Processing math: 100%
Research article

Chaos detection in predator-prey dynamics with delayed interactions and Ivlev-type functional response

  • Received: 06 June 2024 Revised: 21 July 2024 Accepted: 01 August 2024 Published: 21 August 2024
  • MSC : 34C15, 34C23, 37G15, 37N25

  • Regarding delay-induced predator-prey systems, extensive research has focused on the phenomenon of delayed destabilization. However, the question of whether delays contribute to stabilizing or destabilizing the system remains a subtle one. In this paper, the predator-prey interaction with discrete delay involving Ivlev-type functional response is studied by theoretical analysis and numerical simulations. The positivity and boundedness of the solution for the delayed model have been discussed. When time delay is accounted as a bifurcation parameter, stability analysis for the coexistence equilibrium is given in theoretical aspect. Supercritical Hopf bifurcation is detected by numerical simulation. Interestingly, by choosing suitable groups of parameter values, the chaotic solutions appear via a cascade of period-doubling bifurcations, which is also detected. The theoretical analysis and numerical conclusions demonstrate that the delay mechanism plays a crucial role in the exploration of chaotic solutions.

    Citation: Qinghui Liu, Xin Zhang. Chaos detection in predator-prey dynamics with delayed interactions and Ivlev-type functional response[J]. AIMS Mathematics, 2024, 9(9): 24555-24575. doi: 10.3934/math.20241196

    Related Papers:

    [1] Ahmad Suleman, Rizwan Ahmed, Fehaid Salem Alshammari, Nehad Ali Shah . Dynamic complexity of a slow-fast predator-prey model with herd behavior. AIMS Mathematics, 2023, 8(10): 24446-24472. doi: 10.3934/math.20231247
    [2] Aeshah A. Raezah, Jahangir Chowdhury, Fahad Al Basir . Global stability of the interior equilibrium and the stability of Hopf bifurcating limit cycle in a model for crop pest control. AIMS Mathematics, 2024, 9(9): 24229-24246. doi: 10.3934/math.20241179
    [3] A. Q. Khan, Ibraheem M. Alsulami . Complicate dynamical analysis of a discrete predator-prey model with a prey refuge. AIMS Mathematics, 2023, 8(7): 15035-15057. doi: 10.3934/math.2023768
    [4] Jin Liao, André Zegeling, Wentao Huang . The uniqueness of limit cycles in a predator-prey system with Ivlev-type group defense. AIMS Mathematics, 2024, 9(12): 33610-33631. doi: 10.3934/math.20241604
    [5] Guilin Tang, Ning Li . Chaotic behavior and controlling chaos in a fast-slow plankton-fish model. AIMS Mathematics, 2024, 9(6): 14376-14404. doi: 10.3934/math.2024699
    [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] Nazmul Sk, Bapin Mondal, Abhijit Sarkar, Shyam Sundar Santra, Dumitru Baleanu, Mohamed Altanji . Chaos emergence and dissipation in a three-species food web model with intraguild predation and cooperative hunting. AIMS Mathematics, 2024, 9(1): 1023-1045. doi: 10.3934/math.2024051
    [8] Chenxuan Nie, Dan Jin, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and generalist predator. AIMS Mathematics, 2022, 7(7): 13344-13360. doi: 10.3934/math.2022737
    [9] Heping Jiang . Complex dynamics induced by harvesting rate and delay in a diffusive Leslie-Gower predator-prey model. AIMS Mathematics, 2023, 8(9): 20718-20730. doi: 10.3934/math.20231056
    [10] Sahabuddin Sarwardi, Hasanur Mollah, Aeshah A. Raezah, Fahad Al Basir . Direction and stability of Hopf bifurcation in an eco-epidemic model with disease in prey and predator gestation delay using Crowley-Martin functional response. AIMS Mathematics, 2024, 9(10): 27930-27954. doi: 10.3934/math.20241356
  • Regarding delay-induced predator-prey systems, extensive research has focused on the phenomenon of delayed destabilization. However, the question of whether delays contribute to stabilizing or destabilizing the system remains a subtle one. In this paper, the predator-prey interaction with discrete delay involving Ivlev-type functional response is studied by theoretical analysis and numerical simulations. The positivity and boundedness of the solution for the delayed model have been discussed. When time delay is accounted as a bifurcation parameter, stability analysis for the coexistence equilibrium is given in theoretical aspect. Supercritical Hopf bifurcation is detected by numerical simulation. Interestingly, by choosing suitable groups of parameter values, the chaotic solutions appear via a cascade of period-doubling bifurcations, which is also detected. The theoretical analysis and numerical conclusions demonstrate that the delay mechanism plays a crucial role in the exploration of chaotic solutions.



    Investigating the dynamic interaction and interplay between species is essential in mathematical ecology [1,2]. Modeling such systems and analyzing their dynamical behavior may give a prediction on the evolution of populations. Particularly, two-species predator-prey models have led to enthusiasm among many scholars [3,4,5]. The Gause type two-species predator-prey model is given by

    dx(t)dt=rx(t)(1x(t)K)y(t)f(x(t)),dy(t)dt=sy(t)+Yy(t)f(x(t)), (1.1)

    where x(t) and y(t) denote the density of the prey and predator species at time t, respectively. Parameters r,K,s, and Y are positive constants, which denote the intrinsic reproduction rate of the prey, the carrying capacity for the prey species, the death rate of the predator species, and the growth yield constant for the conversion of prey to predator density, respectively [6,7].

    The initial values are x(0)0,y(0)0 due to their biological meanings. In the absence of y(t) in the model (1.1), the prey increases according to the logistic growth ˙x(t)=rx(t)(1x(t)K). The predator y(t) declined exponentially as ˙y(t)=sy(t) and will eventually die in the long-term if the model lacks prey x(t).

    The function f(x) represents the prey-dependent functional response, which is the Ivlev-type functional response [8,9], and takes the form

    f(x)=α(1eβx), (1.2)

    where α>0 is the consumption rate and β>0 is the physiological rate at which saturation is achieved. It is a monotone increasing function that saturates, that is, it has a finite positive limit α as x approaches infinity [1] (see Figure 1).

    Figure 1.  The graph of f(x)=α(1eβx).

    This type functional response is both monotonically increasing and uniformly bounded, which was classified to the Holling-Ⅱ functional response by Garvie [10,11]. Biologically, it was first proposed to describe the increase of the fish. Hence, our results are useful in designing fishing policies for the fishery industry. Other forms of functional responses can be seen in [12,13].

    Considering the Ivlev-type trophic response (1.2) in system (1.1), it changes to

    dx(t)dt=rx(t)(1x(t)K)αy(t)(1eβx(t)),dy(t)dt=sy(t)+Yαy(t)(1eβx(t)). (1.3)

    The dynamical behaviors of the model (1.3) have been investigated extensively. It has the only coexistence equilibrium (x,y) if

    αY>s,1eKβ<sαY, (1.4)

    where

    x=1βln(1sαY),y=rYsx(1xK).

    After applying rescalings in x,y, and t, it can be assumed that K=α=Y=1. Under the assumption (1.4), if

    β>2s+(1s)ln(1s)s+(1s)ln(1s)ln(1s),

    then there exists a unique stable limit cycle. Otherwise, it has no limit cycles. If (1.4) fails, then system (1.3) has no existence equilibrium. Therefore, no limit cycles of system (1.3) exist.

    In 1925, to investigate fish population under harvesting, the predator-prey model with delay was proposed by Volterra. It is described by an integro-differential equation as

    dx(t)dt=rx(t)[11KTG(ts)x(s)ds].

    The above delayed equation is called an integro-differential equation, and such delays are called distributed delays. We can use the linear chain trick to convert systems into systems with discrete delay. Since then, delayed differential equations (DDEs) have been extensively used to model population dynamics [14], neural network [15,16], engineering, the life sciences, etc., including predator-prey interactions.

    By [17], assume the growth rate of the predator species y(t) is proportional to the number of individuals in the population tτ time units in the past that manage to survive until time t. In order to obtain an expression that describes how many predator individuals alive at time tτ are still alive at time t, where τ is the delay due to the gestation of the y(t) [18], we need to solve the following first-order ordinary differential equation for y(t),

    ˙y(t)=sy(t).

    It implies that

    y(t)y(tτ)1sy(t)dy=ttτdt,

    hence

    y(t)=y(tτ)esτ, (1.5)

    where the factor esτ denotes the survival rate of the predator y(t) which was born at time tτ and still remains alive at the time t. When the time delay τ=0, the right side of (1.5) reduces to its prototype y(t). The main difference between y(tτ) and y(tτ)esτ is shown in Figure 2.

    Figure 2.  The difference of consideration of the time delay or not.

    When the time delay τ is small, they are close [19]. However, as the delay increases, the expression y(tτ)esτ could describe practical problems better than the expression y(tτ). Although the predator-prey interaction with y(tτ) has been extensively investigated since its proposal, such systems with the delay term y(tτ)esτ are scarce and are not frequently reported. Therefore, compared to existing studies on the predator-prey system [20,21,22], this is the main contribution and the novelty of this paper in the aspect of establishing the model.

    Similarly, with the well-known Wangersky-Cunningham model, we assume that the change rate of predators depends on the number of prey and of predators present at τ previous time, that is, the delay τ in the interaction term y(t)(1eβx(t)) of the second equation. Therefore, we replace y(t) with the Eq (1.5) in model (1.3), and system (1.3) is reduced to

    dx(t)dt=rx(t)(1x(t)K)αy(t)(1eβx(t)),dy(t)dt=sy(t)+Yαy(tτ)esτ(1eβx(tτ)). (1.6)

    System (1.6) has the initial data

    x(η)=ϕ(η)0,y(η)=ψ(η)0,η[τ,0],ϕ(0)>0,ψ(0)>0, (1.7)

    where (ϕ(η),ψ(η))C([τ,0],R2+0) is the Banach space of continuous functions mapping the interval [τ,0] into R2+0, where R2+0={(x,y):x0,y0} [23]. By the fundamental theory of DDEs, system (1.6) has a unique solution x(t),y(t) satisfying initial data (1.7).

    The main goal of this paper is to show how the delay τ affects the dynamics of model (1.6). This paper is organized as follows. In Section 2, we prove the positivity and boundedness for the solution of system (1.6). In Section 3, when time delay is accounted as a bifurcation parameter, the stability analysis is given for the coexistence equilibrium for model (1.6). We analytically prove that the local Hopf bifurcation critical values are neatly paired. In Section 4, numerical explorations using the numerical continuation software XPPAUT and DDE-Biftool are carried out in order to substantiate the obtained theoretical results. Simulations indicated that as the delay increases, the positive equilibrium loses its stability and bifurcates a family of orbitally asymptotically stable periodic solutions. The coexistence equilibrium undergoes stability switches. For large enough delay, the predator will die out. Before the extinction of the predator, rich dynamics such as Hopf bifurcation, period doubling bifurcation and strange attractor have been demonstrated when time delay is accounted as bifurcation parameter, and the abundance of steady-state chaotic solutions appears via a cascade of period-doubling bifurcations is also detected. The coexistence equilibrium undergos transcritical bifurcation at the die out critical value. We summarize our conclusions in Section 5, especially on the impact of time delay from the biological aspect.

    For the system (1.6) with positive initial data (1.7), we first prove the following two theorems concerning the positivity and boundedness of the solution [24].

    Theorem 2.1. Solutions of system (1.6) with positive initial data (1.7) remain positive for t>0.

    Proof. Assume (x(t),y(t)) is a solution of system (1.6) satisfying initial data (1.7). By the Theorem 2.1 of [25], we solve the following ordinary differential equation (ODE):

    dx(t)x(t)=[r(1x(t)K)α(1eβx(t))y(t)x(t)]dt.

    Integrating between the limit from 0 to t, the solution is

    x(t)=ϕ(0)exp(t0[r(1x(˜s)K)α(1eβx(˜s))y(˜s)x(˜s)]d˜s).

    Obviously, the exponential function is always positive, regardless of the integrand. It implies that x(t) is positive for t>0 and ϕ(0)>0.

    Next, we show that y(t) is positive on t[0,+). Based on the theory of Hale [26], it is obvious that y(t) is well-defined on [τ,+) and

    y(t)=φ(0)est+t0Yαφ(0)y(˜sτ)es(t˜s+τ)(1eβx(˜sτ))d˜s.

    Since φ(0)>0 and initial data (1.7), we have y(t)>0 when t[0,τ], therefore y(t)>0 for t[0,+]. Positivity implies that the cone of the solutions is invariant in the system.

    Theorem 2.2. Solutions of system (1.6) with positive initial data (1.7) are uniformly ultimately bounded.

    Proof. Define the following function:

    ω(t)=Yesτx(t)+y(t+τ).

    The derivative of ω(t) with respect to t is

    ˙ω(t)=Yesτ˙x(t)+˙y(t+τ).

    Substituting ˙x(t) and ˙y(t+τ) into the above expression, we obtain

    ˙ω(t)=Yesτ[rx(t)(1x(t)K)αy(t)(1eβx(t))]sy(t+τ)+Yαy(t)esτ(1eβx(t))=Yesτrx(t)(1x(t)K)sy(t+τ)Yesτrx(t)sy(t+τ)=2Yesτrx(t)sy(t+τ)Yesτrx(t)2Yesτr(x0+ε)sy(t+τ)Yesτrx(t)2Yesτr(x0+ε)min{s,r}[y(t+τ)+Yesτx(t)]=2Yesτr(x0+ε)min{s,r}ω(t),

    where x0 is the upper bound of x(t). By the Lemma 2.1 of [27,28], we obtain

    ω(t)2Yesτr(x0+ε)min{s,r},

    when t is sufficiently big. It implies that x(t) and y(t) are ultimately bounded. The boundedness of the model ensures that there is a restriction on the growth of the species due to limited resources in nature. This completes the proof.

    To begin, we consider the possible equilibria of model (1.6).

    Proposition 3.1. (i) System (1.6) has two distinct equilibria, the trivial equilibrium E0=(0,0) and the semi-trivial equilibrium ˉE=(K,0).

    (ii) If

    (H1)τ<τc=1slnαY(1eβK)s, (3.1)

    holds, the τc is the critical value, then system (1.6) has a coexistence equilibrium E=(x,y), where

    x=1βln(1sesταY),y=rYxsesτ(1xK).

    The existence of E ensures that 1eβK>0. Note that the equilibrium value depends on τ: x is an increasing function with respect to the delay τ, while y is a decreasing function when x>K2, that is: The larger the delay, the higher the number of the prey population, and the lower the number of predators at the equilibrium.

    The characteristic equation corresponding to E0=(0,0) is

    (λr)(λ+sαYe(λ+s)τ)=0,

    whose roots are obtained as λ1=r>0.

    Lemma 3.1. For all τ0, the trivial equilibrium E0 is always unstable.

    The characteristic equation with respect to ˉE=(K,0) is

    (λ+r)[λ+sαYe(λ+s)τ(1eβK)]=0,

    implying that λ1=r<0 and

    λ+sαYe(λ+s)τ(1eβK)=0.

    Let

    f(λ)=λ+sαYe(λ+s)τ(1eβK).

    Therefore,

    f(λ)=1+ταYe(λ+s)τ(1eβK)>0,f(0)=sαYesτ(1eβK),limλf(λ)=,

    for any τ0. Thus, if τ<τc, f(0)<0, and f(λ)=0 has at least one positive root. Thus, when τ<τc, the semi-trivial equilibrium is unstable.

    Lemma 3.2. When τ<τc, ˉE is unstable.

    In this part, assume that (3.1) is satisfied, then the interior (coexistence) equilibrium E exists. The linearized system of (1.6) about E is

    ˙X(t)=A0X(t)+A1X(tτ), (3.2)

    where X(t)=(x(t),y(t))T, X(tτ)=(x(tτ),y(tτ))T, A0=[r(12xK)+αβyeβxα(1eβx)0s], A1=[00αβYyesτβxαYesτ(1eβx)].

    The linearization system (3.2) around E has the following characteristic equation:

    det[λIA0A1eλτ]=0,

    that is,

    λ2+p1(τ)λ+p2(τ)+e(λ+s)τ[p3(τ)λ+p4(τ)]=0, (3.3)

    where

    p1(τ)=sr(12xK)αβyeβx,p2(τ)=s[r(12xK)+αβyeβx],p3(τ)=αY(eβx1),p4(τ)=αY(1eβx)[r(12xK)+αβy(1+Y)eβx]. (3.4)

    Notice that if τ=0, Eq (3.3) reduces to the second-order polynomial equation

    λ2+(p1(0)+p3(0))λ+p2(0)+p4(0)=0, (3.5)

    and it follows that all eigenvalues of Eq (3.5) have negative real parts if, and only if,

    p1(0)+p3(0)>0,p2(0)+p4(0)>0. (3.6)

    The transcendental Eq (3.3) has infinitely many roots. Note that polynomials pi(τ)(i=1,2,3,4) are dependent on τ. The transcendental equation associated with (3.2) at E is

    D(λ):=P(λ,τ)+Q(λ,τ)eλτ=0, (3.7)

    where

    P(λ,τ)=λ2+p1(τ)λ+p2(τ),Q(λ,τ)=esτ[p3(τ)λ+p4(τ)].

    For the characteristic equation, before applying the criterion due to Beretta and Kuang [29] to evaluate the existence of a purely imaginary root, we first verify the following properties for τ[0,τc), where τc is the maximum value when E exists.

    (i)P(0,τ)+Q(0,τ)0;

    (ii)P(iω,τ)+Q(iω,τ)0,ωR;

    (iii)lim|λ|sup{|Q(λ,τ)P(λ,τ)|;Reλ0}<1;

    (iv)F(ω,τ):=|P(iω,τ)|2|Q(iω,τ)|2 has a finite number of zeros;

    (v) Each positive root ω(τ) of F(ω,τ)=0 is continuous and differentiable in τ whenever it exists [18].

    Obviously,

    P(0,τ)+Q(0,τ)=p2(τ)+esτp4(τ)0,τ[0,τc),

    (ⅰ) is satisfied. Assumption ensures that λ=0 is not the root of Eq (3.7).

    Assume that p2(τ)+esτp4(τ)ω2,p1(τ)+esτp3(τ)0, then

    P(iω,τ)+Q(iω,τ)0,ωR.

    It follows from (3.7) that

    lim|λ||Q(λ,τ)P(λ,τ)|=lim|λ||esτ[p3(τ)λ+p4(τ)]λ2+p1(τ)λ+p2(τ)|=0,

    hence (ⅲ) follows.

    For the function F defined in (ⅳ), which follows

    |P(iω,τ)|2=ω4+[p21(τ)2p2(τ)]ω2+p22(τ),

    and

    |Q(iω,τ)|2=e2sτ(p3(τ)2ω2+p4(τ)2),

    such that

    F(ω,τ)=ω4+a1(τ)ω2+a2(τ),

    where

    a1(τ)=p21(τ)2p2(τ)e2sτp23(τ),a2(τ)=p22(τ)e2sτp24(τ).

    F(ω,τ) has at most four roots, therefore assumption (iv) is satisfied. Assume that (ω0,τ0) is a point in its domain such that F(ω0,τ0)=0. In a certain neighborhood of (ω0,τ0), the partial derivatives Fω and Fτ exist and are continuous, and Fω(ω0,τ0)0. Assumption (ⅴ) is satisfied by the implicit function theorem. Assumption (ⅳ) guarantees that Eq (3.7) has at most a finite number of purely imaginary roots [18], i.e., the roots cross the imaginary axis a finite number of times as τ varies.

    Next, we assume that λ=iω(ω>0,i=1) is the pure imaginary root of expression (3.7), then λ=iω satisfies

    |P(iω,τ)|2=|Q(iω,τ)|2,

    i.e., because |eiωτ|=1, ω(τ) is the positive zero root of

    F(ω,τ):=|P(iω,τ)|2|Q(iω,τ)|2=0.

    Define the set

    I={τ|τ0,F(ω,τ)=0has positive zero points}.

    Therefore,

    F(ω,τ)=0, (3.8)

    has positive root ω=ω(τ) if τI. Otherwise, F(ω,τ)=0 does't have a positive zero point. Furthermore, we obtain

    sin(ωτ)=Im(P(iω,τ)Q(iω,τ))=ωesτ[p3(τ)(ω2p2(τ))p1(τ)p4(τ)]ω2p32(τ)+p24(τ),cos(ωτ)=Re(P(iω,τ)Q(iω,τ))=esτ[ω2p1(τ)p3(τ)+p4(τ)(p2(τ)ω2)]ω2p32(τ)+p24(τ). (3.9)

    In addition, define the function θ(τ)[0,2π], which satisfied (3.9) for τI, i.e.,

    sin(θ(τ))=ωesτ[p3(τ)(ω2p2(τ))p1(τ)p4(τ)]ω2p32(τ)+p24(τ),cos(θ(τ))=esτ[ω2p1(τ)p3(τ)+p4(τ)(p2(τ)ω2)]ω2p32(τ)+p24(τ). (3.10)

    The ω(τ)τ in (3.9) and θ(τ) in (3.10) have the following relationship:

    ω(τ)τ=θ(τ)+2nτ,nN0.

    Introduce map τn:IR+0:

    τn(τ)=θ(τ)+2nτω(τ),nN0,τI,

    where ω(τ) is the positive root of (3.8). From I to R, define the continuous and differential function Sn(τ) as

    Sn(τ):=ττn(τ),nN0,τI. (3.11)

    Let λ(τ) be the eigenvalues satisfied by λ(τ)=iω(τ), and the transversality condition is obtained as

    δ(τ):=sign{dReλ(τ)dτ|λ(τ)=iω(τ),τ=τ}=sign{Fω(ω(τ),τ)}×sign{dSn(τ)dτ|λ(τ)=iω(τ),τ=τ}.

    Theorem 3.1. (i) For model (1.6), if either the set I is empty or the function Sn(τ) has no positive zero in I, for 0<τ<τc, the positive equilibrium E is asymptotically stable.

    (ii) If Eq (3.11) has positive roots in I denoted by {τ1,τ2,,τm} with τj<τj+1 and Sn(τ1)>0, the positive equilibrium E is asymptotically stable for τ[0,τ1)(τm,τc) and unstable for τ(τ1,τm), with Hopf bifurcations occurring when τ=τj,j=1,2,,m.

    To illustrate the analytical local Hopf Bifurcation results, we shall present some numerical simulations in this section and will extend them further with the help of numerical bifurcation analysis. The graphs are mainly drawn using DDEBifTool [30,31]. Similar dynamics have been numerically detected in the discrete delay system with Holling type Ⅱ and Beddington-DeAngelis trophic response [19].

    Hereafter, parameters are fixed at the following values,

    r=1,K=1,α=5,s=0.02,Y=0.6 and β=0.1. (4.1)

    According to the biological meaning of the parameters, because r is relatively large, it indicates that the prey has a high breeding rate. The s is relatively small which indicates that the predator has a low death rate. In addition, r is smaller than α to some extent, and it indicates that the changes in the number of prey are less influenced by their own birth rate and more influenced by their ability to evade natural enemies to prevent predation. Yα is larger than s, and it indicates that the changes of the number of predators are more influenced by their ability to prey on food.

    Under (4.1), we consider

    dx(t)dt=x(t)(1x(t))5y(t)(1e0.1x(t)),dy(t)dt=0.02y(t)+3y(tτ)e0.02τ(1e0.1x(tτ)),(ϕ(0),φ(0))=(0.1,0.1). (4.2)

    The ϕ(0) is initial prey population, and it represents the number of prey at the start of the model (4.2). This value can affect the food supply available to predators. If the initial number of prey is low, predators may face a food shortage. In contrary, if the initial number of prey is high, predators may have an abundant food supply. Furthermore, the φ(0) is initial predator population, and it represents the number of predators at the start. This value can influence the initial pressure that predators exert on prey. If the initial number of predators is low, the pressure on prey may be relatively small; conversely, if the initial number of predators is high, the pressure on prey may be greater. The ϕ(0) and φ(0) are chosen as 0.1 here, and they are at the median level relatively. The growth and mortality parameters represent the biological characteristics of predators and prey, such as growth rates and mortality rates. These parameters can influence the population dynamics of predators and prey, thereby affecting the stability of the ecosystem.

    System (4.2) has three equilibria: E0=(0,0), ˉE=(1,0), and the coexistence equilibrium E=(x,y) exists when τ<τc=250.

    Consider Figures 3 and 4: The blue solid line (the red dotted line) represents stable equilibrium (unstable equilibrium) and the filled green circle (open blue circles) indicates stable periodic orbit (unstable periodic orbits).

    Figure 3.  Stability of the three equilibrium of system (4.2) in τx and τy space.
    Figure 4.  The bifurcation diagram for system (4.2) in τx and τy space.

    Figure 3 indicates that the E0=(0,0) is unstable. When τ[0,τ3), ˉE=(1,0) is unstable. When τ>τ3, it is locally asymptotically stable, i.e., at τ=τ3 the predator goes extinct. The coexistence equilibrium E is stable when τ[0,τ1)(τ2,τ3) while it's unstable when τ(τ1,τ2)(τ3,τc). At τ=τ3, the semi-trivial equilibrium ˉE=(1,0) and positive equilibrium E exchange their stability, leading to the appearance of transcritical bifurcation. Note that τ1=4.4730, τ2=69.3061, and τ3=132.9233.

    We produced a corresponding bifurcation diagram as Figure 4, using τ as the primary bifurcation parameter. Figure 4 shows the first critical value τ1 with ω1=0.1215, which the bifurcated Hopf bifurcation and the second critical value is τ2 with ω2=0.0289, generating the supercritical Hopf bifurcation at τ1 and τ2. Biologically, above phenomenon could be interpreted as there being an interval (τ1, τ2) of survival that may exist even though the positive equilibrium is unstable.

    Figures 58 show the trajectories and phase graph of system (4.2) with τ=4.3, τ=4.5, τ=50, and τ=70.5, respectively. Figure 5 illustrates that the coexistence equilibrium E=(0.0669,1.8724) is locally asymptotically with τ=4.3<τ1. It will lose its stability and a bifurcating periodic solution occurs once τ=4.5>τ1 as the time delay increases, as shown in Figure 6. Figure 7 indicates that a stable periodic solution occurs with τ=50. Additionally, Figure 8 shows that the coexistence equilibrium E is locally asymptotically stable with τ=70.5>τ2. The results are coincident with Figures 3 and 4.

    Figure 5.  (Left) When τ=4.3<τ1=4.4730, trajectory of system (4.2) on ty(t) space. (Right) When τ=4.3, phase diagram of the system (4.2) on x(t)y(t) space.
    Figure 6.  (Left) When τ=4.5>τ1=4.4730, trajectory of system (4.2) on ty space. (Right) When τ=4.5, phase diagram of the system (4.2) on xy space.
    Figure 7.  (Left) When τ=50, trajectory of system (4.2) on ty space. (Right) When τ=50, phase diagram of the system (4.2) on xy space.
    Figure 8.  (Left) When τ=70.5, trajectory of system (4.2) on ty space. (Right) When τ=70.5, phase diagram of the system (4.2) on xy space.

    The above simulations indicate that there exists a unique global Hopf bifurcation connecting τ1 and τ2. The global Hopf bifurcation is bounded, and each global Hopf branch connects a pair of Hopf bifurcation values. In the next subsection, we will detect the global Hopf bifurcation [32].

    By the global Hopf bifurcation result of [33,34], it shows that for the following delayed Lotka-Volterra model,

    ˙x(t)=x(t)[r1a11x(tτ)a12y(t)],˙y(t)=y(t)[r2+a21x(t)a22y(t)].

    After the second critical value, the local Hopf bifurcation implies a global Hopf bifurcation. However, for the delayed model (1.6), it is not valid. We will show that as follows.

    Parameters are fixed at the following values,

    r=3.1,K=0.5,α=1.5,s=0.02,Y=0.4 and β=1.2. (4.3)

    According to the biological meaning of the parameters, because r is relatively large, it indicates that the prey has a high breeding rate. The s is relatively small, which indicates that the predator has a low death rate. Because r is larger than α, to some extent, and it indicates that the changes in the number of prey are more influenced by their own birth rate. In addition, Yα is larger than s, and it indicates that the changes in the number of predator are more influenced by their ability to prey on food [35].

    Under (4.3), we consider the following model:

    dx(t)dt=3.1x(t)(12x(t))1.5y(t)(1e1.2x(t)),dy(t)dt=0.02y(t)+0.6y(tτ)e0.02τ(1e1.2x(tτ)),(ϕ(0),φ(0))=(0.1,0.1). (4.4)

    Consider Figures 9 and 10, where the open blue circles stand for unstable periodic orbits. Figure 9 indicates that the E0=(0,0) is always unstable. The ˉE=(1,0) is unstable when τ[0,τ3). When τ>τ3, it is locally asymptotically stable where τ3=130.2649. In τ[0,τ1)(τ2,τ3), the positive equilibrium E is stable, and unstable when τ(τ1,τ2)(τ3,τc) where τ1=2.6330 and τ2=79.8515. By Figure 10, the unstable periodic orbits appear between τ4 and τ5, where τ4=67.2228 and τ5=70.4188. Biologically, since the appearance of unstable bifurcating periodic solutions, the two species could coexist in a chaotic mode.

    Figure 9.  Stability of the three equilibrium of system (4.4) in τx and τy space.
    Figure 10.  (Left) The bifurcation diagram for system (4.4) on τx(t) space. (Right) The larger version of Figure 10 when τ(57,81).

    To see the influence of delay τ on the dynamical behaviors of the model, we detect the complex dynamical behavior when τ(τ4,τ5) by Figures 1114. By Figure 11, for the system (4.4), when τ=65.9, the system has a limit cycle whose period is approximately 250. The periodic orbits are always stable until τ<66.6. When τ>66.6, stable periodic solutions undergo period-doubling bifurcation; as Figure 12 shows, when τ=66.6, the system bifurcates twice the period. When τ=69.5, the system bifurcates with a sequence of period-doubling bifurcations. When τ continues to increase from 65.9 to 70, by Figure 14, the system (4.4) achieves chaotic oscillation through period-doubling bifurcation with a chaotic attractor. In a stable equilibrium, the periodic oscillation by 2, 22, 23 cycles eventually lead to chaos. Eventually, a cascade of period doubling bifurcations leads to chaos, which resembles the chaotic attractor of the following Mackey-Glass equation [36]

    dxdt=βx(tτ)1+(x(tτ))nγx(t).

    The existence of chaotic solutions implies that even a small environmental or parameter perturbation can disrupt the dynamics of the system [37,38]. In current research on the existence of chaos, only the phase diagram or the time course diagram of the system is generally provided, with few discussions on the chaotic path. In fact, the chaotic path can clearly illustrate the dynamic transition process of the system under the influence of parameters. Therefore, compared with existing results on the dynamical behaviors of predator-prey systems [21,39,40], the detection of chaos and the analysis of the chaotic path are the main contributions of this paper in terms of dynamical behaviors.

    Figure 11.  (Left) When τ=65.9, trajectory of system (4.4) on ty space. (Right) When τ=65.9, phase diagram of the system (4.4) on y(tτ)y(t) space.
    Figure 12.  (Left) When τ=66.6, trajectory of system (4.4) on ty space. (Right) When τ=66.6, phase diagram of the system (4.4) on y(tτ)y(t) space.
    Figure 13.  (Left) When τ=69.5, trajectory of system (4.4) on ty space. (Right) When τ=69.5, phase diagram of the system (4.4) on y(tτ)y(t) space.
    Figure 14.  (Top left) When τ=70, trajectory of system (4.4). (Top right) When τ=70, phase diagram embedded with time-delay terms for the system (4.4). (Bottom) When τ=70, projection of the attractor (4.4) into xy sapace.

    Remark 4.1. Since chaos is sensitive to initial conditions, the strange attractor is not obvious under the parameter values as specified in (4.3), which results in the critical values τ4τ6 in Figure 10 being not accurate enough. Furthermore, the interval from τ4 to τ6 is not long enough, and we did not detect the dynamical behaviors in detail within this interval. We hypothesize that within this interval, as the delay τ increases from 70 to τ6, the system (4.4) undergoes four cycle bifurcations, doubling the period twice, leading to a period doubling bifurcation and a limit cycle. The dynamic behavior of the system does not change substantially.

    In this paper, a delayed prey-predator model with an Ivlev-type functional response is investigated, focusing on the effect of the delay on the dynamical behaviors of the model. The supercritical Hopf bifurcation and period doubling types of bifurcations, as well as a strange attractor, can occur at the positive equilibrium when time delay is considered as a bifurcation parameter. The chaotic attractor appears, followed by a sequence of period-doubling bifurcations for small enough of the death rate of the predator species. This study delves into the intricate interplay where time delays and nonlinear responses converge, offering a deeper insight into the chaotic behaviors that may arise within these complex systems.

    From a biological perspective, there are intriguing explanations. If delay is minimal, predator and prey populations stabilize. However, as delay escalates, species exhibit asymptotic, periodic, or quasi-periodic fluctuations, suggesting an oscillatory coexistence of predator and prey. As the time delay continues to increase, the system will exhibit a chaotic phenomenon of 'lose a millimeter, miss a thousand miles'. Consequently, short-term observations can be deceptive in forecasting due to the presence of bifurcation and chaos, highlighting the complexity of long-term ecological dynamics. The results could be very essential for biologists who work with delayed prey-predator systems. In conclusion, this paper makes two contributions: the introduction of an exponential delay term and the detection of chaos.

    In reality, the processing time delay rarely has the same length at every instance; instead, it follows a distribution with some mean value. Our follow-up work will investigate the dynamical behaviors of the model incorporating distributed delay and compare the dynamics resulting from using various distributions, including discrete delay.

    Qinghui Liu: The data analysis and wrote the paper; Xin Zhang: The formal analysis and the validation. All authors have read and approved the final version of the manuscript for publication.

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

    This work was supported by the Basic Science (Natural Science) Research Project of Jiangsu Province (No. 24KJB110010), the Special Reform and Development Project of Nanjing University of Finance and Economics in 2023 (No. XGFB3202311), Teaching Reform Project of Nanjing University of Finance and Economics in 2023 (No. JG23902).

    All authors declare no conflicts of interest in this paper.



    [1] H. P. Zhu, S. A. Campbell, G. S. K. Wolkowicz, Bifurcation analysis of a predator-prey system with nonmonotonic functional response, SIAM J. Appl. Math., 63 (2002), 636–682. https://doi.org/10.1137/S0036139901397285 doi: 10.1137/S0036139901397285
    [2] J. F. Wang, J. P. Shi, J. J Wei, Predator-prey system with strong allee effect in prey, J. Math. Biol., 62 (2011), 291–331. https://doi.org/10.1007/s00285-010-0332-1 doi: 10.1007/s00285-010-0332-1
    [3] W. S. Yang, Global asymptotical stability and persistent property for a diffusive predator-prey system with modified Leslie-Gower functional response, Nonlinear Anal. Real World Appl., 14 (2013), 1323–1330. https://doi.org/10.1016/j.nonrwa.2012.09.020 doi: 10.1016/j.nonrwa.2012.09.020
    [4] S. M. Fu, H. S. Zhang, Effect of hunting cooperation on the dynamic behavior for a diffusive Holling type Ⅱ predator-prey model, Commun. Nonlinear Sci. Numer. Simul., 99 (2021), 105807. https://doi.org/10.1016/j.cnsns.2021.105807 doi: 10.1016/j.cnsns.2021.105807
    [5] S. A. Kashchenko, A. O. Tolbey, Dynamics of a system of two equations with a large delay, Dokl. Math., 108 (2023), 369–373. https://doi.org/10.1134/S1064562423701259 doi: 10.1134/S1064562423701259
    [6] A. Teslya, G. S. K. Wolkowicz, Dynamics of a predator-prey model with distributed delay to represent the conversion process or maturation, Differ. Equ. Dyn. Syst., 31 (2023), 613–649. https://doi.org/10.1007/s12591-020-00546-4 doi: 10.1007/s12591-020-00546-4
    [7] S. Pandey, U. Ghosh, D. Das, S. Chakraborty, A. Sarkar, Rich dynamics of a delay-induced stage-structure prey-predator model with cooperative behaviour in both species and the impact of prey refuge, Math. Comput. Simulation, 216 (2024), 49–76. https://doi.org/10.1016/j.matcom.2023.09.002 doi: 10.1016/j.matcom.2023.09.002
    [8] D. P. Hu, Y. Y. Li, M. Liu, Y. Z. Bai, Stability and Hopf bifurcation for a delayed predator-prey model with stage structure for prey and Ivlev-type functional response, Nonlinear Dyn., 99 (2020), 3323–3350. https://doi.org/10.1007/s11071-020-05467-z doi: 10.1007/s11071-020-05467-z
    [9] Y. Liu, Z. L. Shen, J. J. Wei, Pattern dynamics of a predator-prey system with Ivlev-type functional response, Discrete Cont. Dyn.-B, 29 (2024), 3802–3823. https://doi.org/10.3934/dcdsb.2024024 doi: 10.3934/dcdsb.2024024
    [10] M. R. Garvie, Finite-difference schemes for reaction-diffusion equations modeling predator-prey interactions in Matlab, B. Math. Biol., 69 (2007), 931–956. https://doi.org/10.1007/s11538-006-9062-3 doi: 10.1007/s11538-006-9062-3
    [11] S. Li, C. D. Huang, X. Y. Song, Detection of Hopf bifurcations induced by pregnancy and maturation delays in a spatial predator-prey model via crossing curves method, Chaos Solitons Fract., 175 (2023), 114012. https://doi.org/10.1016/j.chaos.2023.114012 doi: 10.1016/j.chaos.2023.114012
    [12] W. Wang, J. H. Sun, On the predator-prey system with Holling-(n+1) functional response, Acta Math. Sin. Engl. Ser., 23 (2007), 1–6. https://doi.org/10.1007/s10114-005-0603-8 doi: 10.1007/s10114-005-0603-8
    [13] Z. Z. Zhang, W. S. Zhang, P. Anbalagan, M. M. Arjunan, Global dissipativity and adaptive synchronization for fractional-order time-delayed genetic regulatory networks, Asian J. Control, 24 (2022), 3289–3298. https://doi.org/10.1002/asjc.2726 doi: 10.1002/asjc.2726
    [14] X. P. Yan, C. H. Zhang, Bifurcation analysis in a diffusive Logistic population model with two delayed density-dependent feedback terms, Nonlinear Anal. Real World Appl., 63 (2022), 103394. https://doi.org/10.1016/j.nonrwa.2021.103394 doi: 10.1016/j.nonrwa.2021.103394
    [15] T. W. Zhang, H. Z. Qu, J. W. Zhou, Asymptotically almost periodic synchronization in fuzzy competitive neural networks with Caputo-Fabrizio operator, Fuzzy Set. Syst., 471 (2023), 108676. https://doi.org/10.1016/j.fss.2023.108676 doi: 10.1016/j.fss.2023.108676
    [16] T. W. Zhang, Y. K. Li, Global exponential stability of discrete-time almost automorphic caputo-fabrizio bam fuzzy neural networks via exponential Euler technique, Knowl.-Based Syst., 246 (2022), 108675. https://doi.org/10.1016/j.knosys.2022.108675 doi: 10.1016/j.knosys.2022.108675
    [17] J. Arino, L. Wang, G. S. K. Wolkowicz, An alternative formulation for a delayed logistic equation, J. Theor. Biol., 241 (2006), 109–119. https://doi.org/10.1016/j.jtbi.2005.11.007 doi: 10.1016/j.jtbi.2005.11.007
    [18] X. Zhang, Hopf bifurcation in a prey-predator model with constant delay, Int. J. Nonlin. Mech., 117 (2019), 103235. https://doi.org/10.1016/j.ijnonlinmec.2019.103235 doi: 10.1016/j.ijnonlinmec.2019.103235
    [19] X. Zhang, R. X. Shi, R. Z. Yang, Z. Z. Wei, Dynamical behaviors of a delayed prey-predator model with Beddington-Deangelis functional response: Stability and periodicity, Int. J. Bifurcat. Chaos, 30 (2020), 2050244. https://doi.org/10.1142/S0218127420502442 doi: 10.1142/S0218127420502442
    [20] X. H. Wang, H. H. Liu, C. L. Xu, Hopf bifurcations in a predator-prey system of population allelopathy with a discrete delay and a distributed delay, Nonlinear Dyn., 69 (2012), 2155–2167. https://doi.org/10.1007/s11071-012-0416-0 doi: 10.1007/s11071-012-0416-0
    [21] A. Singh, P. Deolia, Dynamical analysis and chaos control in discrete-time prey-predator model, Commun. Nonlinear Sci. Numer. Simul., 90 (2020), 195313. https://doi.org/10.1016/j.cnsns.2020.105313 doi: 10.1016/j.cnsns.2020.105313
    [22] Z. X. Li, A delayed ratio-dependent predator-prey system with stage-structured and impulsive effect, J. Syst. Sci. Complex., 24 (2011), 1118–1129. https://doi.org/10.1007/s11424-011-8198-x doi: 10.1007/s11424-011-8198-x
    [23] B. Y. Xie, F. Xu, Stability analysis for a time-delayed nonlinear predator-prey model, Adv. Differ., 122 (2018), 2018. https://doi.org/10.1186/s13662-018-1564-4 doi: 10.1186/s13662-018-1564-4
    [24] M. R. Xu, S. Liu, Y. Lou, Persistence and extinction in the anti-symmetric Lotka-Volterra systems, J. Differ., 387 (2024), 299–323. https://doi.org/10.1016/j.jde.2023.12.032 doi: 10.1016/j.jde.2023.12.032
    [25] G. Zhu, J. J. Wei, Global stability and bifurcation analysis of a delayed predator-prey system with prey immigration, Electron J. Qual. Theo., 2016 (2016), 1–20. https://doi.org/10.14232/ejqtde.2016.1.13 doi: 10.14232/ejqtde.2016.1.13
    [26] J. K. Hale, Theory of functional differential equations, 2 Eds., New York: Springer Press, 1977.
    [27] F. D. Chen, On a nonlinear nonautonomous predator-prey model with diffusion and distributed delay, J. Comput. Appl. Math., 180 (2005), 33–49. https://doi.org/10.1016/j.cam.2004.10.001 doi: 10.1016/j.cam.2004.10.001
    [28] K. Fang, Z. L. Zhu, F. D. Chen, Z. Li, Qualitative and bifurcation analysis in a Leslie-Gower model with allee effect, Qual. Theory Dyn. Syst., 21 (2022), 33–49. https://doi.org/10.1007/s12346-022-00591-0 doi: 10.1007/s12346-022-00591-0
    [29] E. Beretta, Y. Kuang, Geometric stability switch criteria in delay differential systems with delay dependent parameters, SIAM J. Math. Anal., 33 (2002), 1144–1165. https://doi.org/10.1137/S0036141000376086 doi: 10.1137/S0036141000376086
    [30] L. F. Shampine, S. Thompson, Solving DDEs in MATLAB, Appl. Numer. Math., 37 (2001), 441–458. https://doi.org/10.1016/S0168-9274(00)00055-6 doi: 10.1016/S0168-9274(00)00055-6
    [31] K. Engelborghs, T. Luzyanina, D. Roose, Numerical bifurcation analysis of delay differential equations using DDE-BIFTOOL, ACM Trans. Math. Softw., 28 (2002), 1–21. https://doi.org/10.1145/513001.513002 doi: 10.1145/513001.513002
    [32] H. Y. Shu, L. Wang, J. H. Wu, Bounded global Hopf branches for stage-structured differential equations with unimodal feedback, Nonlinearity, 30 (2017), 943–964. https://doi.org/10.1088/1361-6544/aa5497 doi: 10.1088/1361-6544/aa5497
    [33] J. H. Wu, Symmetric functional differential equations and neural networks with memory, Trans. Am. Math. Soc., 350 (1998), 4799–4838. https://doi.org/10.1090/S0002-9947-98-02083-2 doi: 10.1090/S0002-9947-98-02083-2
    [34] Y. L. Song, J. J. Wei, Local Hopf bifurcation and global periodic solutions in a delayed predator-prey system, J. Math. Anal. Appl., 301 (2005), 1–21. https://doi.org/10.1016/j.jmaa.2004.06.056 doi: 10.1016/j.jmaa.2004.06.056
    [35] H. K. Qi, B. Liu, S. Li, Stability, bifurcation, and chaos of a stage-structured predator-prey model under fear-induced and delay, Appl. Math. Comput., 476 (2024), 128780. https://doi.org/10.1016/j.amc.2024.128780 doi: 10.1016/j.amc.2024.128780
    [36] L. Glass, M. Mackey, Mackey-Glass equation, Scholarpedia, 5 (2009), 6908. https://doi.org/10.4249/scholarpedia.6908 doi: 10.4249/scholarpedia.6908
    [37] W. L. Duan, L. Lin, Noise and delay enhanced stability in tumor-immune responses to chemotherapy system, Chaos Solitons Fract., 148 (2021), 111019. https://doi.org/10.1016/j.chaos.2021.111019 doi: 10.1016/j.chaos.2021.111019
    [38] W. L. Duan, C. H. Zeng, Statistics for anti-synchronization of intracellular calcium dynamics, Appl. Math. Comput., 293 (2017), 611–616. https://doi.org/10.1016/j.amc.2016.07.041 doi: 10.1016/j.amc.2016.07.041
    [39] Q. Din, Complexity and chaos control in a discrete-time prey-predator model, Commun. Nonlinear Sci. Numer. Simul., 49 (2017), 113–134. https://doi.org/10.1016/j.cnsns.2017.01.025 doi: 10.1016/j.cnsns.2017.01.025
    [40] Q. Gao, J. H. Ma, Chaos and Hopf bifurcation of a finance system, Nonlinear Dyn., 58 (2009), 209–216. https://doi.org/10.1007/s11071-009-9472-5 doi: 10.1007/s11071-009-9472-5
  • 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(754) PDF downloads(48) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog