
Recent studies have shown that, in addition to direct predation, fear of predators alters the physiological behavior of prey. Based on this fact, this paper investigates a three-species food chain based on ratio-dependent and Beddington-DeAngelis type functional responses, which incorporates fear effects and two gestation delays. The positivity, boundedness and existence of equilibrium points of the system are investigated, and the local stability behavior of the equilibrium points and the occurrence of Hopf-bifurcation when the time lag parameters exceed the critical values are studied by analyzing the corresponding characteristic equations. The main results show that Hopf-bifurcation occurs when the time delay parameters attain the thresholds. Finally, numerical simulations are performed to verify our main results.
Citation: Yaping Wang, Yuanfu Shao, Chuanfu Chai. Dynamics of a predator-prey model with fear effects and gestation delays[J]. AIMS Mathematics, 2023, 8(3): 7535-7559. doi: 10.3934/math.2023378
[1] | 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 |
[2] | 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 |
[3] | 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. AIMS Mathematics, 2023, 8(12): 29260-29289. doi: 10.3934/math.20231498 |
[4] | San-Xing Wu, Xin-You Meng . Dynamics of a delayed predator-prey system with fear effect, herd behavior and disease in the susceptible prey. AIMS Mathematics, 2021, 6(4): 3654-3685. doi: 10.3934/math.2021218 |
[5] | 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 |
[6] | Weili Kong, Yuanfu Shao . Bifurcations of a Leslie-Gower predator-prey model with fear, strong Allee effect and hunting cooperation. AIMS Mathematics, 2024, 9(11): 31607-31635. doi: 10.3934/math.20241520 |
[7] | 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 |
[8] | Jing Zhang, Shengmao Fu . Hopf bifurcation and Turing pattern of a diffusive Rosenzweig-MacArthur model with fear factor. AIMS Mathematics, 2024, 9(11): 32514-32551. doi: 10.3934/math.20241558 |
[9] | 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 |
[10] | 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 |
Recent studies have shown that, in addition to direct predation, fear of predators alters the physiological behavior of prey. Based on this fact, this paper investigates a three-species food chain based on ratio-dependent and Beddington-DeAngelis type functional responses, which incorporates fear effects and two gestation delays. The positivity, boundedness and existence of equilibrium points of the system are investigated, and the local stability behavior of the equilibrium points and the occurrence of Hopf-bifurcation when the time lag parameters exceed the critical values are studied by analyzing the corresponding characteristic equations. The main results show that Hopf-bifurcation occurs when the time delay parameters attain the thresholds. Finally, numerical simulations are performed to verify our main results.
Predator-prey interactions play an important role in shaping community structure and maintaining ecological diversity. Over the past few decades, food chain models have been extensively studied and a large number of papers have been published. Researchers have used mathematical models to understand the complexity of ecosystems and predict the consequences of ecosystems [1,2,3,4,5,6,7,8].
Generally, predators affect prey populations by killing them directly, but many recent studies [9,10,11] have shown that fear effects play a crucial role in ecology and evolutionary biology. In [9], fear was found to cause prey species to exhibit many anti-predator behaviors and defense mechanisms. In [10], prey birth and death rates were affected by fear of predation risk. In [11], higher levels of fear were found to result in lower equilibrium levels of coexistence and global asymptotic stability of species near equilibrium. Their results all point to the importance of fear in predator-prey systems. The presence of predators forces prey to be more alert and spend less time foraging or choosing new habitats due to the fear of being killed, while fear also reduces the mating rate of prey populations, resulting in a decrease in prey population size. In 2011, Zanette et al. [12] used song sparrows as subjects for their experiments. They left the population of song sparrows in a predator-free area and then manipulated artificial predation risk by playing the sounds of predators. Through experiments, they observed that the song sparrows' perception of predation risk led to a reduction in their breeding population by about 40% compared to the previous population. So this experiment confirms that fear reduces the reproductive rate of prey. In another experiment involving elk, Creel et al. [13] observed that the risk of predation by wolves caused elk to remain alert and move to areas with low predation risk. This suggests that the physiological behavior of elk is influenced by anti-predator behavior. Suraci et al. [14] used sound playback of large carnivores to induce fear in medium carnivores and found that fear of large carnivores increased vigilance and greatly reduced foraging behavior in medium carnivores. These experiments all confirm the important effects of fear on the dynamic behavior of species. In recent years, many researchers have proposed that the predator-prey system should incorporate a fear factor. In 2016, Wang et al. [15] proposed a mathematical model with fear and confirmed the importance of fear on the growth of prey.
In ecology, how predators feed on prey is a very important factor, and the way they interact with each other affects the dynamics of the entire ecosystem. This interaction is called the "functional response", which gives the per capita rate of the predator feeding on prey and is a key component of the mathematical model, allowing us to study the dynamics of the model in a better way. The most widely used functional responses include Holling type Ⅱ, Holling type Ⅲ, Holling type Ⅳ [16,17,18], Monod-Haldane functional response, Beddington-DeAngelis functional response, Crowley-Martin functional response and the ratio-dependent functional responses, etc. [19,20,21,22,23]. The Holling type Ⅱ functional response is the more commonly used functional response, but this functional response depends only on prey density and is not applicable for ecosystems with large population sizes. To achieve the elimination of this situation, we introduce a ratio-dependent functional response, which is determined simultaneously by the relationship between prey density and predator density. In particular, when predators have to go in search of food, they will face a situation of sharing and competition with other predators. Therefore, ratio-dependent functional responses will be more closely related to practicality. Also, considering that predators are disturbed by other predators while handling or searching for prey, the Beddington-DeAngelis functional response deserves further study. Therefore, we will consider introducing ratio-dependent functional response and Beddington-DeAngelis functional response in the model.
Delay differential equations (DDEs) have a long history in modeling predator-prey systems, where delays usually represent the required reaction time, gestation period, maturation, feeding time, etc. [24,25,26,27,28,29]. In terms of mathematical modeling, delayed systems [30,31] make more sense than non-delayed systems. The introduction of time delays in a predator-prey model causes the model to exhibit more complex dynamics. Among them, the gestation delay represents the time delay that occurs when biomass is transferred from prey to predator. For example, Tripathyi et al. [5] investigated the degree of mutual interference between density-dependent predators in a model incorporating discrete gestation delays and Beddington-DeAngelis-type functional responses. They observed that when the delay parameter exceeded a certain threshold, the system experienced Hopf bifurcation and the system lost its stability. Based on this finding, they investigated two different time delay effects in a proportionally dependent predator-prey system and studied the conditions under which Hopf-bifurcation occurred when both delays were present in the model. From the above example, we can determine the importance of delay in stability analysis. Therefore, we add the time lag parameters to the model.
In this paper, we aim to study the effect of time lag parameters on the dynamics of the system. The rest of the paper is organized as follows. In the next section, we construct a predator-prey model with fear effects and multiple delays. In Section 3, we analyze the positivity and boundedness of the solutions of the system. In Section 4, we present the criterion of the existence of equilibrium points. In Section 5, we study the stability of the equilibrium point and the conditions of the occurrence of Hopf bifurcation with time lags as the bifurcation parameter. In Section 6, we use numerical simulations to verify our main analytical results. Finally, we conclude the whole paper.
Let x1(t), x2(t), and x3(t) be the population densities of prey, intermediate predator, and top predator at time t, respectively. We construct the mathematical model through the following assumptions.
(1) In the absence of predators, the prey population increases logically. However, in the presence of intermediate predators, the prey perceives fear, and fear affects the reproductive rate of the prey.
(2) Intermediate predators prey on preys based on a ratio-dependent functional response. However, the interaction between intermediate and top predators follows a Beddington-DeAngelis type functional response. Among them, the expression form of the ratio-dependent functional response is f(x1,x2)=αx1x1+mx2, and the expression form of the Beddington-DeAngelis type functional response is f(x2,x3)=θx2v+px2+qx3.
(3) Most biological processes involve delays, and here we consider two gestation delays, a time interval between the predator preying on its prey and the prey breeding offspring, denoted by τ1, τ2.
Based on these assumptions, we obtain the following model:
dx1dt=rx11+kx2−bx1−ax21−αx1x2x1+mx2,dx2dt=βx1(t−τ1)x2(t−τ1)x1(t−τ1)+mx2(t−τ1)−dx2−θx2x3v+px2+qx3,dx3dt=ηx2(t−τ2)x3(t−τ2)v+px2(t−τ2)+qx3(t−τ2)−ex3, | (2.1) |
where r,k,b,a and α are the intrinsic growth rate of prey, intermediate predator-induced fear, mortality, intraspecific competition rate and prey consumption rate, respectively. Meanwhile, β,d,θ,η and e are prey conversion rate, intermediate predator mortality and consumption rate, intermediate predator conversion rate and top predator mortality rate, respectively. Here m is the half-saturation constant of the ratio-dependent functional response, v is the saturation constant of the Beddington-DeAngelis-type functional response. p and q are the abundance of intermediate predators and disturbance of top predators, respectively. τ1 and τ2 represent gestation time delays.
The initial conditions are
x1(s)=ϕ1(s), x2(s)=ϕ2(s), x3(s)=ϕ3(s), −τ≤s≤0, | (2.2) |
where τ=max{τ1,τ2}, Φ=(ϕ1,ϕ2,ϕ3), and Φ=(ϕ1,ϕ2,ϕ3) belongs to the Banach space of continuous function Φ:[−τ,0]→R3 with
‖Φ‖=sup−τ≤s≤0(|ϕ1(s)|,|ϕ2(s)|,|ϕ3(s)|). |
From the biological perspective, we assume ϕi(s)≥0 and ϕi(0)>0, i=1,2,3.
Theorem 3.1. All solutions of system (2.1) with initial conditions (2.2) are positive for all t>0.
Proof. System (2.1) can be written in the following matrix form:
˙W=F(W), |
where W=(x1x2x3), and F(W)=(F1(W)F2(W)F3(W))=(rx11+kx2−bx1−ax21−αx1x2x1+mx2βx1(t−τ1)x2(t−τ1)x1(t−τ1)+mx2(t−τ1)−dx2−θx2x3v+px2+qx3ηx2(t−τ2)x3(t−τ2)v+px2(t−τ2)+qx3(t−τ2)−ex3) with the initial condition
W(s)=(ϕ1(s),ϕ2(s),ϕ3(s))∈C([−τ,0], R3+) and ϕ1(0),ϕ2(0),ϕ3(0)>0. |
We have easily checked in system (2.1) that whenever choosing W(s)∈R3+, then
Fi(W)|W∈R3+≥0,i=1,2,3. |
Using the Lemma 4 given in [32], we obtain that every solution W(t) of the system (2.1) with initial conditions (2.2) exists in the region R3+ and is positive and invariant for all t>0.
Theorem 3.2. The solutions of system (2.1) which starts in R3+ are uniformly bounded.
Proof. Let U(t)=x1(t−τ1−τ2)+αβx2(t−τ2)+αθβηx3(t).
Calculating the derivative of U(t) along the solution of system (2.1), we have
dUdt=dx1dt(t−τ1−τ2)+αβdx2dt(t−τ2)+αθβηdx3dt(t)=rx1(t−τ1−τ2)1+kx2(t−τ1−τ2)−bx1(t−τ1−τ2)−ax21(t−τ1−τ2)−αx1(t−τ1−τ2)x2(t−τ1−τ2)x1(t−τ1−τ2)+mx2(t−τ1−τ2)+αx1(t−τ1−τ2)x2(t−τ1−τ2)x1(t−τ1−τ2)+mx2(t−τ1−τ2)−αdβx2(t−τ2)−αθβx2(t−τ2)x3(t−τ2)v+px2(t−τ2)+qx3(t−τ2)+αdβx2(t−τ2)−αθβx2(t−τ2)x3(t−τ2)v+px2(t−τ2)+qx3(t−τ2)−αθeβηx3(t)≤rx1(t−τ1−τ2)−bx1(t−τ1−τ2)−ax21(t−τ1−τ2)−αdβx2(t−τ2)−αθeβηx3(t)≤r24a−min{b,d,e}U(t). |
Let z=min{b,d,e} and Q=r24a. Using the theory of differential inequality [33], we obtain
0<U(t)≤Qz(1−e−zt)+U(0)e−zt. |
When t⟶∞, we get 0<U≤Qz. Therefore, all the solutions of system (2.1) starting from R3+ are restricted in the region:
D={(x1,x2,x3)∈R3+:0<U≤Qz}. |
This completes the proof.
The possible equilibrium points of the model are as follows:
(1) The trivial equilibrium point E0(0,0,0) always exists.
(2) The predator free axial equilibrium point E1(r−ba,0,0) exists if r>b.
(3) The top predator free equilibrium point E2(ˉx1,ˉx2,0) where ˉx2=β−dmdˉx1 and ˉx1 is a positive root of −ac(1+cm)x21−[(1+cm)(cb+a)+c2α]x1+(r−b)(1+cm)−cα=0, where c=β−dmd. According to Descarte's rule of sign, the equation must have a unique positive root if (r−b)(1+cm)>cα. Thus, the equilibrium E2 exists if (r−b)(1+cm)>cα.
(4) The interior equilibrium point E∗(x∗1,x∗2,x∗3) exists if the following two isoclines have a positive intersection:
{f(x1,x2)=A1x21+B1x1+C1=0,g(x1,x2)=A2x22+B2x2+C2=0, |
where
x∗3=ρ1x∗2−ρ2, ρ1=η−peqe, ρ2=vq,A1=−a(1+kx2)<0,B1=r−(1+kx2)(b+amx2),C1=x2[rm−(mb+α)(1+kx2)],A2=−m[d(p+ρ1q)+θρ1]<0,B2=βx1(p+ρ1q)+d[(p+ρ1q)x1+mv−ρ2mq]−θ(ρ2x1−m),C2=βx1(v−ρ2q)−dx1(v−ρ2q)+ρ2θx1. |
As x2⟶0, then the above two isoclines become:
{f(x1,0)=−ax21+(r−b)x1=0,g(x1,0)=C2=0, |
where the number of positive roots of the f(x1,0)=0 depends on the sign of r−b, then the above two isoclines have a unique intersection on the first quadrant (the positive quadrant) if the following sufficient condition holds:
{r>b,dx2dx1=−∂f(x1,x2)/∂x1∂f(x1,x2)/∂x2<0,dx2dx1=−∂g(x1,x2)/∂x1∂g(x1,x2)/∂x2>0. |
In this section, our goal is to determine the local behavior of the system (2.1) around the equilibrium point. First, system (2.1) can be expressed as
dWdt=F(W(t),W(t−τ1),W(t−τ2)), |
where W(t)=[x1(t),x2(t),x3(t)]T.
We take E(x1,x2,x3) as an arbitrary equilibrium point of the system (2.1). Assume ˜x1(t)=x1(t)−x1, ˜x2(t)=x2(t)−x2 and ˜x3(t)=x3(t)−x3. Then, the linearized system (2.1) around the equilibrium point E is
dGdt=D1G(t)+D2G(t−τ1)+D3G(t−τ2), |
where
D1=(∂F∂W(t))E,D2=(∂F∂W(t−τ1))E,D3=(∂F∂W(t−τ2))E, |
and G(t)=[˜x1(t),˜x2(t),˜x3(t)]T. The Jacobian matrix of the system (2.1) at E is given by
J=D1+D2e−λτ1+D3e−λτ2. |
Then
J=[a11a120b21e−λτ1a22+b22e−λτ1a230c32e−λτ2a33+c33e−λτ2], |
where
a11=r1+kx2−b−2ax1−αmx22(x1+mx2)2,a12=−krx1(1+kx2)2−αx21(x1+mx2)2,a22=−d−θx3(v+qx3)(v+px2+qx3)2,a23=−θx2(v+px2)(v+px2+qx3)2,a33=−e,b21=βmx22(x1+mx2)2,b22=βx21(x1+mx2)2,c32=ηx3(v+qx3)(v+px2+qx3)2,c33=ηx2(v+px2)(v+px2+qx3)2. |
According to the Jacobian matrix, we analyze the stability of the equilibrium point E0, E1 and E2, and obtain the following theorems:
Theorem 5.1. The trivial equilibrium point E0(0,0,0) is locally asympototically stable if r<b and becomes unstable if r>b.
Theorem 5.2. When τ1≥0, the predator free axial equilibrium point E1(r−ba,0,0) is locally asympototically stable if β<d; otherwise it is unstable.
Theorem 5.3. For system (2.1), according to the value of τ1, there are the following two cases:
(1) When τ1=0, the top predator free equilibrium point E2=(ˉx1,ˉx2,0) is locally asympototically stable for any τ2≥0 if ˉx2<veη−pe, aˉx1+d+βˉx21(ˉx1+mˉx2)2>αˉx1ˉx2(ˉx1+mˉx2)2 and (−aˉx1+αˉx1ˉx2(ˉx1+mˉx2)2)(βˉx21(ˉx1+mˉx2)2−d)+βmˉx22(ˉx1+mˉx2)2(krˉx1(1+kˉx2)2+αˉx21(ˉx1+mˉx2)2)>0; otherwise unstable.
(2) When τ1>0 and ˉx2<veη−pe, if H22−N22<0 and 2φ20+H1−N21−2H2>0 hold, the super predator free equilibrium point E2=(ˉx1,ˉx2,0) is locally asympototically stable for τ1∈[0,^τ1) and unstable for τ1>^τ1. Moreover, the system (2.1) undergoes a Hopf bifurcation when τ1=^τ1, where
^τ1=minj=0,1,2...τ1j=minj=0,1,2...[1φ0arccos[N2(φ20−H2)−H1N1φ20N21φ20+N22]+2πjφ0], |
and φ0 is defined in the proof.
Next, we study the local stability of E∗=(x∗1,x∗2,x∗3). The Jacobian matrix at E∗ is
JE∗=[A11A120B21e−λτ1A22+B22e−λτ1A230C32e−λτ2A33+C33e−λτ2], |
where
A11=−ax∗1+αx∗1x∗2(x∗1+mx∗2)2,A12=−krx∗1(1+kx∗2)2−α(x∗1)2(x∗1+mx∗2)2,A22=−d−θx∗3(v+qx∗3)(v+px∗2+qx∗3)2,A23=−θx∗2(v+px∗2)(v+px∗2+qx∗3)2,A33=−e,B21=βm(x∗2)2(x∗1+mx∗2)2,B22=β(x∗1)2(x∗1+mx∗2)2,C32=ηx∗3(v+qx∗3)(v+px∗2+qx∗3)2,C33=ηx∗2(v+px∗2)(v+px∗2+qx∗3)2. |
The corresponding characteristic equation of the above Jacobian matrix is
λ3+P1λ2+P2λ+P3+e−λτ1(Q1λ2+Q2λ+Q3)+e−λτ2(R1λ2+R2λ+R3)+e−λ(τ1+τ2)(S1λ+S2)=0, | (5.1) |
where
P1=−(A11+A22+A33),P2=A11A22+A22A33+A33A11,P3=−A11A22A33,Q1=−B22,Q2=A11B22+A33B22−A12B21,Q3=A12A33B21−A11A33B22,R1=−C33,R2=A11C33+A22C33−A23C32,R3=A11A23C32−A11A22C33,S1=B22C33,S2=A12B21C33−A11B22C33. |
We divide the following analysis into five cases according to the different ranges of values of the two time delays. Based on the characteristic equation (5.1), we analyze the stability of the interior equilibrium point E∗ and the conditions for the occurrence of Hopf-bifurcation, and obtain the following theorems.
Case 1: τ1=τ2=0.
Theorem 5.4. In absence of both delays, the interior equilibrium point E∗ is locally asymptotically stable if P1+Q1+R1>0, P3+Q3+R3+S2>0 and (P1+Q1+R1)(P2+Q2+R2+S1)>P3+Q3+R3+S2.
Case 2: τ1>0,τ2=0.
Theorem 5.5. For system (2.1), when τ1>0,τ2=0, if (P3+R3)2<(Q3+S2)2 holds, the interior equilibrium point E∗ is locally asymptotically stable for all τ1∈[0,τ10) and unstable for τ1>τ10. Furthermore, when 3φ40+(2E211−4E12−2F211)φ20+(E212−2E11E13+2F11F13−F212)>0, then Hopf-bifurcation occurs at τ1=τ10, where
τ10=mink,jτ(j)1k=mink,j[1φkarccos(E11φ2k−E13)(F13−F11φ2k)+F12φk(φ3k−E12φk)(F13−F11φ2k)2+F212φ2k+2πjφk],k=1,2,3, j=0,1,2..., |
and φk are defined in the proof.
Case 3: τ1=0,τ2>0.
Theorem 5.6. For system (2.1), when τ1=0,τ2>0, if (P3+Q3)2<(R3+S2)2 holds, the interior equilibrium point E∗ is locally asymptotically stable for all τ2∈[0,τ20) and unstable for τ2>τ20. Furthermore, when 3φ40+(2E221−4E22−2F221)φ20+(E222−2E21E23+2F21F23−F222)>0, then Hopf-bifurcation occurs at τ2=τ20, where
τ20=mink,jτ(j)2k=mink,j[1φkarccos(E21φ2k−E23)(F23−F21φ2k)+F22φk(φ3k−E22φk)(F23−F21φ2k)2+F222φ2k+2πjφk],k=1,2,3, j=0,1,2..., |
and φk are defined in the proof.
Case 4: τ1>0,τ2∈(0,τ20).
Theorem 5.7. For system (2.1), when τ1>0,τ2∈(0,τ20), if (P3+R3)2<(Q3+S2)2 holds, the interior equilibrium point E∗ is locally asymptotically stable for all τ1∈[0,ˆτ10) and unstable for τ1>ˆτ10. Furthermore, when M1M3+M2M4>0, then Hopf-bifurcation occurs at τ1=ˆτ10, where
ˆτ10=mink,jτ(j)1k=mink,j1φkarccosΨ1Ψ3+Ψ2Ψ4(Ψ1)2+(Ψ2)2+2πjφk,k=1,2,...,l;j=0,1,2,.... |
and φk are defined in the proof.
Case 5: τ1∈(0,τ10),τ2>0.
Theorem 5.8. For system (2.1), when τ1∈(0,τ10),τ2>0, if (P3+Q3)2<(R3+S2)2 holds, the interior equilibrium point E∗ is locally asymptotically stable for all τ2∈[0,ˆτ20) and unstable for τ2>ˆτ20. Furthermore, when M5M7+M6M8>0, then Hopf-bifurcation occurs at τ2=ˆτ20, where
ˆτ20=mink,jτ(j)2k=mink,j1φkarccosΨ5Ψ7+Ψ6Ψ8(Ψ5)2+(Ψ6)2+2πjφk,k=1,2,...,l;j=0,1,2,..., |
and φk are defined in the proof, where
M5=φ40−P2φ20−Q2φ0cos(φ0τ01)+(−Q1φ20+Q3)sin(φ0τ01),M6=−P1φ30+P3φ0+(−Q1φ20+Q3)cos(φ0τ01)+Q2φ0sin(φ0τ01),M7=3φ30−P2−2Q1φ0sin(φ0τ01)−Q2cos(φ0τ01)−2R1φ0sin(φ0τ2)−R2cos(φ0τ2)+(S2τ01−S1)cos(φ0(τ01+τ2))+S1φ0τ01sin(φ0(τ01+τ2)),M8=−2P1φ0−2Q1φ0cos(φ0τ01)+Q2sin(φ0τ01)−2R1φ0cos(φ0τ2)+R2sin(φ0τ2)+S1φ0τ01cos(φ0(τ01+τ2))+(S1−S2τ01)sin(φ0(τ01+τ2)),Ψ5=−R1φ2k+R3+S2cos(φkτ01)+S1φksin(φkτ01),Ψ6=R2φk−S2sin(φkτ01)+S1φkcos(φkτ01),Ψ7=P1φ2k−P3+(Q1φ2k−Q3)cos(φkτ01)−Q2φksin(φkτ01),Ψ8=φ3k−P2φk−Q2φkcos(φkτ01)+(Q3−Q1φ2k)sin(φkτ01). |
Remark 1. The proofs of above theorems are given in the Appendix.
In this section, we use MATLAB R2020a to perform some numerical simulations to illustrate the results of the analysis in the previous sections and plot the corresponding graphs of system (2.1). Since the problem is not a species-specific case study and no real data are available, some hypothetical data are taken here for the simulation.
We make the parameters values r=4.5, k=6, b=0.2, a=0.06, α=1, m=1, β=1.3, d=0.03, θ=4, v=3.1, p=1.14, q=1.13, η=2.2, e=0.2, and the initial value is (ϕ1(s),ϕ2(s),ϕ3(s))=(2,2,2), s∈[−τ,0]. According to the above parameters, we can obtain all equilibrium points, respectively E0(0,0,0), E1(71.67,0,0), E2(0.0111,0.4699,0) and E∗(14.5,1.73,0.51). In the following, we mainly show several cases of interior equilibrium points.
For τ1=0,τ2=0, the interior equilibrium point E∗ is locally asymptotically stable (see Figure 1).
For τ1>0,τ2=0, we plotted the Hopf-bifurcation diagram Figure 2 concerning τ1. If we change the value of τ1 from 0 to 14, then the system undergoes Hopf-bifurcation. Hence, the interior equilibrium E∗ is locally asymptotically stable for τ1=1<τ10≈10.88 (see Figure 3) and unstable for τ1=13>τ10≈10.88 (see Figure 4).
For τ1=0,τ2>0, if we continuously increase the value of τ2, we can find a critical value of τ20, namely τ20≈0.11 for which system (2.1) undergoes Hopf-bifurcation (see Figure 5. Hence, E∗ is locally asymptotically stable for τ2=0.1<τ20≈0.11 (see Figure 6) and unstable for τ2=0.2>τ20≈0.11 (see Figure 7).
For τ1>0, fixing τ2=0.1∈(0,τ20), we get ˆτ10≈10.28, so when τ1=2<ˆτ10, E∗ is locally asymptotically stable (see Figure 8), when τ1=12>ˆτ10, E∗ is unstable (see Figure 9). Then Hopf-bifurcation occurs at ˆτ10≈10.28 (see Figure 10).
For τ2>0, fixing τ1=1∈(0,τ10), we get ˆτ20≈0.96, so when τ2=0.1<ˆτ20, E∗ is locally asymptotically stable (see Figure 11), when τ2=1.1>ˆτ20, E∗ is unstable (see Figure 12). Then Hopf bifurcation occurs at ˆτ20≈0.96 (see Figure 13).
Based on above analysis, we summarize the dynamics of the interior equilibrium E∗ of system (2.1) in Table 1.
Values of delay | Thresholds of delay | Simulation results | Explanations of simulation |
τ2=0 | τ10≈10.88 | Figure 2 | When τ2=0, E∗ is locally asymptotically stable for τ1<10.88 and unstable for τ1>10.88. |
τ1=0 | τ20≈0.11 | Figure 5 | When τ1=0, E∗ is locally asymptotically stable for τ2<0.11 and unstable for τ2>0.11. |
τ2=0.1∈(0,τ20) | ˆτ10≈10.28 | Figure 10 | When τ2=0.1∈(0,τ20), E∗ is locally asymptotically stable for τ1<10.28 and unstable for τ1>10.28. |
τ1=1∈(0,τ10) | ˆτ20≈0.96 | Figure 13 | When τ1=1∈(0,τ10), E∗ is locally asymptotically stable for τ2<0.96 and unstable for τ2>0.96. |
Finally we investigate how the fear of predators affect the population dynamics. By increasing the fear values and taking k=6,9,15 and 20 respectively, we plot the time series graphs of each species (see Figures 14–16).
(1) The effect of fear on prey. In Figure 14, we observe that fear inhibits prey reproduction and excessive fear leads to prey extinction.
(2) The effect of fear on intermediate predators. In Figure 15, we observe that an increase in the value of fear leads to a shorter duration of intermediate predator turbulence, but little change in the number of species, and too much fear leads to a decrease and stabilization of the number of species.
(3) The effect of fear on the top predator. In Figure 16, an increase in fear leads to a decrease in the turbulence time of the highest predator and a decrease in the number of species, and too much fear leads the highest predator to extinction.
In this paper, we analyze the dynamics of a three-chain model containing indirect predation (fear of predator) and two delays, where the three species are the prey, intermediate predator and top predator, and the two delays represent the gestation delays of the intermediate and top predator, respectively.
First, we investigate the positivity and boundedness of this system, where boundedness can be seen as a natural limit to expansion due to limited resources and positivity implies that species persist. Next, we study the effect of delay on the stability of the model by varying the delay parameters τ1 and τ2. According to the theorem in Section 5, we obtain the conditions when the system is stable. By selecting appropriate parameter values, we draw the Hopf-bifurcation diagrams (Figures 2, 5, 10 and 13), that is, if τ increases continuously, a threshold τ0 will be obtained. If τ<τ0, the system is stable; otherwise, the system becomes unstable, that is, a Hopf-bifurcation occurs at τ=τ0. These observations confirm the important role of delay in the system. Through numerical simulations, we find that fear may reduce the number of species and even lead to extinction of species. Thus, fear has an important influence on the dynamics of predator-prey systems.
In nature, in addition to fear, other factors such as refuge, additional food, and human capture influence the predator-prey system. As a part of future research on the model in this paper, the inclusion of refuge and human capture could be considered to make the model more realistic. Moreover, for the discrete-time model [6,7] and the infectious disease model [8], we believe there will be some intresting findings. All these will be left to our future work.
The authors would like to thank the referees for their suggestions and valuable comments, which have greatly impoved the presentation of this paper.
This work is funded by NSF of China (No. 11861027).
The authors declare that they have no competing interests in this paper.
Here, we give the proofs of Theorems 5.1 to 5.8.
Proof of Theorem 5.1:
Proof. Based on the eigenvalues less than 0, we obtain that E0 is locally asymptotically stable when r<b.
Proof of Theorem 5.2:
Proof. By calculating the Jacobian matrix at the equilibrium E1, we obtain
JE1=[−(r−b)−krr−ba−α00−d+βe−λτ1000−e]. |
Obviously, we get λ1=−(r−b)<0, λ2=−e<0, and the other eigenvalue is the root of the following equation:
−d+βe−λτ1−λ=0. | (A.1) |
In the absence of delay, i.e. when τ1=0, E1 is locally asymptotically stable if β<d.
When τ1>0 and β<d, we assume that λ=ξ+iφ(ξ,φ∈R) is the root of Eq (A.1). Then substituting the value of λ into Eq (A.1) and separating the real and imaginary parts, we obtain
βe−ξτ1cos(φτ1)−d−ξ=0, | (A.2) |
βe−ξτ1+φ=0. | (A.3) |
To eliminate τ1, we square and add the Eqs (A.2) and (A.3), and obtain the following equation
φ2=β2e−2ξτ1−(d−ξ)2, |
i.e., φ=±√(βe−ξτ1+d+ξ)(βe−ξτ1−d−ξ). |
Let ξ≥0, then (βe−ξτ1−d−ξ)≤(β−d−ξ)<0. Thus, if ξ≥0, no such real φ exists, which contradicts the previous assumption φ∈R. So ξ<0, i.e., the Eq (A.1) contains a negative real root and an imaginary root with a negative real part (if exists).
Therefore, E1 is locally asymptotically stable if β<d.
Proof of Theorem 5.3:
Proof. The Jacobian matrix at E2 is:
JE2=[−aˉx1+αˉx1ˉx2(ˉx1+mˉx2)2−krˉx1(1+kˉx2)2−αˉx21(ˉx1+mˉx2)20βmˉx22(ˉx1+mˉx2)2e−λτ1−d+βˉx21(ˉx1+mˉx2)2e−λτ1−θˉx2v+pˉx200−e+ηˉx2v+pˉx2e−λτ2]. |
The characteristic equation of above matrix JE2 is given by det(JE2−λI)=0.
One root of det(JE2−λI)=0 is given by −e+ηˉx2v+pˉx2e−λτ2−λ=0. The analysis process is similar to Eq (A.1). Therefore, if −e+ηˉx2v+pˉx2<0, i.e., ˉx2<veη−pe, then the equation contains negative real roots and imaginary roots with negative real parts (if exists).
The other two eigenvalues are the root of the quadratic equation
λ2+H1λ+H2+e−λτ1(N1λ+N2)=0, | (A.4) |
where
H1=aˉx1+d−αˉx1ˉx2(ˉx1+mˉx2)2,H2=−d(−aˉx1+αˉx1ˉx2(ˉx1+mˉx2)2),N1=βˉx21(ˉx1+mˉx2)2,N2=βˉx21(ˉx1+mˉx2)2(−aˉx1+αˉx1ˉx2(ˉx1+mˉx2)2)+βmˉx22(ˉx1+mˉx2)2(krˉx1(1+kˉx2)2+αˉx21(ˉx1+mˉx2)2). |
Case 1: τ1=0, then Eq (A.4) becomes
λ2+(H1+N1)λ+H2+N2=0. | (A.5) |
According to Routh-Hurwitz criteria, both roots of the Eq (A.5) are negative real parts if (H2+N2)>0 and (H1+N1)>0, i.e. (−aˉx1+αˉx1ˉx2(ˉx1+mˉx2)2)(βˉx21(ˉx1+mˉx2)2−d)+βmˉx22(ˉx1+mˉx2)2(krˉx1(1+kˉx2)2+αˉx21(ˉx1+mˉx2)2)>0 and aˉx1+d+βˉx21(ˉx1+mˉx2)2>αˉx1ˉx2(ˉx1+mˉx2)2.
Case 2: τ1>0 and ˉx2<veη−pe. By substituting λ=ξ+iφ into Eq (A.4) and separating the real and imaginary parts, we have
ξ2−φ2+H1ξ+H2+e−ξτ1[(N1ξ+N1N2)cos(φτ1)+N1φsin(φτ1)]=0, | (A.6) |
2ξφ+H1φ+e−ξτ1[N1φcos(φτ1)−(N1ξ+N1N2)sin(φτ1)]=0. | (A.7) |
Putting ξ=0, Eqs (A.6) and (A.7) become
N2cos(φτ1)+N1sin(φτ1)=φ2−H2, | (A.8) |
N1φcos(φτ1)−N2sin(φτ1)=−H1φ. | (A.9) |
By squaring and adding, we obtain
φ4+(H21−2H2−N21)φ2+H22−N22=0. | (A.10) |
Using Descarte's rule of sign, the equation has at least one positive root φ0 if H22−N22<0. By calculating (A.8) and (A.9), we get
τ1j=1φ0arccos[N2(φ20−H2)−H1N1φ20N21φ20+N22]+2πjφ0j=0,1,2.... | (A.11) |
Let ^τ1=minj=0,1,2...τ1j.
Now the transversality condition Re(dλdτ1)−1λ=iφ0>0 will be verified.
By differentiating Eq (A.4) with respect to τ1, we obtain
(dλdτ1)−1=−2λ+H1λ(λ2+H1λ+H2)+N1λ(N1λ+N2)−τ1λ, |
which leads to
Re(dλdτ1)−1λ=iφ0=Re(−2λ+H1λ(λ2+H1λ+H2))λ=iφ0+Re(N1λ(N1λ+N2))λ=iφ0=H21+2(φ20−H2)H21φ20+(φ20−H2)2−N21N21φ20+N22. |
From (A.10), we have
H21φ20+(φ20−H2)2=N21φ20+N22. |
Then
Re(dλdτ1)−1λ=iφ0=H21−N21+2(φ20−H2)H21φ20+(φ20−H2)2=2φ20+H1−N21−2H2H21φ20+(φ20−H2)2. |
When 2φ20+H1−N21−2H2>0, Re(dλdτ1)−1λ=iφ0>0. Therefore, the transversality condition is satisfied and a Hopf-bifurcation occurs at E2 for τ1=^τ1.
Proof of Theorem 5.4:
Proof. The characteristic (5.1) becomes
λ3+(P1+Q1+R1)λ2+(P2+Q2+R2+S1)λ+P3+Q3+R3+S2=0. | (A.12) |
Therefore, by Routh-Hurwitz Criteria, we obtain all the roots of (A.12) have negative real part, if P1+Q1+R1>0, P3+Q3+R3+S2>0 and (P1+Q1+R1)(P2+Q2+R2+S1)>P3+Q3+R3+S2. This means that the interior equilibrium E∗ is locally asymptotically stable.
Proof of Theorem 5.5:
Proof. The characteristic (5.1) becomes
λ3+E11λ2+E12λ+E13+e−λτ1(F11λ2+F12λ+F13)=0, | (A.13) |
where E11=P1+R1, E12=P2+R2, E13=P3+R3, F11=Q1, F12=Q2+S1, F13=Q3+S2. Let λ=ξ+iφ be the root of (A.13). Then we obtian
ξ3−3ξφ2+E11ξ2−E11φ2+E12ξ+E13+e−ξτ1[(F11ξ2−F11φ2+F12ξ+F13)cos(φτ1)+(2F11ξφ+F12φ)sin(φτ1)]=0,−φ3+3ξ2φ+2E11ξφ+E12φ+e−ξτ1[(2F11ξφ+F12φ)cos(φτ1)−(F11ξ2−F11φ2+F12ξ+F13)sin(φτ1)]=0. |
A necessary condition for changing the stability of the equilibrium point E∗ is that the real part of the root of the Eq (A.13) changes the sign. To find the stable switching point, we consider ξ=0, then we have
(F13−F11φ2)cos(φτ1)+F12φsin(φτ1)=E11φ2−E13, | (A.14) |
F12φcos(φτ1)+(F11φ2−F13)sin(φτ1)=φ3−E12φ. | (A.15) |
By squaring and adding (A.14) and (A.15), we obtain
φ6+K11φ4+K12φ2+K13=0, | (A.16) |
where K11=E211−2E12−F11, K12=E212−2E11E13+2F11F13−F212, K13=E213−F213.
Let φ2=σ, we get
h1(σ)=σ3+K11σ2+K12σ+K13=0. | (A.17) |
Then, h1(0)=K13=E213−F213=(P3+R3)2−(Q3+S2)2 and limσ→∞h1(σ)=+∞. We assume that h1(0)<0⇒(P3+R3)2<(Q3+S2)2. According to Descarte's rule of sign, the Eq (A.17) has at least one positive root and can have at most three positive roots.
Without loss of generality, we assume that it has three positive roots, denoted by φk=√σk and k=1,2,3. By calculating (A.14) and (A.15), we obtain
τj1k=1φkarccos(E11φ2k−E13)(F13−F11φ2k)+F12φk(φ3k−E12φk)(F13−F11φ2k)2+F212φ2k+2πjφk |
where k=1,2,3 and j=0,1,2,.... Let τ10=mink,jτ(j)1k, k=1,2,3, j=0,1,2....
Now the transversality condition Re(dλdτ1)−1λ=iφ0>0 will be verified.
By differentiating Eq (A.13) with respect to τ1, we obtain
(dλdτ1)−1=−3λ2+2E11λ+E12λ(λ3+E11λ2+E12λ+E13)+2F11λ+F12λ(F11λ2+F12λ+F13)−τ1λ, |
which leads to
Re(dλdτ1)−1λ=iφ0=3φ40+(2E211−4E12)φ20+(E212−2E11E13)φ60+(E211−2E12)φ40+(E212−2E11E13)φ20+E213+−2F211φ20+(2F11F13−F212)F211φ40+(F212−2F11F13)φ20+F213. |
From (A.16), we have
φ60+(E211−2E12)φ40+(E212−2E11E13)φ20+E213=F211φ40+(F212−2F11F13)φ20+F213. |
Then
Re(dλdτ1)−1λ=iφ0=3φ40+(2E211−4E12−2F211)φ20+(E212−2E11E13+2F11F13−F212)φ60+(E211−2E12)φ40+(E212−2E11E13)φ20+E213. |
When 3φ40+(2E211−4E12−2F211)φ20+(E212−2E11E13+2F11F13−F212)>0, Re(dλdτ1)−1λ=iφ0>0. Therefore, the transversality condition is satisfied and a Hopf-bifurcation occurs around when τ1 passes through the critical value τ10.
Proof of Theorem 5.6:
Proof. The characteristic (5.1) becomes
λ3+E21λ2+E22λ+E23+e−λτ2(F21λ2+F22λ+F23)=0, |
where E21=P1+Q1, E22=P2+Q2, E23=P3+Q3, F21=Q1, F22=R2+S1, F23=R3+S2.
The calculation process is similar to Theorem 5.5.
Proof of Theorem 5.7:
Proof. We consider that τ2=τ02 in its stable interval, with τ1 as the parameter.
Let λ=ξ+iφ be the root of (5.1), we get
ξ3−3ξφ2+P1ξ2−P1φ2+P2ξ+P3+e−ξτ1[(Q1ξ2−Q1φ2+Q2ξ+Q3)cos(φτ1)+(2Q1ξφ+Q2φ)sin(φτ1)]+e−ξτ02[(R1ξ2−R1φ2+R2ξ+R3)cos(φτ02)+(2R1ξφ+R2φ)sin(φτ02)]+e−ξ(τ1+τ02)[(S1ξ+S2)cos(φ(τ1+τ02))+S1φsin(φ(τ1+τ02))]=0,3ξ2φ−φ3+2P1ξφ+P2φ+e−ξτ1[(2Q1ξφ+Q2φ)cos(φτ1)−(Q1ξ2−Q1φ2+Q2ξ+Q3)×sin(φτ1)]+e−ξτ02[(2R1ξφ+R2φ)cos(φτ02)−(R1ξ2−R1φ2+R2ξ+R3)sin(φτ02)]+e−ξ(τ1+τ02)[S1φcos(φ(τ1+τ02))+(S1ξ+S2)sin(φ(τ1+τ02))]=0. |
A necessary condition for the stability change of the equilibrium point E∗ is that the characteristic Eq (5.1) has a pair of purely imaginary roots, i.e., ξ=0. Let ξ=0, and we obtain
[−Q1(φ)2+Q3+S2cos(φτ02)+S1φsin(φτ02)]cos(φτ1)+[Q2φ−S2sin(φτ02)+S1cos(φτ02)]sin(φτ1)=P1(φ)3−P3+(R1φ−R3)cos(φτ02)−R2φsin(φτ02), | (A.18) |
[Q2φ−S2sin(φτ02)+S1cos(φτ02)]cos(φτ1)+[Q1(φ)2−Q3−S2cos(φτ02)−S1φsin(φτ02)]sin(φτ1)=(φ)3−P2φ−R2φ∗1cos(φτ02)+(R3−R1φ)sin(φτ02). | (A.19) |
Squaring and adding the two equations, and we have
φ6+K21φ4+K22φ2+K23+K24sin(φτ02)+K25cos(φτ02)=0, | (A.20) |
where
K21=P21+R21−Q21−2P2,K23=P23+R23−Q23−S22,K22=P22+R22−Q22−S21−2P1P3−2R1R3+2Q1Q3,K24=−2R1φ5−2P1R2φ4+2(R3+P2R1+Q1S1)φ3+2(P3R2−P2R3+Q2S2−Q3S1),K25=2(−R2+P1R2)φ4−2(P1R3+P3R1−P2R2+Q2S1−Q1S2)φ2+2P3R3−2Q3S2. |
Denote
h2(φ)=φ6+K21φ4+K22φ2+K23+K24sin(φτ02)+K25cos(φτ02). | (A.21) |
Obviously, h2(0)=K23+K25∣φ=0=(P3+R3)2−(Q3+S2)2 and limφ→∞h2(φ)=+∞. We assume that h2(0)<0⇒(P3+R3)2<(Q3+S2)2. Then, the Eq (A.21) has at least one positive root.
Assume that Eq (A.21) has finite positive roots φk(k=1,2,...,l). The critical value
τ(j)1k=1φkarccosΨ1Ψ3+Ψ2Ψ4(Ψ1)2+(Ψ2)2+2πjφk,k=1,2,...,l;j=0,1,2,.... |
where
Ψ1=−Q1φ2k+Q3+S2cos(φkτ02)+S1φksin(φkτ02),Ψ2=Q2φk−S2sin(φkτ02)+S1φkcos(φkτ02),Ψ3=P1φ2k−P3+(R1φ2k−R3)cos(φkτ02)−R2φksin(φkτ02),Ψ4=φ3k−P2φk−R2φkcos(φkτ02)+(R3−R1φ2k)sin(φkτ02). |
Let ˆτ10=mink,jτ(j)1k, k=1,2,...,l;j=0,1,2,....
Next, we differentiate both sides of (5.1) concerning τ1 to verify the transversality condition.
Taking the derivative of λ with respect to τ1 in (5.1) and substituting λ=iφ0, we obtain
Re(dλdτ1)−1λ=iφ0=M1M3+M2M4M21+M22, |
where
M1=φ40−P2φ20−R2φ0cos(φ0τ02)+(−R1φ20+R3)sin(φ0τ02),M2=−P1φ30+P3φ0+(−R1φ20+R3)cos(φ0τ02)+R2φ0sin(φ0τ02),M3=3φ30−P2−2Q1φ0sin(φ0τ1)−Q2cos(φ0τ1)−2R1φ0sin(φ0τ02)−R2cos(φ0τ02)+(S2τ02−S1)cos(φ0(τ1+τ02))+S1φ0τ02sin(φ0(τ1+τ02)),M4=−2P1φ0−2Q1φ0cos(φ0τ1)+Q2sin(φ0τ1)−2R1φ0cos(φ0τ02)+R2sin(φ0τ02)+S1φ0τ02cos(φ0(τ1+τ02))+(S1−S2τ02)sin(φ0(τ1+τ02)). |
If M1M3+M2M4>0 holds, then Re(dλdτ1)−1λ=iφ0>0. Therefore, the transversality condition is satisfied and a Hopf-bifurcation occurs at E∗ for τ1=ˆτ10.
Proof of Theorem 5.8:
Proof. We consider that τ1=τ01 in its stable interval, with τ2 as the parameter.
The calculation process is similar to Theorem 5.7.
[1] |
A. Das, G. P. Samanta, A prey-predator model with refuge for prey and additional food for predator in a fluctuating environment, Physica A, 538 (2019), 427–450. https://doi.org/10.1016/j.physa.2019.122844 doi: 10.1016/j.physa.2019.122844
![]() |
[2] |
D. Pal, G. S. Mahaptra, G. P. Samanta, Optimal harvesting of prey-predator system with interval biological parameters: A bioeconomic model, Math. Biosci., 241 (2013), 181–187. https://doi.org/10.1016/j.mbs.2012.11.007 doi: 10.1016/j.mbs.2012.11.007
![]() |
[3] |
S. Sharma, G. P. Samanta, A Leslie-Gower predator-prey model with disease in prey incorporating a prey refuge, Chaos Soliton. Fract., 70 (2015), 69–84. https://doi.org/10.1016/j.chaos.2014.11.010 doi: 10.1016/j.chaos.2014.11.010
![]() |
[4] |
G. Tang, S. Tang, R. A. Cheke, Global analysis of a holling type Ⅱ predator-prey model with a constant prey refuge, Nonlinear Dynam., 76 (2014), 635–647. https://doi.org/10.1007/s11071-013-1157-4 doi: 10.1007/s11071-013-1157-4
![]() |
[5] |
J. P. Tripathi, S. Abbas, M. Thakur, A density dependent delayed predator-prey model with Beddington-DeAngelis type function response incorporating a prey refuge, Commun. Nonlinear Sci. Numer. Simul., 22 (2015), 427–450. https://doi.org/10.1016/j.cnsns.2014.08.018 doi: 10.1016/j.cnsns.2014.08.018
![]() |
[6] |
A. Q. Khan, S. A. H. Bukhari, M. B. Almatrafi, Global dynamics, Neimark-Sacker bifurcation and hybrid control in a Leslie's prey-predator model, Alex. Eng. J., 61 (2022), 11391–11404. https://doi.org/10.1016/j.aej.2022.04.042 doi: 10.1016/j.aej.2022.04.042
![]() |
[7] |
A. Q. Khan, F. Nazir, M. B. Almatrafi, Bifurcation analysis of a discrete Phytoplankton–Zooplankton model with linear predational response function and toxic substance distribution, Int. J. Biomath., 16 (2022). https://doi.org/10.1142/S1793524522500954 doi: 10.1142/S1793524522500954
![]() |
[8] |
A. Q. Khan, M. Tasneem, M. B. Almatrafi, Discrete-time COVID-19 epidemic model with bifurcation and control, Alex. Eng. J., 19 (2022), 1944–1969. https://doi.org/10.3934/mbe.2022092 doi: 10.3934/mbe.2022092
![]() |
[9] |
N. Mondal, H. Alrabaiah, D. Barman, J. Roy, S. Alam, Influence of predator incited fear and interference competition in the dynamics of prey-predator system where the prey species are protected in a reserved area, Ecol. Environ. Conserv., 28 (2022), 831–852. http://doi.org/10.53550/EEC.2022.v28i02.040 doi: 10.53550/EEC.2022.v28i02.040
![]() |
[10] |
D. Mukherjee, Role of fear in predator-prey system with intraspecific competition, Math. Comput. Simul., 177 (2020), 263–275. https://doi.org/10.1016/j.matcom.2020.04.025 doi: 10.1016/j.matcom.2020.04.025
![]() |
[11] |
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. Simul., 200 (2022), 65–77. https://doi.org/10.1016/j.matcom.2022.04.013 doi: 10.1016/j.matcom.2022.04.013
![]() |
[12] |
L. Y. Zanette1, A. F. White, M. C. Allen, M. Clinchy, Perceived predation risk reduces the number of offspring songbirds produce per year, Science, 334 (2011), 1398–1401. https://doi.org/10.1126/science.1210908 doi: 10.1126/science.1210908
![]() |
[13] |
S. Creel, D. Christianson, S. Liley, J. A. Winnie, Predation risk affects reproductive physiology and demography of elk, Science, 315 (2007), 960. https://doi.org/10.1126/science.1135918 doi: 10.1126/science.1135918
![]() |
[14] |
J. P. Suraci, M. Clinchy, L. M. Dill, D. Roberts, L. Y. Zanette, Fear of large carnivores causes a trophic 585 cascade, Nat. Commun., 7 (2016), 10698. https://doi.org/10.1038/ncomms10698 doi: 10.1038/ncomms10698
![]() |
[15] |
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
![]() |
[16] |
C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the european pine sawfly, Canadian Entomol., 9 (1959), 293–320. https://doi.org/10.4039/Ent91293-5 doi: 10.4039/Ent91293-5
![]() |
[17] |
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
![]() |
[18] |
C. S. Holling, Some characteristics of simple types of predation and parasitism, Canadian Entomol., 91 (1959), 385. https://doi.org/10.4039/Ent91385-7 doi: 10.4039/Ent91385-7
![]() |
[19] |
Z. Liu, R. Yuan, Stability and bifurcation in a delayed predator-prey system with Beddington-DeAngelis functional response, J. Math. Anal. Appl., 296 (2004), 521–537. https://doi.org/10.1016/j.jmaa.2004.04.051 doi: 10.1016/j.jmaa.2004.04.051
![]() |
[20] |
X. Shi, X. Zhou, X. Song, Analysis of a stage-structured predator-prey model with Crowley-Martin function, J. Appl. Math. Comput., 36 (2011), 459–472. https://doi.org/10.1007/s12190-010-0413-8 doi: 10.1007/s12190-010-0413-8
![]() |
[21] |
J. P. Tripathi, S. Tyagi, S. Abbas, Global analysis of a delayed density dependent predator-prey model with Crowley-Martin functional response, Commun. Nonlinear Sci. Numer. Simul., 30 (2016), 45–69. https://doi.org/10.1016/j.cnsns.2015.06.008 doi: 10.1016/j.cnsns.2015.06.008
![]() |
[22] |
R. Chinnathambi, F. A. Rihan, Stability of fractional-order prey-predator system with time-delay and monod-haldane functional response, Nonlinear Dynam., 92 (2018), 1637–1648. https://doi.org/10.1007/s11071-018-4151-z doi: 10.1007/s11071-018-4151-z
![]() |
[23] |
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. https://doi.org/10.1016/j.biosystems.2020.104176 doi: 10.1016/j.biosystems.2020.104176
![]() |
[24] |
M. Banerjee, Y. Takeuchi, Maturation delay for the predators can enhance stable coexistence for a class of prey-predator models, J. Theor. Biol., 412 (2017), 154–171. https://doi.org/10.1016/j.jtbl.2016.10.016 doi: 10.1016/j.jtbl.2016.10.016
![]() |
[25] |
B. Dubey, A. Kumar, A. P. Maiti, Global stability and Hopf-bifurcation of prey-predator system with two discrete delays including habitat complexity and prey refuge, Commun. Nonlinear Sci. Numer. Simul., 67 (2019), 528–554. https://doi.org/10.1016/j.cnsns.2018.07.019 doi: 10.1016/j.cnsns.2018.07.019
![]() |
[26] |
J. Zhang, Bifurcation analysis of a modified Holling-Tanner predator-prey model with time delay, Appl. Math. Model., 36 (2012), 1219–1231. https://doi.org/10.1016/j.apm.2011.07.071 doi: 10.1016/j.apm.2011.07.071
![]() |
[27] |
D. Hu, Y. Li, M. Liu, Y. Bai, Stability and Hopf bifurcation for a delayed predator-prey model with stage structure for prey and ivlev-type functional response, Nonlinear Dynam., 99 (2020), 3323–3350. https://doi.org/10.1007/s11071-020-05467-z doi: 10.1007/s11071-020-05467-z
![]() |
[28] |
Y. Bai, Y. Li, Stability and Hopf bifurcation for a stage-structured predator-prey model incorporating refuge for prey and additional food for predator, Adv. Differ. Equ., 1 (2019), 1–20. https://doi.org/10.1186/s13662-019-1979-6 doi: 10.1186/s13662-019-1979-6
![]() |
[29] |
F. Chen, M. You, Permanence for an integrodifferential model of mutualism, Appl. Math. Comput., 186 (2007), 30–34. https://doi.org/10.1016/j.amc.2006.07.085 doi: 10.1016/j.amc.2006.07.085
![]() |
[30] |
E. Beretta, Y. Kuang, Geometric stability switch criteria in delay differential systems with delay-dependent parameters, SIAM J. Math. Anal., 33 (2002), 1144–1165. https://doi.org/10.1137/S0036141000376086 doi: 10.1137/S0036141000376086
![]() |
[31] |
L. Deng, X. Wang, M. Peng, Hopf bifurcation analysis for a ratio-dependent predator-prey system with two delays and stage structure for the predator, Appl. Math. Comput., 231 (2014), 214–230. https://doi.org/10.1016/j.amc.2014.01.025 doi: 10.1016/j.amc.2014.01.025
![]() |
[32] |
X. Yang, L. Chen, J. Chen, Permanence and positive periodic solution for the single-species nonautonomous delay diffusive models, Comput. Math. Appl., 32 (1996), 109–116. https://doi.org/10.1016/0898-1221(96)00129-0 doi: 10.1016/0898-1221(96)00129-0
![]() |
[33] | S. L. Ross, Differential equations, New York: John Wiley and Sons, 1984. |
1. | Kexin Zhang, Yi Zhang, Lu Liu, 2024, Event-triggered Sliding Mode Control of Biological Systems with Beddington-DeAngelis Functional Response, 978-9-8875-8158-1, 561, 10.23919/CCC63176.2024.10662661 | |
2. | Aditya Bhattacharya, Anindita Bhattacharyya, Bioeconomic and dynamical study of a predator–prey model with age-selective removal of prey during growth according to the Richards type, 2025, 13, 2195-268X, 10.1007/s40435-025-01626-5 | |
3. | Thuy Phuong Nguyen, Thao Dieu Do, Huyen Thi-Thu Nguyen, Dynamical analysis of a stage-structured food-chain model under fear effect and anti-predator behavior, 2025, 1598-5865, 10.1007/s12190-025-02451-x |
Values of delay | Thresholds of delay | Simulation results | Explanations of simulation |
τ2=0 | τ10≈10.88 | Figure 2 | When τ2=0, E∗ is locally asymptotically stable for τ1<10.88 and unstable for τ1>10.88. |
τ1=0 | τ20≈0.11 | Figure 5 | When τ1=0, E∗ is locally asymptotically stable for τ2<0.11 and unstable for τ2>0.11. |
τ2=0.1∈(0,τ20) | ˆτ10≈10.28 | Figure 10 | When τ2=0.1∈(0,τ20), E∗ is locally asymptotically stable for τ1<10.28 and unstable for τ1>10.28. |
τ1=1∈(0,τ10) | ˆτ20≈0.96 | Figure 13 | When τ1=1∈(0,τ10), E∗ is locally asymptotically stable for τ2<0.96 and unstable for τ2>0.96. |
Values of delay | Thresholds of delay | Simulation results | Explanations of simulation |
τ2=0 | τ10≈10.88 | Figure 2 | When τ2=0, E∗ is locally asymptotically stable for τ1<10.88 and unstable for τ1>10.88. |
τ1=0 | τ20≈0.11 | Figure 5 | When τ1=0, E∗ is locally asymptotically stable for τ2<0.11 and unstable for τ2>0.11. |
τ2=0.1∈(0,τ20) | ˆτ10≈10.28 | Figure 10 | When τ2=0.1∈(0,τ20), E∗ is locally asymptotically stable for τ1<10.28 and unstable for τ1>10.28. |
τ1=1∈(0,τ10) | ˆτ20≈0.96 | Figure 13 | When τ1=1∈(0,τ10), E∗ is locally asymptotically stable for τ2<0.96 and unstable for τ2>0.96. |