beneficial to the system stability and the survival of prey. Meanwhile, low anti-predation sensitivity can make the system jump from one equilibrium point to another or make it oscillate between stability and instability frequently, leading to such phenomena as the bubble, or bistability. The fear and mature delays can make the system change from unstable to stable and cause chaos if they are too large. Finally, some ecological suggestions were given to overcome the negative effect induced by fear on the system stability.

    Citation: Weili Kong, Yuanfu Shao. The effects of fear and delay on a predator-prey model with Crowley-Martin functional response and stage structure for predator[J]. AIMS Mathematics, 2023, 8(12): 29260-29289. doi: 10.3934/math.20231498

    Related Papers:

    [1] Weijie Lu, Yonghui Xia . Periodic solution of a stage-structured predator-prey model with Crowley-Martin type functional response. AIMS Mathematics, 2022, 7(5): 8162-8175. doi: 10.3934/math.2022454
    [2] Reshma K P, Ankit Kumar . Stability and bifurcation in a predator-prey system with effect of fear and additional food. AIMS Mathematics, 2024, 9(2): 4211-4240. doi: 10.3934/math.2024208
    [3] Sahabuddin Sarwardi, Hasanur Mollah, Aeshah A. Raezah, Fahad Al Basir . Direction and stability of Hopf bifurcation in an eco-epidemic model with disease in prey and predator gestation delay using Crowley-Martin functional response. AIMS Mathematics, 2024, 9(10): 27930-27954. doi: 10.3934/math.20241356
    [4] Fatao Wang, Ruizhi Yang, Yining Xie, Jing Zhao . Hopf bifurcation in a delayed reaction diffusion predator-prey model with weak Allee effect on prey and fear effect on predator. AIMS Mathematics, 2023, 8(8): 17719-17743. doi: 10.3934/math.2023905
    [5] Vikas Kumar, Nitu Kumari . Controlling chaos in three species food chain model with fear effect. AIMS Mathematics, 2020, 5(2): 828-842. doi: 10.3934/math.2020056
    [6] Xuyang Cao, Qinglong Wang, Jie Liu . Hopf bifurcation in a predator-prey model under fuzzy parameters involving prey refuge and fear effects. AIMS Mathematics, 2024, 9(9): 23945-23970. doi: 10.3934/math.20241164
    [7] Yaping Wang, Yuanfu Shao, Chuanfu Chai . Dynamics of a predator-prey model with fear effects and gestation delays. AIMS Mathematics, 2023, 8(3): 7535-7559. doi: 10.3934/math.2023378
    [8] Binfeng Xie, Na Zhang . Influence of fear effect on a Holling type III prey-predator system with the prey refuge. AIMS Mathematics, 2022, 7(2): 1811-1830. doi: 10.3934/math.2022104
    [9] Jingwen Cui, Hao Liu, Xiaohui Ai . Analysis of a stochastic fear effect predator-prey system with Crowley-Martin functional response and the Ornstein-Uhlenbeck process. AIMS Mathematics, 2024, 9(12): 34981-35003. doi: 10.3934/math.20241665
    [10] Kwadwo Antwi-Fordjour, Rana D. Parshad, Hannah E. Thompson, Stephanie B. Westaway . Fear-driven extinction and (de)stabilization in a predator-prey model incorporating prey herd behavior and mutual interference. AIMS Mathematics, 2023, 8(2): 3353-3377. doi: 10.3934/math.2023173
  • Taking into account the delayed fear induced by predators on the birth rate of prey, the counter-predation sensitiveness of prey, and the direct consumption by predators with stage structure and interference impacts, we proposed a prey-predator model with fear, Crowley-Martin functional response, stage structure and time delays. By use of the functional differential equation theory and Sotomayor's bifurcation theorem, we established some criteria of the local asymptotical stability and bifurcations of the system equilibrium points. Numerically, we validated the theoretical findings and explored the effects of fear, counter-predation sensitivity, direct predation rate and the transversion rate of the immature predator. We found that the functional response as well as the stage structure of predators affected the system stability. The fear and anti-predation sensitivity have positive and negative impacts to the system stability. Low fear level and high anti-predation sensitivity are beneficial to the system stability and the survival of prey. Meanwhile, low anti-predation sensitivity can make the system jump from one equilibrium point to another or make it oscillate between stability and instability frequently, leading to such phenomena as the bubble, or bistability. The fear and mature delays can make the system change from unstable to stable and cause chaos if they are too large. Finally, some ecological suggestions were given to overcome the negative effect induced by fear on the system stability.



    In the study of ecological dynamics, the diversity of populations is a very important index for the continuous development of populations. The dynamics of such systems as the predator-prey system, competition system, co-operation system and more are a vital research area [1,2,3].

    In the real world, the prey-predator system is the most popular, in which the predation is an important feature in evaluating the interactivity between prey and predator since the predator's survival depends on the successful consumption on prey populations, as well as regulating the overall population size. The functional response is such an index that reflects the consumption rate of predator populations on prey populations. It is characterized as the predation efficiency of the prey populations consumed by predator populations and it is usually decided by the biomass of prey populations, predator populations or both of them. Many authors have applied the Holling-type functional response to depict the predation relationship[4,5,6]. The Holling-type functional response is only dependent on the prey's biomass, but when it is at a low level, the number of prey predated by predator usually does not follow a linearly increasing curve with the prey's biomass. Some researches have proved that the consumption of prey by predator depends not only on the prey's biomass, but also on the predator's biomass[7,8,9,10,11,12]. As a result, many predator-dependent functional responses have been introduced, such as Crowley-Martin[9,10], Beddington-DeAngelis[11] and Hassell-Varley[12], which are more appropriate choices in mathematical modeling by experimental evidence. Taking into account the interference among predator populations, the predation rate is often declined when the predator populations are higher, even if the prey populations are higher, too. In this situation, the interactions between predator populations should be considered and the Crowley-Martin functional response is better to describe this kind of interplay than the other two, as it is more realistic in reflecting this biological background[13].

    In the ecological world, almost all kinds of prey and predator species have two stages of structure: The immature stage and the mature stage. In different stages, the individuals exhibit very different biological behaviors, such as the ability of predating and reproduction and susceptibility to predator. Hence, in mathematical modeling, the assumption that every predator has the same skill to predate prey is very impractical, whereas a great quantity of research works were performed on the assumption that every individual undergoes only one stage in its whole living history[14,15,16,17]. Compared with those single-stage systems, the authors in [14,15] showed that it was more realistic for people to predict the system dynamics if the stage structure was incorporated in the mathematical models. Taking the stage structure of species into account, recently some nice results of prey-predator systems with stage structure have been reported[16,17,18]. For example, the authors in [18] presented a predator-prey model where the predator population was divided into two stages and the immature predator was not capable of predation. Many researchers have shown that the parameters representing the stage structure of individuals can change the behaviors of species from stable status to fluctuating status, and vice versa[19]. Therefore, it is necessary to consider the effect of the stage structure of species on the system dynamics.

    Considering the Crowley-Martin type functional response of predator to prey and the stage structure of predators, we have the following system:

    {dP1dt=rP1d1P1c1P21b2P1P31+b1P1+b3P3+b1b3P1P3,dP2dt=ηb2P1P31+b1P1+b3P3+b1b3P1P3d2P2c2P2,dP3dt=c2P2d3P3c3P23, (1.1)

    where P1,P2 and P3 (the simplicity of P1(t) P2(t) and P3(t)) are the biomass of prey, immature predator and mature predator at time t, respectively. r>0 is the growth rate and d1 is the natural death rate of prey. c1 is the intra-specific competition rate between individuals. The immature predator has no predation ability with natural death rate d2, and the transversion rate of predator from immature to mature is c2. The predator functional response to prey is supposed to be the Crowley-Martin type. Parameters b1,b2 and b3 describe the effect of handling time, capture rate and the magnitude of interference among predators, respectively. η is the conversion rate of predators. The intra-specific competition rate between adult predator populations is c3, and the natural death rate is d3.

    The functional response of predator to prey is relatively easily observed[20]. However, besides the direct predation, more and more experimental evidences show that predators can create fear in prey populations by their voice or smell. This non-consumptive action of predators often induces some significant impacts on prey and forces them to change their behaviors [21]. In fact, in some cases, the fear effect of predators can work more efficiently than direct killing, resulting in the reduction of the number of prey or even the extinction of prey, even if the direct predation is absent. The fear effect from predators is a manifestation of sustained psychological stress because the prey species are always worried about the possible attack of a predator. To protect themselves, the prey will take some measures to prevent the predator's attack and present some anti-predator behaviors. Due to the effect of fear, some important aspects of prey behaviors are varied like hunting, foraging, reproduction ability and so on[22,23,24,25]. Empirical observations show that the prey shifts to a relatively safer place with less food and is accompanied by starvation to avoid the predator's attack[26], forages in a less amount due to the predator's fear [27], spends less time hunting to prevent the predation risk of mountain lions[28] or even changes their reproductive physiology due to the predation risk of wolves[29].

    Therefore, it is interesting to take into account the fear effect as well as the counter-predation sensitiveness. We then get the modified version of (1.1) below.

    {dP1dt=rP11+kfP3d1P1c1P21b2P1P31+b1P1+kb3P3+kb1b3P1P3,dP2dt=ηb2P1P31+b1P1+kb3P3+kb1b3P1P3d2P2c2P2,dP3dt=c2P2d3P3c3P23, (1.2)

    where f denotes the fear effect induced by the predator and k is the sensitiveness of prey to predation risk. That is, if the value of parameter k is higher, i.e., the counter-predation level of prey is higher, then the predation rate is to be lower instead.

    On the other hand, except for the effect of direct and indirect predation of a predator, time delay is another important factor that impacts the dynamics of prey-predator interactions[30]. Time delay emerges in almost all ecological processes. For example, there is a time lag for the immature individual to become mature, which is called maturation delay[31]. For predators, they absorb energy by consuming prey to keep their survival, but the conversion of energy is not finished instantly, and there exists a digestion delay[30]. The gestation delay of the newborn is more popular [30,32]. In addition, due to the fear effect induced by the predator, the reproduction efficiency of prey is reduced by many manners, which is all invisible immediately. It takes a relatively long period of time to manifest the impact of a predator's fear on the birth rate of prey. The fear delay refers to the time when prey population starts to change their physiological characteristics due to fear of predators. That is, there is always a time lag for the fear effect on the system dynamics[33,34,35,36]. The delayed biological system has been extensively studied by use of the theory of delay differential equations. Many results indicate that time delays can destroy the system stability. For instance, the authors in [37] considered a predator-prey system with mature delay and established the conditions of global stability and bifurcation. Mortoja et al.[38] considered the dynamics of a predator-prey model with mature delay and disease in prey population. Currently, the studies of how multi-delays affect the system dynamics have been executed and lots of helpful results have been obtained. Particularly the bifurcation thresholds of delays have been derived [35,39,40]. By study of a delayed predator-prey system with no intra-specific competitions, the authors in [39] observed that the system was stable when the time delay was low, while as for the increasing of time delays, the stability was lost and bifurcation appeared, and when the delay was large enough, then chaotic behavior occurred. Beretta et al.[40] involved a kind of geometric method into a delayed differential equation to determine whether the stability was changed or not. All these reveal that time delays play a key role in the analysis of system stability.

    Incorporating the response delay of prey to the fear of predator and mature delay of predator, (1.2) becomes

    {dP1dt=rP11+kfP3(tτ1)d1P1c1P21b2P1P31+b1P1+kb3P3+kb1b3P1P3,dP2dt=ηb2P1P31+b1P1+kb3P3+kb1b3P1P3d2P2c2P2,dP3dt=c2P2(tτ2)d3P3c3P23. (1.3)

    The initial conditions are

    Pi(θ)=φi(θ),θ[τ,0],τ=max{τ1,τ2},i=1,2,3.

    For biological justifications, we assume all parameters in the models are positive and φi(θ)0. In biological research, the system stability is one of the most important issues. We aim to study the local asymptotical stability, parameter bifurcations and the cost and benefit of fear and time delays on the stability of model (1.3) both theoretically and numerically.

    The innovations and contributions of this paper are as follows:

    ● The different predation ability of juvenile and adult predators, the mutual interference between predators, the fear induced by predator on prey as well as the counter-predator sensitivity of prey, and the fear and mature delays are incorporated simultaneously in the model.

    ● The effects of the conversion rate of prey to predator, stage structure, fear and time delays on the system stability are investigated. Particularly, the critical values of some bifurcation parameters are obtained by numerical simulations.

    ● The complicated ecological phenomena like bubble and bistability are presented intuitively.

    The rest of the article is constructed as follows. The existence, boundedness and extinction of species are presented in section two. The local stability and bifurcation of equilibrium points of (1.2) and (1.3) are executed in sections three and four, respectively. Simulations and numerical analysis are given in section five. Finally, a brief conclusion and discussion are given to end this paper in section six.

    By the theory of the delayed functional differential equation[30], if the initial values of system (1.3) are positive, then there is a unique positive solution. That is, system (1.3) is positive invariant. Next, we discuss the boundedness of solutions. From the first equation of (1.3), we have

    dP1dt(rd1)P1c1P21,

    then

    P1rd1c1.

    By the boundedness of P1, together with the second equation of (1.3), we have

    dP2dtηb2P1kb3d2P2c2P2ηb2(rd1)kb3c1(d2+c2)P2.

    Then P2ηb2(rd1)kb3c1(d2+c2) and the boundedness of P2 is derived. By the third equation of (1.3) as well as the boundedness of P2, we obtain

    dP3dtc2ηb2(rd1)kb3c1(d2+c2)d3P3c3P23c2ηb2(rd1)kb3c1(d2+c2)d3P3.

    Then, P3c2ηb2(rd1)kb3c1d3(d2+c2) and the boundedness of P3 is derived. Therefore, the boundedness of all species is derived.

    We proceed with the discussion of the extinction of species. It is clear that if r<d1, then P1 is extinct, which results in the extinction of the juvenile and adult predator. Now, we study the case of r>d1. Define P(t)=P2(t)+P3(t), differentiating P(t) w.r.t t along the solution of (1.3), and we have

    dP(t)dtηb2b1P3d2P2c2P2+c2P2(tτ2)d3P3c3P23;

    that is,

    dP2(t)dt+dP3(t)dt(d2+c2)P2+c2P2(tτ2)+(ηb2b1d3)P3c3P23.

    Consequently, we derive that

    dP3(t)dt(ηb2b1d3)P3c3P23.

    Hence,

    limtP3(t)=ηb2b1d3c3.

    Therefore, we can obtain that the adult predator P3 is extinct if ηb2b1<d3, which means the juvenile predator is also extinct. Here, the prey is stable with limtP1(t)=rd1c1.

    Finally, in this part we give the equilibrium points of (1.2). The equilibrium points can be classified into the following three classes:

    ● The trivial equilibrium point E0(0,0,0), which always exists.

    ● The predator-free equilibrium point ¯E(¯P1,0,0) with ¯P1=rd1c1, which exists under r>d1.

    ● The positive interior equilibrium ˜E(˜P1,˜P2,˜P3), which conditionally exists such that the right sides of (1.2) are equivalent to zero, which cannot be expressed by explicit function and is to be discussed numerically in section five.

    For the discussion of local asymptotic stability (LAS) of system (1.2), we start with the linearized system of (1.2). For simplicity, let g(P1,P3)=1+b1P1+kb3P3+kb1b3P1P3, and rewrite (1.2) as

    P(t)=F(P(t)),P(t)=(P1(t),P2(t),P3(t))T,F=(F1,F2,F3)T,

    then we can linearize system (1.2) as P(t)=JP(t) with Jacobian matrix

    J=(a110a13a21a22a230a32a33),

    where

    a11=r1+kfP3d12c1P1b2P3(1+kb3P3)g2(P1,P3),a13=rkfP1(1+kfP3)2b2P1(1+b1P1)g2(P1,P3),a21=ηb2P3(1+kb3P3)g2(P1,P3),a22=d2c2,a23=ηb2P1(1+b1P1)g2(P1,P3),a32=c2,a33=d32c3P3.

    For convenience of using the Sotomayor's bifurcation theorem[41] latter, we compute the partial derivatives of F on viable Pi(i=1,2,3) as follows:

    FP1P1=(2c1+2b1b2P3(1+kb3P3)2g3(P1,P3)2ηb1b2P3(1+kb3P3)2g3(P1,P3)0),FP1P3=FP3P1=(rkf(1+kfP3)2b2g2(P1,P3)ηb2g2(P1,P3)0),FP3P3=(2rk2f2P1(1+kfP3)3+2kb2b3P1(1+b1P1)2g3(P1,P3)2ηkb2b3P1(1+b1P1)2g3(P1,P3)2c3),FP1P2=FP2P1=FP2P2=FP2P3=FP3P2=(0,0,0)T.

    With these preparations, we analyze the dynamics of (1.2) near the equilibrium points.

    (1) Dynamics near E0

    At E0(0,0,0), the Jacobian matrix J0 of system (1.2) is,

    J0=(rd1000d2c200c2d3).

    It is not difficult to compute that the roots of characteristic equation are λ1=rd1, λ2=d2c2<0, λ3=d3<0. Thus, if r<d1, then it is LAS, otherwise it is unstable[42]. We discuss the existence of parameter bifurcation d1 at rd1=0. Denote dT1 the value of d1r=0. The eigenvectors of λ1=0 corresponding to J0 and JT0 are v1=(1,0,0)T,v2=(1,0,0)T, respectively. The partial derivatives of F(P1,P2,P3) on viable d1 and Fd1 on viable Pi(i=1,2,3) are

    Fd1=(P100),DFd1=(100000000).

    Now, we verify the conditions of bifurcation theorem point by point.

    (i)vT2[Fd1(E0,d1=dT1)]=0.(ii)vT1[DFd1(E0,d1=dT1)v2]=(1,0,0)(100000000)(100)=10.(iii)vT2[D2F(E0,d1=dT1)(v1,v1)]=FP1P1(E0,d1=dT1)=(1,0,0)(2c100)=2c10.

    Consequently, by use of Sotomayor's bifurcation theorem, there exists a transcritical bifurcation near E0 at the threshold dT1=r.

    (2) Dynamics near ¯E

    We linearize (1.2) and get the Jacobian matrix J1 at ¯E as below,

    J1=((rd1)0rkf(rd1)c1b2(rd1)c1+b1(rd1)0d2c2ηb2(rd1)c1+b1(rd1)0c2d3).

    By computation, one characteristic root is λ1=(rd1), and the other two meet

    (d2c2λ)(d3λ)ηc2b2(rd1)c1+b1(rd1)=0,

    namely,

    λ2+(d2+d3+c2)λ+(d2+c2)d3ηc2b2(rd1)c1+b1(rd1)=0.

    If rd1>0 and (d2+c2)d3ηc2b2(rd1)c1+b1(rd1)>0, then all characteristic roots are negative, and system (1.2) is LAS. Otherwise, it is unstable.

    If λ1=rd1=0, by the same manner as before, we can derive that (1.2) undergoes a transcritical bifurcation near ¯E at dT1=r.

    If λ1=(rd1)<0, (d2+c2)d3ηc2b2(rd1)c1+b1(rd1)=0, then λ2=d2d3c2<0,λ3=0. Choose η as a bifurcation parameter and analyze whether the bifurcation happens or not. Let η=ηT such that (d2+c2)d3ηc2b2(rd1)c1+b1(rd1)=0. The eigenvector of zero characteristic value corresponding to J1 is v1=(¯ϱ1,¯ϱ2,1)T, and the eigenvector corresponding to JT1 is v2=(0,¯ϱ3,1)T, where ¯ϱ1=1rd1,¯ϱ2=d3c2,¯ϱ3=d3(c1+b1(rd1))ηb2(rd1). Compute the partial derivatives about parameter η, then

    Fη=(0b2P1P31+b1P1+b3kP3+kb1b3P1P30),DFη=(000b2P3(1+kb3P3)g2(P1,P3)0b2P1(1+b1P1)g2(P1,P3)000).

    The conditions of the bifurcation theorem are validated as follows:

    (i)vT2[Fη(¯E,η=ηT]=0.

    (ii)vT1[DFη(¯E,η=ηT)v2]=(¯ϱ1,¯ϱ2,1)(00000b2(rd1)c1+b1(rd1)000)(0¯ϱ31)=b2d3(rd1)c2(c1+b1(rd1))0.

    (iii)D2F(¯E,η=ηT)(v1,v1)=FP1P1(E1,η=ηT)¯ϱ21+2FP1P2(E1,η=ηT)¯ϱ1¯ϱ2+2FP1P3(E1,η=ηT)¯ϱ1+FP2P2(E1,η=ηT)¯ϱ22+2FP2P3(E1,η=ηT)¯ϱ2+FP3P3(E1,η=ηT)=(2c100)¯ϱ21+2(rkfb2(1+b1¯P1)2ηb2(1+b1¯P1)20)¯ϱ1+(2rk2f2P11+2kb2b3¯P11+b1¯P12ηkb2b3¯P11+b1¯P12c3).

    Consequently,

    vT2[D2F(E1,η=ηT)(v1,v1)]0.

    Therefore, Sotomayor's bifurcation theorem implies that system (1.2) has a transcritical bifurcation near E1 at η=ηT. Similarly we can obtain the bifurcation thresholds of other parameters (for example, c2).

    (3) Dynamics near ˜E

    Next, we research the stability and bifurcation around the coexistence equilibrium state ˜E. The variational matrix J2 at ˜E is

    J2=(˜a110˜a13˜a21˜a22˜a230˜a32˜a33),

    where ˜a11=r1+kf˜P3d12c1˜P1b2˜P3(1+kb3˜P3)g2(˜P1,˜P3),˜a13=rkf˜P1(1+kf˜P3)2b2˜P1(1+b1˜P1)g2(˜P1,˜P3),˜a21=ηb2˜P3(1+kb3˜P3)g2(˜P1,˜P3),a22=d2c2,˜a23=ηb2˜P1(1+b1˜P1)g2(˜P1,˜P3),a32=c2,a33=d32c3˜P3.

    The characteristic equation is

    λ3+δ1λ2+δ2λ+δ3=0, (3.1)

    where δ1=(˜a11+˜a22+˜a33),δ2=˜a11˜a22+˜a11˜a33+˜a22˜a33˜a23˜a32,δ3=˜a11˜a23˜a32˜a11˜a22˜a33˜a13˜a21˜a32. By the Rowth-Hurwitz theorem, if δ1>0,δ3>0,δ1δ2>δ3, then it is LAS, otherwise it is unstable.

    If δ3=0, then there is a zero eigenvalue of (3.1). Choose c2 as a bifurcation parameter and denote cS2 such that δ3=0, then we study the bifurcation about parameter c2. The eigenvectors of the zero eigenvalue corresponding to matrix J2 and JT2 are ˜v1=(˜ϱ1,˜ϱ2,1)T and ˜v2=(˜ϱ3,1,˜ϱ4)T, respectively, where ˜ϱ1=˜a13˜a11,˜ϱ2=˜a33˜a32,˜ϱ3=˜a21˜a11,˜ϱ4=˜a22˜a32. An easy computation results in

    Fc2=(0P2P2),DFc2=(000010010).

    Then, with the help of Matlab2014a, we compute that

    (i)˜vT2[Fc2(˜E,c2=cS2]0.

    (ii)˜vT2[D2F(˜E,c2=cS2)(˜v1,˜v1)]0.

    Therefore, by the bifurcation theorem (Theorem 1 in Page 338 of [41]), (1.2) has a saddle-node bifurcation near ˜E at c2=cS2.

    Next, we investigate the Hopf bifurcation of (1.2) near ˜E. Suppose (1.2) is LAS; that is, δ1>0, δ3>0,δ1δ2>δ3. If δ3=δ1δ2, then the characteristic equation becomes

    (λ2+δ2)(λ+δ1)=0. (3.2)

    It is clear that there is a negative root λ1=δ1. Another two characteristic roots of (3.2) are λ2,3=±δ2i. Choose k as the bifurcation parameter, then the roots of (3.2) can be written as

    λ1(k)=δ1,λ2,3=μ(k)±iν(k).

    Denote k=kH satisfying δ3δ1δ2=0. Now we verify the tranversality condition ddkRe(λi(k))|k=kH0, i=2,3. Putting λ2=μ(k)+iν(k) into (3.1) and calculating the derivative w.r.t k, noting μ(k)=0 and ν(k)=δ2(k), we have

    {1(k)μ(k)2(k)ν(k)=3(k),1(k)μ(k)+2(k)ν(k)=4(k),

    where 1(k)=2δ2(k),2(k)=2δ1(k)δ2(k),3(k)=δ1(k)δ2(k)δ3(k),4(k)=δ2(k)δ2(k). If

    δ22(k)+δ1(k)(δ1(k)δ2(k)δ3(k))0,

    then we compute that

    ddkRe(λi(k))|k=kH=1(k)4(k)2(k)3(k)21(k)+22(k)0.

    Note that λ1(k)=δ10, so by Theorem 2 in Page 353 of [41], a Hopf bifurcation occurs near ˜E at k=kH. Similarly, we can obtain the bifurcation thresholds of other parameters. We summarize our findings as follows.

    Theorem 3.1. For system (1.2), there are the following properties.

    (i) For equilibrium point E0(0,0,0), if r<d1, then system (1.2) is LAS, and there exists a transcritical bifurcation near E0 at the threshold dT1=r.

    (ii) For the predator-free equilibrium state ¯E(¯P1,0,0), if rd1>0 and (d2+c2)d3ηc2b2(rd1)c1+b1(rd1)>0, then system (1.2) is LAS. If rd1=0 or (d2+c2)d3ηc2b2(rd1)c1+b1(rd1)=0, then (1.2) undergoes transcritical bifurcations near ¯E with threshold dT1=r or η=ηT, respectively.

    (iii) For the interior equilibrium point ˜E(˜P1,˜P2,˜P3), if δ1>0,δ3>0,δ1δ2>δ3, then system (1.2) is LAS, otherwise it is unstable. If δ3=0, then a saddle-node bifurcation occurs near ˜E at c2=cS2. If δ1δ2=δ3,δ22(k)+δ1(k)(δ1(k)δ2(k)δ3(k))0, then a Hopf bifurcation occurs near ˜E at k=kH.

    Remark 3.1. The instability of E0 ensures the existence of ¯E and vice versa, but the relationship between the instability of ¯E and the existence of ˜E is not clear. In addition, by use of similar deduction, the bifurcation thresholds of other parameters, such as, the functional response parameters like bi (i=1,2,3), can be obtained similarly.

    As stated before, the time delay usually plays a crucial role in the system dynamics, so we investigate the delayed case now. By applying Taylor's formula, we linearize the delayed system (1.3) and get

    dP(t)dt=M0P(t)+M1P(tτ1)+M2P(tτ2),

    with Jacobian matrix

    M0=(a110˜a13a21a22a2300a33),M1=(00ˆa13000000),M2=(0000000a320),

    where ˘a13=b2P1(1+b1P1)g2(P1,P3),ˆa13=rkfP1(1+kfP3)2, and other parameters are the same as before. The variational matrix of delayed system (1.3) is

    ˜M=M0+eλτ1M1+eλτ2M2.

    Thus the characteristic equation of (1.3) is

    |˜MλI|=|a11λ0˘a13+ˆa13eλτ1a21a22λa230a32eλτ2a33λ|=0,

    i.e.,

    λ3+Υ1λ2+Υ2λ+Υ3+Υ4eλτ2+Υ5eλ(τ1+τ2)=0, (4.1)

    where Υ1=(a11+a22+a33),Υ2=a11a22+a11a33+a22a33a23a32eλτ2, Υ3=a11a23a32, Υ4=a11a23a32˘a13a21a32, Υ5=ˆa13a21a32. Clearly, if τ1=τ2=0, it is accordant with (3.1).

    For E0 and ¯E, the dynamics are the same as before, so we only discuss the dynamics near the coexistence equilibrium state ˜E(˜P1,˜P2,˜P3).

    We suppose (1.3) is LAS in the absence of delays, i.e., conditions δ1>0,δ2>0,δ1δ2>δ23 hold. According to the scenarios of one delay or two delay appearance, we classify our discussion into the following two cases.

    (1) The scenario of one delay.

    τ1>0,τ2=0.

    For system (1.3), if there is only one delay τ1 and the mature delay is free (τ2=0), then we have the following conclusion.

    Theorem 4.1. Suppose the delayed system (1.3) is LAS in the absence of delays. If τ1>0,τ2=0, then there exists a Hopf bifurcation for system (1.3) near ˜E at τ1=˜τ1 under the condition that

    Ξ1Ξ4Ξ2Ξ3>0,

    where ˜τ1,Ξi(i=1,2,3,4) are defined later in the proof.

    Proof. At τ2=0, the characteristic Eq (4.1) becomes

    λ3+Υ1λ2+Υ2λ+Υ3+Υ4+Υ5eλτ1=0. (4.2)

    Take τ1 as a bifurcation parameter and assume the characteristic root of (4.2) is in the general form λ(τ1)=α(τ1)+iβ(τ1). Due to the LAS of (1.2) with no delays, that is, α(τ1)<0, by the continuity of functions there exists a ˜τ1 such that α(˜τ1)=0. When τ1<˜τ1, it remains the LAS, while if τ1>˜τ1, it turns unstable from stable and may undergo bifurcations. Based on the analysis, we substitute the characteristic root λ into (4.2) and then separate the real part and imaginary part respectively, then

    α33αβ2+Υ1(α2β2)+Υ2α+Υ3+Υ4=Υ5eατ1cosβτ1,3α2ββ3+2Υ1αβ+Υ2β=Υ5eατ1sinβτ1.

    If ˜E alters its stability from stable to unstable, then the necessary condition is that the roots of (4.2) are in the following forms: λ(τ1)=iβ(τ1), i.e., α(τ1)=0, then

    {Υ1β2+Υ3+Υ4=Υ5cosβτ1,β3+Υ2β=Υ5sinβτ1. (4.3)

    Squaring both sides of (4.3) and adding them leads to

    β6+(Υ212Υ2)β4+(Υ22Υ1(Υ3+Υ4))β2Υ25=0. (4.4)

    Denote γ1=Υ212Υ2,γ2=Υ22Υ1(Υ3+Υ4),γ3=Υ25. For simplicity we rewrite (4.4) in the below form,

    β6+γ1β4+γ2β2+γ3=0. (4.5)

    Let h(x)=x6+γ1x4+γ2x2+γ3, then h(0)=γ3=Υ25<0,h()=. By the continuity of h(x), we obtain that (4.5) owns one positive solution at least and three positive solutions at most.

    Without loss of generality, we suppose it has three solutions denoted by βi,i=1,2,3, then for some invariant βi, by (4.3), we can seek out a sequence of τ(i,k)1 such that

    τ(i,k)1=1βicos1Υ1β2iΥ3Υ4Υ5+2kπβi,i=1,2,3,k=1,2,.

    Let ˜τ1=mini,kτ(i,k)1, then by Butler's lemma, ˜E remains stable when τ1<˜τ1. Next we verify

    ddτ1Reλ(τ1)|τ1=˜τ1>0,

    i.e., the tranversality condition of Hopf bifurcation holds. Differentiating (4.2) w.r.t. τ1 and substituting τ1=˜τ1,α(˜τ1)=0,βi(˜τ1)=˜βi, we have

    {Ξ1dαdτ1Ξ2dβidτ1=Ξ3,Ξ2dαdτ1+Ξ1dβidτ1=Ξ4, (4.6)

    where

    Ξ1=Υ23˜β2iΥ5˜τ1cos˜βi˜τ1,Ξ2=2Υ1˜βi+Υ5˜τ1sin˜βi˜τ1,Ξ3=Υ5˜βisin˜βi˜τ1,Ξ4=˜βicos˜βi˜τ1.

    Under the condition that Ξ1Ξ4Ξ2Ξ3>0, solving (4.6) results in

    dβidτ1=Ξ1Ξ4Ξ2Ξ3Ξ21+Ξ22>0.

    That is, the tranversality condition holds and, hence, a Hopf bifurcation occurs near ˜E at τ1=˜τ1.

    τ1=0,τ2>0.

    In this case, we take τ2 as a bifurcation parameter, then the characteristic equation is

    λ3+Υ1λ2+Υ2λ+Υ3+(Υ4+Υ5)eλτ2=0. (4.7)

    Repeating the previous process, we have the following result.

    Theorem 4.2. Suppose the delayed system (1.3) is LAS in the absence of delays. If τ1=0,τ2>0, then there exists a Hopf bifurcation for system (1.3) near ˜E at τ2=˜τ2 under the condition that ξ1ξ4ξ2ξ3>0, where

    ˜τ2=mini,kτ(i,k)2,τ(i,k)2=1βicos1Υ3Υ1β2iΥ4+Υ5+2kπβi,i=1,2,3,k=1,2,,ξ1=3˜β2iΥ2(Υ4+Υ5)˜τ2cos˜βi˜τ2,ξ2=2Υ1˜βi(Υ4+Υ5)˜βi˜τ2sin˜βi˜τ2,ξ3=(Υ4+Υ5)˜β2isin˜βi˜τ2,ξ4=(Υ4+Υ5)˜β2icos˜βi˜τ2,i=1,2,3,

    ˜βi=βi(˜τ2),βi(i=1,2,3) is the root of the following equation:

    β6+(Υ212Υ2)β4+(Υ222Υ1Υ3)Υ2(Υ4+Υ5)2=0.

    (2) The scenario of two delays.

    If there are two delays for (1.3), i.e., τi>0(i=1,2), then we fix one delay in the stable interval and choose the other as the Hopf bifurcation parameter. There are the following two scenarios.

    τ2(0,˜τ2) and τ1>0.

    We fix τ2 at some arbitrary point τ2=˘τ2(0,˜τ2), and take τ1 as a bifurcation parameter, then we obtain the criterion of the existence of Hopf bifurcation of system (1.3).

    Theorem 4.3. Suppose the delayed system (1.3) is LAS in the absence of delays. If τ1>0,τ2(0,˜τ2), then there exists a Hopf bifurcation for system (1.3) near ˜E at τ1=ˆτ1 under the conditions that

    (Υ3+Υ4)2Υ25<0and Ω1Ω3Ω2Ω4>0,

    where Υi(i=3,4,5) remains unchanged as those in (4.2), and Ωi(i=1,2,3,4) are defined in the proof.

    Proof. We fix τ2=˘τ2(0,˜τ2), then by (4.1), the characteristic equation is

    λ3+Υ1λ2+Υ2λ+Υ3+Υ4eλ˘τ2+Υ5eλ(τ1+˘τ2)=0. (4.8)

    Based on the same method, let λ(τ1)=α(τ1)+iβ(τ1) be a characteristic root of (4.8), then putting it into (4.8) and separating the real and imaginary parts leads to

    {α33αβ2+Υ1(α2β2)+Υ2α+Υ3+Υ4cosβ˘τ2eα˘τ2=Υ5cosβ(˘τ2+τ1)eα(˘τ2+τ1),3α2ββ3+2Υ1αβ+Υ2βΥ4sinβ˘τ2eα˘τ2=Υ5sinβ(˘τ2+τ1)eα(˘τ2+τ1). (4.9)

    The stability near ˜E changes from stable to unstable under the condition that there is a pair of purely imaginary roots for (4.8), i.e., α(τ1)=0, and whence (4.9) becomes

    {Υ1β2+Υ3+Υ4cosβ˘τ2=Υ5cosβ(˘τ2+τ1),β3+Υ2βΥ4sinβ˘τ2=Υ5sinβ(˘τ2+τ1). (4.10)

    Squaring the left and right sides of (4.10) and then adding them, we have

    β6+Λ1β4+Λ2β2+Λ3sinβ˘τ2+Λ4cosβ˘τ2+Λ5=0, (4.11)

    where

    Λ1=Υ212Υ2,Λ2=Υ222Υ1Υ3,Λ3=2Υ4β32Υ2Υ4β,Λ4=2Υ3Υ42Υ1Υ4β2,Λ5=Υ23+Υ24Υ25.

    Let h(x)=x6+Λ1x4+Λ2x2+Λ3sin˘τ2x+Λ4cos˘τ2x+Λ5. It is clear that h()=, then under the condition that h(0)=(Υ3+Υ4)2Υ25<0, by the continuity of h(x), (4.13) has at least one positive root and at most a finite number of positive solutions denoted by βj,j=1,,N. Thus, for some fixed solution βj(j1,2,,N), by (4.10) we can find a sequence of τ(j,k)1 such that

    τ(j,k)1=1βjcos1Υ1β2jΥ3Υ4cosβj˘τ2Υ5˘τ2+2kπβj,j=1,2,,N,k=1,2,.

    Let ˆτ1=minj,kτ(j,k)1, then by Butler's lemma, ˜E remains stable when τ1<ˆτ1. Again, we verify the following tranversality condition

    ddτ1Reλ(τ1)|τ1=ˆτ1>0.

    Differentiating (4.9) w.r.t. τ1 and putting τ1=ˆτ1,α(ˆτ1)=0,ˆβj=βj(ˆτ1) for some j{1,2,,N}, we have

    Ω1dαdτ1Ω2dβjdτ1=Ω3,
    Ω2dαdτ1+Ω1dβjdτ1=Ω4,

    where

    Ω1=3ˆβ2j+Υ2Υ4˘τ2cosˆβj˘τ2Υ5(˘τ2+ˆτ1)cosˆβj(ˆτ1+˘τ2),Ω2=2Υ1ˆβj+Υ4˘τ2sinˆβj˘τ2+γ5(τ1+˘τ2)sinˆβj(ˆτ1+˘τ2),Ω3=γ5ˆβjsinˆβj(ˆτ1+˘τ2),Ω4=γ5ˆβjcosˆβj(ˆτ1+˘τ2).

    Solving them results in

    dβjdτ1=Ω1Ω4Ω2Ω3Ω21+Ω21>0

    provided that Ω1Ω3Ω2Ω4>0. That is, the tranversality condition holds and, hence, a Hopf bifurcation occurs near ˜E at τ1=ˆτ1.

    τ1(0,˜τ1) and τ2>0.

    Fix τ1 at τ1=˘τ1(0,˜τ1) and take τ2 as a free parameter, then we have the similar theorem.

    Theorem 4.4. Suppose the delayed system (1.3) is LAS in the absence of delays. If ˘τ1(0,˜τ1),τ2>0, there exists a Hopf bifurcation for system (1.3) near ˜E at τ2=ˆτ2 under the conditions that

    Υ23Υ25(Υ4+Υ5cosβj˘τ1)2<0,Π1Π3Π2Π4>0,

    where

    ˆτ2=minj,kτ(j,k)2,τ(j,k)2=1βjcos1(Υ1β2jΥ3)(Υ4+Υ5cosβj˘τ1)+(Υ2βjβ3j)Υ5sinβj˘τ1(Υ4+Υ5cosβj˘τ1)2+Υ25sin2βj˘τ1+2kπβj,Π1=3ˆβ2j+Υ2Υ4ˆτ2cosˆβjˆτ2)Υ5(˘τ1+ˆτ2)cosˆβj(˘τ1+ˆτ2),Π2=2Υ1ˆβjΥ4ˆτ2sinˆβjˆτ2Υ5(˘τ1+ˆτ2)sinˆβj(˘τ1+ˆτ2),Π3=γ4ˆβjsinˆβjˆτ2+Υ5ˆβj(˘τ1+ˆτ2)sinˆβj(˘τ1+ˆτ2),Π4=γ4ˆβjcosˆβjˆτ2+Υ5ˆβj(˘τ1+ˆτ2)cosˆβj(˘τ1+ˆτ2),j=1,2,,N,k=1,2,,

    and ˆβj=βj(ˆτ2), βj(j=1,2,,N) is the root of the following equation:

    β6+(Υ212Υ2)β4+(Υ222Υ1Υ3)β2+Υ23Υ25(Υ4+Υ5cosβ˘τ1)2=0.

    In this part, we give some examples to simulate and validate the previous theoretical analysis and take the discussion further to discuss how the fear and time delays affect the stability of the equilibrium points. For this purpose, we fix a set of parameters and change the values of fear level f induced by the predator, the counter-predation parameter k of prey, the transversion level η of predator from prey and the maturing efficiency c2 from immature to mature predator. Without the others stated, the rest of the parameters are chosen as follows:

    r=3,d1=0.1,c1=0.1,b1=0.5,b2=0.6,b3=0.2,d2=0.1,d3=0.1,c3=0.02. (5.1)

    We make the numerical simulations from the following four scenarios.

    For system (1.2), let r=0.08,f=k=η=c2=0 and other parameter values are the same as (5.1), then it is not difficult to find that all species will become extinct and the trivial equilibrium point is E0(0,0,0). Keep the parameters unchanged and r=3, then the prey can survive while the predator remains extinct. The equilibrium state is ¯E(29,0,0). By Theorem 3.1, both E0 and ¯E are LAS. See Figures 1 and 2, respectively.

    Figure 1.  The time series diagram of species of (1.2). It shows that all species are extinct and the trivial equilibrium point E0(0,0,0) is LAS.
    Figure 2.  The dynamics of species of (1.2); (a) is the time series diagram of species; (b) is the phase diagram. It shows that the predator-free equilibrium point ¯E(¯P1,0,0) is LAS.

    Due to the complexity of the equations of equilibrium points, next we numerically validate that there exists a coexistence equilibrium point. Choose f=0.5,k=1,η=0.5,c2=0.8 and the rest of the parameters are taken from (5.1), then by computation we get the unique interior equilibrium point as ˜E(3.6540,0.9939,4.2827). It is LAS by Theorem 3.1 (see Figure 3). That is to say, all the species can coexist together.

    Figure 3.  The dynamics of species of (1.2); (a) is the time series diagram of species; (b) is the phase diagram. It indicates that the coexisting equilibrium point ˜E(3.6540,0.9939,4.2827) is LAS.

    In order to find whether the functional response and the stage structure affect the system dynamics, we let η=0.9 or c2=0.3, and the other parameter values are identical with Figure 3. By simulation we get the time series graph of prey and predator species (see Figure 4). Comparing Figure 3 with Figure 4, we find that when the parameter values of η or c2 are changed, the interior equilibrium point of (1.2) becomes unstable from stable, i.e., the conversion rate of prey to predator as well as the stage structure of predators affects the system stability.

    Figure 4.  The time series diagrams of species of system (1.2) with variable parameters η and c2, the other parameter values are identical with Figure 3; (a) is with η=0.9; (b) is with c2=0.3.

    Due to the predator's fear on prey, the prey presents some counter-predation features, then it is necessary to explore how the anti-predation sensitivity k affects the stability of the system. In this subsection, we mainly pay our attention on the effect of parameters k and f on the system dynamics. By resetting the parameter values of f=2 or k=0.2 and keeping the rest the same as Figure 3, we get the time series diagrams of model (1.2) (Figure 5). Simulation figures reveal that system (1.2) becomes oscillatory accompanied with a period fluctuation, which means that the system bifurcation about parameters k and f may occur, so we further explore the system bifurcation phenomenon.

    Figure 5.  The time series diagrams of system (1.2); the values of fear f and anti-predation sensitivity k are reset behind, the other parameters are the same as Figure 3(a) is with f=2; (b) is with k=0.2. It shows that system (1.2) becomes unstable accompanied with periodic fluctuation.

    (1) Bifurcation analysis of parameter k.

    In view of the importance of parameter k, which indirectly reveals how the fear level affects the whole system dynamics, we start with the bifurcation analysis of k.

    (i) Bifurcation of k for different fear level f.

    For system (1.2) with the parameter values given in Figure 5(a), by use of Matlab and Moncont [43], we depict the continuous trajectories of the coexistence equilibrium point of prey with k in different regions (see Figure 6). We find from Figure 6 that when k=0.002041 or k=1.8719, the Hopf bifurcation appears, and when k=0.002169, the saddle-node bifurcation appears. That is, when k is smaller, the system dynamics are more complex.

    Figure 6.  The bifurcation diagrams of prey P1 w.r.t parameter k of system (1.2) with f=2; (a) is with k[0,0.05]; (b) is with k[0,2].

    To better visualize the effect of k on the stability for different fear f, we change the value of f and draw the bifurcation picture of prey about parameter k with f=0.5 (see Figure 7). Compared with Figure 6(a), we find that the saddle-node bifurcation region is enlarged when the fear is reduced, which indicates that the fear indirectly affects the prey stability by the anti-predation sensitivity k.

    Figure 7.  The bifurcation diagram of prey P1 w.r.t parameter k of system (1.2) with f=0.5 and k[0,0.05].

    Figures 6 and 7 show that if the fear level of prey is high but the anti-predation sensitivity is very low, then the system changes its stability from stable to unstable and even makes the prey extinct. When the anti-predation sensitivity is high enough, like k>1.8719, then the system becomes stable again and the prey and predator can coexist. If the fear level is lower, then the system remains a stable and larger region. The survival region of prey is also enlarged.

    (ii) Two parameter bifurcation.

    The interplay of k and f on the dynamics of (1.2).

    As previously stated, both the parameters k and f represent the effects of fear on the system dynamics, and we now study their joint effects. By Matcont, we get the bi-parameter bifurcation picture (see Figure 8). Figure 8(a) is with k[0,2] and Figure 8(b) is the enlarged part of Figure 8 (a) with k[0,0.8], where R1 is the region of stable ¯E, R2 is the region of bistability between stable ¯E and oscillatory ˜E, R3 is the region of bistability between stable ¯E and ˜E, R4 is the region of oscillatory ˜E, R5 is the region of stable ˜E, R6 is the region of bistability between two limit cycles and R7 is the region of bistability between stable ˜E and oscillatory ˜E. Figure 7 shows that as the changes of k and f occur, the cush point, homoclinic bifurcation curve (HC)), saddle-node bifurcation of limit cycle (LPC) and generalized Hopf point appear. The stability of ˜E is changed and many bistability regions appear.

    Figure 8.  The dynamics of interior equilibrium point of system (1.2) about parameters k and f.

    To better visualize the dynamics of system switching, keep the parameter values given in (5.1) unchanged. Take f=0.5,η=0.5,c2=0.8 and change the value of k, or keep k=1,η=0.5,c2=0.8 and change the value of f, then we draw the phase graph of the system with three groups of different initial data (see Figure 9) where the initial data for the blue line, green line and red line are (5, 4, 2), (1, 2, 0.5) and (1, 1, 4), respectively.

    Figure 9.  The phase diagrams with three different initial data and varying k and f. The initial data for the blue trajectory is (5, 4, 2), the green trajectory is (1, 2, 0.5) and the red trajectory is (1, 1, 4). The previous four diagrams are with f=0.5 and the latter four diagrams are with k=1, and the changing parameter values see the bottom of each diagram.

    Figures 8 and 9 show that the fear level and anti-predation sensitivity can jointly affect the equilibrium stability of (1.2). Specifically, when the anti-predation sensitivity is low, the coexistence state of prey and predator will lose. The system will jump from one equilibrium point to another and oscillate between stability and instability frequently, even leading to bistability. The system dynamics are very complex.

    The interplay of k and c2 on the dynamics.

    As a stage structure system, the transversion rate is a very important indictor that decides on the whole system dynamics, so we investigate the interrelationship of anti-predation parameter k and the transversion rate c2. By simulation we get the graph of interplay effects between k and c2 on the system dynamics (see Figure 10) where R1 is the region of stable ˜E, R2 is the region of bistable between stable ˜E and oscillatory ˜E and R3 is the region of oscillatory ˜E. Clearly, when k is small, the dynamics of (1.2) are very rich. The cush point produces at the point (k,c2)=(0.008273,0.355558) and the bistability regions between oscillatory ˜E and stable ˜E are presented. It shows that the anti-predation behaviors of prey affect the predation rate of the predator, resulting in a change of the transversion rate from immature predator to mature predator, and the system presents some complicated dynamics.

    Figure 10.  The bifurcation diagram of two-parameters k and c2 of system (1.2).

    (2) Bubble phenomenon.

    In the biological system dynamics, we often see a closed loop-like structure where oscillation appears from one Hopf bifurcation point and disappears from another Hopf bifurcation point. From the diagram it seems like a bubble in the real world, so we call it a bubble phenomenon [32,35]. By drawing the Hopf bifurcation picture w.r.t. fear level f, we see the interesting bubble phenomenon (see Figure 11). It shows that as the change of parameter f occurs, the system (1.2) changes its stability from stable to unstable, then from unstable to stable again. That is, if the fear induced by the predator is low, then prey and predator can coexist and the system remains stable, but when it becomes large (f>0.649), then the system becomes unstable. If the fear level is sufficiently large, for example, f>21.104, then the prey and predator coexist again and the system remains stable.

    Figure 11.  The bifurcation diagram of species P1 of system (1.2) about parameter f.

    (3) Bistability.

    For a system with fixed parameter values, if the initial data is changed, then the solution curves of the system is convergent with different attractors and we say the system has a bistability phenomenon. From the bifurcation graphs of two parameters (Figures 8 and 10), we find that (1.2) has some different bistability behaviors. Given some different initial data, we depict the diagrams of time series and phase of system (1.2) to better visualize the bistability phenomenon (see Figure 12) where the parameter values are same as given in (5.1), and f=2,k=1.87,η=0.5 and c2=0.8. Figure 12(a) shows that the trajectories with different initial data converge to different cycles respectively. The trajectories of Figure 12(c) corresponding to different initial data converges to a cycle and a stable equilibrium respectively. However, the trajectories of Figure 12(e) converge to the same cycle although the initial data is different. Biologically, it means that under different conditions, the prey and predator can experience multiple stable coexistence states, which is valuable for people to keep the species alive and serve the economic development.

    Figure 12.  The bistability phenomena of system (1.2) with a set of fixed parameter values but different initial data; (a) is the time series diagram with initial data (1.03, 0.19, 1.22) for the green line and (1, 0.1, 1.2) for the blue line; (b) is the phase diagram of (a); (c) is the time series diagram with initial data (1, 0.1, 1) for the green line and (1, 1, 1) for the blue line; (d) is the phase diagram of (c); (e) is the time series diagram with initial data (1.03, 0.19, 1.22) for the green line and (3, 1, 1) for the blue line; (f) is the phase diagram of (e).

    In a word, the fear of predator brings some important influences to system (1.2), which leads to richer dynamical properties of the system. It makes the prey produce anti-predation behaviors and affects the predation rate of the predator, as well as the transversion rate of mature predator from immature predator. Particularly, the change of anti-predation sensitivity brings large influences to the system dynamics, such as changing the system stability from stable to unstable or from unstable to stable, accompanied by bifurcations, limit cycles and the bubble or bistability phenomena.

    Now we investigate how the time delay affects the system dynamics of system (1.3). For system (1.3), we keep the parameter values the same as given in Figure 3; that is, it is LAS when the delays are absent. Then, by choosing two small time delays as τ1=τ2=0.08, we find it is still stable (see Figure 13(a) and (b)). However, if we increase the delay values, for example, let τ1=τ2=5, then we observe that it is unstable and presents periodic fluctuation behaviors (see Figure 13(c) and (d)).

    Figure 13.  The dynamics of system (1.3) with time delays. (a) is with τ1=τ2=0.08, (b) is the phase diagram of (a); (c) is with τ1=τ2=5, (d) is the phase diagram of (c).

    By the simulation Figure 13, we conjecture that there may be a Hopf bifurcation. First, we consider the case of one delay. Let τ2=0 (or τ1=0) and depict the bifurcation picture w.r.t. τ1 (or τ2), then we get the bifurcation pictures (see Figure 14(a) or Figure 14(b)). Second, we consider the case of two delays. Keep one delay fixed in the range of assuring the system stability, and take the other delay as bifurcation parameter, then we get the delay bifurcation pictures (see Figure 14(c) and (d)). Similarly, as the change of one delay, Hopf bifurcations occur and system stability are changed from stable to unstable. If they are large enough, then the system will become unstable and produce chaos (see Figure 14(e) and (f)).

    Figure 14.  The Hopf bifurcation diagram of time delays of system (1.3); (a) is the bifurcation diagram of τ1 with τ2=0; (b) is the bifurcation diagram of τ2 with τ1=0; (c) is the bifurcation diagram of τ1 with τ2=3; (d) is the bifurcation diagram of τ2 with τ1=3; (e) is chaos phenomenon of (2.3) with τ1[50,150] and τ2=0,T=10000; (f) is the chaos phenomenon of (1.3) with τ2[10,50] and τ2=0,T=5000.

    All simulation figures indicate that if the prey can respond timely to the fear of predator and the mature period of the predator is very short, then the system can remain stable. Meanwhile, if the fear delay or mature delay is large enough, then the system stability is lost and accompanied with periodic fluctuation. If one of them becomes too large, then chaos occurs and the system becomes unstable. In a word, the fear and mature delays bring some significant influences to the system stability.

    Combining the real ecological environment, we formulated a delayed predator-prey system with a Crowley-Martin functional response and stage structure for predators. The existence of equilibrium points, the locally asymptotical stability and bifurcation behaviors of the system were investigated. All theoretical results were summarized in Theorems 3.1 and 4.1–4.4. In order to validate the theoretical results and intuitively reveal the influence on system dynamics of fear, stage structure, anti-predation sensitivity and time delays, some numerical analyses were given. Figures 13 validated the stability of equilibrium points. Figure 4 showed the functional response and stage structure of how predators affected the system stability. Figures 512 revealed how the fear and anti-predation sensitivity affected the system stability numerically. Figures 13 and 14 explained the effects of fear delay and mature delay on the system stability. We summarize our findings as follows:

    (1) The conversion rate from prey to predator and the stage structure of predator bring some influence on the stability of the equilibrium point of the system.

    (2) The fear induced by the predator seriously affects the stability of the interior equilibrium point. It makes the prey produce anti-predator behaviors, which affects the predation rate and transversion rate of mature predator from immature predator. The fear and anti-predation sensitivity jointly affects the system stability and produces complex dynamical phenomena such as bubble, or bistability.

    (3) The fear delay and mature delay will bring some important influence on the system dynamics. They can change the system stability from stable to unstable and even lead to chaos if they are large enough.

    By these findings, from the angle of protecting the species from extinction to maintaining the system permanence and the biological balance, we should make some measures such as

    ● Constructing some refuge zones for prey to prevent the fear of predators and increase the anti-predation sensitivity of prey.

    ● Supply some additional food for predators and decrease the predation rate to decline the negative effect of predators fear on prey.

    ● Shorten the mature delay and accelerate the mature speed of predators from immature predator by using some biological control strategies, like changing their genes.

    There are still some topics to be investigated. For example, the environments are always fluctuated by random noise, so incorporating the effects of environmental stochastic noise on the system dynamics is very interesting[44]. On the other hand, how the refuge and additional food supplement affects the stability is another interesting topic[45,46]. In addition, whether the mature delay can be efficiently controlled by some reasonable ways is also necessary to be considered. All of these are left for our future research.

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

    This research is supported by the National Science Foundation of China (No. 11861027, No. 12161062) and the Science and Technology Program of Inner Mongolia Autonomous Region (No. 2023YFHH0024, No. 2022YFHH0017, No. 2022YFHH0063).

    The authors would like to thank the editor and the anonymous reviewers for their valuable comments, which helped to improve the presentation of the paper.

    There is no competing interests in this paper.



    [1] A. Lotka, Elements physical biology, Baltimore: Williams and Wilkins, 1924.
    [2] V. Volterra, Fluctuations in the abundance of a species considered mathematically, Nature, 118 (1926), 558–560. http://dx.doi.org/10.1038/118558a0 doi: 10.1038/118558a0
    [3] J. Collings, The effects of the functional response on the bifurcation behavior of a mite predator-prey interaction model, J. Math. Biol., 36 (1997), 149–168. http://dx.doi.org/10.1007/s002850050095 doi: 10.1007/s002850050095
    [4] T. Kar, Modelling and analysis of a harvested prey-predator system incorporating a prey refuge, J. Comput. Appl. Math., 185 (2006), 19–33. http://dx.doi.org/10.1016/j.cam.2005.01.035 doi: 10.1016/j.cam.2005.01.035
    [5] Y. Huang, F. Chen, Z. Li, Stability analysis of a prey-predator model with holling type III response function incorporating a prey refuge, Appl. Math. Comput., 182 (2006), 672–683. http://dx.doi.org/10.1016/j.amc.2006.04.030 doi: 10.1016/j.amc.2006.04.030
    [6] J. Dawes, M. Souza, A derivation of Hollings type I, II and III functional responses in predator-prey systems, J. Theor. Biol., 327 (2017), 11–22. http://dx.doi.org/10.1016/j.jtbi.2013.02.017 doi: 10.1016/j.jtbi.2013.02.017
    [7] K. Antwi-Fordjour, R. Parshad, M. Beauregard, Dynamics of a predator-prey model with generalized functional response and mutual interference, Math. Biosci., 360 (2020), 108407. http://dx.doi.org/10.1016/j.mbs.2020.108407 doi: 10.1016/j.mbs.2020.108407
    [8] J. Roy, D. Barman, S. Alam, Role of fear in a predator-prey system with ratio-dependent functional response in deterministic and stochastic environment, Biosystems, 197 (2020), 104176. http://dx.doi.org/10.1016/j.biosystems.2020.104176 doi: 10.1016/j.biosystems.2020.104176
    [9] P. Panja, Dynamics of a predator-prey model with Crowley-Martin functional response, refuge on predator and harvesting of super-predator, J. Biol. Syst., 29 (2021), 631–646. http://dx.doi.org/10.1142/S0218339021500121 doi: 10.1142/S0218339021500121
    [10] J. Danane, Stochastic predator-prey Lévy jump model with Crowley-Martin functional response and stage structure, J. Appl. Math. Comput., 67 (2021), 41–67. http://dx.doi.org/10.1007/s12190-020-01490-w doi: 10.1007/s12190-020-01490-w
    [11] J. Tripathi, S. Abbas, M. Thakur, Dynamical analysis of a prey-predator model with Beddington-DeAngelis type function response incorporating a prey refuge, Nonlinear Dyn., 80 (2015), 177–196. http://dx.doi.org/10.1007/s11071-014-1859-2 doi: 10.1007/s11071-014-1859-2
    [12] C. Xu, P. Li, Oscillations for a delayed predator-prey modelwith Hassell-Varley-type functional response, C. R. Biol., 338 (2015), 227–240. http://dx.doi.org/10.1016/j.crvi.2015.01.002 doi: 10.1016/j.crvi.2015.01.002
    [13] J. Tripathi, S. Tyagi, S. Abbas, Global analysis of a delayed density dependent predator-prey model with Crowley-Martin functional response, Commun. Nonlinear Sci., 30 (2016), 45–69. http://dx.doi.org/10.1016/j.cnsns.2015.06.008 doi: 10.1016/j.cnsns.2015.06.008
    [14] A. De Roos, L. Persson, H. Thieme, Emergent allee effects in top predators feeding on structured prey populations, Proc. R. Soc. Lond. B, 270 (2003), 611–618. http://dx.doi.org/10.1098/rspb.2002.2286 doi: 10.1098/rspb.2002.2286
    [15] V. Pavlovˊa, L. Berec, Impacts of predation on dynamics of age-structured prey: Allee effects and multi-stability, Theor. Ecol., 5 (2012), 533–544. http://dx.doi.org/10.1007/s12080-011-0144-y doi: 10.1007/s12080-011-0144-y
    [16] J. Cui, L. Chen, W. Wang, The effect of dispersal on population growth with stage-structure, Comput. Math. Appl., 39 (2000), 91–102. http://dx.doi.org/10.1016/S0898-1221(99)00316-8 doi: 10.1016/S0898-1221(99)00316-8
    [17] P. Panday, N. Pal, S. Samanta, P. Tryjanowski, J. Chattopadhyay, Dynamics of a stage-structured predator-prey model: cost and benefit of fear-induced group defense, J. Theor. Biol., 528 (2021), 110846. http://dx.doi.org/10.1016/j.jtbi.2021.110846 doi: 10.1016/j.jtbi.2021.110846
    [18] X. Sun, H. Huo, H. Xiang, Bifurcation and stability analysis in predator-prey model with a stage-structure for predator, Nonlinear Dyn., 58 (2009), 497–513. http://dx.doi.org/10.1007/s11071-009-9495-y doi: 10.1007/s11071-009-9495-y
    [19] T. Kostova, J. Li, M. Friedman, Two models for competition between age classes, Math. Biosci., 157 (1999), 65–89. http://dx.doi.org/10.1016/S0025-5564(98)10077-9 doi: 10.1016/S0025-5564(98)10077-9
    [20] R. Taylor, Predation, New York: Chapman and Hall, 1984.
    [21] S. Lima, L. Dill, Behavioral decisions made under the risk of predation: a review and prospectus, Can. J. Zool., 68 (1990), 619–640. http://dx.doi.org/10.1139/z90-092 doi: 10.1139/z90-092
    [22] J. Brown, Vigilance, patch use and habitat selection: foraging under predation risk, Evol. Ecol. Res., 1 (1999), 49–71.
    [23] W. Cresswell, Predation in bird populations, J. Ornithol., 152 (2011), 251–263. http://dx.doi.org/10.1007/s10336-010-0638-1 doi: 10.1007/s10336-010-0638-1
    [24] A. Wirsing, W. Ripple, A comparison of shark and wolf research reveals similar behavioral responses by prey, Front. Ecol. Environ., 9 (2011), 335–341. http://dx.doi.org/10.1890/090226 doi: 10.1890/090226
    [25] S. Mortoja, P. Panja, S. Mondal, Dynamics of a predator-prey model with stage-structure on both species and anti-predator behavior, Informatics in Medicine Unlocked, 10 (2018), 50–57. http://dx.doi.org/10.1016/j.imu.2017.12.004 doi: 10.1016/j.imu.2017.12.004
    [26] W. Ripple, R. Beschta, Wolves and the ecology of fear: can predation risk structure ecosystems? BioScience, 54 (2004), 755–766. http://dx.doi.org/10.1641/0006-3568(2004)054[0755: WATEOF]2.0.CO; 2
    [27] A. Sih, Optimal behavior: can foragers balance two conflicting demands, Science, 210 (1980), 1041–1043. http://dx.doi.org/10.1126/science.210.4473.1041 doi: 10.1126/science.210.4473.1041
    [28] B. Pierce, R. Bowyer, V. Bleich, Habitat selection by mule deer: forage benefits or risk of predation? J. Wildl. Manage., 68 (2004), 533–541. http://dx.doi.org/10.2193/0022-541X(2004)068[0533: HSBMDF]2.0.CO; 2
    [29] S. Creel, D. Christianson, S. Liley, J. Winnie, Predation risk affects reproductive physiology and demography of elk, Science, 315 (2007), 960. http://dx.doi.org/10.1126/science.1135918 doi: 10.1126/science.1135918
    [30] Y. Kuang, Delay differential equations: with applications in population dynamics, New York: Academic Press, 1993.
    [31] S. Gourley, Y. Kuang, A stage structured predator-prey model and its dependence on maturation delay and death rate, J. Math. Biol., 49 (2019), 188–200. http://dx.doi.org/10.1007/s00285-004-0278-2 doi: 10.1007/s00285-004-0278-2
    [32] N. Pal, S. Samanta, S. Biswas, M. Alquran, K. Al-Khaled, J. Chattopadhyay, Stability and bifurcation analysis of a three-species food chain model with delay, Int. J. Bifurcat. Chaos, 25 (2015), 1550123. http://dx.doi.org/10.1142/S0218127415501230 doi: 10.1142/S0218127415501230
    [33] S. Pal, N. Pal, S. Samanta, J. Chattopadhyay, Effect of hunting cooperation and fear in a predator-prey model, Ecol. Complex., 39 (2019), 100770. http://dx.doi.org/10.1016/j.ecocom.2019.100770 doi: 10.1016/j.ecocom.2019.100770
    [34] Y. Shao, Global stability of a delayed predator-prey system with fear and Holling-type II functional response in deterministic and stochastic environments, Math. Comput. Simulat., 200 (2022), 65–77. http://dx.doi.org/10.1016/j.matcom.2022.04.013 doi: 10.1016/j.matcom.2022.04.013
    [35] B. Kumar Das, D. Sahoo, G. Samanta, Impact of fear in a delayed-induced predator-prey system with intraspecific competition within predator species, Math. Comput. Simulat., 191 (2022), 134–156. http://dx.doi.org/10.1016/j.matcom.2021.08.005 doi: 10.1016/j.matcom.2021.08.005
    [36] B. Dubey, S. Ankit Kumar, Stability switching and chaos in a multiple delayed prey-predator model with fear effect and anti-predator behavor, Math. Comput. Simulat., 188 (2021), 164–192. http://dx.doi.org/10.1016/j.matcom.2021.03.037 doi: 10.1016/j.matcom.2021.03.037
    [37] K. Chakraborty, S. Haldar, T. Kar, Global stability and bifurcation analysis of a delay induced prey-predator system with stage structure, Nonlinear Dyn., 73 (2013), 1307–1325. http://dx.doi.org/10.1007/s11071-013-0864-1 doi: 10.1007/s11071-013-0864-1
    [38] S. Mortoja, P. Panja, S. Mondal, Dynamics of a predator-prey model with nonlinear incidence rate, Crowley-Martin type functional response and disease in prey population, Ecological Genetics and Genomics, 10 (2019), 100035. http://dx.doi.org/10.1016/j.egg.2018.100035 doi: 10.1016/j.egg.2018.100035
    [39] Y. Shao, W. Kong, A predator-prey model with Beddington-DeAngelis functional response and multiple delays in deterministic and stochastic environments, Mathematics, 10 (2022), 3378. http://dx.doi.org/10.3390/math10183378 doi: 10.3390/math10183378
    [40] E. Beretta, Y. Kuang, Geometric stability switch criteria in delay differential systems with delay-dependent parameters, SIAM J. Math. Anal., 33 (2002), 1144–1165. http://dx.doi.org/10.1137/S0036141000376086 doi: 10.1137/S0036141000376086
    [41] L. Perko, Differential equations and dynamical systems, New York: Springer, 2001. http://dx.doi.org/10.1007/978-1-4613-0003-8
    [42] V. Kolmanovskii, A. Myshkis, Applied theory of functional differential differential equations, Dordrecht: Springer, 1992. http://dx.doi.org/10.1007/978-94-015-8084-7
    [43] A. Dhooge, W. Govaerts, Y. Kuznetsov, H. Meijer, B. Sautois, New features of the software matcont for bifurcation analysis of dynamical systems, Math. Comp. Model. Dyn., 14 (2007), 147–175. http://dx.doi.org/10.1080/13873950701742754 doi: 10.1080/13873950701742754
    [44] Y. Zhao, S. Yuan, Stability in distribution of a stochastic hybrid competitive Lotka-Volterra model with Lévy jumps, Chaos Soliton. Fract., 85 (2016), 98–109. http://dx.doi.org/10.1016/j.chaos.2016.01.015 doi: 10.1016/j.chaos.2016.01.015
    [45] S. Mondal, A. Maiti, G. Samanta, Effects of fear and additional food in a delayed predator-prey model, Biophysical Reviews and Letters, 13 (2018), 157–177. http://dx.doi.org/10.1142/S1793048018500091 doi: 10.1142/S1793048018500091
    [46] A. Thirthar, S. Majeed, M. Alqudah, P. Panja, T. Abdeljawad, Fear effect in a predator-prey model with additional food, prey refuge and harvesting on super predator, Chaos Soliton. Fract., 159 (2022), 112091. http://dx.doi.org/10.1016/j.chaos.2022.112091 doi: 10.1016/j.chaos.2022.112091
  • This article has been cited by:

    1. Bushra E. Kashem, Hassan F. Al-Husseiny, The dynamic of two prey–One predator food web model with fear and harvesting, 2024, 11, 26668181, 100875, 10.1016/j.padiff.2024.100875
    2. Weili Kong, Yuanfu Shao, Bifurcations of a Leslie-Gower predator-prey model with fear, strong Allee effect and hunting cooperation, 2024, 9, 2473-6988, 31607, 10.3934/math.20241520
    3. 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(2611) PDF downloads(162) Cited by(3)

Figures and Tables

Figures(14)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog