Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article

Stability and bifurcation in a predator-prey system with effect of fear and additional food

  • Received: 20 October 2023 Revised: 01 December 2023 Accepted: 10 December 2023 Published: 15 January 2024
  • MSC : 34D20, 34D23, 37N25, 92B05

  • In the present study, we propose and analyze a three-dimensional prey-predator model. The prey grows logistically in the absence of the predator and their relationship follows the Crowley-Martin type functional response. In this paper, we examine the impact of supply of the additional food to the predators and the influence of fear in the prey population. Since the predator depends partially on the provided other resources, we incorporate a novel parameter, the degree of dependence, which basically demonstrates how dependent the predator is on the prey population. We investigate the steady-state solutions, and their local and global behavior, which contributes to understanding the long-term dynamics of the interaction. We have shown that the degree of dependence and the cost of fear both can cause periodic orbits to appear in the system via a Hopf-bifurcation. Our findings show that with the newly introduced parameter, we can control the oscillations from the system, which helps to balance the ecosystem. The direction and stability have also been investigated using the center manifold theorem and normal form theory. Last, we perform an extensive numerical simulation to validate our theoretical findings. Our main goal of this work is to maintain the ecological balance in the presence of fear effect and additional food for predators.

    Citation: Reshma K P, Ankit Kumar. Stability and bifurcation in a predator-prey system with effect of fear and additional food[J]. AIMS Mathematics, 2024, 9(2): 4211-4240. doi: 10.3934/math.2024208

    Related Papers:

    [1] 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
    [2] Vikas Kumar, Nitu Kumari . Controlling chaos in three species food chain model with fear effect. AIMS Mathematics, 2020, 5(2): 828-842. doi: 10.3934/math.2020056
    [3] 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
    [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] Binfeng Xie, Na Zhang . Influence of fear effect on a Holling type III prey-predator system with the prey refuge. AIMS Mathematics, 2022, 7(2): 1811-1830. doi: 10.3934/math.2022104
    [6] 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
    [7] San-Xing Wu, Xin-You Meng . Dynamics of a delayed predator-prey system with fear effect, herd behavior and disease in the susceptible prey. AIMS Mathematics, 2021, 6(4): 3654-3685. doi: 10.3934/math.2021218
    [8] 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
    [9] Weili Kong, Yuanfu Shao . Bifurcations of a Leslie-Gower predator-prey model with fear, strong Allee effect and hunting cooperation. AIMS Mathematics, 2024, 9(11): 31607-31635. doi: 10.3934/math.20241520
    [10] 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
  • In the present study, we propose and analyze a three-dimensional prey-predator model. The prey grows logistically in the absence of the predator and their relationship follows the Crowley-Martin type functional response. In this paper, we examine the impact of supply of the additional food to the predators and the influence of fear in the prey population. Since the predator depends partially on the provided other resources, we incorporate a novel parameter, the degree of dependence, which basically demonstrates how dependent the predator is on the prey population. We investigate the steady-state solutions, and their local and global behavior, which contributes to understanding the long-term dynamics of the interaction. We have shown that the degree of dependence and the cost of fear both can cause periodic orbits to appear in the system via a Hopf-bifurcation. Our findings show that with the newly introduced parameter, we can control the oscillations from the system, which helps to balance the ecosystem. The direction and stability have also been investigated using the center manifold theorem and normal form theory. Last, we perform an extensive numerical simulation to validate our theoretical findings. Our main goal of this work is to maintain the ecological balance in the presence of fear effect and additional food for predators.



    Predation is one of the most common phenomena for survival in the wilderness that connects all the elements of a multispecies food web. It plays an essential role in maintaining stability and balance in the nature. Mathematical models are the most effective tools to illustrate the relationship between two or more species. Initially, the mathematical models for prey-predator interactions were explored by two mathematicians, Alfred J. Lotla [1] and Vito Volterra [2], independently in the mid 1920s. They found periodic fluctuations of species around their steady states. Since then, ecologists and mathematical biologists have been in practice with various models considering different essential parameters [3,4,5,6]. The dynamics of the prey-predator system are significantly influenced and become more realistic by the incorporation of some notions such as functional response [7,8,9,10,11], stage structure [12,13], effect of fear [14,15] and time delay [5].

    Understanding the complex dynamics of a population model depends on the functional response of predators, which quantifies the prey captured by each predator during certain periods of time. It significantly influences the dynamical features of the system. From an ecological perspective, incorporating the functional response enhances the realism and accuracy of the model. In recent years, numerous studies have been conducted to explore the effects of different functional responses on the prey-predator system. These studies include the Holling type-II [7,16,17], Holling type-III [8,18], Holling type-IV [6,12,19], Beddington-DeAngelis [9,20,21,22] and Crowley-Martin functional responses [10,11,13,23,24,25,26,27].

    In recent decades, several researchers have deeply investigated the qualitative behavior of prey-predator systems when supplementary nourishment sources are available for the predators [17,20,28,29,30,31,32]. In particular, Sahoo [32] put forward and explored a model to investigate the significance of supplementary nourishment within an eco-epidemiological system. The research focused on managing chaotic population dynamics of predator-prey interactions by introducing supplementary nourishment resources for top predators. To accomplish this, they developed a three-species predator-prey model that specifically incorporates the provision of supplementary food to the top predators. In another study, Thirthar et al. [20] focused on prey, predator and super predator within a predator-prey system. In this study, predation interactions between prey and predator are regulated by the Beddington-DeAngelis functional response, and the impact of predation fear in prey species was also carried out. The authors assumed that the predator species offers protection to the super predator species. It was anticipated that the super predators, driven by fear of predators and their refuge behavior, would receive a constant supply of additional food, which in turn would benefit both the super predators and the predators. Furthermore, the latest research by Kumar and Dubey [17] examined the intricacies of a prey-predator model with the effect of supplementary nourishment, multiple delays, and the interaction between the prey and predator was modeled using a Holling type-II response. The provision of extra nourishment was specifically targeted toward the predator population, aiming to reduce its dependence on the prey species. Srinivasu et al. [30] examine a modified predator-prey model with the type II functional response to study how providing supplementary nourishment to the predator impacts the behavior of the system. The research emphasizes that the handling times of the available foods significantly influence the eventual state of the ecosystem. Ghosh et al. [31] investigate the influence of providing supplementary nourishment to predators within an ecological model that includes a prey refuge. The authors illustrate theoretically and numerically that species can co-exist in oscillatory patterns when there is an availability of both abundant and high-quality supplementary nourishment and prevent predator extinction in systems with high prey refuge, offering insights for conservation biologists in real-world ecological systems.

    Since predation events are obvious, our knowledge of the prey-predator system in forest provinces has mostly concentrated on the direct killing of predators on prey populations. The effects of predators on their prey can be physical as well as mental. Some recent studies have revealed a new perspective, showing that predators not only reduce prey populations by directly killing prey but also by inducing fear in prey individuals, which results in a fall in the prey birth rate. Zanette et al. [33] carried out research involving song sparrows, manipulating predation risk by playing predator sounds to some species and non-predator sounds to others. Despite protecting nests from direct killing, the fear of predators alone led to a significant 40% reduction in offspring production due to decreased egg laying, lower hatching success and increased nestling mortality. Various anti-predator responses contributed to these demographic effects. In another study by Cresswell [34], the impact of predators on population size extends beyond direct lethal effects, as fear of predation leads to costly behavioral changes with potentially significant population and community effects. The dominance of lethal or non-lethal effects depends on factors such as resource constraints, prey vulnerability, trophic interactions, and evolutionary history, making cases without any population effects due to predation or cases dominated solely by lethal effects rare and specialized.

    As a result of the fear induced by predators, prey individuals experience a restricted environment for their activities, including mating, preventing them from freely accessing open habitats. The fear inhibits them from venturing into open spaces, which reduces their ability to reproduce [33]. We construct a function that represents the fear effect on the prey population. The reduction of birth rate in the prey population depends on the number of predator individuals y and their nature. We quantify the nature (aggressiveness, agility, etc.) of the predator in terms of a parameter k, which is known as the cost of fear. Thus, the fear function f(y,k) is a function of y and k, which follows the following characteristics:

    (1) In the absence of predators or when the cost of fear is negligible, the impact on the reproduction rate of the prey species remains same

    f(0,k)=f(y,0)=1.

    (2) The reproduction rate of prey approches zero at the high predator density or at the large cost of fear

    limyf(y,k)=limkf(y,k)=0.

    (3) As the density of predators and the cost of the induced fear increase, the reproductive rate of prey decreases

    fy<0,fk<0.

    There may be several functions which follow all the conditions of a fear function, but commonly we use following functions due to their simplicity

    f(y,k)=11+ky,f(y,k)=eky,f(y,k)=11+k1y+k2y2.

    The proposed prey-predator model incorporates the influence of fear on prey reproduction, suggesting that high levels of fear can contribute to stabilizing the system by eliminating the periodic solutions. In the study conducted by Wang et al. [14] on a terrestrial vertebrate, it was observed that fear of predators alone had a significant impact on anti-predator defense, leading to a notable reduction in prey reproduction. To capture the influence of fear on the population dynamics of terrestrial animals, the authors developed a predator-prey model that incorporates the reproductive cost associated with fear. Dubey et al. [35] analyze the predator-prey model that accounts for fear-induced responses in prey towards predators and anti-predator behavior exhibited by the prey. Additionally, the model incorporates delays in fear response and gestation. They state that the system experiences Hopf-bifurcation in the context of both delays.

    Furthermore, Chen et al. [36] examined how the dynamics of predator-prey model behavior is affected by the fear effect and the Leslie-Gower function. When fear intensifies, the system will experience multiple dynamic behaviors, switching until the prey population is wiped out. However, because there are availability of other prey sources, the predator population manages to endure. The concept of effect of fear on the species dynamics has been utilized by various researchers to show the efficacy of their respective models [15,37,38,39]. Zhang et al. [38] examined anti-predator behavior and fear on a Holling type-II predator-prey model with prey refuge. The study shows that fear reduces the predator's population density and stabilizes the system by preventing the existence of periodic solutions, while the presence of prey refuge significantly affects the persistence of the predator population.

    The Beddington-DeAngelis functional response, formulated by Beddington [21] and DeAngelis [22], introduces a predator-dependent functional response that incorporates predator interference. It shares similarities with the Holling type II response, as it considers the interactions between predators in addition to prey searching and processing activities. In this response, predators form groups consisting of two or more individuals that not only engage in prey-related tasks but also interact with each other. The Beddington-DeAngelis functional response suggests that as prey abundance increases, the impact of predator interference on the feeding rate diminishes, implying that the collective behavior of predators becomes less influential. In the study conducted by Crowley and Martin [23], it was anticipated that predation would decrease as a result of high predator density, which can lead to interference among individual predators. This interference occurs when predator individuals interact with each other and it can reduce the overall predation rate even when there is an abundance of prey available. The Crowley-Martin functional response describes the relationship between prey biomass and the population of birds. In the Crowley-Martin functional response, predator interference is taken into account, meaning that the presence of one predator can impede the activity of another predator. This interference affects both the prey and predator population. Many authors have used the Crowley-Martin functional response to demonstrate the effectiveness of their models [11,13,21,22,24,25,26,27]. The difference between the Beddington-DeAngelis and Crowley-Martin functional responses is how they consider the effects of predator interference. The Beddington-DeAngelis type response suggests that in the case of abundant prey, the impact of predator interference on the feeding rate becomes negligible. In contrast, the Crowley-Martin functional response consistently considers the impact of interference on the feeding rate, regardless of prey abundance. Experimental evidence supports the notion that mutual predator interference decreases the feeding rate of predators [40,41].

    Maiti et al. [23] investigated a prey-predator model that includes prey refuge and a Crowley-Martin type response to describe the interaction between unreserved prey and predators. The study focuses on analyzing the dynamic of the system while considering delay as a parameter that can cause bifurcation. Tripathi et al. [26] examined a two-dimensional predator-prey model with discrete delay and the Crowley-Martin functional response, which incorporates mutual interference by predators at high prey density. The authors analyze the impact of predator density dependence on system permanence and numerically describe Hopf-bifurcation in the non-delayed system. Moreover, Papanikolaou et al. [27] found that the Crowley-Martin model outperforms other functional responses, including the Beddington-DeAngelis, Hassell-Varley and Holling type-II functional responses. This finding further highlights the effectiveness and relevance of the Crowley-Martin functional response in capturing the dynamics of predator-prey interactions. Considering these insights, investigating the Crowley-Martin type interaction between prey and predator can potentially reveal rich and complex dynamics within the system.

    As we mentioned earlier, fear is a psychological activity of prey in the presence of predators, which was ignored in classical models. The prey shows anti-predator nature under fear, such as reducing the foraging time, changing the living locality, adjusting reproduction strategies, etc. Due to all these, the reproduction rate decreases. Therefore, it is reasonable to consider fear in the model. The additional food is provided to predators to avoid the overexploitation of prey resources and maintain the ecological balance. In the existing models, the additional food is provided at a constant rate, but it must vary with the number of predators. In the present model, we introduce the additional food in a separate equation and keep its quantity depending on the number of predators. Since the effect of fear varies with the dependency of the predator on the provided food, we incorporated a new parameter, the degree of dependency. To the best of our understanding a mathematical model with all the above points has not been investigated. Keeping all these in view, in Section 2, we discuss the model construction. In Section 3, we show our model is biologically well-behaved by proving the positivity and boundedness of the solutions. We discuss various dynamical features of our system in Section 4. In Section 5, an extensive computer simulation is carried out to validate our results. The study concludes with a few final remarks.

    We consider an ecological habitat where two species survive, one is prey and another is predator. The prey population follows the logistic growth behavior when predators are absent. This is governed by the equation below:

    dxdt=rxr0xr1x2,

    where x is the number of individuals in the prey population at time t, r and r0 are the birth rate and the natural mortality rate of prey and r1 is the intra-specific interference among the prey individuals. In the latest research, it was discoverd that the presence of predators induces fear within the prey population which leads to a reduction in their reproduction rate. This fear-induced reduction in reproduction rate is influenced by both the predator population y and a parameter k representing the cost of fear. Consequently, the modified form of the prey population growth equation is as stated:

    dxdt=rf(y,k)xr0xr1x2,

    where f(y,k) is referred as fear function and rf(y,k) is the effective growth rate of prey under the fear effect. In this study, we consider the influence of predation fear on the prey population when predators are present. The cost of anti-predator defense is related to the predator population density. To account for this, we replace the prey reproduction term rx by multiplying it with the fear function f(y,k)=11+ky, emphasizing the costs carried on by predation fear. This modification allows us to explore how predation fear affects the dynamics of the prey-predator system in the presence of predators. The following system governs the prey-predator interaction in presence of fear effect.

    dxdt=rx1+kyr0xr1x2Q(x,y)y,dydt=c1Q(x,y)yδyey2, (2.1)

    where Q(x,y) is functional response of predator and c1 (0c11) is the conversion efficiency of y on x. The parameters δ and e are the natural death rate of predators and the intra-specific interference among predators.

    The Holling type-II response is prey-dependent, indicating that the predator feeding rate declines with increasing prey density due to handling and searching time. The Crowley-Martin response characterizes the prey-predator interaction, predicting a decrease in predator density at high predator concentrations due to resource competition. The Crowley-Martin response is formulated as

    Q(x,y)=αx(1+ax)(1+by),

    where α is the capture rate, a is the handling time and b is the degree of interference among predator individuals. Additionally, it changes into mass action law (when a and b are zero) and Holling type-II response (when a>0,b=0 or a=0,b>0). With the Crowley-Martin functional response, our model implies the forthcoming form.

    dxdt=rx1+kyr0xr1x2αxy(1+ax)(1+by),dydt=c1αxy(1+ax)(1+by)δyey2. (2.2)

    Additional food in a prey-predator system refers to providing extra nourishment resources to the predator population, potentially beyond what is naturally available from their primary prey. This additional food can have various effects on the dynamics of predator-prey interactions and may influence population sizes, stability and coexistence patterns in the ecosystem. As a result, the equation describing the relationship between the additional food resource and the predator population has been modified as follows:

    dAdt=λyβAϕAy,dydt=c2ϕAyδyey2,

    where A is the quantity of additional food provided to the predator at time t. The parameters λ and β are the maximum supply rate and the natural depletion rate of additional food, respectively. The parameter ϕ is the consumption rate of additional food by predators and c2 (0c21) is the conversion efficiency of y on A. The three-dimensional predator-prey model with additional food can be expressed as follows:

    dxdt=rx1+kyr0xr1x2αxy(1+ax)(1+by),dAdt=λyβAϕAy,dydt=c1αxy(1+ax)(1+by)+c2ϕAyδyey2.

    Further, we introduce the degree of dependency in the prey-predator system, which refers to the reliance or dependence of the predator population on its primary prey as a food source. The parameter A0 (0A01) is the degree of dependency of predators on prey or the dependency factor. If A0=1, predators are fully dependent on prey and no additional food is necessary. If A0=0, predators are completely dependent on supplementary nourishment, and the prey species increases logistically. If A0 is in between 0 and 1, then the predator partially depends on prey and supplementary nourishment. The predator depends partially on the prey and the degree of dependency A0, which depends on the nature of the prey and provided additional food. Therefore, in the presence of additional food, the fear function f(y,k) also depends on the degree of dependency A0. To modify the fear function f(y,k), we consider the following points:

    (1) When the predator completely depends on the additional food then no fear will be induced by the predator, i.e., when A0=0 then f(y,k,A0)=1.

    (2) When the predators are fully dependent on prey then induced fear is maximum, i.e., A0=1 then f(y,k,A0) is maximum and same as the fear function discussed in the introduction.

    (3) The amount of fear monotonically increases with respect to A0, i.e., fA0>0.

    These points lead us to the following modified fear function, f(y,k,A0)=11+kA0y.

    Considering the points mentioned above, our mathematical model can be described as follows:

    dxdt=rx1+kA0yr0xr1x2αA0xy(1+ax)(1+by),dAdt=λ(1A0)yβAϕAy,dydt=c1αA0xy(1+ax)(1+by)+c2ϕAyδyey2,x(0)0,A(0)0,y(0)0. (2.3)

    The biological description of the parameters in the system (2.3), including their units and values, is provided in Table 1.

    Table 1.  Parameters and their biological meaning.
    Parameter Biological Description Units Values
    x Density of prey population biomass -
    A Quantity of additional food biomass -
    y Density of predator population biomass -
    r Birth rate of prey time1 3
    r0 Natural mortality rate of prey time1 0.2
    r1 Intra-specific interference among prey individuals biomass1.time1 0.01
    k Cost of fear in prey biomass1 5
    α Attack rate of predator on prey biomass1.time1 0.8
    a Handling time biomass1 0.1
    b Magnitude to interference among predator individuals biomass1 0.1
    A0 Degree of dependency factor of predators on prey Constant and 0<A0<1 0.2
    λ Maximum supply rate of additional food time1 1.5
    β Natural depletion rate of additional food time1 0.32
    ϕ Consumption rate of additional food by predators biomass1.time1 0.7
    c1 Conversion efficiency of y on x Constant and 0<c1<1 0.5
    c2 Conversion efficiency of y on A Constant and 0<c2<1 0.4
    δ Mortality rate of predators time1 0.35
    e Intra-specific interference among predators biomass1.time1 0.025

     | Show Table
    DownLoad: CSV

    Remark 1. From the first equation of the above system, dxdt(rr0)x, which implies limtx(t)=0 if r<r0 and consequently, limty(t)=0 at low supply rate of additional food. This is a trivial case from a biological perspective, to avoid this case we consider r>r0 throughtout this paper.

    A biological model must carry some essential characteristics such as positivity and boundedness of the solutions, persistence of the species. The existence of the species is demonstrated by the positivity and the boundedness of the solutions represents a natural restriction on species to grow exponentially. Both these characteristics are followed by the system (2.3) and have been shown by the theorem below.

    Theorem 1. The set

    Ω={(x,A,y):0xrr1,0c1x+c2A+y1σ[c1r24r1+c22λ2(1A0)24e]},

    where σ=min{r0,β,δ}, is a positive invariant set for all solutions of the system (2.3) that start in the interior of the positive quadrant.

    Proof. We write the system (2.3) in matrix form

    ˙X=G(X),

    where X=(x,A,y)TR3 and G(X) is given by

    G(X)=[G1(X)G2(X)G3(X)]=[rx1+kA0yr0xr1x2αA0xy(1+ax)(1+by)λ(1A0)yβAϕAyc1αA0xy(1+ax)(1+by)+c2ϕAyδyey2].

    The fundamental theorem of ordinary differential equation ensures the local existence and uniqueness of the solution because X(0)=X0R3+ and G:R3R3+ are locally Lipschitz-continuous in Ω. It follows that X(t)0 for all t0 because [Gi(X)]xi(t)=0,XR3+0. In fact, from the first and third equations of system (2.3), it can easily be seen that ˙x|x=00, ˙y|y=00 and hence x(t)0, y(t)0 for all t0, second, ˙A|A=0=λ(1A0)y0 for all t0 (as y(t)0 for all t0) and hence A(t)0 for all t0. Therefore, for system (2.3) x(t),A(t),y(t) all are positive.

    Now, we express the first equation of system (2.3) as follows:

    dxdtrxr1x2,

    which implies

    limtsupx(t)rr1.

    Now, we construct a function

    W(t)=c1x(t)+c2A(t)+y(t),

    which on differentiation

    dW(t)dt=c1dxdt+c2dAdt+dydtc1rxc1r1x2+c2λ(1A0)yey2σWc1rr4r1+c22λ2(1A0)24eσW,

    where σ=min{r0,β,δ}. Thus, it is evident that

    limtsupW(t)1σ(c1r24r1+c22λ2(1A0)24e)=:ys,

    and we observe that, if x(t)rr1 and W(t)1σ[c1r24r1+c22λ2(1A0)24e], then dx(t)dt0, dW(t)dt0. This demonstrates that the system (2.3) solutions are bounded and remain in Ω for all t>0 if (x(0),A(0),y(0))Ω.

    The persistence of a system indicates that all species will survive in the future and none will wipe out if they are initially present in the system. In the following theorem, we will establish some requirement for the uniform persistence of the system (2.3).

    Theorem 2. The system (2.3) is uniformly persistent under the following conditions and is satisfied.

    r>(r0+αA0ys)(1+kA0ys),c1αA0xi(1+axi)(1+bys)>δ,

    where xi and yi are defined in the proof.

    Proof. If M1 and M2 are positive constants that satisfy each positive solution X(t)=(x(t),A(t),y(t)) of the system with positive initial conditions, then the system (2.3) is said to have uniform persistence.

    M1limtinfX(t)limtsupX(t)M2.

    Now, we define

    M2=max{rr1,ysc2,ys}.

    As a result, Theorem 1 implies that

    limtsupX(t)M2.

    This further demonstrates that for any sufficiently small ϵ>0, there exists a T>0 such that for all tT, the following hold

    x(t)<rr1+ϵ,A(t)<ysc2+ϵ,y(t)<ys+ϵ.

    For any tT, we can write the first equation of the system (2.3).

    dxdtrx1+kA0(ys+ϵ)r0xr1x2αA0x(ys+ϵ)(r1+kA0(ys+ϵ)r0αA0(ys+ϵ))xr1x2,

    consequently, it follows

    limtinfx(t)1r1(r1+kA0(ys+ϵ)r0αA0(ys+ϵ)).

    In the case of arbitrarily small ϵ>0, we get

    limtinfx(t)1r1(r1+kA0ysr0αA0ys)=:xi.

    The prey species of the model (2.3) is persistent, if the inequality below holds true.

    r>(r0+αA0ys)(1+kA0ys).

    Now, the third equation of the model (2.3) implies

    dydtc1αA0(xi+ϵ)y(1+a(xi+ϵ))(1+b(ys+ϵ))δyey2,

    it follows that

    limtinfy(t)1e(c1αA0(xi+ϵ)(1+a(xi+ϵ))(1+b(ys+ϵ))δ).

    In the case of arbitrarily small ϵ>0, we get

    limtinfy(t)1e(c1αA0xi(1+axi)(1+bys)δ)=:yi.

    The predator species of the system (2.3) persists if the forthcoming conditions holds,

    c1αA0xi(1+axi)(1+bys)>δ.

    The second equation of the system (2.3) implies,

    dAdtλ(1A0)(yi+ϵ)βAϕ(ys+ϵ)A,

    consequently, it follows

    limtinfA(t)λ(1A0)(yi+ϵ)β+ϕ(ys+ϵ).

    Again if ϵ is arbitrarily small then we have

    limtinfA(t)λ(1A0)yiβ+ϕys=:Ai.

    Taking M1=min{xi,Ai,yi}, the theorem follows.

    The equilibrium points are the steady-state solutions for a system. From the ecological aspects, their presence may determine whether the populations coexist, one species becomes extinct, or both species collapse. We investigated that, the system (2.3) may have four equilibrium points: trivial equilibrium E0(0,0,0), axial equilibrium E1(x,0,0), planar equilibrium E2(0,˜A,˜y) and positive equilibrium E(x,A,y). The equilibrium points E0(0,0,0) and E1(x,0,0) are always present in the system without any restiction, where x=1r1(rr0).

    Existence of E2(0,˜A,˜y): ˜A and ˜y are positive solutions of the forthcoming system of non-linear equations

    λ(1A0)yβAϕAy=0,c2ϕAδey=0. (4.1)

    According to the second equation of the above system

    A=δ+eyϕc2.

    When we plug the value of A into the first equation of the system (4.1), we obtain

    ϕey2+(ϕδ+βeϕc2λ(1A0))y+βδ=0. (4.2)

    We observed that, the Eq (4.2) has two positive roots, if

    ϕc2λ(1A0)>ϕδ+βe,(ϕc2λ(1A0)ϕδβe)2>4ϕeβδ, (4.3)

    hold.

    System (2.3) has two prey-free equilibriums: ~E2(0,˜A,˜y) and ^E2(0,ˆA,ˆy) under the conditions given in the Eq (4.3). Again, if ϕc2λ(1A0)<ϕδ+βe, the Eq (4.2) has no positive roots. As a result, in this case E2 does not exist.

    Existence of positive equilibrium E(x,A,y): It is evident that x,A and y are the positive solutions of the forthcoming system of non-linear equations:

    r1+kA0yr0r1xαA0y(1+ax)(1+by)=0,λ(1A0)yβAϕAy=0,c1αA0x(1+ax)(1+by)+c2ϕAδey=0. (4.4)

    We can obtain from the second equation of the system (4.4)

    A=λ(1A0)yβ+ϕy.

    On substituting A in the first and third equations of the system (4.4) the forthcoming equations are obtained:

    r1+kA0yr0r1xαA0y(1+ax)(1+by)=0, (4.5)
    c1αA0x(1+ax)(1+by)+c2ϕλ(1A0)yβ+ϕyδey=0. (4.6)

    From the Eq (4.5), we observe the following points:

    (1)When y=0, then

    x=rr0r1=:x0.

    (2)When x=0, then the quadratic equation follows

    m1y2+m2y+m3=0, (4.7)

    where

    m1=kA0(αA0+br0)>0,m2=(r0(kA0+b)(brA0α)),m3=(rr0)<0.

    Equation (4.7) always has exactly one positive root (say y0). (3)The slope of the curve is

    dydx=r1αA0ay(1+ax)2(1+by)(rkA0(1+kA0y)2+αA0(1+ax)(1+by)2).

    It is evident that, dydx<0 exists under the following inequlity

    r1>αA0ays(1+bys). (4.8)

    From the Eq (4.6), we observe the following points:

    (1)When y=0, then

    x=δc1αA0aδ=:x1.

    (2)The slope of the Eq (4.6) at any point (x,y) is

    dydx=c1αA0(1+ax)2(1+by)(c1αA0bx(1+by)2(1+ax)c2βϕλ(1A0)(β+ϕy)2+e)>0.

    It is evident that, dydx>0 under the following inequlity,

    c2ϕλ(1A0)β<e(β+ϕyi)2. (4.9)

    From the above analysis, it is concluded that the curve of the Eq (4.5) comes into the first quadrant through the y-axis intersecting at (0,y0) and it is monotonically decreasing under the inequality (4.8) and then it leaves the first quadrant by intersecting the x-axis at (x0,0). The curve of the Eq (4.6) enters in the first qudrant through the x-axis intersecting at (x1,0). Under the inequality (4.9), the curve increases monotonically and approaches to its horizontal asymptote which is depited in the Figure 1 by the dashed line. Both the curves (4.5) and (4.6) intersect in the interior of first quadrant if

    x0>x1 (4.10)
    Figure 1.  The existence of unique interior equilibrium under the inequalities stated in Theorem 3.

    hold. From the above discussion, the system (2.3) has a unique positive solution E if the inequalities (4.8) to (4.10) hold.

    Consequently, we can formulate the following theorem.

    Theorem 3. If (4.8)–(4.10) hold, the system (2.3) has a unique interior equilibrium, E(x,A,y).

    Remark 2. The number of positive equilibrium for the system (2.3) depends on values of parameters, which we have chosen. Several possibilities are depicted in Figure 2.

    Figure 2.  Four possibilities of the prey and predator zero growth isoclines. We have selected the forthcoming parameter values. r=2,k=0.01,A0=0.5,r0=0.5,r1=0.043,α=0.3,a=0.15,b=0.02,c1=0.4,c2=0.5,δ=0.3,e=0.02,λ=2,β=0.32,ϕ=0.7. We vary the parameter δ. (a) For the parameter value δ=0.2, there is no existence of a positive equilibrium. (b) Positive equilibrium point exists uniquely when the parameter value δ=0.3. (c) There exists two positive equilibrium points for the parameter value δ=0.271. (d) Three positive equilibrium for parameter value δ=0.26.

    In ecology, an equilibrium point is said to be stable if a population settle and if the initial population is near to the equilibrium point. It is known that the local behavior of a system near an equilibrium point is the same as the corresponding linearized system. Hence, we investigate the stability behavior of the equilibrium points of the non-linear system (2.3).

    Theorem 4. A necessary and sufficient condition for the negativity of the real parts of all the roots of the polynomial L(ω)=ωn1+a1ωn11+.....+an with real coefficient is the positivity of all the principal diagonals of the minors of the Hurwitz matrix

    Hn=(a1100000....0a3a2a11000....0a5a4a3a2a100....0........................0000000....an).

    For a proof of this theorem, see Bellman [42].

    We calculate the variational matrix of the system and conclued the following point.

    ● The equilibrium point E0(0,0,0) is always saddle point having stable manifold along the A and y axes and unstable manifold along the x axis.

    ● The equilibrium point E1(x,0,0) is

    - locally asymptotically stable if δ>c1αA0x(1+ax) hold.

    - saddle point having stable manifold along the A and x-axes and unstable manifold along the y-axis if δ<c1αA0x(1+ax) hold.

    - stable if δ=c1αA0x(1+ax) hold.

    ● The Jacobian matrix examined at the prey-free equilibrium point E2(0,˜A,˜y) can be obtained by

    J|E2=[r1+kA0˜yr0αA0˜y(1+b˜y)000λ(1A0)˜y˜Aλ(1A0)ϕ˜Ac1αA0˜y(1+b˜y)c2ϕ˜ye˜y].

    The characteristic equation of the above matrix can be obtained by

    (ζ(r1+kA0˜yr0αA0˜y(1+b˜y)))(ζ2+(λ(1A0)˜y˜A+e˜y)ζ+((λ(1A0)˜y˜A(e˜y))c2ϕ˜y(λ(1A0)ϕ˜A)))=0. (4.11)

    The roots of the Eq (4.11) have negative real parts, under the following conditions

    r<(1+kA0˜y)(r0+αA0˜y(1+b˜y)),λ(1A0)<ϕ˜A. (4.12)

    Hence, using the Routh-Hurwitz criterion ~E2(0,˜A,˜y) is locally asymptotically stable under the conditions in (4.12).

    The roots of the Eq (4.11) have both positive and negative real parts if

    λ(1A0)˜y˜A(e˜y)<c2ϕ˜y(λ(1A0)ϕ˜A). (4.13)

    Therefore, ~E2(0,˜A,˜y) is a saddle point under the inequlity in (4.13).

    Remark 3. Similar analysis supports the stability behavior of ^E2(0,ˆA,ˆy) on replacing ˜A by ˆA and ˜y by ˆy.

    ● The Jacobian matrix at the unique positive equilibrium point E(x,A,y) is given by

    J|E=[r1x+αA0axy(1+ax)2(1+by)0rxkA0(1+kA0y)2αA0x(1+ax)(1+by)20λ(1A0)yAλ(1A0)ϕAc1αA0y(1+ax)2(1+by)c2ϕyc1αA0bxy(1+ax)(1+by)2ey].

    The characteristic equation of the above matrix is,

    η3+A1η2+A2η+A3=0, (4.14)

    where

    A1=r1xαA0axy(1+ax)2(1+by)+λ(1A0)yA+c1αA0bxy(1+ax)(1+by)2+ey,A2=(λ(1A0)yA)[c1αA0bxy(1+ax)(1+by)2+ey]c2ϕy(λ(1A0)ϕA)+(r1xαA0axy(1+ax)2(1+by))[c1αA0bxy(1+ax)(1+by)2+ey]+(rxkA0(1+kA0y)2+αA0x(1+ax)(1+by)2)[c1αA0y(1+ax)2(1+by)]+(r1xαA0axy(1+ax)2(1+by))(λ(1A0)yA),A3=(r1xαA0axy(1+ax)2(1+by))[(λ(1A0)yA)(c1αA0bxy(1+ax)(1+by)2+ey)c2ϕy(λ(1A0)ϕA)]+(rxkA0(1+kA0y)2+αA0x(1+ax)(1+by)2)((λ(1A0)yA)(c1αA0y(1+ax)2(1+by))).

    By the Routh-Hurwitz criterion [42,43], all the characteristic roots of J|E have negative real part iff

    A1>0,A3>0,A1A2>A3. (4.15)

    Thus, the following theorem can be stated.

    Theorem 5. The positive equilibrium point E(x,A,y) is locally asymptotically stable, if the conditions in (4.15) hold.

    It is shown that conditions in (4.15) hold true if

    λ(1A0)ϕ<A,A0<r1(1+ax)2(1+by)αay. (4.16)

    Remark 4. The model (2.3) is stable around its interior equilibrium E if the conditions in (4.16) hold.

    In the next theorem, we provide a criterion for global asymptotic stability of the unique positive equilibrium E(x,A,y) of the system (2.3).

    Theorem 6. The positive equilibrium E(x,A,y) of the system (2.3) is globally asymptotically stable below the forthcoming inequlities:

    A0<r1(1+ax)(1+by)αay,(rkA0(1+kA0y))2<4(r1αA0ay(1+ax)(1+by))(L2(c1αA0bx(1+bys)(1+ax)(1+by)+e)). (4.17)

    Proof. We select an appropriate positive definite function about E as

    V(x,A,y)=(xxxln(xx))+L12(AA)2+L2(yyyln(yy)),

    where L1 and L2 are positive constants and will be defined later. Now, we differentiate V in context of t along the solutions of system (2.3), we obtain

    dVdt=(xxx)dxdt+L1(AA)dAdt+L2(yyy)dydt=[r1αA0ay(1+ax)(1+by)(1+ax)](xx)2L1(β+ϕy)(AA)2L2[c1αA0bx(1+by)(1+by)(1+ax)+e](yy)2+(L1(λ(1A0)ϕA)+L2c2ϕ)(AA)(yy)+[rkA0(1+kA0y)(1+kA0y)αA0(1+ax)(1+by)(1+by)+L2c1αA0(1+by)(1+ax)(1+ax)](xx)(yy).

    Choosing L2=1+axc1(1+by) and L1=L2c2ϕϕAλ(1A0), we get

    dVdt=[r1αA0ay(1+ax)(1+by)(1+ax)](xx)2(1+ax)c2ϕc1(1+by)(ϕAλ(1A0))(β+ϕy)(AA)21+axc1(1+by)[c1αA0bx(1+by)(1+by)(1+ax)+e](yy)2rkA0(1+kA0y)(1+kA0y)(xx)(yy).

    Using the Sylvester criterion [43], if the inequalities in (4.17) hold, dVdt is negative definite. Hence, E is globally asymptotically stable above the inequalities in (4.17) [44].

    In dynamical system, Hopf-bifurcation is a occurrence in which stability of an equilibrium point bifurcates into instability and periodic solutions on varing a parameter. The parameter A0 is crucial in the system (2.3), we perform a Hopf-bifurcation analysis with A0 as the bifurcation parameter and obtain a threshold value A0=A0 where bifurcation occurs. The following are the necessary and sufficient criteria for Hopf-bifurcation to occur at A0=A0.

    (a) A1|A0>0,A3|A0>0,

    (b) f(A0)(A1A2A3)|A0=0,

    (c) Re[dηjdA0]A0=A00, where ηj,j=1,2,3 are the roots of Eq (4.14).

    We obtain an equation in A0 from the expression A1A2A3=0 and assume it has at least one positive root A0. Second, for some ϵ>0, there is an interval containing A0, (A0ϵ,A0+ϵ) such that A0ϵ>0 and A2>0 for A0(A0ϵ,A0+ϵ). Therefore, Eq (4.14) is unable to have any real positive root for A0(A0ϵ,A0+ϵ).

    Therefore, when A0=A0, we can write the Eq (4.14) as follows:

    (η+A1)(η2+A2)=0

    the above equation has the following three roots

    η1,2=±ιμ,η3=θ,

    where, μ=A2 and θ=A1.

    For A0(A0ϵ,A0+ϵ), above roots can be written as

    η1,2=k1(A0)±ιk2(A0),η=A1(A0).

    The transversality criterion need to be verified. On differentiation of Eq (4.14) in context of the bifurcation parameter A0, we get

    [dηdA0]A0=A0=[˙A1η2+˙A2η+˙A33η2+2A1η+A2]η=ιA2=dRdA02(A21+A2)+ι[A2˙A22A2A1A2dRdA02A2(A21+A2)],

    where R=A1A2A3 and ˙As,s=1,2,3 denote the derivative of As in context of time. Thus

    Re[dηdA0]A0=A0=dRdA02(A21+A2).

    Hence Re[dηdA0]A0=A00 if [dRdA0]A0=A00. This condition implies that the rate of change in the real part of characteristic roots at the bifurcation point is non-zero. The forthcoming theorem can be expressed as a result.

    Theorem 7. The system experiences a Hopf-bifurcation near the positive equilibrium E(x,A,y), if the necessary and sufficient conditions (a)(c) are satisfied. The threshold value of the bifurcation parameter A0 is provided by the equation f(A0)=0.

    From Theorem 7, the system (2.3) experiences a Hopf-bifurcation at A0=A0. Now we will discuss the direction and stability of bifurcating periodic orbit [45]. The Eigenvectors v1 and v3 of the variational matrix J|E corresponding to the characteristic roots η1,2=±ιμ, and η3=A1

    v1=[r11ιr12r21ιr22r31ιr32],v3=[r13r23r33],

    where

    r11=1,r12=0,r13=1,r21=(a11a22+μ2)a23a13(a13a22)2+(a13μ)2,r22=(a22a11)a23a13μ(a13a22)2+(a13μ)2,r23=(c11+C1)c23(c22+C1)c13,r31=c11c13,r32=μc13,r33=(c11+C1)c13.

    Now, we consider the transformation in system (2.3) in terms of X, Y and Z with x=x+r11X+r12Y+r13Z, A=A+r21X+r22Y+r23Z and y=y+r31X+r32Y+r33Z, the system (2.3) simplifies

    dXdt=(r22r33r32r23)Λ1+(r23r31r21r33)Λ2+(r21r32r31r22)Λ3ΛΘ1(X,Y,Z),dYdt=(r23r31r21r33)Λ1+(r33r11r31r13)Λ2+(r21r13r23r11)Λ3ΛΘ2(X,Y,Z),dZdt=(r21r32r22r31)Λ1+(r12r31r11r32)Λ2+(r11r22r12r21)Λ3ΛΘ1(X,Y,Z), (4.18)

    where Λ=r22r33r32r23+r21r32r31r22,

    Λ1=r(x+X+Z)1+kA0(y+r31X+r32Y+r33Z)r0(x+X+Z)r1(x+X+Z)2αA0(y+r31X+r32Y+r33Z)(x+X+Z)(1+a(x+X+Z))(1+b(y+r31X+r32Y+r33Z)),Λ2=(λ(1A0)(y+r31X+r32Y+r33Z))β(A+r21X+r22Y+r23Z)(ϕ(A+r21X+r22Y+r23Z)(y+r31X+r32Y+r33Z)),Λ3=αc1A0(y+r31X+r32Y+r33Z)(x+X+Z)(1+a(x+X+Z))(1+b(y+r31X+r32Y+r33Z))+(c2ϕ(A+r21X+r22Y+r23Z)(y+r31X+r32Y+r33Z))δ(y+r31X+r32Y+r33Z)e(y+r31X+r32Y+r33Z)2.

    The system (4.18) has an equilibrium point (0,0,0), and the variational matrix calculated at this point is as follows

    J(0,0,0)=[0μ0μ0000θ],

    where θ=Θ3y. At A0=A0 and (x,A,y)=(0,0,0), The values of g11, g02, g20, G21, e11, H101, H110, e20, μ, κ20, κ11 and g21 can be calculated by following relations

    g11=14(2Θ1X2+2Θ1Y2+ι(2Θ2X2+2Θ2Y2)),g02=14(2Θ1X22Θ1Y222Θ2XY+ι(2Θ2X22Θ2Y2+22Θ1XY)),g20=14(2Θ1X22Θ1Y2+22Θ2XY+ι(2Θ2X22Θ2Y222Θ1XY)),G21=18(3Θ1X3+3Θ1XY2+3Θ2X2Y+3Θ2Y3+ι(3Θ2X3+3Θ2XY23Θ1X2Y3Θ2Y3)),μ=Θ1Y,e11=14(2Θ3X2+2Θ3Y2),e20=14(2Θ3X22Θ3Y22ι2Θ3XY),
    H110=12(2Θ1XZ+2Θ2YZ+ι(2Θ2XZ2Θ1YZ)),H101=12(2Θ1XZ2Θ2YZ+ι(2Θ2XZ+2Θ1YZ)),g21=G21+2H110κ11+H101κ20.

    We can calculate the value for κ11 and κ20 by solving (θ2ιμ)κ20=e20 and θκ11=e11. Based on the above analysis, we can compute the following quantities

    C1(0)=ι2μ(g20g112|g11|2|g02|23)+g212,μ2=Re[C1(0)][dk1dA0]A0=A0,β2=2Re[C1(0)],T2=Im[C1(0)]+μ2[dk2dA0]A0=A0μ.

    From the above analysis of Hopf-bifurcation, we now state the following theorem.

    Theorem 8. (1) The Hopf-bifurcation curve exists for A0>A0 (A0<A0) if μ2>0 (μ2<0).

    (2) The bifurcating perodic solutions are stable (unstable) if β2<0 (β2>0).

    (3) The bifurcating perodic solution increases (decreases) if T2>0 (T2<0).

    In this section, we will conduct computer simulations using MATLAB R2022a to confirm the validity of our theoretical finidings discussed in the preceding sections. For the simulation, we have selected the forthcoming parameter values.

    r=3,k=5,A0=0.2,r0=0.2,r1=0.01,α=0.8,a=0.1,b=0.1,c1=0.5,c2=0.4,δ=0.35,e=0.025,λ=1.5,β=0.32,ϕ=0.7. (5.1)

    For the above set of parameters, the inequalities in Theorem 3 hold true, therefore the positive equilibrium point E(0.2863,1.5289,3.771) exist uniquely. Since the conditions in Theorem 5 are also followed by E (A1=3.0494>0,A3=0.0361>0,A1A2A3=0.3917>0), hence it is locally asymptotically stable. The stability nature of E is depicted in Figure 3. The left panel illustrates the time series evolution of population and the right panel illustrates phase portrait in the xAy-space. The figure illustrates that both species fluctuate around their equilibrium level for a finite time before settling. It illustrates the local asymptotic stability of the system (2.3) around the interior equilibrium.

    Figure 3.  Evolution of time series (a) and species phase portrait (b) for the set of parameters selected in (5.1). Interior equilibrium point E is locally asymptotically stable.

    Now, to investigate the effect of the degree of dependency A0, we increase it keeping all other parameters unchanged. We noticed that the transient period is increasing with A0, which indicates a route to Hopf-bifurcation. At A0=0.2872(A0) the system undergoes a Hopf-bifurcation and populations face perpetual oscillations. The necessary and sufficient conditions given in Theorem 5 are also satisfied. The bifurcation value of A0 can be calculated by solving the equation f(A0)=0. For further values of A0, the system has a limit cycle. In Figure 4, the unstable nature of E and limit cycle are plotted. A solution trajectory, which is initiated from near to the positive equilibrium is attracted by the limit cycle around it (see Figure 4c). This shows that the Hopf-bifurcation is supercritical. If we further increase the parameter A0 then at the higher dependency of predators on prey the system again get back to its stability near E via another Hopf-bifurcation. Beyond another critical value of A0, A0=0.9291, the limit cycles leave the system and E becomes stable. The stable behavior of E at A0=0.955>A0 is illustrated in the Figure 5.

    Figure 4.  Unstable behavior of positive equilibrium point and existence of stable limit cycle, when A0=0.5>A0.
    Figure 5.  Evolution of time series (a) and species phase portrait (b), when A0=0.955>A0, the interior equilibrium E is stable.

    As we vary parameter A0, we can see that the system changes its stability twice. Thus, the Hopf-bifurcation occurs multiple times in the full range of the dependency factor. In order to provide a comprehensive representation of the Hopf-bifurcation, we have constructed a three-dimensional bifurcation diagram as depicted in Figure 6. This diagram illustrates the attractors of the system in the prey and predator plane for many values of parameter A0. It is remarkable that the positive equilibrium point remains stable below the first critical value of A0. However, once the critical value is exceeded, the trajectories are attracted toward a limit cycles. These stable limit cycles are represented by the colored closed orbits displayed in the diagram. When A0=0.2872=A0, we can see that all the conditions of Theorem 7 are satisfied. Also we comutate that

    C1(0)=0.00090.0225i,μ2=0.016221>0,β2=0.0018<0,T2=0.1112>0.
    Figure 6.  Bifurcation diagram with respect to parameter A0, representing attractors (equilibrium points and limit cycles).

    By Theorem 8, this result implies that the Hopf-bifurcation taking place, is of supercritical nature, and bifurcating periodic solutions are stable. When A0=0.9291=A0, Hopf-bifucation occurs and the limit cycle disappears. We can see that the all conditions of Theorem 7 are satisfied and

    C1(0)=0.00040.0337i,μ2=0.02581<0,β2=0.000817<0,T2=0.06389>0.

    By Theorem 8, this result implies that the Hopf-bifurcation taking place is of supercritical nature, and bifurcating periodic solutions are stable.

    From Figure 6, we observe that at the low degree of dependency, the predator has a higher dependency on additional food, resulting in the system (2.3) being stable. However, as the dependency factor gradually increases, the system (2.3) is unstable and forms a limit cycle. This indicates that the predator's dependence extends to both the prey and additional food resources. Finally, as we further increase the value of A0, the predator's reliance shifts more towards the prey, leading to the system (2.3) returning to a stable state. Overall, the bifurcation diagram visually demonstrates how the system (2.3) dynamics change as the dependency of predator on prey varies, showing the changes from stable to unstable states as well as the effect of predator dependency factor on both prey and supplementary nourishment.

    The fear parameter k is also a vital parameter that significantly influences the dynamics of the system. For the best illustration of varing k, we change the parameter A0=0.4, whereas other parameters are the same as in (5.1). Initially, for lower values of k, the system is stable near its positive equilibrium. Now, we start increasing the fear parameter, and this change in k opposes the stability and at k=2.5441(k), the system falls into the never ending oscillations via a Hopf-bifurcation. If we choose k beyond k, then E becomes unstable and the trajectory started from near by points approaches a stable limit cycle. The unstable equilibrium point and stable limit cycle at k=6>k are shown in the Figure 7.

    Figure 7.  Existence of limit cycle for k=6>k.

    Figure 8 provides a comprehensive bifurcation diagram that illustrates the behavior changes as the fear parameter k rises. It is evident from the figure that the system remains stable for lower values of k. However, as k approaches its threshold value (k), the system destabilizes and experiences a Hopf-bifurcation. Beyond the critical value, the equilibrium point becomes unstable and a periodic solution appears in the system. When k=2.552=k, we can see that all the conditions of Theorem 7 are satisfied. Also, we computate that

    C1(0)=0.00070.01665i,μ2=15.02683>0,β2=0.00145<0,T2=0.807699>0.
    Figure 8.  Prey and predator species bifurcation diagram in context of parameter k.

    By Theorem 8, the bifurcating periodic solution exists for k<k, the bifurcating periodic solution increases and bifurcating periodic solutions are stable.

    The behavior of system (2.3) experiences Hopf-bifurcation in the context of the parameters A0 and k. Hence, A0k-plane can be divided into two types of regions, region of stability where the equilibrium point is stable and the region of instability where the equilibrium point is unstable and limit cycle presents in the system. Both the regions are separated by a curve, where the characteristic roots of variational matrix are purely imaginary, known as Hopf-bifurcation curve. Previously, we noticed that the system switches its stability twice on varying parameter A0, while only once on varying the fear parameter k. Therefore, there are two stability regions and one instability region in the A0k-plane separated by two Hopf-bifurcation curves. The stability and instability regions are represented in Figure 9. When we consider a particular value of the cost of fear k=3, there are two bifurcation points for the parameter A0: A0=0.3735 and A0. System (2.3) remains stable until A_{0}^{'} as depicted in the Figure 9. However, when A_{0} exceeds A_{0}^{'} ( A_{0} > A_{0}^{'} ), the system becomes unstable and enters a limit cycle. If we further increase A_{0} beyond A_{0}^{''} ( A_{0} > A_{0}^{''} ), the system (2.3) regains its stability.

    Figure 9.  Stability and instability regions of the system (2.3) in A_{0}k -plane.

    Now, to maintain the equilibrium levels of prey ( x^* = 2 ) and predator ( y^* = 2 ), we vary the parameters k and A_{0} whereas the other parameters are fixed (same as in (5.1)). We notice that we have to decrease the value of parameter A_{0} as the parameter k increases to maintain the population density of prey or predator (refer the Figure 10).

    Figure 10.  A graph in values of parameters A_{0} and k maintaining the prey and predators level constant.

    The prey-predator system in ecology describes the relationship between species, where the prey population supports the predator population as a food source, and the predator population consumes the prey population through predation. When additional food is introduced into a prey-predator system, it can lead to growth in the prey population, which subsequently supports a larger predator population due to more available food resources. This can result in a more stable or balanced system compared to a system without additional food. Numerous studies [20,30,31,32] have investigated the effects of additional food in prey-predator systems, providing valuable insights into how this factor can influence population dynamics and ecological interactions between the species involved. Recent studies [37,38,39] have examined the effect of predator-induced fear on prey individuals, leading to a reduction in their reproducion rate. The separation of the fear effect from direct killing by predators was made feasible due to some extensive experimental efforts by various ecologists [33]. The Crowley-Martin functional response model investigates how predator predation rates change with variations in prey density and shows more realistic dynamics [13,26,27]. In this research, we examine a prey-predator model that incorporates three important factors: Fear effect in the prey species induced by predators, predator-prey interaction by the Crowley-Martin functional response, and the influence of providing additional food to the predator. The presence of additional food may influence the predator's dependency on prey, which may lead to a significant difference in the dynamics of the system. It is essential to understand how much the predator depends on prey and additional food. We introduced a new parameter, the degree of dependency, which indicates the predator's dependency on prey. After conducting a thorough analysis, we have demonstrated that the proposed system is characterized by positivity, boundedness and persistence, highlighting its ecological significance and well-posedness. We conducted an analysis to examine how the number of interior steady states and their stability characteristics change with variations in parameters. Our system exhibits four types of equilibria: The trivial equilibrium E_{0}(0, 0, 0) , the axial equilibrium E_{1}(x^*, 0, 0) , two prey-free equilibria \tilde {E_{2}} and \hat {E_{2}} under condition (4.3) and a unique positive equilibrium E^* under conditions (4.8)–(4.10). We have carried out the local and global stability of all existing equilibria in the system. With the stability results, we conclude that the coexistence of both species depends on the equilibrium level of provided additional food and the degree of dependency (refer to the set of inequalities in (4.16) and (4.17)). This analysis carries various conditions with the stability of the equilibrium being influenced by the degree of dependency.

    We have shown that the degree of dependency plays a vital role in stabilize the system through Hopf-bifurcation. The Hopf-bifurcation phenomenon has also been validated by a computer simulation. We noticed, for a particular range of cost of fear, that the system switches its stability around the positive equilibrium point multiple times (see Figure 9). This indicates that the system's stability relies on the absence or minimal degree of dependency between the two species. If the dependency exceeds a certain initial threshold, both species begin oscillating around unstable steady states. However, if the dependency attains another threshold, the system regains stability. This provides a clear demonstration of how changes in the degree of dependency impact the ecological dynamics in the predator-prey system.

    Additionally, we explored the Hopf-bifurcation with respect to the fear parameter k and obtained the threshold value k^* = 2.58 , where the system destabilizes. This shows that the system is stable if the degree of fear is low, but if the fear is increased beyond a level, then both the species start oscillating around their unstable steady state. From the ecological point of view, this nature can be justified. Initially, when k is very small, both populations coexist with stability. Now, we start increasing the cost of fear, which leads to a reduction in the reproduction rate of prey. Due to this, the prey population starts decreasing and consequently, the predators also start decreasing due to the scarcity of food. At the low density of predators, the fear effect gets weak and the prey population starts growing. Now, the abundance of food predators also increases. After a certain level of cost of fear, both the populations oscillate perpetually. This leads to limit cycles in the system at the higher values of cost of fear.

    In conclusion, we find that predator-induced fear and the predator's dependency on prey both are effective in controlling the oscillations in the system, thereby playing a crucial role in stabilizing or balancing the ecosystem. These factors contribute significantly to the regulation and persistence of predator-prey dynamics, highlighting their importance in ecological stability. The outcome of our study has an important effect on the conservation of prey populations, as they are supported by a system with additional food for predators. We believe that our research will deepen the understanding of how ecological dynamics change in the presence of fear effect, additional food resources, and degree of dependency are added.

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

    Author's are thankful to anonymous reviewers for careful reading and constructive suggestions that improved the quality and presentation of the paper.

    The authors declare that they have no competing interests regarding the present study.



    [1] A. J. Lotka, Elements of physical biology, 1925, New York: Williams & Wilkins.
    [2] V. Volterra, Variazioni e fluttuazioni del numero d'individui in specie animali conviventi, Societá anonima tipografica "Leonardo da Vinci", 1927.
    [3] Z. H. Ma, S. F. Wang, A delay-induced predator-prey model with Holling type functional response and habitat complexity, Nonlinear Dynam., 93 (2018), 1519–1544. https://doi.org/10.1007/s11071-018-4274-2 doi: 10.1007/s11071-018-4274-2
    [4] C. S. Holling, The components of predation as revealed by a study of small mammal predation of the {E}uropean pine sawfly, Can. Entomol., 91 (1959), 293–320.
    [5] R. Xu, M. A. J. Chaplain, F. A. Davidson, Global stability of a Lotka-Volterra type predator-prey model with stage structure and time delay, Appl. Math. Comput., 159 (2004), 863–880. https://doi.org/10.1016/j.amc.2003.11.008 doi: 10.1016/j.amc.2003.11.008
    [6] S. K. Sasmal, Y. Takeuchi, Dynamics of a predator-prey system with fear and group defense, J. Math. Anal. Appl., 481 (2020), 123471. https://doi.org/10.1016/j.jmaa.2019.123471 doi: 10.1016/j.jmaa.2019.123471
    [7] G. R. Jiang, Q. S. Lu, L. N. Qian, Complex dynamics of a Holling type II prey-predator system with state feedback control, Chaos Soliton. Fract., 31 (2007), 448–461. https://doi.org/10.1016/j.chaos.2005.09.077 doi: 10.1016/j.chaos.2005.09.077
    [8] Y. J. Huang, F. D. Chen, L. Zhong, Stability analysis of a prey-predator model with Holling type III response function incorporating a prey refuge, Appl. Math. Comput., 182 (2006), 672–683. https://doi.org/10.1016/j.amc.2006.04.030 doi: 10.1016/j.amc.2006.04.030
    [9] Z. H. Liu, R. Yuan, Stability and bifurcation in a delayed predator-prey system with Beddington-DeAngelis functional response, J. Math. Anal. Appl., 296 (2004), 521–537. https://doi.org/10.1016/j.jmaa.2004.04.051 doi: 10.1016/j.jmaa.2004.04.051
    [10] R. K. Upadhyay, R. K. Naji, Dynamics of a three species food chain model with Crowley-Martin type functional response, Chaos Soliton, Fract., 42 (2009), 1337–1346. https://doi.org/10.1016/j.chaos.2009.03.020 doi: 10.1016/j.chaos.2009.03.020
    [11] W. j. Lu, Y. H. Xia, Periodic solution of a stage-structured predator-prey model with Crowley-Martin type functional response, AIMS Math., 7 (2022), 8162–8175. https://www.aimspress.com/article/doi/10.3934/math.2022454 doi: 10.3934/math.2022454
    [12] M. Falconi, M. Huenchucona, C. Vidal, Stability and global dynamic of a stage-structured predator-prey model with group defense mechanism of the prey, Appl. Math. Comput., 270 (2015), 47–61. https://doi.org/10.1016/j.amc.2015.07.109 doi: 10.1016/j.amc.2015.07.109
    [13] X. Y. Meng, H. F. Huo, H. Xiang, Q. Y. Yin, Stability in a predator-prey model with Crowley-Martin function and stage structure for prey, Appl. Math. Comput., 232 (2014), 810–819. https://doi.org/10.1016/j.amc.2014.01.139 doi: 10.1016/j.amc.2014.01.139
    [14] X. Y. Wang, L. Zanette, X. F. Zou, Modelling the fear effect in predator-prey interactions, J. Math. Biol., 73 (2016), 1179–1204. https://doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
    [15] B. F. Xie, N. Zhang, Influence of fear effect on a Holling type III prey-predator system with the prey refuge, AIMS Math., 7 (2022), 1811–1830. https://doi.org/10.3934/math.2022104 doi: 10.3934/math.2022104
    [16] Q. Liu, D. Q. Jiang, T. Hayat, A. Alsaedi, Dynamics of a stochastic predator-prey model with stage structure for predator and Holling type II functional response, J. Nonlinear Sci., 28 (2018), 1151–1187. https://doi.org/10.1007/s00332-018-9444-3 doi: 10.1007/s00332-018-9444-3
    [17] A. Kumar, B. Dubey, Stability and bifurcation of a prey-predator system with additional food and two discrete delays, CMES-Comp. Model. Eng., 126 (2021), 505–547. https://doi.org/10.32604/cmes.2021.013206 doi: 10.32604/cmes.2021.013206
    [18] J. C. Huang, S. G. Ruan, J. Song, Bifurcations in a predator-prey system of Leslie type with generalized Holling type III functional response, J. Differ. Equations, 257 (2014), 1721–1752. https://doi.org/10.1016/j.jde.2014.04.024 doi: 10.1016/j.jde.2014.04.024
    [19] S. Belvisi, E. Venturino, An ecoepidemic model with diseased predators and prey group defense, Simul. Model. Pract. Th., 34 (2013), 144–155. https://doi.org/10.1016/j.simpat.2013.02.004 doi: 10.1016/j.simpat.2013.02.004
    [20] A. A. Thirthar, S. J. Majeed, M. A. Alqudah, P. Panja, T. Abdeljawad, Fear effect in a predator-prey model with additional food, prey refuge and harvesting on super predator, Chaos Soliton. Fract., 159 (2022), 112091. https://doi.org/10.1016/j.chaos.2022.112091 doi: 10.1016/j.chaos.2022.112091
    [21] J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 44 (1975), 331–340. https://doi.org/10.2307/3866 doi: 10.2307/3866
    [22] D. L. DeAngelis, R. A. Goldstein, R. V. O'Neill, A model for tropic interaction, Ecology, 56 (1975), 881–892. https://doi.org/10.2307/1936298 doi: 10.2307/1936298
    [23] A. P. Maiti, B. Dubey, J. Tushar, A delayed prey-predator model with Crowley-Martin-type functional response including prey refuge, Math. Method. Appl. Sci., 40 (2017), 5792–5809. https://doi.org/10.1002/mma.4429 doi: 10.1002/mma.4429
    [24] P. H. Crowley, E. K. Martin, Functional responses and interference within and between year classes of a dragonfly population, J. N. Am. Benthol. Soc., 8 (1989), 211–221. https://doi.org/10.2307/1467324 doi: 10.2307/1467324
    [25] X Y. Zhou, J. G. Cui, Global stability of the viral dynamics with Crowley-Martin functional response, Bull. Korean Math. Soc., 48 (2011), 555–574. https://doi.org/10.4134/BKMS.2011.48.3.555 doi: 10.4134/BKMS.2011.48.3.555
    [26] J. P. Tripathi, S. Tyagi, S. Abbas, Global analysis of a delayed density dependent predator-prey model with Crowley-Martin functional response, Commun. Nonlinear Sci. Numer. Simul., 30 (2016), 45–69. https://doi.org/10.1016/j.cnsns.2015.06.008 doi: 10.1016/j.cnsns.2015.06.008
    [27] N. E. Papanikolaou, N. Demiris, P. G. Milonas, S. Preston, T. Kypraios, Does mutual interference affect the feeding rate of aphidophagous coccinellids? A modeling perspective, Plos One, 11 (2016), e0146168. https://doi.org/10.1371/journal.pone.0146168 doi: 10.1371/journal.pone.0146168
    [28] B. S. R. V. Prasad, M. Banerjee, P. D. N. Srinivasu, Dynamics of additional food provided predator-prey system with mutually interfering predators, Math. Biosci., 246 (2013), 176–190. https://doi.org/10.1016/j.mbs.2013.08.013 doi: 10.1016/j.mbs.2013.08.013
    [29] B. Sahoo, S. Poria, The chaos and control of a food chain model supplying additional food to top-predator, Chaos Soliton. Fract., 58 (2014), 52–64. https://doi.org/10.1016/j.chaos.2013.11.008 doi: 10.1016/j.chaos.2013.11.008
    [30] P. D. N. Srinivasu, B. S. R. V. Prasad, M. Venkatesulu, Biological control through provision of additional food to predators: A theoretical study, Theor. Popul. Biol., 72 (2007), 111–120. https://doi.org/10.1016/j.tpb.2007.03.011 doi: 10.1016/j.tpb.2007.03.011
    [31] J. Ghosh, B. Sahoo, S. Poria, Prey-predator dynamics with prey refuge providing additional food to predator, Chaos Soliton. Fract., 96 (2017), 110–119. https://doi.org/10.1016/j.chaos.2017.01.010 doi: 10.1016/j.chaos.2017.01.010
    [32] B. Sahoo, Role of additional food in eco-epidemiological system with disease in the prey, Appl. Math. Comput., 259 (2015), 61–79. https://doi.org/10.1016/j.amc.2015.02.038 doi: 10.1016/j.amc.2015.02.038
    [33] L. Y. Zanette, A. F. White, M. C. Allen, M. Clinchy, Perceived predation risk reduces the number of offspring songbirds produce per year, Science, 334 (2011), 1398–1401. https://doi.org/10.1126/science.1210908 doi: 10.1126/science.1210908
    [34] W. Cresswell, Predation in bird populations, J. Ornithol., 152 (2011), 251–263. https://doi.org/10.1007/s10336-010-0638-1 doi: 10.1007/s10336-010-0638-1
    [35] B. Dubey, N, Sajan, A. Kumar, Stability switching and chaos in a multiple delayed prey-predator model with fear effect and anti-predator behavior, Math. Comput. Simulat., 188 (2021), 164–192. https://doi.org/10.1016/j.matcom.2021.03.037 doi: 10.1016/j.matcom.2021.03.037
    [36] M. M. Chen, Y. Takeuchi, J. F. Zhang, Dynamic complexity of a modified Leslie-Gower predator-prey system with fear effect, Commun. Nonlinear Sci. Numer. Simul., 119 (2023), 107109. https://doi.org/10.1016/j.cnsns.2023.107109 doi: 10.1016/j.cnsns.2023.107109
    [37] K. Sarkar, S. Khajanchi, Impact of fear effect on the growth of prey in a predator-prey interaction model, Ecol. Complex., 42 (2020), 100826. https://doi.org/10.1016/j.ecocom.2020.100826 doi: 10.1016/j.ecocom.2020.100826
    [38] H. S. Zhang, Y. L. Cai, S. M. Fu, W. M. Wang, Impact of the fear effect in a prey-predator model incorporating a prey refuge, Appl. Math. Comput., 356 (2019), 328–337. https://doi.org/10.1016/j.amc.2019.03.034 doi: 10.1016/j.amc.2019.03.034
    [39] M. Clinchy, M. J. Sheriff, L. Y. Zanette, Predator-induced stress and the ecology of fear, Funct. Ecol., 27 (2013), 56–65. https://doi.org/10.1111/1365-2435.12007 doi: 10.1111/1365-2435.12007
    [40] M. P. Hassell, Mutual interference between searching insect parasites, J. Anim. Ecol., 40 (1971), 473–486. https://doi.org/10.2307/3256 doi: 10.2307/3256
    [41] J. P. Tripathi, S. Abbas, M. Thakur, A density dependent delayed predator-prey model with Beddington-Deangelis type function response incorporating a prey refuge, Commun. Nonlinear Sci. Numer. Simul., 22 (2015), 427–450. https://doi.org/10.1016/j.cnsns.2014.08.018 doi: 10.1016/j.cnsns.2014.08.018
    [42] R. Bellman, Introduction to matrix analysis, SIAM, New York, 1997.
    [43] S. Ahmad, M. R. M. Rao, Theory of ordinary differential equations: With applications of biology and engineering, Affiliated East-West Private Limited, New Delhi, 1999.
    [44] L. Perko, Differential equations and dynamical systems, Springer Science & Business Media, 2013.
    [45] B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and applications of Hopf bifurcation, CUP Archive, London, UK, 1981.
  • This article has been cited by:

    1. Shri Harine P, Ankit Kumar, Reshma K P, Local and global dynamics of a prey–predator system with fear, Allee effect, and variable attack rate, 2024, 34, 1054-1500, 10.1063/5.0227458
    2. Ankit Kumar, Reshma K P, Shri Harine P, Global dynamics of an ecological model in presences of fear and group defense in prey and Allee effect in predator, 2025, 0924-090X, 10.1007/s11071-024-10706-8
    3. Reshma K P, Ankit Kumar, Global Dynamics of a Stage Structure Prey–Predator Model With Fear, Group Defense, and Antipredator Behavior, 2025, 0170-4214, 10.1002/mma.10845
  • Reader Comments
  • © 2024 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(1826) PDF downloads(124) Cited by(3)

Figures and Tables

Figures(10)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog