Research article Special Issues

Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model

  • The predation strategy for predators and the avoidance strategy of prey are important topics in ecology and evolutionary biology. Both prey and predators adjust their behaviours in order to gain the maximal benefits and to increase their biomass for each. In the present paper, we consider a modified Leslie-Gower predator-prey model where predators cooperate during hunting and due to fear of predation risk, prey populations show anti-predator behaviour. We investigate step by step the impact of hunting cooperation and fear effect on the dynamics of the system. We observe that in the absence of fear effect, hunting cooperation can induce both supercritical and subcritical Hopf- bifurcations. It is also observed that fear factor can stabilize the predator-prey system by excluding the existence of periodic solutions and makes the system more robust compared to hunting cooperation. Moreover, the system shows two different types of bi-stabilities behaviour: one is between coexisting equilibrium and limit cycle oscillation, and another is between prey-free equilibrium and coexisting equilibrium. We also observe generalized Hopf-bifurcation and Bogdanov-Takens bifurcation in two parameter bifurcation analysis. We perform extensive numerical simulations for supporting evidence of our analytical findings.

    Citation: Saheb Pal, Nikhil Pal, Sudip Samanta, Joydev Chattopadhyay. Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 5146-5179. doi: 10.3934/mbe.2019258

    Related Papers:

    [1] Hongqiuxue Wu, Zhong Li, Mengxin He . Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting. Mathematical Biosciences and Engineering, 2023, 20(10): 18592-18629. doi: 10.3934/mbe.2023825
    [2] Zhenliang Zhu, Yuming Chen, Zhong Li, Fengde Chen . Dynamic behaviors of a Leslie-Gower model with strong Allee effect and fear effect in prey. Mathematical Biosciences and Engineering, 2023, 20(6): 10977-10999. doi: 10.3934/mbe.2023486
    [3] Mengyun Xing, Mengxin He, Zhong Li . Dynamics of a modified Leslie-Gower predator-prey model with double Allee effects. Mathematical Biosciences and Engineering, 2024, 21(1): 792-831. doi: 10.3934/mbe.2024034
    [4] Shuo Yao, Jingen Yang, Sanling Yuan . Bifurcation analysis in a modified Leslie-Gower predator-prey model with fear effect and multiple delays. Mathematical Biosciences and Engineering, 2024, 21(4): 5658-5685. doi: 10.3934/mbe.2024249
    [5] Rongjie Yu, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao . Bifurcation analysis of Leslie-Gower predator-prey system with harvesting and fear effect. Mathematical Biosciences and Engineering, 2023, 20(10): 18267-18300. doi: 10.3934/mbe.2023812
    [6] Manoj K. Singh, Brajesh K. Singh, Poonam, Carlo Cattani . Under nonlinear prey-harvesting, effect of strong Allee effect on the dynamics of a modified Leslie-Gower predator-prey model. Mathematical Biosciences and Engineering, 2023, 20(6): 9625-9644. doi: 10.3934/mbe.2023422
    [7] Zhenzhen Shi, Huidong Cheng, Yu Liu, Yanhui Wang . Optimization of an integrated feedback control for a pest management predator-prey model. Mathematical Biosciences and Engineering, 2019, 16(6): 7963-7981. doi: 10.3934/mbe.2019401
    [8] Yangyang Li, Fengxue Zhang, Xianglai Zhuo . Flip bifurcation of a discrete predator-prey model with modified Leslie-Gower and Holling-type III schemes. Mathematical Biosciences and Engineering, 2020, 17(3): 2003-2015. doi: 10.3934/mbe.2020106
    [9] Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar . Dynamical analysis of a predator-prey system with prey vigilance and hunting cooperation in predators. Mathematical Biosciences and Engineering, 2024, 21(2): 2768-2786. doi: 10.3934/mbe.2024123
    [10] Eduardo Gonzalez-Olivares, Alejandro Rojas-Palma . Global stability in a modified Leslie-Gower type predation model assuming mutual interference among generalist predators. Mathematical Biosciences and Engineering, 2020, 17(6): 7708-7731. doi: 10.3934/mbe.2020392
  • The predation strategy for predators and the avoidance strategy of prey are important topics in ecology and evolutionary biology. Both prey and predators adjust their behaviours in order to gain the maximal benefits and to increase their biomass for each. In the present paper, we consider a modified Leslie-Gower predator-prey model where predators cooperate during hunting and due to fear of predation risk, prey populations show anti-predator behaviour. We investigate step by step the impact of hunting cooperation and fear effect on the dynamics of the system. We observe that in the absence of fear effect, hunting cooperation can induce both supercritical and subcritical Hopf- bifurcations. It is also observed that fear factor can stabilize the predator-prey system by excluding the existence of periodic solutions and makes the system more robust compared to hunting cooperation. Moreover, the system shows two different types of bi-stabilities behaviour: one is between coexisting equilibrium and limit cycle oscillation, and another is between prey-free equilibrium and coexisting equilibrium. We also observe generalized Hopf-bifurcation and Bogdanov-Takens bifurcation in two parameter bifurcation analysis. We perform extensive numerical simulations for supporting evidence of our analytical findings.


    Dynamical complexities of interacting predator-prey model are extensively studied by several researchers to understand the long-time behaviour of the species. Most of the predator-prey models are based upon classical Lotka-Volterra formalism, where the prey consumption rate of predators is the growth rate of predators with a conversion factor. However, in this paper we confine ourselves with Leslie-Gower model, whose main feature is logistic type growth function for predators [1]. Leslie [2] introduced that the environmental carrying capacity (KP) of the predators is proportional to the number of prey abundance (N) i.e. KP=kN, for some positive constant k, which is the conversion factor of prey into predators. Also, a positive constant term added with KP to avoid a mathematical singularity when prey population becomes zero. From a biological point of view, for the case of severe scarcity of prey, predators can switch over other populations (alternative food), but its growth will be restricted due to non availability of its favorite food. Also, adding this type of positive constant introduces a maximum decrease rate, which stands for environment protection.

    Thus we consider the following modified Leslie-Gower model within homogeneous environment with Holling functional response of type Ⅱ corresponds to the prey consumption rate of predators:

    {dNdT=N[b0d0a0NρP1+hρN],dPdT=P[c0PkN+m0], (1.1)

    where all variables and parameters are positive, and they are defined in Table 1.

    Table 1.  Interpretation of variables and parameters used in the model (1.1).
    Variable/ Parameters Interpretation
    N Prey population density
    P Predator population density
    T Time
    b0 Per capita birth rate of prey
    d0 Per capita death rate of prey
    a0 Intra-specific competition parameter
    ρ Encounter rate of predators with their prey
    h Predator handling time of a prey individual
    c0 Growth rate of predator
    k Conversion factor of prey biomass into predator biomass
    m0 A positive constant related to alternative food

     | Show Table
    DownLoad: CSV

    Many authors studied the Leslie-Gower type predator-prey model and its modified version. Aziz-Alaoui and Okiye [3] showed the global stability of the positive interior equilibrium point in a modified Leslie-Gower model. Nindjin et al. [4] analyzed the modified Leslie-Gower model with time delay. Gupta and Chandra [5] studied the modified Leslie-Gower model with Michaelis-Menten type prey harvesting and show rich complex-dynamical behaviours. Zhu and Wang [6] showed the global attractive behaviour of positive periodic solutions for the modified Leslie-Gower model, where predator-prey interaction follows Holling type Ⅱ scheme.

    Now, we consider two different aspects to describe the predator-prey interactions:

    1. Predators cooperate during hunting and as a result, success of capturing a prey increases with predator density.

    2. Due to fear of predation risk, the birth rate of prey population reduces.

    The role of mutual beneficial interactions (e.g. cooperation) among social animals is a rapidly growing research field in population dynamics and conservation biology. Cooperative hunting is the most widely distributed form of cooperative behaviour in animals. Cooperative hunting is widespread among vertebrates (e.g., lions [7], African wild dogs [8], chimpanzees [9], wolves [10]), birds [11], primates [12] etc.

    Different animals adopt different strategies for making a group. For instance, wolves have the strategies to enclose the prey and catch it, for the acquisition of larger and faster prey [13]. Group hunting in lionesses generally involves a formation whereby some lionesses circle prey while others wait for prey to move towards them [7]. For wild chimpanzees, when a large prey is subdued, all the chimpanzees scream with excitement and trying to reach it [9]. Painted dog forms a rally, rousing one another and getting each other for group hunting [14]. Bailey et al. [15] discussed different strategies of cooperative hunting, mutual benefit for living in a social group and how predators form a cooperative group with active individuals.

    In general, animals of same species compete with each other for resources, but because of some specific mechanism is at work, they help each other for mutual benefit and show cooperative behaviour. Two main constraints that lead to cooperative behaviour of the individuals are need for assistance in hunting and killing large prey, and the need for defence against attack by other predators [16,17]. There are several benefits of group hunting, such as the probability of capturing large prey increases [18], rate of hunting success increases with the number of adults [8], chasing distance decreases [8], locate food more quickly with increasing group size [19], helps to prevent the carcass from being stolen by other predators [20,21] etc. Packer and Ruttan [22] enlisted 61 species of mammals, birds, vertebrates and invertebrates, their hunting strategy, which prey or types of prey to predate, single or multiple preys to predate at a time and percentage of their hunting success, from different sources.

    Predators can impact the ecology and evolution of their prey directly by eating them, but also indirectly by influencing the behaviour and physiology of survivors [23]. This types of behavioural and physiological changes due to the fear of predation are more powerful and longer-lasting evolutionary consequence compared to direct predation [23,24]. For example, birds respond to the sounds of predators with anti-predator defences and flee from their nests at the first sign of danger [24]. Mule deer spends less time for foraging due to the predation risk of mountain lions [25]. Creel et al. [26] observed that reproductive physiology of elk changes due to the predation risk of wolves.

    Many experimental studies show that manipulation of fear is strong enough to affects the population dynamics of ecological systems. Zanette et al. [27] manipulated fear of song sparrows during the entire breeding season by eliminating direct predation and observed that perception of predation risk alone reduces the number of offspring produced per year by 40%. Hua et al. [28] observed that perception of predation risk negatively affect blue bird's reproduction by manipulating the vocal cues of their predators. Suraci et al. [29] experimentally showed that manipulation of fear of large carnivore's vocalization in free living mesocarnivores reduces their foraging, which is beneficial for mesocarnivore's prey. This type of anti-predator behaviour although beneficial for increasing the probability of survival but driving long-term costs on reproduction that may affect population numbers [24]. However, very recently, Peers et al. [30] argued that predator cue experiments may not be realistically comparable to those of the prey's natural sensory environment, and the understanding of fear effects will continue to rely on untested assumptions in the wild. How prey population sustains potentially risky activities such as feeding or mating against anti-predator demands, has been a major issue in modern behavioural ecology [31,32,33]. In the presence of predators, prey change their behaviour due to predation risk. These behavioural responses to predation risk include reduce foraging time [34], increase vigilance [35], reduce movement [36], retreat to low-risk areas or refuge [37] and change in group size [38].

    In mathematical modelling approach, many authors investigated the impact of hunting cooperation [39,40,41,42,43] and the impact of fear [44,45,46,47,48] in the predator-prey systems separately. In the context of hunting cooperation, Duarte et al. [39] investigated population dynamics of cooperative hunting in a McCann-Yodzis model by considering a fraction of predators cooperates during hunting. Berec [40] studied predators foraging facilitation or hunting cooperation in a Rosenzweig-MacArthur model and proposed a family of functional responses by considering attack rate and handling time of predators varies with predator density, in a Holling type Ⅱ functional response. Teixeira Alves and Hilker [41] showed that cooperative hunting in predators enhances the stability scenario of the coexistence steady state and it produces oscillatory dynamics in a predator-prey model, which is not possible in the absence of hunting cooperation. Recently, Pal et al. [43] investigated a discrete-time predator-prey model with hunting cooperation and observed that hunting cooperation has the potential to modify the well-known period-doubling route to chaos by reverse period-halving bifurcations and makes the system stable, but very high hunting cooperation can be detrimental for the system since it ultimately leads to extinction. On the other hand, recently, few researchers have investigated the impact of fear effect in a predator-prey system with the help of mathematical modeling. In this context, Wang et al. [44] first proposed a mathematical model by incorporating fear of the predator on prey, where the cost of fear reduces in the birth rate of prey. Their mathematical analysis suggests that high levels of fear can stabilize the predator-prey dynamics by excluding the existence of periodic solution. Wang and Zhou [45] investigated a predator-prey model with maturation delay for prey under the cost of fear and adaptive avoidance of predators. Their numerical findings showed that both strong adaptation of adult prey and the large cost of fear destabilize the predator-prey dynamics but higher population density stabilize the system. Panday et al. [46] studied a three-species food chain model, by considering the growth rate of prey and middle predator reduced due to the cost of fear of middle predator and top predator, respectively. Their numerical findings showed that fear has the potential to stabilize a chaotic system. Pal et al. [48] studied the impact of fear in a predator-prey model with Beddington-DeAngelis functional response and observed that the model exhibits multiple Hopf-bifurcations. Sasmal [47] studied ecological and epidemiological models with strong Allee effect and cost of fear in prey reproduction. Recently, Sha et al. [49] investigated an eco-epidemiological system with fear effects and observed that fear can produce backward bifurcation and chaotic oscillations in a simple eco-epidemic model.

    In the present paper, we are interested to observe the dynamics of the such system where predators show cooperative behaviour during hunting and also create fear upon their prey [50,51,52]. For example, wolves cooperate during hunting, and when wolves are present, elk used anti-predator strategies and avoided areas frequented by the wolves [53]. Elk responds in the presence of wolves by altering patterns of agammaegation, habitat selection, vigilance, foraging, and sensitivity to environmental conditions [38,54,55,56]. Lioness shows cooperative behaviours during hunting [7]. Due to the fear of predation risk, zebras reached areas where the encounter with lionesses less frequent [51]. Thus by both killing and frightening, predators could have a dual impact upon their prey. The above observations motivate us to study the dynamics of predator-prey system including cooperative hunting by predators and a reduction of prey birth rate due to fear of predation risk. To the best of our knowledge, the effect of hunting cooperation and fear in the predator-prey system simultaneously has not yet been studied, and is the first attempt in this direction. The main focus of our present study is to explore the following ecological issues:

    1. How does cooperative hunting impact (stabilize or destabilize) on the dynamics of a modified Leslie-Gower system?

    2. Can fear effect stabilize the predator-prey system?

    3. Can cooperative hunting and fear effect simultaneously promote species persistence and enhance ecosystem stability?

    To incorporate the cooperative behaviour of predators in the system (1.1), we briefly outline Berec's approach for cooperative behaviours among predators [40]. Berec discussed cooperative phenomena in a Holling type Ⅱ functional response with two different approaches.

    Firstly, when encounter rate of predators (ρ) increases with increasing predator density and handling time (h) is constant i.e.

    ρ(P)=ρ0(e+P)w,h(P)=h,e0andw<0,

    then type Ⅱ functional response with this ρ and h is called encounter-driven functional response.

    Secondly, if the encounter rate of predators ρ is constant and handling time h decreases with increasing predator density i.e.

    ρ(P)=ρ,h(P)=h0(e+P)w,e0andw<0,

    then type Ⅱ functional response with this ρ and h is called handling-driven functional response.

    In this paper, we confine ourselves with Berec's encounter-driven functional response. In particular, we choose w=1 and so ρ(P)=ρ0(e+P)=μ+νP, where μ corresponds to ρ0e and ν corresponds to ρ0. Also for the Leslie-Gower model (1.1), positive constant k which is stated earlier as the conversion factor of prey into predators, affected with the predator density. Thus we consider k=k0(μ+νP).

    In Leslie-Gower or modified Leslie-Gower model, the carrying capacity of the predator population is variable and grows linearly with prey population density. However, in the present paper, we assume that the environment carrying capacity of the predator grows with both prey and predators abundances. In terms of foraging theory, resource supply rates likely increase due to cooperative foraging, such as reduced search costs and increased encounter rates [57]. So, due to cooperation, predators can consume more prey, which can increase the carrying capacity for the predator population. Zhang [58] theoretically demonstrated that cooperation or mutualism can promote the competitive ability of competitors through increasing their carrying capacities. Hamilton et al. [59] suggested a new mechanism for the biogeographic expansion of modern humans by a modified density-dependent model with varying intraspecific competition for finite resources. They showed that cooperative behaviors reduce intraspecific competition and increase population carrying capacity. They also suggested that cooperation and sociality in modern humans led to increased regional carrying capacities of humans.

    Following the above observations, while formulating the following mathematical model, we use the formalism of increasing the cooperative behavior of predator increases its carrying capacity. Therefore, the model (1.1) with the Berec's encounter-driven functional response becomes:

    {dNdT=N[b0d0a0N(μ+νP)P1+h(μ+νP)N],dPdT=P[c0Pk0(μ+νP)N+m0]. (1.2)

    It is to be noted that in the absence of hunting cooperation (ν=0), the above model (1.2) will be exactly same as the modified Leslie-Gower model (1.1).

    Our second assumption is that birth rate of prey population is reduced due to fear induced by predators. To incorporate the fear phenomenon, we multiply the reproduction term i.e., birth rate (b0) of prey population with a decreasing function of the predator population size, ϕ(τ,P)=1(1+τP). Similar fear function was considered by many authors [44,46,47,60]. The parameter τ(>0) represents the level of fear that reduces the growth rate of prey. Expression of ϕ(τ,P) is appropriate since it satisfies the following biological assumptions:

    1. ϕ(0,P)=1: if there is no anti-predator behaviours of prey, then the birth rate of prey remains unchanged.

    2. ϕ(τ,0)=1: if the predator population becomes zero, then no reduction in prey reproduction due to anti-predator behaviours.

    3. limτϕ(τ,P)=0: when anti-predator behaviour is very large and predator density P>0, then prey reproduction decreases and ultimately becomes zero.

    4. limPϕ(τ,P)=0: when predator population is very large and strength of fear τ>0, then prey reproduction decreases and ultimately becomes zero, due to large anti-predator behaviours.

    5. ϕ(τ,P)τ<0: the reproduction of prey decreases with the increasing of anti-predator behaviours.

    6. ϕ(τ,P)P<0: the reproduction of prey decreases with the increasing of predator populations.

    Thus the model (1.2) with fear phenomena becomes

    {dNdT=N[b01+τPd0a0N(μ+νP)P1+h(μ+νP)N],dPdT=P[c0Pk0(μ+νP)N+m0]. (1.3)

    To reduce the number of parameters, we rescale the variables and parameters and the system (1.3) gets transformed to

    {dxdt=x[b1+βydx(1+αy)yp+(1+αy)x]xf(x,y),dydt=y[1qyr+(1+αy)x]yg(x,y), (1.4)

    This is the model we will analyze in the following. The new variables and parameters are given as follows: x=a0c0N, y=a0hc20P, t=c0T, b=b0c0, d=d0c0, α=νhc20μa0, β=hc20τa0, p=a0μhc0, q=hμk0, r=m0a0μc0k0.

    In this section, we present some basic results, such as positivity, boundedness and permanence of the solutions of system (1.4).

    To prove positivity of the model (1.4), we follow the theorem given in H.R. Thieme's book [61].

    From the standpoint of ecological meaning, we consider the following region R2+={(x,y):x0,y0}. Now the function F1(x,y)=xf(x,y) and F2(x,y)=yg(x,y) of the system (1.4) are continuously differentiable and locally Lipschitz in R2+={(x,y):x0,y0}. Therefore, Theorem A.4, page 423 in H. R. Thieme's book [61] implies that the solutions of the initial value problem with non-negative initial conditions exist on the interval [0,S) and unique, where S is a sufficiently large number.

    In theoretical ecology, boundedness of the system (1.4) implies that the system is well behaved. Boundedness of the solutions entails that none of the interacting species grow abruptly or exponentially for a long-time interval. The number/abundance of each species is bounded due to limited resource. To prove the boundedness of the system (1.4), we state the following lemma whose proof is given in [62].

    Lemma 2.1. Assume that ξ,η>0 with u(0)>0. Then for the differential inequation dudtu(t)(ξηu(t)),

    lim suptu(t)ξη,

    and also for the differential inequation dudtu(t)(ξηu(t)),

    lim inftu(t)ξη.

    Proposition 2.2. Assume that b>d and q>α(bd+ϵ1), where ϵ1>0 is a arbitrary small number. Then all the solutions (x(t),y(t)) of the system (1.4) with initial conditions x(0)>0,y(0)>0 are bounded.

    Proof. From the first equation of the system (1.4)

    dxdt=x[b1+βydx(1+αy)yp+(1+αy)x]
    dxdtx(bdx).

    Assume that b>d. Now applying the Lemma 2.1, we have

    lim suptx(t)bd. (2.1)

    Thus for arbitrary small ϵ1>0, there exists a positive real number T1>0 such that

    x(t)bd+ϵ1,t>T1.

    Also, the second equation of the system (1.4) can be written as

    dydt=y[r+(1+αy)xqy]r+(1+αy)x.

    Thus, for all t>T1, we have

    dydtyr[r+bd+ϵ1(qα(bd+ϵ1))y].

    Assume that q>α(bd+ϵ1). Then again applying the Lemma 2.1, we have

    lim supty(t)r+bd+ϵ1qα(bd+ϵ1). (2.2)

    Thus, for arbitrary small ϵ2>0, there exist a positive real number T2>T1 such that

    y(t)r+bd+ϵ1qα(bd+ϵ1)+ϵ2,t>T2.

    Therefore, the solutions of the system (1.4) are bounded.

    Geometrically, uniform permanence means the existence of a region in the phase plane at a non-zero distance from the boundary in which population species enter and must lie ultimately that ensures the long time survival of species in biological sense. Analytically the system (1.4) is said to be permanent if there exist positive constants m and M such that each positive solution (x(t,x0,y0),y(t,x0,y0)) of the system (1.4) with initial condition (x0,y0)int(R2+) satisfies

    min{lim inftx(t,x0,y0),lim infty(t,x0,y0)}m,max{lim suptx(t,x0,y0),lim supty(t,x0,y0)}M.

    Proposition 2.3. The system (1.4) with initial conditions x(0)>0, y(0)>0 is permanent if b1+βM2d(1+αM2)M2p>0.

    Proof. From the prey equation of the system (1.4), we have

    dxdtx[b1+βM2dx(1+αM2)M2p],x(w0x),whereM2is the upper bound ofypopulationandw0=b1+βM2d(1+αM2)M2p. (2.3)

    Assume that w0>0. Then by applying the Lemma 2.1, we have

    lim inftx(t)w0.

    Also, from the predator equation of system (1.4), we have

    dydtyqy2r=y(1qyr).

    Again, by applying the Lemma 2.1, we get the following result

    lim infty(t)rq.

    Choose m=min{w0,rq} and from (2.1) and (2.2), choose M=max{bd,r+bd+ϵ1qα(bd+ϵ1)}. Then the condition of permanence of the system (1.4) follows.

    The equilibria of the system (1.4) occur at the intersection of prey nullclines and predator nullclines governed by the following equations:

    xf(x,y)=0;
    yg(x,y)=0.

    The non-trivial prey nullcline f(x,y)=0 represents a cubic equation of x and y and non-trivial predator nullcline g(x,y)=0 represents a hyperbolic curve of the form

    (xqα)(y+qα)=(rα+qα2).

    The predator-prey model (1.4) has following trivial and boundary equilibria:

    (ⅰ) Trivial equilibrium point E0=(0,0).

    (ⅱ) Prey only equilibrium point E1=(bd,0).

    (ⅲ) Predator only equilibrium point E2=(0,rq).

    Now, the positive interior equilibrium points of the system (1.4) are the solution of the following system of equations:

    {b1+βydx(1+αy)yp+(1+αy)x=0,1qyr+(1+αy)x=0. (3.1)

    Let (ˉx,ˉy) be a positive solution of the system of equations (3.1). Then we have

    ˉx=qˉyr1+αˉy, (3.2)

    and ˉy is a positive root of the equation

    σ0y4+σ1y3+σ2y2+σ3y+σ4=0, (3.3)

    where

    σ0=α2β,σ1=α2+2αβ+βq2+αβqd,σ2=2α+β+q2+βpq2βqr+(α+β)qd+αβd(pr)bqα,σ3=1+(pr)(qβrbα)+d(α+β)(pr)+q(drb),σ4=(pr)(drb).

    Here σ0 and σ1 are positive, so the above equation (3.3) can have atmost three positive roots depending on the sign of σ2, σ3 and σ4.

    Now using the linearized stability analysis, we can obtain the local stability criteria of the above mentioned equilibrium points. The community matrix at E0=(0,0), E1=(bd,0) and E2=(0,rq) are respectively given by

    JE0=(bd001),JE1=(db(db)(bβ+bdp+bd)01)andJE2=(bqβr+qdr(αr+q)pq20αr+qq21).

    Thus the equilibrium points E0=(0,0) and E1=(bd,0) both are unstable. The prey free equilibrium point E2=(0,rq) is locally asymptotically stable if the cooperation strength α greater than the threshold value bpq3r2(βr+q)dpq2r2qr, or equivalently, the fear strength β greater than the threshold value bpq3r2(αr+q)+dpq2rqr. Also the community matrix at ˉE=(ˉx,ˉy) is given by

    JˉE=(ˉx+ˉxˉy(1+αˉy)2(qˉy+pr)2bβˉx(1+βˉy)2(qˉyr)(p+ˉx+αˉxˉy)+pαˉxˉy(qˉy+pr)21+αˉyqαˉxq1). (3.4)

    The trace and determinant value of the above community matrix are given by

    T(ˉx,ˉy)=trace(JˉE)=1ˉx+αˉxq+ˉxˉy(1+αˉy)2(qˉy+pr)2,D(ˉx,ˉy)=det(JˉE)=(αˉxq1)[ˉx+ˉxˉy(1+αˉy)2(qˉy+pr)2]+(1+αˉyq)[bβˉx(1+βˉy)2+(qˉyr)(p+ˉx+αˉxˉy)+pαˉxˉy(qˉy+pr)2].

    Both of T and D can be positive or negative. So the stability of interior equilibrium point depends on both the trace and determinant value of the community matrix. Therefore, by using Routh-Hurwitz criterion, the interior equilibrium ˉE(ˉx,ˉy) is locally asymptotically stable if T(ˉx,ˉy)<0 and D(ˉx,ˉy)>0.

    Nonlinear mathematical models of interacting populations show rich and complex dynamical behaviours even when system complexity is low (two or three species). Oscillating behaviour is a most frequent dynamical behaviours in population dynamics. Oscillating behaviour or the existence of limit cycle lead to the Hopf-bifurcation of the system. Hopf-bifurcation is defined as the appearance or disappearance of a periodic orbit through a local change in the stability properties of an equilibrium point. Here we explore the possibility of occurrence of Hopf-bifurcation and the direction of Hopf-bifurcation around the interior equilibrium point ˜E(˜x,˜y) with respect to bifurcating parameter α.

    The following theorem gives the necessary and sufficient condition for Hopf-bifurcation of the system (1.4).

    Theorem 4.1. The necessary and sufficient conditions for the system (1.4) undergoes Hopf-bifurcation at α=αh around the interior equilibrium point ˜E are that D(˜x(α),˜y(α))>0 and ddαT(˜x(α),˜y(α))0 at α=αh.

    Proof. Consider the community matrix (3.4) evaluated at ˜E. For Hopf-bifurcation, the matrix J˜E must have a pair of purely imaginary eigenvalues. At a critical value α=αh, T(˜x(α),˜y(α))=0. So the characteristic equation of the community matrix becomes

    λ2+D(˜x(α),˜y(α))|α=αh=0.

    The above equation must have a pair of purely imaginary roots λ1,2=±iθ0, where θ0=D(˜x(α),˜y(α))|α=αh, if D(˜x(α),˜y(α))>0 at α=αh.

    Now we check the transversality condition which confirms that the eigenvalues crosses the imaginary axis of the complex plane with non-zero speed. For this, let at any neighbouring point α of αh, the eigenvalues of the community matrix are λ1,2=χ(α)±iθ(α), where χ(α)=T(˜x(α),˜y(α))2 and θ(α)=D(˜x(α),˜y(α))T2(˜x(α),˜y(α))4. Now,

    ddαχ(α)|α=αh=12[ddαT(˜x(α),˜y(α))]α=αh.

    Thus, the transversality condition satisfied if ddαT(˜x(α),˜y(α))0 at α=αh.

    Therefore, the system (1.4) undergoes Hopf-bifurcation at α=αh.

    Here, we discuss the direction and stability properties of the bifurcating periodic solutions originating from the coexisting equilibrium point ˜E(˜x,˜y) via Hopf-bifurcation. To investigate the stability and direction of Hopf-bifurcation, we calculate the 1st Lyapunov coefficient by using the results given in [63].

    First, we transform the equilibrium point ˜E(˜x,˜y) of the system (1.4) into the origin by letting z1=x˜x,z2=y˜y. Then the system (1.4) becomes

    {dz1dt=b(z1+˜x)1+β(z2+˜y)d(z1+˜x)(z1+˜x)2(1+α(z2+˜y))(z1+˜x)(z2+˜y)p+(1+α(z2+˜y))(z1+˜x),dz2dt=(z2+˜y)q(z2+˜y)2r+(1+α(z2+˜y))(z1+˜x).

    Expanding Taylor's series of the above system at (z1,z2)=(0,0) up to terms of order 3 produces the following system:

    {˙z1=p10z1+p01z2+p20z21+p11z1z2+p02z22+p30z31+p21z21z2+p12z1z22+p03z32+O(|z|4),˙z2=q10z1+q01z2+q20z21+q11z1z2+q02z22+q30z31+q21z21z2+q12z1z22+q03z32+O(|z|4), (4.1)

    where

    (p10p01q10q01)=J˜E,p20=1+p˜y(1+α˜y)2(p+˜x+α˜x˜y)3,p11=bβ(1+β˜y)21+2α˜yp+˜x+α˜x˜y+˜x(1+α˜y)(p+˜x+3α˜x˜y+4pα˜y+2α2˜x˜y2)(p+˜x+α˜x˜y)3,p02=bβ2˜x(1+β˜y)3α˜xp+˜x+α˜x˜y+αx2(p+˜x+2pα˜y+2α˜x˜y+α2˜x˜y2)(p+˜x+α˜x˜y)3,p30=p˜y(1+α˜y)3(p+˜x+α˜x˜y)4,p21=(1+α˜y)(1+3α˜y)(p+˜x+α˜x˜y)2x(1+α˜y)2(1+6α˜y)(p+˜x+α˜x˜y)3+3α˜x2˜y(1+α˜y)(p+˜x+α˜x˜y)4,p12=bβ2(1+β˜y)3αp+˜x+α˜x˜y+α˜x(3+5α˜y)(p+˜x+α˜x˜y)2α˜x2(1+α˜y)(2+5α˜y)(p+˜x+α˜x˜y)3+α2˜x2˜y(1+α˜y)(˜x2p+α˜x˜y)(p+˜x+α˜x˜y)4,p03=β3b˜x(1+β˜y)4+α2˜x2(p+˜x+α˜x˜y)2α2˜x3(1+2α˜y)(p+˜x+α˜x˜y)3+α3˜x4˜y(1+α˜y)(p+˜x+α˜x˜y)4,q20=q˜y2(1+α˜y)2(r+˜x+α˜x˜y)3,q11=q˜y(2+3α˜y)(r+˜x+α˜x˜y)22qα˜x˜y2(1+α˜y)(r+˜x+α˜x˜y)3,q02=q(r+˜x+α˜x˜y)+qα˜x˜y(2r+2˜x+α˜x˜y)(r+˜x+α˜x˜y)3,q30=q˜y2(1+α˜y)3(r+˜x+α˜x˜y)4,q12=q(1+3α˜y)(r+˜x+α˜x˜y)22qα˜x˜y(2+3α˜y)(r+˜x+α˜x˜y)3+3qα2˜x2˜y2(1+α˜y)(r+˜x+α˜x˜y)4,q21=2q˜y(1+α˜y)(1+2α˜y)(r+˜x+α˜x˜y)3+3qα˜x˜y2(1+α˜y)2(r+˜x+α˜x˜y)4,q03=qα˜x(r+˜xα˜x˜y)(r+˜x+α˜x˜y)3+qα3˜x3˜y2(r+˜x+α˜x˜y)4.

    By neglecting the higher order terms of degree 4 and above, the system (4.1) can be written as

    ˙Z=J˜EZ+A(Z), (4.2)

    where

    Z=(z1z2)andA=(A1A2)=(p20z21+p11z1z2+p02z22+p30z31+p21z21z2+p12z1z22+p03z32q20z21+q11z1z2+q02z22+q30z31+q21z21z2+q12z1z22+q03z32).

    The eigenvector ˜v of community matrix J˜E corresponding to the eigenvalue iθ0 at α=αh is ˜v=(p01,iθ0p10)T. Now define

    Q=(Re(˜v),Im(˜v))=(p010p10θ0).

    Let Z=QU or U=Q1Z, where U=(u,v)T. Under this transformation, system (4.2) becomes

    ˙U=(Q1J˜EQ)U+Q1A(QU).

    This can be written as,

    (˙u˙v)=(0θ0θ00)(uv)+(S1(u,v;α=αh)S2(u,v;α=αh)), (4.3)

    where S1 and S2 are non-linear in u and v given by,

    S1(u,v;α=αh)=1p01A1,S2(u,v;α=αh)=1θ0p01(p10A1+p01A2),

    with

    A1=(p20p201p11p01p10+p02p210)u2+θ0(2p02p10p11p01)uv+θ20p02v2+(p12p01p210p03p310+p30p301p21p201p10)u3+θ0(2p12p10p01p21p2013p03p210)u2v+θ20(p12p013p03p10)uv2θ30p03v3,A2=(q20p201q11p01p10+q02p210)u2+θ0(2q02p10q11p01)uv+θ20q02v2+(q30p301+q12p01p210q21p201p10q03p310)u3+θ0(2q12p01p10q21p2013q03p210)u2v+θ20(q12p013q03p10)uv2θ30q03v3.

    Now, we calculate the 1st Lyapunov coefficient, based on the normal form (4.3), which determines the stability and direction of periodic solution. The 1st Lyapunov coefficient is

    l1=116[S1uuu+S1uvv+S2uuv+S2vvv]+116θ0[S1uv(S1uu+S1vv)S2uv(S2uu+S2vv)S1uuS2uu+S1vvS2vv],

    where all the partial derivatives are calculated at the bifurcation point, i.e., (u,v;α)=(0,0;αh).

    By applying the result given in [63] (see equation (20.2.14), page 385), Hopf-bifurcation is supercritical if l1<0 and it is subcritical if l1>0. It is to be noted that when l1=0, the system (1.4) exhibits generalized Hopf-bifurcation (or, Bautin bifurcation) at which the interior equilibrium has a pair of purely imaginary eigenvalues. The generalized Hopf-bifurcation point separates branches of subcritical and supercritical Hopf-bifurcations in the parameter plane.

    In this section, we discuss Bogdanov-Takens (BT) bifurcation of the system (1.4) in a small neighbourhood of the interior equilibrium point ˜E(˜x,˜y). BT bifurcation is a bifurcation of an equilibrium point in a two-parameter family of autonomous system and occurs for a system when the community matrix evaluated at the equilibrium has a zero eigenvalue of algebraic multiplicity 2. In such a situation the community matrix is similar to the Jordan block of the form (0100). We follow the techniques and steps of Kuznetsov's book [64] to derive a normal form by using a series of transformations. Now we consider the system (1.4) and choose α and p as an arbitrary family of parameters and keeping all the other parameters are fixed. Without loss of generality, we define ˜x(α,p)˜x and ˜y(α,p)˜y. It is to be noted that BT bifurcation may occur for other parameters also. Since the all the parameters are positive and y is non-negative, so the functions F1(x,y)=xf(x,y) and F2(x,y)=yg(x,y) in the system of equations (1.4) are smooth function of x and y. Assume that at α=αbt and p=pbt, T(˜x,˜y)=0,D(˜x,˜y)=0. Then the community matrix in (3.4) evaluated at ˜E(˜x,˜y) has a zero eigenvalue of algebraic multiplicity 2. Now we consider the following parametric region

    ˉΩ={(b,β,d,α,p,q,r)R7+:T(˜x,˜y)=0,D(˜x,˜y)=0}.

    Mathematically, the surface represented by ˉΩ is called a BT surface. In the following, we reduce the system (1.4) in the canonical form of a BT bifurcation by employing a series of C change of coordinates [64] in a small neighborhood of (0,0). For these, first we use the transformation x=u1+˜x and y=u2+˜y, so that the equilibrium point of the system (1.4) shifted to the origin. Then the model system (1.4) can be written as

    {du1dt=(u1+˜x)[b1+β(u2+˜y)d(u1+˜x)(1+α(u2+˜y))(u2+˜y)p+(1+α(u2+˜y))(u1+˜x)],du2dt=(u2+˜y)[1q(u2+˜y)r+(1+α(u2+˜y))(u1+˜x)]. (5.1)

    Taking the small perturbation of the model (5.1) at α=αbt and p=pbt, we get

    {du1dt=(u1+˜x)[b1+β(u2+˜y)d(u1+˜x)(1+(αbt+δ1)(u2+˜y))(u2+˜y)(pbt+δ2)+(1+(αbt+δ1)(u2+˜y))(u1+˜x)],du2dt=(u2+˜y)[1q(u2+˜y)r+(1+(αbt+δ1)(u2+˜y))(u1+˜x)],

    where δ=(δ1,δ2) is a parameter vector vary in a small neighborhood of (0,0). Expanding the Taylor's series of the above system at (u1,u2)=(0,0) up to the terms of order 2, produces the following system:

    {du1dt=a00(δ)+a10(δ)u1+a01(δ)u2+a20(δ)u21+a11(δ)u1u2+a02(δ)u22+O(|u|3),du2dt=b00(δ)+b10(δ)u1+b01(δ)u2+b20(δ)u21+b11(δ)u1u2+b02(δ)u22+O(|u|3), (5.2)

    where

    a00=b˜x1+β˜yd˜x˜x2(1+(αbt+δ1)˜x˜y)˜y(pbt+δ2)+(1+(αbt+δ1)˜y)˜x,a10=b1+βyd2˜x˜y+(α+δ1)˜y2(p+δ2)+˜x+(α+δ1)˜x˜y+˜x˜y(1+(α+δ1)˜y)((p+δ2)+˜x+(α+δ1)˜x˜y)2,a01=bβ˜x(1+βy)2˜x+2(α+δ1)˜x˜y(p+δ2)+˜x+(α+δ1)˜x˜y+2˜x2˜y(α+δ1)(1+(α+δ1)˜y)((p+δ2)+˜x+(α+δ1)˜x˜y)2,a20=1+r˜y(1+(α+δ1)˜y)2((p+δ2)+˜x+(α+δ1)˜x˜y)3,a11=bβ(1+β˜y)21+2(α+δ1)˜y(p+δ2)+˜x+(α+δ1)˜x˜y+˜x(1+(α+δ1)˜y)((p+δ2)+˜x+(α+δ1)˜x˜y)2+2˜x2˜y(α+δ1)(1+(α+δ1)˜y)(2(p+δ2)+1+(α+δ1)˜y)((p+δ2)+˜x+(α+δ1)˜x˜y)3,a02=bβ2˜x(1+β˜y)3(α+δ1)˜x(p+δ2)+˜x+(α+δ1)˜x˜y+(α+δ1)˜x2((p+δ2)+˜x+(α+δ1)˜x˜y)2+˜x2˜y(α+δ1)(2(p+δ2)+˜x+(α+δ1)˜x˜y)((p+δ2)+˜x+(α+δ1)˜x˜y)3,b00=˜yq˜y2r+(1+(αbt+δ1)˜y)˜x,b10=q˜y2(1+(α+δ1)˜y)(r+˜x+(α+δ1)˜x˜y)2,b01=1q˜y(2r+2˜x+(α+δ1)˜x˜y)(r+˜x+(α+δ1)˜x˜y)3,b20=q˜y2(1+(α+δ1)˜y)2(r+˜x+(α+δ1)˜x˜y)3,b11=q˜y(2+3(α+δ1)˜y)(r+˜x+(α+δ1)˜x˜y)22q(α+δ1)˜x˜y2(1+(α+δ1)˜y)(r+˜x+(α+δ1)˜x˜y)3,b02=q(r+˜x+(α+δ1)˜x˜y)+q(α+δ1)˜x˜y(2r+2˜x+(α+δ1)˜x˜y)(r+˜x+(α+δ1)˜x˜y)3.

    Here note that a00(0)=0=b00(0) and

    a10(0)+b01(0)=0,a10(0)b01(0)a01(0)b10(0)=0.

    Now we introduce affine transformation

    v1=u1,v2=a10u1+a01u2,

    which reduce the system (5.2) to

    {dv1dt=a00(δ)+v2+c20(δ)v21+c11(δ)v1v2+c02(δ)v22+O||v||3),dv2dt=d00(δ)+d10(δ)v1+d01(δ)v2+d20(δ)v21+d11(δ)v1v2+d02(δ)v21+O(||v||3), (5.3)

    where

    c20(δ)=a20a11a10a01+a02a210a201,c11(δ)=a11a012a10a02a201,c02(δ)=a02a210,d00(δ)=a10a00+a01b00,d10(δ)=a01b10a10b01,d01(δ)=a10+b01,d20(δ)=a10a20a11a210a01+a02a310a201+a01b20a10b11+b02a210a01,d11(δ)=b11+a11a10a012a210a02a2012a10b02a01,d02(δ)=a10a02a201+b02a01.

    Next, under the following C change of coordinates in a small neighbourhood of (0,0)

    x1=v1,x2=d00(δ)+d10(δ)v1+d01(δ)v2+d20(δ)v21+d11(δ)v1v2+d02(δ)v21+O(||v||3),

    the system (5.3) becomes

    {dx1dt=x2,dx2dt=e00(δ)+e10(δ)x1+e01(δ)x2+e20(δ)x21+e11(δ)x1x2+e02(δ)x22+O(||x||3), (5.4)

    where

    e00(δ)=d00(δ)a00(δ)d01(δ)+a200(δ)(d02(δ)c02(δ)d01(δ))2a00(δ)c02(δ)(d00(δ)a00(δ)d01(δ))+,=d00(δ)a00(δ)d01(δ)+a200(δ)c02(δ)d01(δ)+a200(δ)d02(δ)+,e10(δ)=d10(δ)+a00(δ)c11(δ)d01(δ)a00(δ)d11(δ)+c11(δ)(d00(δ)a00(δ)d01(δ))2a00(δ)c02(δ)d10(δ)+,=d10(δ)+c11(δ)d00(δ)a00(δ)d11(δ)2a00(δ)c02(δ)d10(δ)+,e01(δ)=d01(δ)a00(δ)c11(δ)+2a00(δ)c02(δ)d01(δ)2a00(δ)d02(δ)+2c02(δ)d00(δ)+,e20(δ)=d20(δ)+c11(δ)d10(δ)d01(δ)c20(δ)+,e11(δ)=c11(δ)d01(δ)+d11(δ)+2c20(δ)+c11(δ)d01(δ)+2c02(δ)d10(δ)=d11(δ)+2c20(δ)+2c02(δ)d10(δ)+,e02(δ)=d02(δ)c02(δ)d01(δ)+c11(δ)+2c02(δ)d01(δ)+.

    It is easily observed that e00(0)=0=e10(0)=e01(0), since d00(0)=0=d10(0)=d01(0) and

    e20(0)=d20(0),e11(0)=d11(0)+2c20(0),e02(0)=d02(0)+c11(0).

    We now use the following parameter-dependent shift of coordinates in the x1-direction to annihilate the x2 term on the RHS of the second equation of (5.4)

    x1=y1+ϕ(δ),x2=y2.

    Using the above transformation system (5.4) reduces to

    {dy1dt=y2,dy2dt=f00(δ)+f10(δ)y1+f01(δ)y2+f20(δ)y21+f11(δ)y1y2+f02(δ)y22+O(||y||3), (5.5)

    where

    f00=e00+e10ϕ+e20ϕ2+,f10=e10+2e20ϕ+,f01=e01+e11ϕ+,f20=e20+,f11=e11+,f02=e02+.

    Assume that

    e11(0)=d11(0)+2c20(0)0. (BT.1)

    Using standard arguments based on the Implicit Function Theorem gives local existence of a smooth function

    ϕ=ϕ(δ)e01(δ)e11(0)

    annihilating the term proportional to y2 in the 2nd equation of (5.5). Here we note that ϕ(0)=0, thus f00(0)=0=f10(0)=f01(0), f11(0)=e11(0) and f20(0)=e20(0). Now we introduce a new time scale, defined by dt=(1+Λz1)dθ, where Λ=Λ(δ) is a smooth function to be defined later. Also the direction of time is preserved near the origin for small ||δ||. This transformation reduces the system (5.5) to

    {dy1dθ=y2+Λy1y2,dy2dθ=f00+(f10+f00Λ)y1+(f20+f10Λ)y21+f11y1y2+f02y22+O(||y||3). (5.6)

    Now assuming z1=y1,z2=y2(1+Λy1), the system (5.6) becomes

    {dz1dθ=z2,dz2dθ=f00+(f10+2f00Λ)z1+(f20+2f10Λ+f00Λ2)z21+f11z1z2+(f02+Λ)z22+O(||z||3).

    To remove the z22 term we specify the time reparametrization as Λ(δ)=f02(δ). Consequently, we have

    {dz1dθ=z2,dz2dθ=f00(δ)+L(δ)z1+M(δ)z21+N(δ)z1z2+O(||z||3), (5.7)

    where

    L(δ)=f10(δ)2f00(δ)f02(δ),M(δ)=f20(δ)2f10(δ)f02(δ)+f00(δ)f202(δ),N(δ)=f11(δ).

    We now introduce a new time scale and denote it by t again as follows t=|N(δ)M(δ)|θ. Since N(0)=f11(0)=e11(0)0 due to (BT.1), the time scaling above will be well defined if we further assume

    M(0)=f20(0)2f10(0)f02(0)+f00(0)f202(0)=f20(0)=e20(0)=d20(0)0. (BT.2)

    Simultaneously, we introduce new variables χ=N2(δ)M(δ)z1, ψ=sign(N(δ)M(δ))N3(δ)M2(δ)z2, which takes the system (5.7) into the following form

    {dχdt=ψ,dψdt=β1+β2χ+χ2+sχψ+O((χ,ψ)3), (5.8)

    where

    β1(δ)=N4(δ)M3(δ)f00(δ),β2(δ)=N2(δ)M2(δ)Land
    s=sign(N(0)M(0))=sign(d11(0)+2c20(0)d20(0))=±1.

    In order to define an invertible smooth change of parameters near the origin, we assume that rank(βδ)|δ=0=2, i.e., at δ=(δ1,δ2)=(0,0)

    |β1δ1β1δ2β2δ1β2δ2|0. (BT.3)

    The system (5.8) is called normal form of Bogdanov-Takens bifurcation. Thus under the assumptions of non-degeneracy conditions (BT.1), (BT.2) and (BT.3), the system (1.4) can be reduced to the system (5.8). Also by using Lemma 8.8, page 324 in [64], the system (5.8) is locally topologically equivalent near the origin to the system

    {dχdt=ψ,dψdt=β1+β2χ+χ2±χψ. (5.9)

    Thus the system (1.4) undergoes BT bifurcation by choosing α and p as a bifurcation parameter when the parameters δ1 and δ2 vary in a small neighbourhood of the origin.

    To justify our analytical findings and to depict the effect of cooperation and fear on the dynamics of the system (1.4), we perform extensive numerical simulations. In the present paper, all the numerical findings are performed in Matlab by using ode45 method and Matcont package with suitably chosen small time steps. In order to understand the role of cooperation and fear on the stability and bifurcation of the coexistence equilibrium point, first we consider the system (1.4) without hunting cooperation (α=0) and without fear effect (β=0). We choose the parameter values as

    b=5,d=0.7,p=5,q=1,r=1, (6.1)

    with initial condition (x0,y0)=(4,1). First, we observe that the system (1.4) for the above set of parameter choice shows stable dynamics (Figure 1).

    Figure 1.  Figure illustrating the stable behaviour of the system (1.4), around the unique interior equilibrium point ˜E(˜x,˜y)=(3.7568,4.7568), where α=0, β=0 and other parameters are given in (6.1).

    Now, our aim is to investigate the dynamics of the system with hunting cooperation (α0) but without fear effect (β=0). We observe that an increase in the strength of hunting cooperation makes the system unstable, and both the predator and prey populations show periodic oscillations (Figure 2). The system (1.4) loses its stability via Hopf-bifurcation at α=0.512218(αl). We also observe that further increase of the strength of α, the system (1.4) undergoes another Hopf-bifurcation at α=6.510087(αu) and becomes stable. However, for very large values of α(α>αe=20.5) prey population goes to extinction due to overexploitation by the predator population (Figure 3). Therefore, the system (1.4) undergoes multiple Hopf-bifurcations at αl=0.512218 and αu=6.510087, and shows periodic oscillation when strength of hunting cooperation is in the range αl<α<αu. We notice that there is a hump-shaped relationship between predator density at the equilibrium and the parameter α (Figure 3). The predator density increases when the strength of hunting cooperation low, but declines when the strength of hunting cooperation large due to overexploitation of prey by the predators. Such type of relationship between predator density at the equilibrium (˜y) and the cooperation strength (α) was also observed by Teixeira Alves and Hilker [41] in a Lotka-Volterra predator-prey model with hunting cooperation among predators. The biological implications for predator population increases for small values of α because of predators better forage on prey, but for large values of α predator population decreases as decrease in prey density. Also, due to overexploitation of predators, extinction of prey population occurs at the threshold value α=20.5 (Figure 4).

    Figure 2.  Figure depicts that cooperative behaviour of predators destabilize the predator-prey dynamics of the system (1.4), where α=1, β=0 and other parameters are given in (6.1). In the right figure, black asterisk denotes the unique interior equilibrium point ˜E(˜x,˜y)=(0.6840,5.3298).
    Figure 3.  The system (1.4) exhibits multiple Hopf-bifurcations at α=0.512218(αl) and at α=6.510087(αu) respectively, where β=0 and other parameters are given in (6.1).
    Figure 4.  Figure depicts the extinction of prey population of the system (1.4) at α=20.5, where β=0 and other parameters are given in (6.1).

    A supercritical Hopf-bifurcation means that the stable periodic solution overlaps an unstable equilibrium point. In the Subsection 4.2, we have obtained the sufficient condition for supercritical Hopf-bifurcation of the system (1.4) theoretically. Now for the set of parameter values (6.1), we calculate the 1st Lyapunov coefficient corresponding to the threshold value α=αl. The 1st Lyapunov coefficient is Ll=0.072533, which confirms that the limit cycle obtained in our system is stable i.e. the trajectories starting from different initial conditions converge to the same limit cycle. We draw the limit cycle when α=1 and observe that the stable interior equilibrium point becomes unstable and a stable limit cycle is born (Figure 5). Therefore, the system undergoes supercritical Hopf-bifurcation.

    Figure 5.  Existence of stable limit cycle of the system (1.4), where α=1, β=0 and other parameters are given in (6.1). Blue and magenta trajectories start from two different initial conditions (4,1) and (2,4), respectively. Red asterisk denotes the unstable equilibrium point (0.6840,5.3298).

    A subcritical bifurcation means that the unstable periodic solution overlaps a stable equilibrium point. Now, we calculate the first Lyapunov coefficient for the system (1.4) at the threshold value α=αu. The first Lyapunov coefficient is Lu=0.069969, which confirms that there exists a subcritical Hopf-bifurcation. We draw the solution trajectories when α=6.7 and observe that the unstable equilibrium point becomes stable and there is an unstable limit cycle present around the stable equilibrium point (0.0635,1.8501) (Figure 6). We also observe that the system shows a stable limit cycle outside of the unstable limit cycle. So, the system (1.4) also exhibits subcritical Hopf-bifurcation and hunting cooperation can change the stability property of the limit cycle. The subcritical Hopf- bifurcation leads to a significant domain of bi-stability, where stable steady state and time-periodic state coexist. Here, the prey and predator population tend to the interior equilibrium point, if the initial populations start inside the unstable limit cycle and if the initial populations start outside the unstable limit cycle, then the prey and predator population converge to the stable limit cycle. So, in the absence of fear effect (β=0) the system (1.4) shows stable dynamics and limit cycle (stable/unstable) oscillations depending on the strength of hunting cooperation α.

    Figure 6.  Multiple limit cycles (both stable and unstable) of the system (1.4), where α=6.7, β=0 and other parameters are given in (6.1). Green trajectory start from the initial condition (0.2,2.73). Blue and magenta trajectories start from two different initial conditions (0.6,6.6) and (0.6,4), respectively. Red asterisk denotes the stable equilibrium point (0.0635,1.8501).

    Now, we are going to investigate the impact of fear effect on the dynamics of the system (1.4). We observe that for a fixed value of α, if we increase the strength of fear parameter (β), then the system (1.4) remains stable or becomes stable. Here, we choose two different values of α, where in the absence of fear effect (β=0) the system (1.4) shows stable focus (α=0.1) and periodic oscillations (α=1). Now, if we incorporate fear phenomenon into the system with fear strength β=0.1, then stable system remains stable and oscillating system becomes stable (Figure 7). We also draw two parametric bifurcation diagram in αβ parametric space and observe that fear factor (β) has a stabilizing effect on the dynamics of the system (Figure 8). It is also observed that the system (1.4) exhibits GH bifurcation (generalized Hopf bifurcation) at the threshold value α=4.7255712(αgh) and β=0.062264(βgh), where the 1st Lyapunov coefficient becomes zero. Further, for large values of hunting cooperation (α) and fear factor (β), the prey population goes to extinction from the system whereas predator population converges to a fixed value (Figure 9).

    Figure 7.  Figure shows the system (1.4) with fear phenomenon is more stable than the system without fear phenomenon. Red curves denote the solution trajectories of the system without fear effect and green curves denote the solution trajectories of the system with fear effect.
    Figure 8.  Figure depicts the two parametric bifurcation diagram in αβ parametric space. Blue curve denotes the Hopf-bifurcation curve of the system (1.4), where the switching of stability occurs. Other parameters are same as in (6.1).
    Figure 9.  Figure shows region of survival and extinction of prey population of the system (1.4) in α-β parametric space, where other parameters are same as in (6.1).

    Previously, we have observed the bi-stability behaviour between coexistence equilibrium and periodic oscillations. Here, we also observe bi-stability behaviour between prey-free equilibrium point and coexistence equilibrium point. First, we draw the bifurcation diagram of the system (1.4) with respect to the parameter α (Figure 10), where p=0.3, β=0.1 and other parameters are same as in (6.1). Here, we draw black curve for stable branch and magenta curve for unstable branch of the interior equilibrium, respectively. We also draw green line for stable branch and red line for unstable branch of the prey-free equilibrium, respectively. We observe that there exists an interval (0.1536<α<0.2212), where both the interior equilibrium point and the prey-free equilibrium point are stable (Figure 10). Further, by considering parameters from the bistable region, we draw the basin of attraction of the prey-free and interior attractors. We observe that for initial conditions in the black region, the trajectories converge to the interior equilibrium point (1.1371,2.7662) and for initial conditions in green region the trajectories converge to the prey-free equilibrium point (0,1) (Figure 11(a)). For better visualization, we also draw phase portrait of the system (1.4) by considering different initial conditions from green and black regions (Figure 11(b)).

    Figure 10.  Figure shows bifurcation diagram of the system (1.4) with respect to the bifurcation parameter α. Here, p=0.3, β=0.1 and other parameters are same as in (6.1). Black curve and green line denote the stable branches of coexisting equilibrium point and prey-free equilibrium (PFE) point, respectively.
    Figure 11.  Figure depicts the bi-stability behaviour of the system (1.4). Left figure shows basin of attraction of two different attractors. In the right figure, green trajectories converge to the prey-free equilibrium point (0,1) and black trajectories converge to the coexisting equilibrium point (1.1371,2.7662). Here, α=0.2, p=0.3, β=0.1 and other parameters are same as in (6.1).

    The Bogdanov-Takens bifurcation (BT bifurcation) is a two-parametric bifurcation of an equilibrium point and at the bifurcation point, the community matrix evaluated at the interior equilibrium point has a zero eigenvalue of algebraic multiplicity two. In the neighbourhood of the bifurcation point, the system has two equilibrium points (one saddle and other non-saddle), which will collide and disappear via a saddle-node bifurcation. The non-saddle equilibrium undergoes Hopf-bifurcation, which degenerates into an orbit homoclinic to the saddle and disappears via a saddle homoclinic bifurcation. We draw two parametric bifurcation diagram in αp parametric space, when β=0.1 and other parameters are same as in (6.1). We observe that the system undergoes BT bifurcation at α=0.8993835(αbt) and p=0.4800637(pbt) (Figure 12). It is also observed that when both the eigenvalues are zero at (αbt,pbt), there exists interior equilibrium point (0.0878,1.1810), which is a cusp of codimension 2 (Figure 13).

    Figure 12.  Figure shows the system (1.4) undergoes Bogdanov-Taken bifurcation in αp parametric space, where β=0.1 and other parameters are given in (6.1).
    Figure 13.  Figure shows that the interior equilibrium of the system (1.4) is a cusp of codimension 2, where both the eigenvalues becomes zero. Here, α=αbt=0.8993835 and p=pbt=0.4800637, β=0.1 and other parameters are given in (6.1).

    Again, we consider the parameter values as

    b=5,d=1,p=0.5,q=0.25,r=0.25, (6.2)

    so that the system (1.4) without hunting cooperation (α=0) and without fear effect (β=0) shows periodic oscillations (Figure 14). First, we draw the bifurcation diagram of the system (1.4) w.r.t. cooperation parameter α (with β=0) and observe that hunting cooperation alone can stabilize the oscillating system via Hopf-bifurcation (Figure 15). Again, we draw the bifurcation diagram of the system (1.4) w.r.t. fear parameter β (α=0) and observe that fear phenomenon alone can also stabilize the oscillating system via Hopf-bifurcation (Figure 16). It is also observed that fear factor (β) has more stabilizing effect compare to hunting cooperation (α).

    Figure 14.  Figure shows periodic oscillation of the system (1.4) for α=0, β=0 and the other parameters are given in (6.2). Black asterisk denotes the unique interior equilibrium point ˜E(˜x,˜y)=(0.7808,4.1231).
    Figure 15.  Bifurcation diagram shows that hunting cooperation (α) can stabilize the system (1.4), where β=0 and other parameters are given in (6.2).
    Figure 16.  Bifurcation diagram shows that fear factor (β) can stabilize the system (1.4), where α=0 and other parameters are given in (6.2).

    In ecology and evolutionary biology, the mechanism of predator-prey interaction is a central topic. Both prey and predators use different strategies to maximize their biomass. To increase the success rate of catching a prey, many predators cooperate during hunting. On the other hand, due to fear of predation, prey populations show different types of anti-predator behaviours as their survival strategies. Prey population show anti-predator defences such as changes its habitat zone, forage less, increase vigilance, adjust their reproductive strategies etc. Such type of anti-predator behaviour, reduces the birth rate of prey populations.

    In the present paper, we have considered a modified Leslie-Gower model with hunting cooperation and fear. Our main assumptions are: (ⅰ) predators can cooperate during hunting which increases their effectual consumption and therefore increases the carrying capacity of the predator population; (ⅱ) fear of predation can reduce the birth rate of prey population. We investigated the model step by step by considering different strengths of hunting cooperation and fear factor. To study the impact of hunting cooperation and fear in the predator-prey system, we considered two situations where the ecosystem showed stable dynamics and unstable (limit cycle oscillations) dynamics. Then we gradually increased the strength of hunting cooperation and the cost of fear and explored the impact of cooperation and fear in the ecological system. We observed that in the absence of fear if we increase the strength of hunting cooperation, then the system becomes unstable around the interior equilibrium and undergoes multiple Hopf-bifurcations, where the first Hopf-bifurcation is supercritical and the second one is a subcritical. So, the coexisting equilibrium point may lose stability via supercritical Hopf-bifurcation with an intermediate value of hunting cooperation and regain stability via subcritical Hopf-bifurcation if the strength of hunting cooperation is large. We have also observed the existence of multiple limit cycles and bi-stability phenomena via subcritical Hopf-bifurcation. Here, we found that the solution trajectories tend to a coexisting equilibrium point or oscillate periodically depending on the initial population size. We also investigated the impact of hunting cooperation and fear on the dynamics of the system when the system shows limit cycle oscillations around the interior equilibrium. We observed that if we increase the strength of hunting cooperation or cost of fear, then the population oscillations will be replaced by stable focus. Further, we have demonstrated the impact of the fear phenomenon that can have in the predator-prey interactions, when predators show cooperative behaviour during hunting. We have observed that fear factor has a stabilizing effect and with fear effect, the stable system remains stable and the oscillating system becomes stable by excluding the existence of periodic solutions. We also found that fear factor has more stabilizing effect compared to hunting cooperation and makes the system more robust. It is interesting to note that with increase in hunting cooperation and fear factor, the density of both prey and predator populations eventually will decrease. From the two parametric bifurcation diagram, we observed that for any cooperation strength, if the fear parameter is increased, then the system becomes stable. However, for low level of fear, if the cooperation level is increased then the system switches it's dynamics from stable to limit cycle oscillation and from limit cycle oscillation to stable focus. It is also to be noted that when the strengths of hunting cooperation and fear factor are very high, prey population goes to extinction due to overexploitation by the predators.

    Cooperative hunting is a form of foraging facilitation, whose special feature is that it generates strong Allee effect in predators [40,41] in Lotka-Volterra model. Recently, in another manuscript, we have modified the Lotka-Volterra model by assuming that cooperation among predators induces fear in prey population [65]. We have observed strong demographic Allee phenomenon among predators, backward bifurcation and many complex dynamical behaviors. In the present paper, since we have considered a modified Leslie-Gower model (logistic-type growth for predators), Allee phenomenon is not possible in our model, although the predators show cooperative behaviour during hunting. We like to point out that in Lotka-Volterra model, hunting cooperation among predators enhances the growth rate of predator whereas in the modified Leslie-Gower model, predator hunting cooperation has a positive effect on the carrying capacity of predator. Allee phenomenon has ecological significance on the perspective of the persistence of the species. In Lotka-Volterra model with predator hunting cooperation Allee phenomenon is common [40,41,65] whereas for modified Leslie-Gower model this is not obvious. Thus from a predator persistence point of view modified Leslie-Gower model is more robust compared to Lotka-Volterra model. The system with hunting cooperation and fear phenomena exhibits rich dynamical behaviours including both supercritical and subcritical Hopf-bifurcation, bi-stability between steady state and limit cycle, bi-stability between prey-free equilibrium point and coexisting equilibrium point, generalized Hopf-bifurcation, Bogdanov-Takens bifurcation. So, hunting cooperation and fear effect are important factors and play vital roles in determining the long-term population dynamics of ecosystems stability.

    Research work of Saheb Pal is supported by the Junior Research Fellowship from the UGC, Government of India. The authors are thankful to the three anonymous reviewers and the Editor for their valuable comments and suggestions, which helped us to improve the paper.

    The authors declare there is no conflict of interest in this paper.



    [1] P. Leslie and J. Gower, The properties of a stochastic model for the predator-prey type of interac-tion between two species, Biometrika, 47 (1960), 219–234.
    [2] P. Leslie, Some further notes on the use of matrices in population mathematics, Biometrika, 35 (1948), 213–245.
    [3] M. Aziz-Alaouiand M.D. Okiye, Boundednessand globalstability for apredator-prey modelwith modified LeslieGower and Holling-type II schemes, Appl. Math. Lett., 16 (2003), 1069–1075.
    [4] A. Nindjin, M. Aziz-Alaoui and M. Cadivel, Analysis of a predator–prey model with modified Leslie–Gower and Holling-type II schemes with time delay, Nonlinear Anal. R. World Appl., 7 (2006), 1104–1118.
    [5] R. Gupta and P. Chandra, Bifurcation analysis of modified Leslie–Gower predator–prey model with Michaelis–Menten type prey harvesting, J. Math. Anal. Appl., 398 (2013), 278-295.
    [6] Y. Zhu and K. Wang, Existence and global attractivity of positive periodic solutions for a preda-torprey model with modified Leslie–Gower Holling-type II schemes, J. Math. Anal. Appl., 384 (2011), 400–408.
    [7] P. E. Stander, Cooperative hunting in lions: the role of the individual, Behav. Ecol. Sociobiol., 29 (1992), 445–454.
    [8] S. Creel and N. M. Creel, Communal hunting and pack size in African wild dogs, Lycaon pictus, Animal Behav., 50 (1995), 1325–1339.
    [9] C. Boesch, Cooperative hunting in wild chimpanzees, Animal Behav., 48 (1994), 653–667.
    [10] D. L. Mech, The Wolf, Natural History Press, New York, 1970.
    [11] D. P. Hector, Cooperative hunting and its relationship to foraging success and prey size in an avian predator, Ethology, 73 (1986), 247–257.
    [12] P. M. Kappeler and C. P. Van Schaik, Cooperation in Primates and Humans: Mechanisms and Evolution, Springer, Berlin, 2006.
    [13] P. S. Rodman, Inclusive fitness and group size with a reconsideration of group sizes in lions and wolves, Am. Nat., 118 (1981), 275–283.
    [14] J. McNutt, L. Boggs, H. Heldring, et al., Running wild: dispelling the myths of the African wild dog, Smithsonian Institution Press, Washington D.C., 1996.
    [15] I. Bailey, J. P. Myatt and A. M. Wilson, Group hunting within the carnivora: physiological, cognitive and environmental influences on strategy and cooperation, Behav. Ecol. Sociobiol., 67 (2013), 1–17.
    [16] M. Fox, Behaviour of wolves dogs and related canids, Harper and Row, New York, 1971.
    [17] D. W. Macdonald, The ecology of carnivore social behaviour, Nature, 301 (1983), 379.
    [18] J. C. Bednarz, Cooperative hunting Harris' hawks (Parabuteo unicinctus), Science, 239 (1988), 1525–1527.
    [19] T. Pitcher, A. Magurran and I. Winfield, Fish in larger shoals find food faster, Behav. Ecol. Sociobiol., 10 (1982), 149–151.
    [20] H. J. Brockmann and C. Barnard, Kleptoparasitism in birds, Animal Behav., 27 (1979), 487–514.
    [21] J. A. Vucetich, R. O. Peterson and T. A. Waite, Raven scavenging favours group foraging in wolves, Animal Behav., 67 (2004), 1117–1126.
    [22] C. Packer and L. Ruttan, The evolution of cooperative hunting, Am. Nat., 132 (1988), 159–198.
    [23] S. L. Lima, Nonlethal effects in the ecology of predator-prey interactions, Bioscience, 48 (1998), 25–34.
    [24] W. Cresswell, Predation in bird populations, J. Ornithol., 152 (2011), 251–263.
    [25] 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.
    [26] S. Creel, D. Christianson, S. Liley, et al., Predation risk affects reproductive physiology and demography of elk, Science, 315 (2007), 960.
    [27] L. Y. Zanette, A. F. White, M. C. Allen, et al. Perceived predation risk reduces the number of offspring songbirds produce per year, Science, 334 (2011), 1398–1401.
    [28] F. Hua, K. E. Sieving, R. J. Fletcher, et al., Increased perception of predation risk to adults and offspring alters avian reproductive strategy and performance, Behav. Ecol., 25 (2014), 509–519.
    [29] J. P. Suraci, M. Clinchy, L. M. Dill, et al., Fear of large carnivores causes a trophic cascade, Nat. Commun., 7 (2016), 10698.
    [30] M. J. Peers, Y. N. Majchrzak, E. Neilson, et al., Quantifying fear effects on prey demography in nature, Ecology, 99 (2018), 1716–1723.
    [31] A. Sih, Optimal behavior: can foragers balance two conflicting demands?, Science, 210 (1980), 1041–1043.
    [32] A. Sih, Prey uncertainty and the balancing of antipredator and feeding needs, Am. Nat., 139 (1992), 1052–1069.
    [33] A. I. Houston and J. M. McNamara, Models of adaptive behaviour: an approach based on state, Cambridge University Press, Cambridge, 1999.
    [34] Z. Abramsky, M. L. Rosenzweig and A. Subach, The costs of apprehensive foraging, Ecology, 83 (2002), 1330 –1340.
    [35] M. A. Elgar, Predator vigilance and group size in mammals and birds: a critical review of the empirical evidence, Biol. Rev., 64 (1989), 13–33.
    [36] A. Sih and T. M. McCarthy, Prey responses to pulses of risk and safety: testing the risk allocation hypothesis, Animal Behav., 63 (2002), 437–443.
    [37] D. T. Blumstein and J. C. Daniel, Isolation from mammalian predators differentially affects two congeners, Behav. Ecol., 13 (2002), 657–663.
    [38] S. Creel and J. A. Winnie Jr., Responses of elk herd size to fine-scale spatial and temporal variation in the risk of predation by wolves, Animal Behav., 69 (2005), 1181–1189.
    [39] J. Duarte, C. Januário, N. Martins, et al., Chaos and crises in a model for cooperative hunting: A symbolic dynamics approach, Chaos, 19 (2009), 043102.
    [40] L. Berec, Impacts of foraging facilitation among predators on predator–prey dynamics, Bull. Math. Biol., 72 (2010), 94–121.
    [41] M. Teixeira Alves and F. M. Hilker, Hunting cooperation and Allee effects in predators, J. Theor. Biol., 419 (2017), 13–22.
    [42] L. Pˇ ribylová and A. Peniaˇ sková, Foraging facilitation among predators and its impact on the stability of predator–prey dynamics, Ecol. Complex., 29 (2017), 30–39.
    [43] S. Pal, N. Pal and J. Chattopadhyay, Hunting cooperation in a discrete-time predator-prey system, Int. J. Bifurc. Chaos Appl. Sci. Eng., 28 (2018), 1850083.
    [44] X. Wang, L. Y. Zanette and X. Zou, Modelling the fear effect in predator–prey interactions, J. Math. Biol, 73 (2016), 1179–1204.
    [45] X. Wang and X. Zou, Modeling the fear effect in predator-prey interactions with adaptive avoid-ance of predators, Bull. Math. Biol., 79 (2017), 1325–1359.
    [46] P. Panday, N. Pal, S. Samanta, et al., Stability and bifurcation analysis of a three-species food chain model with fear, Int. J. Bifurc. Chaos Appl. Sci. Eng., 28 (2018), 1850009.
    [47] S. K. Sasmal, Population dynamics with multiple Allee effects induced by fear factors–A mathe-matical study on prey–predator interactions, Appl. Math Model., 64 (2018), 1–14.
    [48] S. Pal, S. Majhi, S. Mandal, et al., Role of fear in a predator–prey model with Beddington-DeAngelis functional response, Z. Naturforsch. A, 74 (2019) DOI: 10.1515/zna-2018-0449.
    [49] A. Sha, S. Samanta, M. Martcheva, et al., Backward bifurcation, oscillations and chaos in an eco-epidemiological model with fear effect, J. Biol. Dyn., 13 (2019), 301–327.
    [50] P. A. Schmidt and L. D. Mech, Wolf pack size and food acquisition, Am. Nat., 150 (1997), 513–517.
    [51] N. Courbin, A. J. Loveridge, D. Macdonald, et al. Reactive responses of zebras to lion encounters shape their predator–prey space game at large scale, Oikos, 125 (2016), 829–838.
    [52] S. Periquet, M. Valeix, A. J. Loveridge, et al., Individual vigilance of African herbivores while drinking: the role of immediate predation risk and context, Animal Behav., 79 (2010), 665–671.
    [53] W. J. Ripple and E. J. Larsen, Historic aspen recruitment, elk, and wolves in northern Yellowstone National Park, USA, Biol. Conserv., 95 (2000), 361–370.
    [54] S. Creel, J. A. Winnie Jr, B. Maxwell, et al., Elk alter habitat selection as an antipredator response to wolves, Ecology, 86 (2005), 3387–3397.
    [55] J. A. Winnie Jr, D. Christianson, S. Creel, et al., Elk decision-making rules are simplified in the presence of wolves, Behav. Ecol. Sociobiol., 61 (2006), 277.
    [56] J. A. Winnie Jr, and S. Creel, Sex-specific behavioural responses of elk to spatial and temporal variation in the threat of wolf predation, Animal Behav., 73 (2007), 215–225.
    [57] D. W. Stephens and J. R. Krebs, Foraging theory, Princeton University Press, Princeton, New Jersey, 1986.
    [58] Z. Zhang, Mutualism or cooperation among competitors promotes coexistence and competitive ability, Ecol. Model., 164 (2003), 271–282.
    [59] M. J. Hamilton, O. Burger, J. P. DeLong, et al., Population stability, cooperation, and the invasi-bility of the human species, Proc. Natl. Acad. Sci. U.S.A., 106 (2009), 12255–12260.
    [60] 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.
    [61] H. R. Thieme, Mathematics in population biology, Princeton University Press, Princeton, New Jersey, 2003.
    [62] F. Chen, On a nonlinear nonautonomous predator–prey model with diffusion and distributed delay, J. Comput. Appl. Math., 180 (2005), 33–49.
    [63] S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Vol. 2, Springer, New York, 1990.
    [64] Y. A. Kuznetsov, Elements of Applied Bifurcation Theory, Vol. 112, Springer, New York, USA, 1998.
    [65] S. Pal, N. Pal, S. Samanta, et al., A predator-prey model with hunting cooperation and fear, Ecol. Complex., Under review, (M.S. no: ECOCOM-2019-21).
  • This article has been cited by:

    1. Mainul Hossain, Nikhil Pal, Sudip Samanta, Impact of fear on an eco-epidemiological model, 2020, 134, 09600779, 109718, 10.1016/j.chaos.2020.109718
    2. Debasis Mukherjee, Role of fear in predator–prey system with intraspecific competition, 2020, 177, 03784754, 263, 10.1016/j.matcom.2020.04.025
    3. Ying Huang, Zhenliang Zhu, Zhong Li, Modeling the Allee effect and fear effect in predator–prey system incorporating a prey refuge, 2020, 2020, 1687-1847, 10.1186/s13662-020-02727-5
    4. Saheb Pal, Mainul Hossain, Pijush Panday, N. C. Pati, Nikhil Pal, Joydev Chattopadhyay, Cooperation delay induced chaos in an ecological system, 2020, 30, 1054-1500, 083124, 10.1063/5.0012880
    5. Pankaj Kumar Tiwari, Kawkab Abdullah Nabhan Al Amri, Sudip Samanta, Qamar Jalil Ahmad Khan, Joydev Chattopadhyay, A systematic study of autonomous and nonautonomous predator–prey models with combined effects of fear, migration and switching, 2021, 103, 0924-090X, 2125, 10.1007/s11071-021-06210-y
    6. Krishnanand Vishwakarma, Moitri Sen, Influence of Allee effect in prey and hunting cooperation in predator with Holling type-III functional response, 2021, 1598-5865, 10.1007/s12190-021-01520-1
    7. Pijush Panday, Sudip Samanta, Nikhil Pal, Joydev Chattopadhyay, Delay induced multiple stability switch and chaos in a predator–prey model with fear effect, 2020, 172, 03784754, 134, 10.1016/j.matcom.2019.12.015
    8. Junli Liu, Bairu Liu, Pan Lv, Tailei Zhang, An eco-epidemiological model with fear effect and hunting cooperation, 2021, 142, 09600779, 110494, 10.1016/j.chaos.2020.110494
    9. Jyotirmoy Roy, Dipesh Barman, Shariful Alam, Role of fear in a predator-prey system with ratio-dependent functional response in deterministic and stochastic environment, 2020, 197, 03032647, 104176, 10.1016/j.biosystems.2020.104176
    10. Mainul Hossain, Nikhil Pal, Sudip Samanta, Joydev Chattopadhyay, Fear Induced Stabilization in an Intraguild Predation Model, 2020, 30, 0218-1274, 2050053, 10.1142/S0218127420500534
    11. Xiaoqin Wang, Yiping Tan, Yongli Cai, Weiming Wang, Impact of the Fear Effect on the Stability and Bifurcation of a Leslie–Gower Predator–Prey Model, 2020, 30, 0218-1274, 2050210, 10.1142/S0218127420502107
    12. Dalia Yousri, Dalia Allam, Magdy. B. Eteiba, Optimal photovoltaic array reconfiguration for alleviating the partial shading influence based on a modified harris hawks optimizer, 2020, 206, 01968904, 112470, 10.1016/j.enconman.2020.112470
    13. Jialin Chen, Xiaqing He, Fengde Chen, The Influence of Fear Effect to a Discrete-Time Predator-Prey System with Predator Has Other Food Resource, 2021, 9, 2227-7390, 865, 10.3390/math9080865
    14. PANKAJ KUMAR TIWARI, MAITRI VERMA, SOUMITRA PAL, YUN KANG, ARVIND KUMAR MISRA, A DELAY NONAUTONOMOUS PREDATOR–PREY MODEL FOR THE EFFECTS OF FEAR, REFUGE AND HUNTING COOPERATION, 2021, 29, 0218-3390, 927, 10.1142/S0218339021500236
    15. NAZMUL SK, PANKAJ KUMAR TIWARI, YUN KANG, SAMARES PAL, A NONAUTONOMOUS MODEL FOR THE INTERACTIVE EFFECTS OF FEAR, REFUGE AND ADDITIONAL FOOD IN A PREY–PREDATOR SYSTEM, 2021, 29, 0218-3390, 107, 10.1142/S0218339021500054
    16. Aytül Gökçe, A mathematical model of population dynamics revisited with fear factor, maturation delay, and spatial coefficients, 2022, 45, 0170-4214, 11828, 10.1002/mma.8483
    17. Yuke Zhang, Xinzhu Meng, Dynamics Analysis of a Predator–Prey Model with Hunting Cooperative and Nonlinear Stochastic Disturbance, 2022, 10, 2227-7390, 2890, 10.3390/math10162890
    18. Xuebing Zhang, Qi An, Ling Wang, Spatiotemporal dynamics of a delayed diffusive ratio-dependent predator–prey model with fear effect, 2021, 105, 0924-090X, 3775, 10.1007/s11071-021-06780-x
    19. Debjit Pal, Dipak Kesh, Debasis Mukherjee, Qualitative study of cross-diffusion and pattern formation in Leslie–Gower predator–prey model with fear and Allee effects, 2023, 167, 09600779, 113033, 10.1016/j.chaos.2022.113033
    20. Haokun Qi, Xinzhu Meng, Dynamics of a stochastic predator-prey model with fear effect and hunting cooperation, 2022, 1598-5865, 10.1007/s12190-022-01746-7
    21. Érika Diz-Pita, M. Victoria Otero-Espinar, Predator–Prey Models: A Review of Some Recent Advances, 2021, 9, 2227-7390, 1783, 10.3390/math9151783
    22. Sourav Kumar Sasmal, Yasuhiro Takeuchi, Editorial: Mathematical Modeling to Solve the Problems in Life Sciences, 2020, 17, 1551-0018, 2967, 10.3934/mbe.2020167
    23. Subarna Roy, Pankaj Kumar Tiwari, Himadri Nayak, Maia Martcheva, Effects of fear, refuge and hunting cooperation in a seasonally forced eco-epidemic model with selective predation, 2022, 137, 2190-5444, 10.1140/epjp/s13360-022-02751-2
    24. Yong Yao, Lingling Liu, Dynamics of a Leslie-Gower predator-prey system with hunting cooperation and prey harvesting, 2022, 27, 1531-3492, 4787, 10.3934/dcdsb.2021252
    25. Sijia Lin, Fengde Chen, Zhong Li, Lijuan Chen, Complex Dynamic Behaviors of a Modified Discrete Leslie–Gower Predator–Prey System with Fear Effect on Prey Species, 2022, 11, 2075-1680, 520, 10.3390/axioms11100520
    26. S. Vinoth, R. Sivasamy, K. Sathiyanathan, Bundit Unyong, Grienggrai Rajchakit, R. Vadivel, Nallappan Gunasekaran, The dynamics of a Leslie type predator–prey model with fear and Allee effect, 2021, 2021, 1687-1847, 10.1186/s13662-021-03490-x
    27. N. S. N. V. K. Vyshnavi Devi, Debaldev Jana, The role of fear in a time-variant prey–predator model with multiple delays and alternative food source to predator, 2022, 10, 2195-268X, 630, 10.1007/s40435-021-00809-0
    28. Aytül GÖKÇE, Exploring a simple stochastic mathematical model including fear with linear functional response, 2021, 2645-8845, 10.33401/fujma.981385
    29. Basaznew Belew, Dawit Melese, Qiankun Song, Modeling and Analysis of Predator-Prey Model with Fear Effect in Prey and Hunting Cooperation among Predators and Harvesting, 2022, 2022, 1687-0042, 1, 10.1155/2022/2776698
    30. Qamar Din, Muhammad Arfan Zulfiqar, Qualitative behavior of a discrete predator–prey system under fear effects, 2022, 77, 0932-0784, 1023, 10.1515/zna-2022-0129
    31. Aytül Gökçe, Dynamical behaviour of a predator–prey system encapsulating the fear affecting death rate of prey and intra-specific competition: Revisited in a fluctuating environment, 2023, 421, 03770427, 114849, 10.1016/j.cam.2022.114849
    32. Kawkab Abdullah Nabhan Al Amri, Qamar J. A. Khan, Combining impact of velocity, fear and refuge for the predator–prey dynamics, 2023, 17, 1751-3758, 10.1080/17513758.2023.2181989
    33. Ankur Jyoti Kashyap, Debasish Bhattacharjee, Hemanta Kumar Sarmah, A fractional model in exploring the role of fear in mass mortality of pelicans in the Salton Sea, 2021, 11, 2146-5703, 28, 10.11121/ijocta.2021.1123
    34. Narayan Mondal, Dipesh Barman, Jyotirmoy Roy, Shariful Alam, Mohammad Sajid, A MODIFIED LESLIE-GOWER FRACTIONAL ORDER PREY-PREDATOR INTERACTION MODEL INCORPORATING THE EFFECT OF FEAR ON PREY, 2023, 13, 2156-907X, 198, 10.11948/20220011
    35. Sourav Kumar Sasmal, Balram Dubey, Diffusive patterns in a predator–prey system with fear and hunting cooperation, 2022, 137, 2190-5444, 10.1140/epjp/s13360-022-02497-x
    36. Seonguk Kim, Kwadwo Antwi-Fordjour, Prey group defense to predator aggregated induced fear, 2022, 137, 2190-5444, 10.1140/epjp/s13360-022-02926-x
    37. Xuebing Zhang, Honglan Zhu, Qi An, Dynamics analysis of a diffusive predator-prey model with spatial memory and nonlocal fear effect, 2023, 525, 0022247X, 127123, 10.1016/j.jmaa.2023.127123
    38. Miaomiao Chen, Yasuhiro Takeuchi, Jia-Fang Zhang, Dynamic complexity of a modified Leslie–Gower predator–prey system with fear effect, 2023, 119, 10075704, 107109, 10.1016/j.cnsns.2023.107109
    39. Ibtissam Benamara, Abderrahim El Abdllaoui, Bifurcation in a delayed predator–prey model with Holling type IV functional response incorporating hunting cooperation and fear effect, 2023, 2195-268X, 10.1007/s40435-023-01123-7
    40. Ankit Parwaliya, Anuraj Singh, Ajay Kumar, Dipesh Barman, The impact of delays on prey-predator dynamics with predation-induced fear, 2024, 70, 1598-5865, 4877, 10.1007/s12190-024-02129-w
    41. Tanuja Das, Prashant K. Srivastava, Stability and codimension 2 bifurcations in an SIR model with incubation delay, 2024, 79, 14681218, 104127, 10.1016/j.nonrwa.2024.104127
    42. Anuj Kumar Umrao, Prashant K. Srivastava, 2024, Chapter 3, 978-3-031-59071-9, 67, 10.1007/978-3-031-59072-6_3
    43. Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh, Combined impact of fear and Allee effect in predator-prey interaction models on their growth, 2024, 21, 1551-0018, 7211, 10.3934/mbe.2024319
    44. Rongjie Yu, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao, Bifurcation analysis of Leslie-Gower predator-prey system with harvesting and fear effect, 2023, 20, 1551-0018, 18267, 10.3934/mbe.2023812
    45. SHILPA SAMADDAR, MAUSUMI DHAR, PARITOSH BHATTACHARYA, UTTAM GHOSH, BIFURCATION ANALYSIS OF A MODIFIED LESLIE–GOWER PREDATOR–PREY MODEL WITH HUNTING COOPERATION AND FAVORABLE ADDITIONAL FOOD FOR PREDATOR, 2023, 31, 0218-3390, 1015, 10.1142/S0218339023500353
    46. Zhenliang Zhu, Yuming Chen, Zhong Li, Fengde Chen, Dynamic behaviors of a Leslie-Gower model with strong Allee effect and fear effect in prey, 2023, 20, 1551-0018, 10977, 10.3934/mbe.2023486
    47. Shangming Chen, Fengde Chen, Vaibhava Srivastava, Rana D. Parshad, Dynamical analysis of a Lotka–Volterra competition model with both Allee and fear effects, 2024, 17, 1793-5245, 10.1142/S1793524523500778
    48. Soumitra Pal, Pijush Panday, Nikhil Pal, A. K. Misra, Joydev Chattopadhyay, Dynamical behaviors of a constant prey refuge ratio-dependent prey–predator model with Allee and fear effects, 2024, 17, 1793-5245, 10.1142/S1793524523500109
    49. Biswajit Paul, Bapin Mondal, Uttam Ghosh, Mathematical Modeling of Regime Shifts in Fluctuating Environments: The Impact of Allee Effects and Cooperation, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05282-0
    50. Qamar Din, Khansa Jameel, Muhammad Sajjad Shabbir, Discrete-Time Predator–Prey System Incorporating Fear Effect: Stability, Bifurcation, and Chaos Control, 2024, 23, 1575-5460, 10.1007/s12346-024-01145-2
    51. Yalong Xue, Fengde Chen, Xiangdong Xie, Shengjiang Chen, An analysis of a predator-prey model in which fear reduces prey birth and death rates, 2024, 9, 2473-6988, 12906, 10.3934/math.2024630
    52. Dingyong Bai, Jiaoxin Tang, Global Dynamics of a Predator–Prey System with Cooperative Hunting, 2023, 13, 2076-3417, 8178, 10.3390/app13148178
    53. Xiaoke Ma, Ying Su, Xingfu Zou, Joint Impact of Maturation Delay and Fear Effect on the Population Dynamics of a Predator-Prey System, 2024, 84, 0036-1399, 1557, 10.1137/23M1596569
    54. DIPESH BARMAN, SUBARNA ROY, PANKAJ KUMAR TIWARI, SHARIFUL ALAM, TWO-FOLD IMPACTS OF FEAR IN A SEASONALLY FORCED PREDATOR–PREY SYSTEM WITH COSNER FUNCTIONAL RESPONSE, 2023, 31, 0218-3390, 517, 10.1142/S0218339023500183
    55. Subarna Roy, Nazmul Sk, Pankaj Kumar Tiwari, Bifurcation analysis of autonomous and nonautonomous modified Leslie–Gower models, 2024, 34, 1054-1500, 10.1063/5.0171936
    56. Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar, Dynamical analysis of a predator-prey system with prey vigilance and hunting cooperation in predators, 2024, 21, 1551-0018, 2768, 10.3934/mbe.2024123
    57. 娇欣 唐, Hopf Bifurcation of a Predator-Prey System with Cooperative Hunting and Intraspecific Competition, 2023, 12, 2324-7991, 2113, 10.12677/AAM.2023.125215
    58. Huatao Wang, Yan Zhang, Li Ma, Bifurcation and stability of a diffusive predator–prey model with the fear effect and time delay, 2023, 33, 1054-1500, 10.1063/5.0157410
    59. Hongyan Sun, Pengmiao Hao, Jianzhi Cao, Dynamics analysis of a diffusive prey-taxis system with memory and maturation delays, 2024, 14173875, 1, 10.14232/ejqtde.2024.1.40
    60. Anuj Kumar Umrao, Subarna Roy, Pankaj Kumar Tiwari, Prashant K. Srivastava, Dynamical behaviors of autonomous and nonautonomous models of generalist predator–prey system with fear, mutual interference and nonlinear harvesting, 2024, 183, 09600779, 114891, 10.1016/j.chaos.2024.114891
    61. Jianzhi Cao, Fang Li, Pengmiao Hao, Bifurcation analysis of a diffusive predator–prey model with fear effect, 2024, 47, 0170-4214, 13404, 10.1002/mma.10198
    62. Deepak Tripathi, Jai Prakash Tripathi, Satish Kumar Tiwari, Debaldev Jana, Li-Feng Hou, Yu Shi, Gui-Quan Sun, Vandana Tiwari, Joshua Kiddy K. Asamoah, Modified Holling Tanner diffusive and non-diffusive predator–prey models: The impact of prey refuge and fear effect, 2024, 65, 22113797, 107995, 10.1016/j.rinp.2024.107995
    63. G Ranjith Kumar, K Ramesh, Aziz Khan, K. Lakshminarayan, Thabet Abdeljawad, Dynamical study of fractional order Leslie-Gower model of predator-prey with fear, Allee effect, and inter-species rivalry, 2024, 14, 26667207, 100403, 10.1016/j.rico.2024.100403
    64. Sunil Kumar, Teekam Singh, Role of Allee Effect, Hunting Cooperation, and Dispersal to Prey–Predator Model, 2023, 33, 0218-1274, 10.1142/S0218127423501559
    65. Soumitra Pal, Pankaj Kumar Tiwari, Arvind Kumar Misra, Hao Wang, Fear effect in a three-species food chain model with generalist predator, 2023, 21, 1551-0018, 1, 10.3934/mbe.2024001
    66. Mengting Sui, Yanfei Du, Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay, 2023, 31, 2688-1594, 5124, 10.3934/era.2023262
    67. Rasha M. Yaseen, May M. Helal, Kaushik Dehingia, Ahmed A. Mohsen, Effect of the Fear Factor and Prey Refuge in an Asymmetric Predator–Prey Model, 2024, 54, 0103-9733, 10.1007/s13538-024-01594-9
    68. Anuj Kumar Umrao, Sonu Lamba, Prashant K. Srivastava, Nonlinear dynamics in a fear‐driven predator–prey system: Bistability, bifurcations, hydra effect, and optimal harvesting, 2024, 0170-4214, 10.1002/mma.10431
    69. Yuan Tian, Huanmeng Li, Kaibiao Sun, Complex dynamics of a fishery model: Impact of the triple effects of fear, cooperative hunting and intermittent harvesting, 2024, 218, 03784754, 31, 10.1016/j.matcom.2023.11.024
    70. Debjit Pal, Dipak Kesh, Debasis Mukherjee, Cross-diffusion mediated Spatiotemporal patterns in a predator–prey system with hunting cooperation and fear effect, 2024, 220, 03784754, 128, 10.1016/j.matcom.2024.01.003
    71. Shankha Narayan Chattopadhyay, Arvind Kumar Gupta, Tipping points, multistability, and stochasticity in a two-dimensional traffic network dynamics, 2024, 34, 1054-1500, 10.1063/5.0202785
    72. Sushil Kumar Singh, Harshit Chauhan, Yash Vardhan, Savinder Kaur, Prey–predator realities: unveiling competition, cooperation, and shelter dynamics, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05171-6
    73. Sayan Mandal, Pankaj Kumar Tiwari, Schooling behavior in a generalist predator–prey system: exploring fear, refuge and cooperative strategies in a stochastic environment, 2024, 139, 2190-5444, 10.1140/epjp/s13360-023-04787-4
    74. Anuj Kumar Umrao, Prashant K. Srivastava, Bifurcation Analysis of a Predator–Prey Model with Allee Effect and Fear Effect in Prey and Hunting Cooperation in Predator, 2023, 0971-3514, 10.1007/s12591-023-00663-w
    75. Bapin Mondal, Abhijit Sarkar, Nazmul Sk, Shyam Sundar Santra, Taseer Muhammad, Exploring resilience, chaos, and bifurcations in a discrete food web model incorporating the mate finding Allee effect, 2023, 138, 2190-5444, 10.1140/epjp/s13360-023-04651-5
    76. Ankit Parwaliya, Anuraj Singh, Ajay Kumar, Hopf bifurcation in a delayed prey–predator model with prey refuge involving fear effect, 2024, 17, 1793-5245, 10.1142/S1793524523500420
    77. Yalong Xue, Impact of both-density-dependent fear effect in a Leslie–Gower predator–prey model with Beddington–DeAngelis functional response, 2024, 185, 09600779, 115055, 10.1016/j.chaos.2024.115055
    78. Nauman Ahmed, Muhammad Waqas Yasin, Ali Akgül, Dumitru Baleanu, Ovidiu Tintareanu-Mircea, Mathematical analysis and pattern formation in diffusive predator–prey system, 2025, 1598-5865, 10.1007/s12190-024-02341-8
    79. Fangyu Zhu, Ruizhi Yang, Bifurcation in a modified Leslie-Gower model with nonlocal competition and fear effect, 2025, 0, 1531-3492, 0, 10.3934/dcdsb.2024195
    80. 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
    81. Xuebing Zhang, Jia Liu, Guanglan Wang, Impacts of nonlocal fear effect and delay on a reaction–diffusion predator–prey model, 2025, 18, 1793-5245, 10.1142/S1793524523500894
    82. Christian Cortés-García, Dynamics in a Filippov–Gause Predator–Prey Model with Hunting Cooperation or Competition Among Predators for High or Low Prey Densities, 2025, 35, 0218-1274, 10.1142/S0218127425500142
    83. Jawdat Alebraheem, Asymptotic stability of deterministic and stochastic prey-predator models with prey herd immigration, 2025, 10, 2473-6988, 4620, 10.3934/math.2025214
    84. Jyotirmoy Roy, Bapin Mondal, Animesh Mahata, Shariful Alam, Sankar Prasad Mondal, Influence of Breeding Delays and Memory Effects on Predator-Prey Model Amidst Fear, 2025, 55, 0103-9733, 10.1007/s13538-025-01737-6
    85. Siti Nurul Afiyah, , 2025, 3302, 0094-243X, 030001, 10.1063/5.0262460
    86. Md Golam Mortuja, Mithilesh Kumar Chaube, Santosh Kumar, Dynamic analysis of a predator-prey model with Michaelis-Menten prey harvesting and hunting cooperation in predators, 2025, 1598-5865, 10.1007/s12190-025-02482-4
  • Reader Comments
  • © 2019 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(9289) PDF downloads(1773) Cited by(86)

Figures and Tables

Figures(16)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog