
Citation: Vikas Kumar, Nitu Kumari. Controlling chaos in three species food chain model with fear effect[J]. AIMS Mathematics, 2020, 5(2): 828-842. doi: 10.3934/math.2020056
[1] | Xiongxiong Du, Xiaoling Han, Ceyu Lei . Dynamics of a nonlinear discrete predator-prey system with fear effect. AIMS Mathematics, 2023, 8(10): 23953-23973. doi: 10.3934/math.20231221 |
[2] | Xiaoming Su, Jiahui Wang, Adiya Bao . Stability analysis and chaos control in a discrete predator-prey system with Allee effect, fear effect, and refuge. AIMS Mathematics, 2024, 9(5): 13462-13491. doi: 10.3934/math.2024656 |
[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] | 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 |
[5] | 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 |
[6] | 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 |
[7] | 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 |
[8] | 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 |
[9] | 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 |
[10] | Yao Shi, Zhenyu Wang . Bifurcation analysis and chaos control of a discrete fractional-order Leslie-Gower model with fear factor. AIMS Mathematics, 2024, 9(11): 30298-30319. doi: 10.3934/math.20241462 |
Chaotic behaviour exists in many natural systems, such as weather and climate. We can study this behaviour using a nonlinear mathematical model. Controlling chaos or stabilizing chaotic population dynamics becomes a challenge. Small manipulation in a chaotic system can control chaos, for example, stabilize an unstable periodic orbit. This manuscript mainly concerns with the chaotic dynamics of three species food chain model. Hastings and Powell [1] shows the appearance of chaos in a natural food chain or web models. Many researchers [2,3,4,5] have studied the Hastings and Powell model [1] by incorporating some biological factors such as alternative food, Allee effects, refugia, disease, fear.
In ecology, the predator-prey interaction happens at higher trophic levels and predators have an impact on prey populations which may be direct and indirect or both [6]. In the direct effect, the predator predates prey while in the indirect effect, the predator induces fear in the prey population and this fear can change the prey's behaviour [7]. The changing in the behaviour of prey includes foraging and reproduction [8,9]. Zanette et al. [9] have done fieldwork on song sparrows during the breeding season while eliminating the direct predation to observe the impact of fear. A reduction of 40% reproduction has been found in the number of offspring due to the fear of predation alone. Due to the fear effect, the prey shows a variety of antipredator responses, including different psychological changes, habitat changes, and foraging. For example, birds respond to their antipredator defences, by escaping from their homes [10] and mule deer reduces foraging time due to predation danger of mountain lions [11].
The proposed model of Wang et al. [12] shows a reduction in the production of prey due to fear of the predator. They observed that the high level of fear could stabilize the system (by excluding population oscillations). Further, they found that the fear effect could change population oscillations from a supercritical Hopf bifurcation to subcritical Hopf bifurcation, which induces multiple limit cycles. The impact of fear has been investigated on prey-predator systems with refuge and allee effect [13,14]. Fear effect also has been observed on discrete, delayed, diffusive, and eco-epidemiological models [15,16,17,18,19,20]. In the fear-induced prey-predator models, most of the prey-predator systems are studied with Holling type-Ⅱ functional response. Pal et al. [21] proposed a prey-predator model, where the interaction between prey and predator has been followed by Beddington-DeAngelis functional response. Based on their results, they conclude that the fear effect has both stabilizing and destabilizing effects.
Chaotic dynamics in tritrophic food chain model has recently been incorporated by Panday et al. [4]. For the first time, they assumed that the growth rate of prey and the middle predator is suppressed due to the cost of fear of middle and top predator, respectively. Further, they conclude that fear factors can control chaotic dynamics.
Our goal in this manuscript is to incorporate the fear factor in an earlier studied model [22,23] and since the system is chaotic so our effort would be to stabilize the dynamics of a proposed model system with the help of fear factors.
Here we consider a chaotic model [22,23], describes the interaction of the species in the food chain of prey, specialist predator and top specialist predator. The nonlinear model system is described as follows:
du1dt=a0u1−b0u21−w0u1u2d0+u1,du2dt=w1u1u2d1+u1−w2u2u3d2+u2−a1u2,du3dt=−c3u3+w3u2u3d3+u2. | (2.1) |
Here, u1,u2 and u3 are the respective population densities of prey, intermediate predator and top predator. a0,a1,b0,d0, d1,d2,d3,w0,w1,w2,w3 and c3 are positive constants. a0 represents the intrinsic growth rate of prey u1, a1 is the mortality rate of the predator u2 in the absence of u1 only. The parameters w0,w1,w2 are the maximum value which per capita growth rate can attain. d0,d1,d2, and d3 are half saturation constants corresponding to Holling type-Ⅱ functional response, b0 is the rate of competition among individuals of prey, parameter c3 is the mortality rate of the top predator in the absence of intermediate predator, and w3 is measure of its assimilation efficiency.
It is observed from the field experiments that the production reduces due to fear effects. The following assumptions are made to incorporate fear effect in model system (2.1):
(ⅰ) We assume that due to fear of top predator u3, the growth rate of intermediate predator u2 reduces. Therefore the modified growth rate change to w11+f2u3 (a monotonically decreasing function of both f2 and u3).
(ⅱ) We also assume that the growth rate of prey reduces due to fear of intermediate predator and the growth rate change to a01+f1u2, (a monotonically decreasing function of both f1 and u2).
Here g1(f1,u2)=11+f1u2 and g2(f2,u3)=11+f2u3 are fear functions, which account for the cost of anti-predator defence due to fear, where f1 and f2 represent the fear parameters (level of fear) of prey and intermediate predator, respectively. From the biological point of view, it is reasonable to assume that
g1(0,u2)=1,g1(f1,0)=1,limf1→∞g1(f1,u2)=0,limu2→∞g1(f1,u2)=0,∂g1(f1,u2)∂f1<0,∂g1(f1,u2)∂u2<0, | (2.2) |
and
g2(0,u3)=1,g2(f2,0)=1,limf2→∞g2(f2,u3)=0,limu3→∞g2(f2,u3)=0,∂g2(f2,u3)∂f2<0,∂g2(f2,u3)∂u3<0. | (2.3) |
By using above assumptions, the model system (2.1) takes the form:
du1dt=a0u11+f1u2−b0u21−w0u1u2d0+u1,du2dt=w1u1u2d1+u1⋅11+f2u3−w2u2u3d2+u2−a1u2,du3dt=−c3u3+w3u2u3d3+u2. | (2.4) |
We analyze the model system (2.4) with the following positive initial conditions
u1(0)>0,u2(0)>0,u3(0)>0. | (2.5) |
The rest of the paper is organised as follows: In section 3, the existence of equilibria is discussed. The local stability and bifurcation analysis are investigated in sections 4 and 5, respectively. Numerical simulation is discussed in section 6. Finally, the manuscript ends with a discussion in section 7.
In this section, the existence of non-negative equilibria are discussed and further, stability analysis of these equilibrium points are established. The following non-negative equilibrium points are obtained:
(ⅰ) Trivial equilibrium point E0(0,0,0), corresponds to the total extinction of prey and predator species.
(ⅱ) Axial equilibrium point E1(a0/b0,0,0), corresponds to the extinction of the predator species.
(ⅲ) Boundary equilibrium point E2(~u1,~u2,0), corresponds to the extinction of the top predator species, where
~u1=a1d1w1−a1, |
and ~u2 is the positive root of the quadratic equation
f1w0~u22+(w0+b0~u1f1(d0+~u1))~u2−(a0−b0~u1)(d0+~u1)=0. | (3.1) |
The planar equilibrium E2(~u1,~u2,0) exists if the following conditions are satisfied
w1>a1anda0+b0d1<a0w1a1. | (3.2) |
(ⅳ) We obtain positive equilibrium point E∗(u∗1,u∗2,u∗3), by solving these system of nonlinear equations:
a01+f1u2−b0u1−w0u2d0+u1=0,w1u1d1+u1⋅11+f2u3−w2u3d2+u2−a1=0,−c3+w3u2d3+u2=0. | (3.3) |
Solving the following Eq (3.3), we obtain,
u∗2=c3d3w3−c3, | (3.4) |
value of u∗1 is obtained by solving the equation
b0u12+(b0d0−a01+f1u∗2)u1+(w0u∗2−a0d01+f1u∗2)=0, | (3.5) |
and u∗3 is obtained from the equation
f2w2d2+u∗2u23+(w2d2+u∗2+a1f2)u3+(a1−w1u∗1d1+u∗1)=0. | (3.6) |
From Eqs (3.4), (3.5) and (3.6), it is clear that, for the existence of positive equilibrium point following conditions should be satisfied:
w3>c3,u∗1>a1d1w1−a1,w0u∗2<a0d01+f1u∗2andb0d0>a01+f1u∗2. | (3.7) |
Now, in order to study the behavior of solution near the equilibrium points, we compute the Jacobian matrix of the model system (2.4) at any point (u1,u2,u3), which is given by
JE=[A11A120A21A22A230A32A33] | (3.8) |
where,
A11=a01+f1u2−2b0u1−w0u2d0+u1+w0u1u2(d0+u1)2,A12=−w0u1d0+u1−a0f1u1(1+f1u2)2,A21=w1d1u2(d1+u1)2(1+f2u3),A22=w2u2u3(d2+u2)2−w2u3d2+u2−a1+w1u1(d1+u1)(1+f2u3),A23=−w2u2d2+u2−f2w1u1u2(d1+u1)(1+f2u3)2,A32=w3d3u3(d3+u2)2andA33=−c3+w3u2d3+u2. |
For local asymptotic stability, the solutions must approach to an equilibrium point under initial conditions close to the equilibrium point. Here, the criterion for local stability of the equilibria are obtained by linearizing the model system around the corresponding equilibrium point.
Theorem 1.
(a) The trivial equilibrium point E0(0,0,0) is always unstable.
(b) Axial equilibrium point E1(a0/b0,0,0) is locally asymptotically stable if a1>a0w1a0+b0d1.
(c) The equilibrium point E2(~u1,~u2,0) is locally asymptotically stable if b0>w0˜u2(d0+˜u1)2 and c3>w3~u2d3+~u2.
Proof.
(a) The Jacobian matrix at trivial equilibrium point E0(0,0,0) is given by
JE0=[a0000−a1000−c3] |
and the eigenvalues of the Jacobian matrix JE0 are a0,−a1 and −c3. So the equilibrium point E0(0,0,0) is always unstable.
(b) The Jacobian matrix at the axial equilibrium point E1(a0/b0,0,0) is
JE1=[−a0−(a0w0a0+b0d0+a20f1b0)00−a1+a0w1a0+b0d1000−c3] |
and the eigenvalues of JE1 are −a0,−c3 and −a1+a0w1a0+b0d1. Hence E1(a0/b0,0,0) is local asymptotically stable if a1>a0w1a0+b0d1.
(c) The Jacobian matrix associated with E2(~u1,~u2,0) is
JE2=[P1−P20P30−P400P5] |
where,
P1=˜u1(−b0+w0˜u2(d0+˜u1)2),P2=w0~u1d0+~u1+a0f1~u1(1+f1~u2)2,P3=w1d1~u2(d1+~u1)2,P4=w2~u2d2+~u2+f2w1~u1~u2d1+~u1andP5=−c3+w3~u2d3+~u2. |
The corresponding characteristic equation of the matrix JE2 is
(P5−λ)(λ2−P1λ+P2P3)=0. | (4.1) |
The characteristic Eq (4.1) have negative real parts if P5<0 and P1<0. Hence the equilibrium point E2(~u1,~u2,0) is locally asymptotically stable if the conditions stated in the Theorem 1(c) are satisfied.
Theorem 2. The positive equilibrium E∗(u∗1,u∗2,u∗3) is locally asymptotically stable if the following conditions hold:
b0u∗1>w0u∗1u∗2(d0+u∗1)2+w2u∗2u∗3(d2+u∗2)2 | (4.2) |
b0u∗1α+b0w2βu∗1u∗2u∗3(d2+u∗2)2>b20w2u∗12u∗2u∗3(d2+u∗2)2+w2w3d3u∗2u∗32γ(d2+u∗2)2(d3+u∗2)2+αβ, | (4.3) |
where, α,β and γ are given in (4.6).
Proof. The Jacobian matrix associated with the positive equilibrium E∗(u∗1,u∗2,u∗3) is
JE∗=[J11J120J21J22J230J320]. | (4.4) |
and corresponding characteristic equation is
σ3+ρ1σ2+ρ2σ+ρ3=0, | (4.5) |
where,
ρ1=−(J11+J22),ρ2=J11J22−J12J21−J23J32,ρ3=J11J23J32,ρ1ρ2−ρ3=(J11+J22)(J12J21−J11J22)+J22J23J32, |
with
J11=−b0u∗1+w0u∗1u∗2(d0+u∗1)2,J12=−w0u∗1d0+u∗1−a0f1u∗1(1+f1u∗2)2,J21=w1d1u∗2(d1+u∗1)2(1+f2u∗3),J22=w2u∗2u∗3(d2+u∗2)2,J23=−w2u∗2d2+u∗2−f2w1u∗1u∗2(d1+u∗1)(1+f2u∗3)2andJ32=w3d3u∗3(d3+u∗2)2. |
The interior equilibrium will be locally asymptotically stable if the coefficients of the characteristic Eq (4.5), ρ1,ρ2 and ρ3, are satisfy the Routh-Hurwitz stability criterion, i.e. ρ1>0,ρ3>0 and ρ1ρ2−ρ3>0.
Straightforward computation shows that ρ1>0 and ρ3>0, if
−b0u∗1+w0u∗1u∗2(d0+u∗1)2+w2u∗2u∗3(d2+u∗2)2<0, |
which is true from the condition (4.2).
Also since,
ρ1ρ2−ρ3=(J11+J22)(J12J21−J11J22)+J22J23J32={−b0u∗1+w0u∗1u∗2(d0+u∗1)2+w2u∗2u∗3(d2+u∗2)2}×{−(b0u∗1+w0u∗1u∗2(d0+u∗1)2)×w2u∗2u∗3(d2+u∗2)2−(w0u∗1d0+u∗1+a0f1u∗1(1+f1u∗2)2)×w1d1u∗2(d1+u∗1)2(1+f2u∗3)}−w2u∗2u∗3(d2+u∗2)2⋅w3d3u∗3(d3+u∗2)2(w2u∗2d2+u∗2+f2w1u∗1u∗2(d1+u∗1)(1+f2u∗3)2)=b0u∗1α+b0w2βu∗1u∗2u∗3(d2+u∗2)2−b20w2u∗12u∗2u∗3(d2+u∗2)2−αβ−w2w3d3u∗2u∗32γ(d2+u∗2)2(d3+u∗2)2>0, |
which is true, using the condition (4.3).
Where,
α=w0w2u∗1u∗22u∗3(d0+u∗1)2(d2+u∗2)2+w1d1u∗2(d1+u∗1)2(1+f2u∗3)(w0u∗1d0+u∗1+a0f1u∗1(1+f1u∗2)2),β=w0u∗1d0+u∗1+a0f1u∗1(1+f1u∗2)2andγ=w2u∗2d2+u∗2+f2w1u∗1u∗2(d1+u∗1)(1+f2u∗3)2. | (4.6) |
In this work, our primary attention is to stabilize the chaotic dynamics using fear parameters. We observe from the numerical simulation that the model system obtains stable dynamics after a limit cycle oscillation. Therefore the occurrence of Hopf bifurcation is obvious.
Hopf bifurcation is a point or a critical value of a parameter where the system stability changes and periodic solution arises. In the next theorem, we prove the existence of Hopf bifurcation, and for this, we choose fear parameter f1 as a bifurcation parameter.
Theorem 3. When the fear parameter f1 crosses a critical value f∗1, then the system enters into Hopf-bifurcation around the positive equilibrium E∗ if the following conditions hold:
ρ1(f1∗)>0,ρ3(f1∗)>0,ρ1(f1∗)ρ2(f1∗)−ρ3(f1∗)=0and[ρ1(f1∗)ρ2(f1∗)]′≠ρ′3(f1∗). | (5.1) |
Proof. Since we have seen that the interior equilibrium is locally asymptotically stable and we know that the model system loses its stability due to change in some parameter value. Hence, we choose the fear parameter f1 as the bifurcation parameter. If there exists a critical value f1∗ such that
ρ1(f1∗)ρ2(f1∗)−ρ3(f1∗)=0. |
For f1=f1∗ the characteristic Eq (4.5) must be of the form
(σ2(f1∗)+ρ2(f1∗))(σ(f1∗)+ρ1(f1∗))=0, | (5.2) |
the above equation has three roots −ρ1(f1∗),i√ρ2(f1∗) and −i√ρ2(f1∗). To show Hopf bifurcation occurs at f1=f1∗, we need to satisfy transversality condition
dRe(σ(f1))df1|f1=f1∗≠0. |
For all f1, the roots are in general of the form
σ1(f1)=μ(f1)+iν(f1),σ2(f1)=μ(f1)−iν(f1),σ3(f1)=−ρ1(f1). |
Now, we shall verify the transversality condition
dRe(σj(f1))df1|f1=f1∗≠0,j=1,2. |
Substituting σj(f1)=μ(f1)±iν(f1), into (5.2) and calculating the derivative, we have
K(f1)μ′(f1)−L(f1)ν′(f1)+M(f1)=0,K(f1)μ′(f1)+L(f1)ν′(f1)+N(f1)=0, | (5.3) |
where
K(f1)=3μ2(f1)+2ρ1(f1)μ(f1)+ρ2(f1)−3ν2(f1),L(f1)=6μ(f1)ν(f1)+2ρ1(f1)ν(f1),M(f1)=μ2(f1)ρ′1(f1)+ρ′2(f1)μ(f1)+ρ′3(f1)−ρ′1(f1)ν2(f1),N(f1)=2μ(f1)ν(f1)ρ′1(f1)+ρ′2(f1)ν(f1). |
Notices that μ(f1∗)=0,ν(f1∗)=√ρ2(f1∗), then we have
K(f1∗)=−2ρ2(f1∗),L(f1∗)=2ρ1(f1∗)√ρ2(f1∗),M(f1∗)=ρ′3(f1∗)−ρ′1(f1∗)ρ2(f1∗),N(f1∗)=ρ′2(f1∗)√ρ2(f1∗). |
Solving μ′(f1∗) from Eq (5.3) we get,
dRe(σj(f1))df1|f1=f1∗=μ′(f1)f1=f1∗=−L(f1∗)N(f1∗)+K(f1∗)M(f1∗)K2(f1∗)+L2(f1∗) |
=12ρ′3(f1∗)−(ρ1(f1∗)ρ2(f1∗))′ρ21(f1∗)+ρ2(f1∗)≠0. |
If [ρ1(f1∗)ρ2(f1∗)]′≠ρ′3(f1∗) and
σ3(f1∗)=−ρ1(f1∗)<0. |
Thus the transversality conditions hold and hence Hopf-bifurcation occurs at f1=f1∗.
From numerical simulation, we observed that the Hopf bifurcation occurs with respect to the fear parameter f1 at the critical value f∗1=0.02769 around the equilibrium point (24.58, 10.77, 5.839) and also occurs with respect to the fear parameter f2 at the critical value f∗2=0.5268 around the equilibrium (35.24, 10.77, 0.8819). The parameters used in the calculation, for the bifurcation points, are fixed as given in Figure 1.
The main goal of this manuscript is to control the chaotic dynamic of the model system (2.1). For this purpose, we have proposed a mathematical model (2.4) by incorporating a biological factor of fear. Model system (2.1) has been studied in literature by many authors [22,23]. The model system (2.1) exhibits chaotic attractor, which can be seen in Figure 1. When we increase the value of the intrinsic growth rate of prey a0, the system shows a chaotic dynamic in the range 1.96≤a0≤2. These changes can be seen from a bifurcation diagram plotted in Figure 2. From Figure 3, we observe that the model system enters into a chaotic dynamic from a stable focus via limit cycle oscillation and period-doubling. It is noticed that when f1=f2=0 then our proposed model (2.4) is same as an earlier studied model [22,23,24].
Next, we studied the impact of fear on the dynamics of the model system (2.4) and observed that in the absence of fear (i.e., f1=f2=0), the model system exhibits chaotic dynamics. Further, by varying f1 and f2, we investigate the effect of fear on the system dynamics. Firstly, we investigate the impact of fear of intermediate predator (f1) on prey while the fear of top predator on the intermediate predator is absent, i.e., f2=0. We observe that the increase in fear effect f1 makes the system dynamics (2.4) stable from chaotic, which can be seen by a bifurcation diagram plotted in Figure 4. The system remains chaotic for the low cost of fear in prey growth. However, an increased cost of fear (f1≥0.03) stabilizes the system dynamics from chaotic to stable focus, see Figure 5. From Figure 6, it is observed that above the threshold value of f1=0.08, the extinction of top predator population is possible, whereas prey and intermediate predator shows stable coexistence. Next, we consider the fear of top predator (f2), which reduces the growth rate of an intermediate predator where the fear of intermediate predator is absent, i.e., f1=0. It is observed that an increase in the fear parameter f2 makes system (2.4) stable from a chaotic dynamics, see Figures 7 and 8. The bifurcation diagram illustrates that the chaotic dynamics observed for the low cost of fear while the stable dynamics observed at a higher cost of fear (f2≥0.6). We have also observed that the large value of f2(f2=500), top predator goes to be extinct, whereas prey and intermediate predator population remain stable (can be seen in Figure 9).
Next, our investigation is to observe the dynamics of the model system (2.4) in the presence of both fear factors (f1 and f2). From Figure 10, we observe that for small values of fear factors the system dynamics remain chaotic and the increased value of f1 or f2 or both changes system dynamics from chaos to stable focus. The chaotic or higher periodic oscillations change to period-doubling, period-doubling to limit cycle and limit cycle to stable focus, as we increase the cost of fear factors.
In this work, we have analysed a three species food chain model (2.4) incorporating fear factors in the classical model (2.1). Next, analytically, we have proved the existence of biological feasible equilibrium points. Further, the local stability analysis of the proposed model system (2.4) is done corresponding to existing equilibria. Since the model system dynamics transits from the limit cycle to stable focus, hence there occurs Hopf bifurcation. Therefore we have also done Hopf bifurcation analysis, where we have proved that the system enters into Hopf bifurcation with the fear parameters f1 and f2 as bifurcation parameters.
The model system (2.4) exhibits chaotic dynamics, and a chaotic attractor is observed (Figure 1). From Figures 2 and 3, we observe that increase in the intrinsic growth rate of prey (a0) leads to chaotic dynamics which can be controlled by fear factors. Further, from Figures 4 and 5, it is observed that the increase in fear factor f1 makes the system stable from chaotic dynamics (even in the absence of cost of fear of top predator in intermediate predator). Even the large cost of fear causes a population to become extinct, see Figure 6. Similarly, increase in fear factor f2 makes system (2.4) stable, which has been shown in Figures 7 and 8. Also, the large cost of fear causes the top predator population to become extinct (Figure 9).
It should be noted that these chaotic dynamics hold for a small cost of fear. If we increase the cost of fear, then the system dynamics change from chaos to stable focus. The model system remains chaotic for the low cost of fear, whereas when we increase the value of f1 or f2 or both, then the system tends to stable dynamics (see, Figure 10). Therefore, we can conclude that the fear parameters control the chaotic dynamics in a food chain model.
The research work of the first author is supported by the Council of Scientific & Industrial Research (CSIR, India) under the file no: 09/1058(0006)/2016-EMR-I. The authors are very grateful to anonymous referees for reading their paper carefully and for several constructive remarks and suggestions.
The authors declare that there is no conflicts of interest in this paper.
[1] |
A. Hastings, T. Powell, Chaos in a three-species food chain, Ecol., 72 (1991), 896-903. doi: 10.2307/1940591
![]() |
[2] |
J. N. Eisenberg, D. R. Maszle, The structural stability of a three-species food chain model, J. Theor. Biol., 176 (1995), 501-510. doi: 10.1006/jtbi.1995.0216
![]() |
[3] |
B. Sahoo, S. Poria, The chaos and control of a food chain model supplying additional food to top-predator, Chaos. Solitons Fractals., 58 (2014), 52-64. doi: 10.1016/j.chaos.2013.11.008
![]() |
[4] | P. Panday, N. Pal, S. Samanta, et al. Stability and bifurcation analysis of a three-species food chain model with fear, Internat. J. Bifur. Chaos Appl. Sci. Engrg., 28 (2018), 1850009. |
[5] | B. Nath, N. Kumari, V. Kumar, et al. Refugia and allee effect in prey species stabilize chaos in a tri-trophic food chain model, Differ. Equ. Dyn. Syst., (2019), 1-27. |
[6] | R. J. Taylor, Predation, New York: Chapman and Hall Press, (1984). |
[7] |
S. L. Lima, L. M. Dill, Behavioral decisions made under the risk of predation: A review and prospectus, Can. J. Zool., 68 (1990), 619-640. doi: 10.1139/z90-092
![]() |
[8] |
O. J. Schmitz, A. P. Beckerman, K. M. O'Brien, Behaviorally mediated trophic cascades: Effects of predation risk on food web interactions, Ecol., 78 (1997), 1388-1399. doi: 10.1890/0012-9658(1997)078[1388:BMTCEO]2.0.CO;2
![]() |
[9] |
L. Y. Zanette, A. F. White, M. C. Allen, et al. Perceived predation risk reduces the number of offspring songbirds produce per year, Sci., 334 (2011), 1398-1401. doi: 10.1126/science.1210908
![]() |
[10] |
W. Cresswell, Predation in bird populations, J. Ornitho., 152 (2011), 251-263. doi: 10.1007/s10336-010-0638-1
![]() |
[11] |
K. B. Altendorf, J. W. Laundré, C. A. López González, et al. Assessing effects of predation risk on foraging behavior of mule deer, J. Mammal., 82 (2001), 430-439. doi: 10.1644/1545-1542(2001)082<0430:AEOPRO>2.0.CO;2
![]() |
[12] |
X. Wang, L. Zanette, X. Zou, Modelling the fear effect in predator-prey interactions, J. Math. Biol., 73 (2016), 1179-1204. doi: 10.1007/s00285-016-0989-1
![]() |
[13] |
S. K. Sasmal, Population dynamics with multiple allee effects induced by fear factors-a mathematical study on prey-predator interactions, Appl. Math. Model., 64 (2018), 1-14. doi: 10.1016/j.apm.2018.07.021
![]() |
[14] |
H. Zhang, Y. Cai, S. Fu, et al. Impact of the fear effect in a prey-predator model incorporating a prey refuge, Appl. Math. Computation, 356 (2019), 328-337. doi: 10.1016/j.amc.2019.03.034
![]() |
[15] | K. Kundu, S. Pal, S. Samanta, et al. Impact of fear effect in a discrete-time predator-prey system, Bull. Calcutta Math. Soc., 110 (2018), 245-264. |
[16] |
S. Mondal, A. Maiti, G. Samanta, Effects of fear and additional food in a delayed predator-prey model, Biophys. Rev. Lett., 13 (2018), 157-177. doi: 10.1142/S1793048018500091
![]() |
[17] |
D. Duan, B. Niu, J. Wei, Hopf-hopf bifurcation and chaotic attractors in a delayed diffusive predator-prey model with fear effect, Chaos. Solitons Fractals., 123 (2019), 206-216. doi: 10.1016/j.chaos.2019.04.012
![]() |
[18] | X. Wang, X. Zou, Pattern formation of a predator-prey model with the cost of anti-predator behaviors, Math. Biosci. Eng., 15 (2017), 775-805. |
[19] | S. Chen, Z. Liu, J. Shi, Nonexistence of nonconstant positive steady states of a diffusive predatorprey model with fear effect, J. Nonlinear Model. Anal., 1 (2019), 47-56. |
[20] |
A. Sha, S. Samanta, M. Martcheva, et al. Backward bifurcation, oscillations and chaos in an ecoepidemiological model with fear effect, J. Biol. Dyn., 13 (2019), 301-327. doi: 10.1080/17513758.2019.1593525
![]() |
[21] | S. Pal, S. Majhi, S. Mandal, et al. Role of fear in a predator-prey model with beddington-deangelis functional response, Z. für Naturforsch. A., 74 (2019), 581-595. |
[22] |
V. Rai, R. K. Upadhyay, Chaotic population dynamics and biology of the top-predator, Chaos, Solitons Fractals., 21 (2004), 1195-1204. doi: 10.1016/j.chaos.2003.12.065
![]() |
[23] |
R. Upadhyay, R. Naji, N. Kumari, Dynamical complexity in some ecological models: Effect of toxin production by phytoplankton, Nonlinear Anal. Model Control., 12 (2007), 123-138. doi: 10.15388/NA.2007.12.1.14726
![]() |
[24] | M. A. Aziz-Alaoui, Study of a Leslie-Gower-type tritrophic population model. Chaos Solitons Fractals., 14 (2002), 1275-1293. |
1. | Vikas Kumar, Nitu Kumari, Stability and Bifurcation Analysis of Hassell–Varley Prey–Predator System with Fear Effect, 2020, 6, 2349-5103, 10.1007/s40819-020-00899-y | |
2. | Pingping Cong, Meng Fan, Xingfu Zou, Dynamics of a three-species food chain model with fear effect, 2021, 99, 10075704, 105809, 10.1016/j.cnsns.2021.105809 | |
3. | Ashutosh Maurya, Anupam Priyadarshi, Complexity to order: impact of antipedator behaviour on chaotic and non-chaotic intraguild predation model , 2021, 1598-5865, 10.1007/s12190-021-01541-w | |
4. | Vikas Kumar, Nitu Kumari, Bifurcation study and pattern formation analysis of a tritrophic food chain model with group defense and Ivlev-like nonmonotonic functional response, 2021, 147, 09600779, 110964, 10.1016/j.chaos.2021.110964 | |
5. | Firas Hussean Maghool, Raid Kamel Naji, Oluwole D. Makinde, The Dynamics of a Tritrophic Leslie-Gower Food-Web System with the Effect of Fear, 2021, 2021, 1687-0042, 1, 10.1155/2021/2112814 | |
6. | Jinxing Zhao, Yuanfu Shao, Bifurcations of a prey-predator system with fear, refuge and additional food, 2022, 20, 1551-0018, 3700, 10.3934/mbe.2023173 | |
7. | Vikas Kumar, Nitu Kumari, Stability and Bifurcation Analysis of Fractional-Order Delayed Prey–Predator System and the Effect of Diffusion, 2022, 32, 0218-1274, 10.1142/S021812742250002X | |
8. | Zainab Saeed Abbas, Raid Kamel Naji, Modeling and Analysis of the Influence of Fear on a Harvested Food Web System, 2022, 10, 2227-7390, 3300, 10.3390/math10183300 | |
9. | Khushbu Singh, K Kaladhar, Population dynamic study of interaction between two predators and one prey, 2024, 99, 0031-8949, 025023, 10.1088/1402-4896/ad1dd2 | |
10. | K. Akshaya, S. Muthukumar, N. P. Deepak, M. Siva Pradeep, 2024, 3122, 0094-243X, 040018, 10.1063/5.0220288 | |
11. | Harendra Pal Singh, Suruchi Singh, Effect of fear with saturated fear cost and harvesting on aquatic food chain model (plankton–fish model) in the presence of nanoparticles, 2024, 226, 03784754, 283, 10.1016/j.matcom.2024.07.010 | |
12. | Duo Hua, Xingbo Liu, Computer‐assisted analysis of chaos in a three‐species food chain model, 2023, 151, 0022-2526, 1166, 10.1111/sapm.12624 | |
13. | Zuchong Shang, Yuanhua Qiao, Complex dynamics of a four-species food web model with nonlinear top predator harvesting and fear effect, 2024, 223, 03784754, 458, 10.1016/j.matcom.2024.04.024 | |
14. | Anuj Kumar Umrao, Prashant K. Srivastava, 2024, Chapter 3, 978-3-031-59071-9, 67, 10.1007/978-3-031-59072-6_3 | |
15. | JUAN LIU, RANJIT KUMAR UPADHYAY, RASHMI AGRAWAL, ANWAR ZEB, TAREQ SAEED, ZAHID AHMAD, HOPF BIFURCATION AND GLOBAL STABILITY OF A PREDATOR–PREY MODEL WITH ALTERNATIVE RESOURCE AND CANNIBALISM, 2023, 31, 0218-348X, 10.1142/S0218348X23400844 | |
16. | JUAN LIU, RANJIT KUMAR UPADHYAY, RASHMI AGRAWAL, ANWAR ZEB, TAREQ SAEED, DYNAMICS OF A PREY–PREDATOR MODEL INCORPORATING GESTATION TIME DELAY AND FEAR EFFECT, 2023, 31, 0218-348X, 10.1142/S0218348X23401552 | |
17. | Balram Dubey, Bifurcation Analysis and Spatiotemporal Dynamics in a Diffusive Predator–Prey System Incorporating a Holling Type II Functional Response, 2024, 34, 0218-1274, 10.1142/S0218127424501050 | |
18. | Pingping Cong, Meng Fan, Xingfu Zou, Fear effect exerted by carnivore in grassland ecosystem, 2024, 17, 1793-5245, 10.1142/S1793524523500584 | |
19. | Zuchong Shang, Yuanhua Qiao, Complex Dynamics of a Quad-Trophic Food Chain Model with Beddington–DeAngelis Functional Response, Fear Effect and Prey Refuge, 2025, 24, 1575-5460, 10.1007/s12346-024-01208-4 | |
20. | ANUJ KUMAR UMRAO, PRASHANT K. SRIVASTAVA, IMPACT OF HUNTING COOPERATION AND FEAR EFFECT IN A GENERALIST PREDATOR–PREY MODEL, 2024, 32, 0218-3390, 1261, 10.1142/S021833902340003X |