Loading [MathJax]/jax/output/SVG/jax.js
Research article

Dynamics of a nonlinear discrete predator-prey system with fear effect

  • Received: 18 May 2023 Revised: 16 July 2023 Accepted: 24 July 2023 Published: 07 August 2023
  • MSC : 37G15, 37N25, 39A28, 92D25, 93C10

  • In this paper, we investigate a nonlinear discrete prey-predator system with fear effects. The existence, local stability and boundedness of positive equilibrium point are discussed. Using the center manifold theorem and bifurcation theory, the conditions for the existence of flip bifurcation and Neimark-Sacker bifurcation in the interior of R2+ are established. Furthermore, the numerical simulations not only show complex dynamical behaviors, but also verify our analysis results. A feedback control strategy is employed to control bifurcation and chaos in the system.

    Citation: Xiongxiong Du, Xiaoling Han, Ceyu Lei. Dynamics of a nonlinear discrete predator-prey system with fear effect[J]. AIMS Mathematics, 2023, 8(10): 23953-23973. doi: 10.3934/math.20231221

    Related Papers:

    [1] Xiaoming Su, Jiahui Wang, Adiya Bao . Stability analysis and chaos control in a discrete predator-prey system with Allee effect, fear effect, and refuge. AIMS Mathematics, 2024, 9(5): 13462-13491. doi: 10.3934/math.2024656
    [2] Jie Liu, Qinglong Wang, Xuyang Cao, Ting Yu . Bifurcation and optimal harvesting analysis of a discrete-time predator–prey model with fear and prey refuge effects. AIMS Mathematics, 2024, 9(10): 26283-26306. doi: 10.3934/math.20241281
    [3] Figen Kangalgil, Seval Isșık . Effect of immigration in a predator-prey system: Stability, bifurcation and chaos. AIMS Mathematics, 2022, 7(8): 14354-14375. doi: 10.3934/math.2022791
    [4] 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
    [5] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
    [6] Ali Yousef, Ashraf Adnan Thirthar, Abdesslem Larmani Alaoui, Prabir Panja, Thabet Abdeljawad . The hunting cooperation of a predator under two prey's competition and fear-effect in the prey-predator fractional-order model. AIMS Mathematics, 2022, 7(4): 5463-5479. doi: 10.3934/math.2022303
    [7] 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
    [8] A. Q. Khan, Ibraheem M. Alsulami, S. K. A. Hamdani . Controlling the chaos and bifurcations of a discrete prey-predator model. AIMS Mathematics, 2024, 9(1): 1783-1818. doi: 10.3934/math.2024087
    [9] Xuyang Cao, Qinglong Wang, Jie Liu . Hopf bifurcation in a predator-prey model under fuzzy parameters involving prey refuge and fear effects. AIMS Mathematics, 2024, 9(9): 23945-23970. doi: 10.3934/math.20241164
    [10] Ali Al Khabyah, Rizwan Ahmed, Muhammad Saeed Akram, Shehraz Akhtar . Stability, bifurcation, and chaos control in a discrete predator-prey model with strong Allee effect. AIMS Mathematics, 2023, 8(4): 8060-8081. doi: 10.3934/math.2023408
  • In this paper, we investigate a nonlinear discrete prey-predator system with fear effects. The existence, local stability and boundedness of positive equilibrium point are discussed. Using the center manifold theorem and bifurcation theory, the conditions for the existence of flip bifurcation and Neimark-Sacker bifurcation in the interior of R2+ are established. Furthermore, the numerical simulations not only show complex dynamical behaviors, but also verify our analysis results. A feedback control strategy is employed to control bifurcation and chaos in the system.



    Populations in nature are rarely isolated, and intensively interact with others in the biological community. All kinds of organisms are divided into different levels according to their physiological characteristics, food sources, etc. and different levels of populations have a variety of connections. The predator-prey process is the most fundamental, important and universal process in the study of population dynamic behavior. Many researchers have studied the dynamic behavior of many prey-predator systems in ecology and behavioral phenomena between species [1,2,3,4,5,6,7,8,9,10,11]. Some authors have also explored the complexity, stability, and conditional requirements for spatial pattern formation in prey-predator systems [12,13,14].

    Numerous studies have shown that discrete-time systems are more suitable than the continuous system of small populations, and provide valid evidence for these [15,16,17,18]. Cheng et al. [19] studied a discrete-time ratio-dependent prey-predator system with Allee effect, and obtained the model with logistic growth function that have somewhat similar bifurcation structures. In the past few years, a large number of research studies have indicated that discrete prey-predator systems have more abundant dynamic behaviors than continuous systems, such as chaos. Scientists have also analyzed the corresponding dynamic behaviors between populations by numerical simulation [20,21,22,23,24,25,26,27,28,29,30].

    Qamar Din [31] studied the following discrete-time system:

    {un+1=unexp[r(1unK)βvnun+γ],vn+1=vnexp(1avnbun+cd), (1.1)

    where r,K,β,γ,a,b,c,n and d are greater than zero, r is the intrinsic growth rate of the prey u population, K denotes environmental carrying capacity of the prey u in a particular habitat, β and a represent the maximum value of the per capita reduction rate of the prey u and predator v, respectively. γ and c indicate the extent to which the environment provides protection to prey and predator. b denotes the quality of food that the prey provides for conversion into predator births, and d measures the death rate of the predator. n stands for time.

    In 2016, Wang et al. [21] showed, through experiments, that preys fear of predators would lead to a decrease in the birth rate of prey, and F(k,v)=11+kv was used to denote the fear factor. Here, k reflects the degree of fear that drives prey anti-predator behavior. In the past, many researchers have only studied the effects of direct killing, no matter how they improve the predator-prey model. In this paper, we combine fear (indirect effects) and investigate the effects of fear on population dynamics.

    To study the effects of fear on population dynamics, on the basis of system (1.1), we introduce the fear factor F(k,v)=11+kv and the growth rate α of the predator v population, and consider the discrete-time predator-prey system:

    {un+1=unexp[r1+kvn(1unK)βvnun+γ],vn+1=vnexp(αavnbun+cd). (1.2)

    This article is organized as follows: in Section 2, the existence, stability and boundedness of the system at different equilibrium points are analyzed. In Section 3, we discuss the specific conditions for the existence of Neimark-Sacker bifurcation and flip bifurcation. In Section 4, chaos is controlled by the feedback control method. In Section 5, we carry out numerical simulations, including the bifurcation diagrams, phase portraits and solution diagrams. Finally, a brief conclusion is given in the last section.

    In this section, we consider the discrete-time system (1.2) in the closed first quadrant R2+ of the (u,v) plane. We study the existence, stability and boundedness of the equilibrium points by the eigenvalues for the Jacobian matrix of (1.2) at the equilibrium points.

    To obtain the equilibrium points of (1.2), we calculate the following equations:

    {u=uexp[r1+kv(1uK)βvu+γ],v=vexp(αavbu+cd).

    Through calculation, the following results can be gained directly:

    Proposition 1. (i) For all parameter values, system (1.2) has two equilibrium points H0=(0,0),H1=(K,0);

    (ii) If α>d, then system (1.2) has a boundary positive equilibrium point H2=(0,(αd)ca);

    (iii) System (1.2) has a unique positive equilibrium point H3=(u,v)=(u,1a(αd)(bu+c)), where α>d, and u is the only positive solution to the quadratic equation of one variabie

    C0u2+C1u+C2=0,

    where

    C0=Kkβa2(αd)2b2+r,

    C1=2bcKkβa2(αd)2+bKβa(αd)+rγrK,

    C2=Kkβa2(αd)2c2+cKβa(αd)rKγ.

    Definition 1. [11] Suppose that λ1 and λ2 are two roots of the characteristic equation F(λ)=λ2+Mλ+N=0, where M and N are constants. Then equilibrium point (u, v) is called

    (i) sink if |λ1|<1 and |λ2|<1, and it is locally asymptotically stable;

    (ii) source or repeller if |λ1|>1 and |λ2|>1, and it is locally unstable;

    (iii) saddle if either (|λ1|<1 and |λ2|>1) or (|λ1|>1 and |λ2|<1);

    (iv) non-hyperbolic if either |λ1|=1 or |λ2|=1.

    The Jacobian matrix for equilibrium point H0(0,0) is:

    QH0=[er00eαd]. (2.1)

    Then, λ1=er,λ2=eαd. Thus, the following proposition holds.

    Proposition 2. Equilibrium point H0(0,0) is

    (i) source and it is locally unstable if α>d;

    (ii) saddle if α<d.

    Proof. According to (2.1), the two eigenvalues of (1.2) at the equilibrium point H0(0,0) are λ1=er,λ2=eαd. If αd>0 and r>0, then |λ1|>1, |λ2|>1. Thus from Definition 1, H0=(0,0) is a source. If α<d, then 0<|λ2|<1. H0=(0,0) is a saddle. This completes the proof.

    For equilibrium point H1=(K,0), the Jacobian matrix is described as follows:

    QH1=[1rβKK+γ0eαd], (2.2)

    the corresponding characteristic roots are λ1=1r,λ2=eαd. Thus, the following proposition holds.

    Proposition 3. The eigenvalues at the boundary equilibrium point H1=(K,0) are λ1=1r,λ2=eαd, then

    (i) H1=(K,0) is sink, if 0<r<2 and αd<0;

    (ii) H1=(K,0) is saddle, if one of the following conditions is true:

        (ii-1) αd>0 and 0<r<2;

        (ii-2) αd<0 and r>2;

    (iii) H1=(K,0) is non-hyperbolic, if either r=2 or αd=0;

    (iv) H1=(K,0) is source, if r>2 and αd>0.

    Proof. According to (2.2), the two eigenvalues of (1.2) at the boundary equilibrium point are λ1=1r,λ2=eαd. If αd are greater than zero, then |λ2|>1. Thus from Definition 1, when |λ1|<1, then 0<r<2. Thus, H1=(K,0) is a saddle. Similarly, when |λ1|>1, then r>2, H1=(K,0) is a source. Similarly, we can prove (i), (iii) by the same way.

    For equilibrium point H2=(0,(αd)ca), the Jacobian matrix is evaluated as follows:

    QH2=[exp[ara+kc(αd)β(αd)caγ]0bc(αd)2a1(αd)], (2.3)

    the corresponding characteristic roots are λ1=exp[ara+kc(αd)β(αd)caγ],λ2=1(αd). Thus, the following proposition holds.

    Proposition 4. The eigenvalues of QH2 are λ1=exp[ara+kc(αd)β(αd)caγ] and λ2=1(αd), then

    (i) H2=(0,(αd)ca) is sink if 0<αd<2 and 0<r<βc(αd)[a+kc(αd)]a2γ;

    (ii) H2=(0,(αd)ca) is source if αd>2 and r>βc(αd)[a+kc(αd)]a2γ;

    (iii) H2=(0,(αd)ca) is saddle if one of the following conditions is true:

        (iii-1) αd=2 and 0<r<βc(αd)[a+kc(αd)]a2γ;

        (iii-2) 0<αd<2 and r>βc(αd)[a+kc(αd)]a2γ;

    (iv) H2=(0,(αd)ca) is non-hyperbolic if either αd=2 or r=βc(αd)[a+kc(αd)]a2γ.

    Proof. (i) According to (2.3), the two eigenvalues of (1.2) at the boundary equilibrium point H2 are λ1=exp[ara+kc(αd)β(αd)caγ],λ2=1(αd). H2=(0,(αd)ca) is sink if and only if |λ1|<1 and |λ2|<1. When |λ1|<1, then 0<αd<2. When |λ2|<1, then 0<r<βc(αd)[a+kc(αd)]a2γ. In conclusion, H2=(0,(αd)ca) is sink if 0<αd<2 and 0<r<βc(αd)[a+kc(αd)]a2γ; Similarly, Proposition 4 (ii)–(iv) can be proved.

    Lemma 1. [11] Suppose that F(λ)=λ2Mλ+N, and F(1)>0, λ1 and λ2 are roots of F(λ)=0. Then the following results hold true:

    (i) |λ1|<1 and |λ2|<1 if and only if F(1)>0 and N<1;

    (ii) |λ1|<1 and |λ2|>1 (or |λ1|>1 and |λ2|<1) if and only if F(1)<0;

    (iii) |λ1|>1 and |λ2|>1 if and only if F(1)>0 and N>1;

    (iv) λ1=1 and |λ2|1 if and only if F(1)=0 and N0,2;

    (v) λ1 and λ2 are complex and |λ1|=|λ2|=1 if and only if M24N<0 and N=1.

    The Jacobian matrix Q(u,v) relevant system (1.2) at the positive equilibrium point H3(u,v) is as follows:

    QH3=[1ru(1+kv)K+βuv(u+γ)2(1uK)rku(1+kv)2βuu+γabu2(bu+c)21(αd)].

    Then the characteristic equation related to QH3 is

    F(λ)=λ2m(u,v)λ+n(u,v)=0,

    where

    m(u,v)=trQH3=2ru(1+kv)K+βuv(u+γ)2(αd)

    and

    n(u,v)=detQH3=[1ru(1+kv)K+βuv(u+γ)2][1(αd)]+[(1uK)rku(1+kv)2+βuu+γ]abu2(bu+c)2.

    Therefore

    F(1)=1m(u,v)+n(u,v),F(1)=1+m(u,v)+n(u,v).

    Using Lemma 1, we have the following proposition:

    Proposition 5. Let H3(u,v) be the unique positive equilibrium point of system (1.2), then the following propositions hold:

    (i) H3(u,v) is sink if and only if |m(u,v)|<1+n(u,v)<2;

    (ii) H3(u,v) is saddle if and only if m2(u,v)>4n(u,v) and |m(u,v)|>|1+n(u,v)|;

    (iii) H3(u,v) is source if and only if |n(u,v)|>1 and |m(u,v)|<|1+n(u,v)|;

    (iv) H3(u,v) is non-hyperbolic if and only if |m(u,v)|=|1+n(u,v)| or n(u,v)=1 and |m(u,v)|2.

    Proof. (i) According to Lemma 1, H3(u,v) is a sink point if and only if F(1)>0,F(1)>0 and N<1, it can be acquired by calculation |m(u,v)|<1+n(u,v)<2. Consequently, Proposition 5 (i) holds. Similarly, Proposition 5 (ii)–(iv) can be established.

    Lemma 2. [20] Assume that ut satisfies u0>0, and ut+1utexp[A(1But)] for t[t1,), where B is a positive constant. Then limtsup ut1ABexp(A1).

    Theorem 1. Every positive solution {(un,vn)} of system (1.2) is uniformly bounded.

    Proof. Suppose that {(un,vn)} be an arbitrary positive solution corresponding to system (1.2). Then, by the first part of (1.2), it is known

    un+1unexp[r1+kvn(1unK)]unexp[r(1unK)]

    for all n=0,1,2,. Suppose that u0>0, then according to Lemma 2, we gain

    limnsup unKrexp(r1):=M1.

    From the second part of (1.2), we acquire

    vn+1=vnexp(αavnbun+cd)vnexp(αavnbun+c)vnexp(αavnbM1+c).

    Assume that v0>0, then using Lemma 2, we gain

    limnsup vnbM1+caexp(α1):=M2.

    That is to say that limnsup (un,vn)M, where M=max{M1,M2}. This completes the proof.

    The characteristic equation related to system (1.2) at the positive interior equilibrium point H3 is

    F(λ)=λ2m(u,v)λ+n(u,v)=0, (3.1)

    where

    m(u,v)=1ru(1+kv)K+Φ+Θ,n(u,v)=Θ[1ru(1+kv)K+Φ]+Ψ,Θ:=1(αd),Φ:=βuv(u+γ)2,Ψ:=[(1uK)rku(1+kv)2+βuu+γ]abu2(bu+c)2.

    Assume that m2(u,v)>4n(u,v), that is,

    (1ru(1+kv)K+Φ+Θ)2>4Θ[1ru(1+kv)K+Φ]+4Ψ (3.2)

    and m(u,v)+n(u,v)=1, that is to say

    r=(1+kv)Ku(1+Θ)(2+2Θ+Φ(1+Θ)+Ψ). (3.3)

    Then eigenvalue of F(λ)=0 are λ1=1 and λ2=2+Φ+Θru(1+kv)K. The condition |λ2|1 indicates that

    Θ[1ru(1+kv)K+Φ]+Ψ±1. (3.4)

    Consider the following set

    A1={(a,b,c,d,K,r,k,α,β,γ)R10+:(3.2), (3.3) and (3.4) are satisfied }

    Based on the above analysis, we can obtain that when the parameters change on set A1, system (1.2) will occur flip bifurcation at H3(u,v).

    We consider the following system

    (uv)=(uexp[r11+kv(1uK)βvu+γ]vexp(αavbu+cd)), (3.5)

    here (a,b,c,d,K,r1,k,α,β,γ)A1.

    Consider a perturbation corresponding to system (3.5) as follows:

    (uv)=(uexp[(r1+¯r)1+kv(1uK)βvu+γ]vexp(αavbu+cd)), (3.6)

    where ¯r is a small perturbation parameter and |¯r|1.

    Let p=uu and q=vv. Then we gain

    (pq)=(S11S12S21S22)(pq)+(f1(p,q,¯r)f2(p,q,¯r)), (3.7)

    where

    f1(p,q,¯r)=S13p2+S14pq+S15q2+T11p3+T12p2q+T13pq2+T14q3+W1p¯r+W2q¯r+W3¯r2+W4pq¯r+W5p2¯r+W6q2¯r+W7p¯r2+W8q¯r2+W9¯r3+O((|p|,|q|,|¯r|)4),f2(p,q,¯r)=S23p2+S24pq+S25q2+T21p3+T22p2q+T23pq2+T24q3+O((|p|,|q|,|¯r|)4),S11=1r1u(1+kv)K+βuv(u+γ)2,S12=(1uK)r1ku(1+kv)2βuu+γ,S13=r1K(1+kv)r1βuv(u+γ)2K(1+kv)+βv(u+γ)2βuv(u+γ)3+r21u2K2(1+kv)2+β2uv22(u+γ)4,S14=r1k(Ku)K(1+kv)2βu+γr21k2(Ku)uK2(1+kv)4r1kβ(Ku)uK(1+kv)2(u+γ)2β2u(u+γ)3r1kβu(u+γ)K(1+kv)2,S15=u2(r1k(Ku)K(1+kv)2+βu+γ)2+r1k2(Ku)uK(1+kv)3,T11=r212K2(1+kv)2r1βv2K(1+kv)(u+γ)2+β2v22βv(γu)6(u+γ)4r31u6K3(1+kv)32βv3(u+γ)3r1βv(uγ)6K(1+kv)(u+γ)3+β2v2(γ2uγ)3(u+γ)5+r21βuv2K2(1+kv)2(u+γ)2r1β2uv22K(1+kv)(u+γ)4+β3uv3(u+γ)6,T12=r21k(K2u)K2(1+kv)3r1kβv(Ku)2r1βu2K(1+kv)2(u+γ)2+r1βK(1+kv)(u+γ)+r1kK(1+kv)2+β2(u+γ)2+β(γu)2(1+kv)3r1kβv(Ku)(γu)2K(1+kv)2(u+γ)3β2v2(u+γ)3r31k(Ku)u2K3(1+kv)3r21βu2K2(1+kv)2(u+γ)+r21kβuv(Ku)K2(1+kv)3(u+γ)2+r1β2uvK(1+kv)(u+γ)3r1kβ2uv2(Ku)2K(1+kv)2(u+γ)4β3uv22(u+γ)5,T13=r21k2(Ku)(K(2+K)u)2K2(1+kv)4+r1kβ(K3u)K(1+kv)2r21kβ(Ku)u2K2(1+kv)3(u+γ)+β2(u+γ)2+β3uv2β2u(u+γ)2(u+γ)4+r1k2(K2u)K(1+kv)3r31k3(Ku)2u2K3(1+kv)5+r21k2β(Ku)2uv2K2(1+kv)4(u+γ)2+r1kβ2(Ku)2uv2K(1+kv)2(u+γ)3r1kβ(Ku)2u(1+kv)2K(1+kv)3(u+γ)2r1β2u2K(1+kv)(u+γ)2r1kβ(Ku)u2K(1+kv)2(u+γ)2,
    T14=r31k3(Ku)3u6K3(1+kv)6r1kβ2(Ku)u2K(1+kv)2(u+γ)2r21k2β(Ku)2u2K2(1+kv)4(u+γ)β3u6(u+γ)3r21k3(Ku)2uK2(1+kv)5r1k2β(Ku)uK(1+kv)3(u+γ)r1k3(Ku)u(1+kv)4,W1=K2uK(1+kv)+(Ku)uK(1+kv)(βv(u+γ)2r1K(1+kv)),W3=(Ku)2u2K2(1+kv)2,W2=(Ku)uK(1+kv)(r1k(Ku)K(1+kv)2βu+γ)k(Ku)uK(1+kv)2,W4=β(Ku)uβ(K2u)(u+γ)K(1+kv)(u+γ)2+r21k(Ku)2ukK2(K2u)(1+kv)2K3(1+kv)4+3r1k(Ku)ur1k(Ku)2K2(1+kv)3r1kβ(Ku)2uvK2(1+kv)3(u+γ)2+r1β(Ku)uK2(1+kv)2(u+γ)β2(Ku)uvK(1+kv)(u+γ)3kβ(Ku)uvK(1+kv)2(u+γ)2,W5=r1(2uK)K2(1+kv)2+β(K3u)v2K(1+kv)(u+γ)21K(1+kv)+β(Ku)(γu)v2K(1+kv)(u+γ)3+r21(Ku)u2K3(1+kv)3r1β(Ku)uvK2(1+kv)2(u+γ)2+β2(Ku)uv22K(1+kv)(u+γ)4,W6=r21k2(Ku)3u2K3(1+kv)5+β2(Ku)u2K(1+kv)(u+γ)2+r1kβ(Ku)2u2K2(1+kv)3(u+γ)+(Ku)u2K(1+kv)+r1k2(Ku)2uK2(1+kv)4+kβ(Ku)u2K(1+kv)2(u+γ)+k2(Ku)uK(1+kv)3,W7=(Ku)22(Ku)u2K2(1+kv)2r1(Ku)2u2K3(1+kv)3+β(Ku)2uv2K2(1+kv)2(u+γ)2,W8=r1k(Ku)3u2K3(1+kv)4β(Ku)2u2K2(1+kv)2(u+γ)k(Ku)2uK2(1+kv)3,W9=(Ku)3u6K3(1+kv)3,S21=abv2(bu+c)2,S22=1(αd),S23=ab2v2(av2bu2c)2(bu+c)4,S25=a(av2bu2c)2(bu+c)2,T21=ab3v(6b2u26abuv+a2v2+12bcu6acv+6c2)6(bu+c)6,S24=abv(2bu+2cav)(bu+c)3,T22=ab2v(4b2u25abuv+a2v2+8bcu5acv+4c2)2(bu+c)5,T23=ab(2b2u24abuv+a2v2+4bcu4acv+2c2)2(bu+c)4,T24=a2(3buav+3c)6(bu+c)3.

    We construct a nonsingular matrix D1 and translate it as follows:

    (pq)=D1(uv), (3.8)

    where

    D1=(S12S121S11λ2S11).

    Taking D11 on both sides of Eq (3.8), we obtain

    (uv)=(100λ2)(uv)+(g1(p,q,¯r)g2(p,q,¯r)), (3.9)

    where

    g1(p,q,¯r)=[S13(λ2S11)S12S23]p2S12(λ2+1)+[S14(λ2S11)S12S24]pqS12(λ2+1)+W1(λ2S11)p¯rS12(λ2+1)+[S15(λ2S11)S12S25]q2S12(λ2+1)+[T11(λ2S11)S12T21]p3S12(λ2+1)+W2(λ2S11)q¯rS12(λ2+1)+[T12(λ2S11)S12T22]p2qS12(λ2+1)+[T13(λ2S11)S12T23]pq2S12(λ2+1)+W3(λ2S11)¯r2S12(λ2+1)+[T14(λ2S11)S12T24]q3S12(λ2+1)+W4(λ2S11)pq¯rS12(λ2+1)+W5(λ2S11)p2¯rS12(λ2+1)+W6(λ2S11)q2¯rS12(λ2+1)+W7(λ2S11)p¯r2S12(λ2+1)+W8(λ2S11)q¯r2S12(λ2+1)+W9(λ2S11)¯r3S12(λ2+1)+O((|p|,|q|,|¯r|)4),g2(p,q,¯r)=[S13(λ2+S11)+S12S23]p2S12(λ2+1)+[S14(λ2+S11)+S12S24]pqS12(λ2+1)+W1(λ2+S11)p¯rS12(λ2+1)+[S15(λ2+S11)+S12S25]q2S12(λ2+1)+[T11(λ2+S11)+S12T21]p3S12(λ2+1)+W2(λ2+S11)q¯rS12(λ2+1)+[T12(λ2+S11)+S12T22]p2qS12(λ2+1)+[T13(λ2+S11)+S12T23]pq2S12(λ2+1)+W3(λ2+S11)¯r2S12(λ2+1)+[T14(λ2+S11)+S12T24]q3S12(λ2+1)+W4(λ2+S11)pq¯rS12(λ2+1)+W5(λ2+S11)p2¯rS12(λ2+1)+W6(λ2+S11)q2¯rS12(λ2+1)+W7(λ2+S11)p¯r2S12(λ2+1)+W8(λ2+S11)q¯r2S12(λ2+1)+W9(λ2+S11)¯r3S12(λ2+1)+O((|p|,|q|,|¯r|)4),p=S12(u+v),q=(λ2+S11)v(1+S11)u.

    Applying the center manifold theorem Wc(0) of system (3.9) at the trivial equilibrium point (0, 0) in a limited field of ¯r=0. Then there exists a center manifold Wc(0) as follows:

    Wc(0)={(p,q,¯r)R3:q(p,¯r)=e0¯r+e1p2+e2p¯r+e3¯r2+O((|p|+|¯r|)3)}

    and satisfies

    H(q(p,¯r))=q(u+g1(p,q(p,¯r),¯r))λ2q(p,¯r)g2(p,q(p,¯r),¯r)=0,

    and we have

    e0=0,e1=[S13(1+S11)+S12S23]S12[S14(1+S11)+S12S24](1+S11)1λ22+[S15(1+S11)+S12S25](1+S11)2(1λ22)S12,e2=[S12W1W2(1+S11)](1+S11)(1λ2)2,e3=W3(1+S11)S12(1λ2)2.

    Therefore, we consider the map restricted to the center manifold Wc(0) as below:

    G:pp+n1p2+n2p¯r+n3p2¯r+n4p¯r2+n5p3+O((|p|+|¯r|)4),

    where

    n1=[S13(λ2S11)S12S23]S121+λ2[S14(λ2S11)S12S24](1+S11)1+λ2+[S15(λ2S11)S12S25](1+S11)2S12(1+λ2),n2=W1(λ2S11)1+λ2W2(λ2S11)(1+S11)S12(1+λ2),n3=[S13(λ2S11)S12S23]2e2S121+λ2+[S14(λ2S11)S12S24](λ2S11)e21+λ22[S15(λ2S11)S12S25](1+S11)(λ2S11)e2S12(1+λ2)+W1(λ2S11)e11+λ2+(λ2S11)2e1S12(1+λ2)+S12(λ2S11)W51+λ2+W6(λ2S11)(1+S11)2(1+λ2)S12[S14(λ2S11)S12S24]e2(1+S11)+W4(λ2S11)(1+S11)1+λ2,n4=[S13(λ2S11)S12S23]2e2S121+λ2+[S14(λ2S11)S12S24](λ2S11)e3λ2+1[S14(λ2S11)S12S24](1+S11)e3λ2+1+(λ2S11)(W1e2+W7)λ2+1[S15(λ2S11)S12S25](1+S11)(λ2S11)e3λ2+1+(λ2S11)2W2e2S12(λ2+1)W8(λ2S11)(S11+1)S12(λ2+1),
    n5=[T11(λ2S11)S12T21]S2121+λ2[T12(λ2S11)S12T22](1+S11)S12λ2+12[S15(λ2S11)S12S25](1+S11)(λ2S11)e1(1+λ2)S12+[S13(λ2S11)S12S23]2e2S121+λ2[S14(λ2S11)S12S24](1+S11)e1λ2+1+[S14(λ2S11)S12S24](λ2S11)e1λ2+1+[T13(λ2S11)S12T23](1+S11)2λ2+1[T14(λ2S11)S12T24](1+S11)3S12(λ2+1).

    According to flip bifurcation, we define the following two nonzero real numbers δ1 and δ2, where

    δ1=(2Gp¯r+12G¯r2Gp2)|(0,0)=n2,δ2=(163Gp3+(122Gp2)2)|(0,0)=n21+n5.

    From the above analysis, we get the following theorem:

    Theorem 2. If δ10,δ20, then system (1.2) passes through a flip bifurcation at the fixed point H3(u,v) when the parameter r alters in the small region of r1. In addition, if δ2>0 (resp., δ2<0), then the period-two orbits that bifurcate from fixed point H3(u,v) are stable (resp., unstable).

    When the parameters change on set A2, system (1.2) will undergo Neimark-Sacker bifurcation at the unique positive interior equilibrium point H3(u,v), where

    A2={(a,b,c,d,K,α,β,γ,k,r2):r2=(1+kv)KuΘ(Θ+ΦΘ+Ψ1),|1ru(1+kv)K+Φ+Θ|<2}.

    Consider a perturbation related to system (1.2) as follows:

    (uv)=(uexp[(r2+¯r)1+kv(1uK)βvu+γ]vexp(αavbu+cd)), (3.10)

    where ¯r is a limited perturbation parameter and |¯r|1.

    The characteristic equation of system (3.10) at H3(u,v) is as follows:

    λ2m(¯r)λ+n(¯r)=0,

    where

    m(¯r)=1(r2+¯r)u(1+kv)K+Φ+Θ

    and

    n(¯r)=Θ[1(r2+¯r)u(1+kv)K+Φ]+Ψ.

    Since parameters (a,b,c,d,K,α,β,γ,k,r2)A2, the characteristic values of system (3.10) at H3(u,v) are a pair of complex conjugate numbers λ and ˉλ with |λ|=|ˉλ|=1 as follows

    λ,ˉλ=m(¯r)±i4n(¯r)m2(¯r)2.

    Therefore we have

    |λ|=|ˉλ|=n(¯r))1/2,d|λ|d¯r|¯r=0=d|ˉλ|d¯r|¯r=0=Θu2(1+kv)n(0)<0.

    When ¯r changes in limited field of ¯r=0, then λ,ˉλ=x±iy, where

    x=m(0)2,y=4n(0)m2(0)2.

    In addition, Neimark-Sacker bifurcation requires that ¯r=0, λz,ˉλz1 (z = 1, 2, 3, 4), which is equivalent to m(0) -2, 0, -1, 2. Because parameters (a,b,c,d,K,α,β,γ,k,r2)A2, therefore m(0) -2, 2. We only require m(0) 0, -1, so that

    1+Φ+Θr2u(1+kv)K,1+Φ+Θr2u(1+kv)K1. (3.11)

    Let p=uu and q=vv.

    After the transformation of the equilibrium point H3(u,v) of system (3.10) to the origin, we have

    (pq)=(S11S12S21S22)(pq)+(f1(p,q)f2(p,q)), (3.12)

    where

    f1(p,q)=S13p2+S14pq+S15q2+T11p3+T12p2q+T13pq2+T14q3+O((|p|,|q|)4),f2(p,q)=S23p2+S24pq+S25q2+T21p3+T22p2q+T23pq2+T24q3+O((|p|,|q|)4),

    and S11,S12,S13,S14,S15,T11,T12,T13,T14,S21,S22,S23,S24,S25,T21,T22,T23,T24, are given in (3.7) by substituting r1 for r2+¯r.

    Besides that, we analyse the normal form of system (3.12) when ¯r=0.

    Consider the translation as follows:

    (pq)=D2(uv),

    where

    D2=(S120xS11y).

    Taking D12 on both sides of system (3.12), we acquire

    (uv)=(xyyx)(uv)+(f(u,v)g(u,v)),

    where

    f(u,v)=S13p2S12+S14pqS12+S15q2S12+T11p3S12+T12p2qS12+T13pq2S12+T14q3S12+O((|p|,|q|)4),g(u,v)=[S13(xS11)S12S23]p2S12y+[S14(xS11)S12S24]pqS12y+[S15(xS11)S12S25]q2S12y+[T11(xS11)S12T21]p3S12y+[T12(xS11)S12T22]p2qS12y+[T13(xS11)S12T23]pq2S12y+[T14(xS11)S12T24]q3S12y+O((|p|,|q|)4),p=S12u,q=(xS11)uyv.

    System (1.2) occurs the Neimark-Sacker bifurcation if the following quantity ϑ is not zero,

    ϑ=Re[(12ˉλ)ˉλ21λϱ11ϱ20]12ϱ112ϱ022+Re(ˉλϱ21),

    where

    ϱ11=14[(fpp+fqq)+i(gpp+gqq)],ϱ20=18[(fppfqq+2gpq)+i(gppgqq2fpq)],ϱ02=18[(fppfqq2gpq)+i(gppgqq+2fpq)],ϱ21=116[(fppp+fpqq+gppq+gqqq)+i(gppp+gpqqfppqfqqq)].

    If ϑ0, Neimark-Sacker bifurcation will undergo in system (1.2), and the following theorem holds:

    Theorem 3. System (1.2) undergoes a Neimark-Sacker bifurcation at the positive equilibrium point H3(u,v) if conditions (3.11) are satisfied and ϑ0. In addition, if ϑ>0 (resp., ϑ<0), then an repelling (resp., attracting) invariant closed curve bifurcates from fixed point H3(u,v) for r<r2 (resp., r>r2).

    In this section, we will adopt the feedback control method [26,27,28] to stabilize the chaotic orbit at an unstable equilibrium point by adding a feedback control term to the system (1.2). Therefore, system (1.2) makes the following form:

    {un+1=unexp[r1+kvn(1unK)βvnun+γ]x(un,vn)=f(un,vn),vn+1=vnexp(αavnbun+cd)=g(un,vn), (4.1)

    where x(un,vn)=h1(unu)+h2(vnv) is feedback controlling force, h1 and h2 are feedback gains, and (u,v) is the unique positive equilibrium point of (1.2). Furthermore, f(u,v)=u, and g(u,v)=v.

    The Jacobian matrix of system (4.1) at positive equilibrium point (u,v) is as follows:

    J(u,v)=[S11h1S12h2S21S22],

    where

    S11=1r1u(1+kv)K+βuv(u+γ)2,S12=(1uK)r1ku(1+kv)2βuu+γ,S21=abv2(bu+c)2,S22=1(αd).

    Thus, the characteristic equation related to J(u,v) is:

    λ2(S11+S22h1)λ+(S11h1)S22(S12h2)S21=0. (4.2)

    Let λ1 and λ2 be the eigenvalues of characteristic equation (4.2), then

    λ1+λ2=S11+S22h1,λ1λ2=(S11h1)S22(S12h2)S21. (4.3)

    Next, we must solve equations λ1=±1 and λ1λ2=1 to gain the critical stability line. At the same time, it also ensures that the absolute value λ1 and λ2 are less than one.

    Suppose that λ1λ2=1, then we gain

    L1:S11S22S12S211=S22h1S21h2.

    Assume that λ1=1, then we have

    L2:S11+S22S11S22+S12S211=(1S22)h1+S21h2.

    Assume that λ1=1, then we obtain

    L3:S11+S22+S11S22S12S21+1=(1+S22)h1S21h2.

    Thus, the stable eigenvalues lie within the triangular area with the boundaries of the straight lines L1,L2,L3. In addition, when the control parameters h1 and h2 take values in the triangular region, system (4.1) will not generate chaos.

    In this section, we draw the bifurcation diagrams, phase portraits, solution of the figures and maximum Lyapunov exponents for system (1.2) to verify the above theoretical analysis and show the new interesting complex dynamical behaviors and the stability of the predator-prey system at the equilibrium point by using numerical simulations.

    First, in Figure 1, we consider that the fear factor k=0 and take r as the bifurcation parameter to discuss the dynamic behavior of (1.2) at H3(u,v). We consider the parameter values as (a,b,c,d,α,β,γ,K)=(1.8,2.8,3.5,0.6,1.27,1.1,2.7,1.5)A1 with the initial value of (u0,v0) = (2, 1) and r [2.8, 4.6]. Flip bifurcation emerges from the unique positive equilibrium point and loses its stability as r goes through a critical value r=3.102, and it is stable when r<3.102, and when r>3.102, system (1.2) oscillates with periods of 2,22,23,. It can be obtained from Figure 1(c) and Figure 2(a–c) that chaos will happen in system (1.2) as the bifurcation parameters r continue to increase.

    Figure 1.  (a, b) Bifurcation diagram of system (1.2) with r [2.8, 4.6], a=1.8,b=2.8,c=3.5,d=0.6,α=1.27,β=1.1, γ=2.7,k=0,K=1.5 the initial value is (u0,v0) = (2, 1). (c) Maximum Lyapunov exponents corresponding to (a, b).
    Figure 2.  Phase portraits and solution portraits for various values of r corresponding to Figure 1.

    In Figure 3, taking (a,b,c,d,α,β,γ,K)=(1.7,1.6,3.1,0.01,1.2,2.6,1.2,3.5)A2 with the initial value of (u0,v0) = (1, 2) and r [4.6, 6.2]. Neimark-Sacker bifurcation emerges from the unique positive equilibrium point and loses its stability as r goes through a critical value r=5.05. It can be seen from Figure 3(a) that when r<4.7, the equilibrium point of (1.2) with respect to prey does not exist. Not only does the system have an attractive-invariant loop and periodic solutions, but it also exhibits dynamical chaos as the bifurcation parameters r continue to increase. Figure 3(c) is the maximum Lyapunov exponent diagram related to Figure 3(a, b). It can be seen from the MLE that system (1.2) will appear chaotic. By observing the Figure 4(a–c), it can be found that when r>5.05, a limit cycle, a periodic window and chaos appear in system (1.2).

    Figure 3.  (a, b) Bifurcation diagram of system (1.2) with r [4.6, 6.2], a=1.7,b=1.6,c=3.1,d=0.01,α=1.2,β=2.6,γ=1.2,k=0,K=3.5 the initial value is (u0,v0) = (1, 2). (c) Maximum Lyapunov exponents related to (a, b).
    Figure 4.  Phase portraits and solution portraits for various values of r corresponding to Figure 3.

    In Figure 5, we consider that the fear factor k>0. Taking (a,b,c,d,α,β,γ,k,K)=(2,2.8,3.5,0.6,1.1,1.1,3,0.5,2)A1 with the initial value of (u0,v0) = (2, 1) and r [5.2, 7]. Flip bifurcation appears from the unique positive equilibrium point and loses its stability as r goes through a critical value r=5.63, and it is stable when r<5.63 and when r>5.63, system (1.2) oscillates with periods of 2,22,23,. It can be acquired from Figure 5(c) that chaos will happen in system (1.2) as the bifurcation parameters r continue to increase.

    Figure 5.  (a, b) Bifurcation diagram of system (1.2) with r [5.2, 7], a=2,b=2.8,c=3.5,d=0.6,α=1.1,β=1.1,γ=3,k=0.5,K=2 the initial value is (u0,v0) = (2, 1). (c) Maximum Lyapunov exponents related to (a, b).

    In Figure 6, taking (a,b,c,d,α,β,γ,k,K)=(1.7,1.6,3.1,0.01,1.2,2.6,1.2,0.3,3.5)A2 with the initial value of (u0,v0) = (1, 2) and r [8,11]. Neimark-Sacker bifurcation emerges from the unique positive equilibrium point and loses its stability as r goes through a critical value r=8.63. We notice that the equilibrium point of (1.2) is stable for r<8.63, loses its stability at r=8.63 and not only a limit cycle but also periodic solution emerge when the bifurcation parameter r>8.63. Other than that, the value of the MLE related to (1.2) is greater than 0 as r continues to increase, and thus chaos will occur, i.e., the solution of (1.2) is arbitrarily periodic.

    Figure 6.  (a, b) Bifurcation diagram relevent u and v in system (1.2) with r [8,11], a=1.7,b=1.6,c=3.1,d=0.01,α=1.2,β=2.6, γ=1.2,k=0.3,K=3.5 the initial value is (u0,v0) = (1, 2). (c) Maximum Lyapunov exponents related to (a, b).

    In Figure 7, taking (a,b,c,d,α,β,γ,r,K)=(1.5,2.8,3.5,0.6,1,1.1,3,2,1.5) with the initial value of (u0,v0) = (2, 1) and k [0, 6], k is a bifurcation parameter. At this time, the bifurcation phenomenon of (1.2) will not occur. The population density of prey and predator will continue to decrease and tend to 0 with the increase of fear factor k. It is important to note that the cost of fear does not lead to the extinction of predators, but rather to the extinction of prey.

    Figure 7.  Bifurcation diagram of system (1.2) with k [0, 6], a=1.5,b=2.8,c=3.5,d=0.6,α=1,β=1.1,γ=3,K=1.5,r=2 the initial value is (u0,v0) = (2, 1).

    In Figure 8, when the parameter value is (a,b,c,d,α,β,γ,k,r,K)=(1.7,1.6,3.1,0.01,1.2,2.6,1.2,0.3,10.8,3.5) with the initial value of (u0,v0) = (1, 2). In Figure 6(c) when the bifurcation parameter r=10.8, system (1.2) will produce chaos. When the h1 and h2 are controlled in the triangular region surrounded by three straight lines L1, L2, and L3, the chaos generated by system (4.1) will be controlled near the equilibrium point and become an asymptotically stable state.

    Figure 8.  The bounded region for the eigenvalues of the controlled system (4.1) in the (h1,h2) plane.

    Studies have shown that discrete systems have richer and more complex dynamic behaviors than continuous systems. Hence, on the basis of previous research work, this paper discusses the stability, bifurcation and chaos control of a nonlinear discrete prey-predator system with fear effect. Based on the results of the study, we can draw the following conclusions:

    (a) System (1.2) has four equilibrium points, where the stable equilibrium point is positive, and depicts the coexistence of prey and predators.

    (b) System (1.2) has flip bifurcation and Neimark-Sacker bifurcation happen at the positive interior equilibrium point when r alters in A1 and A2 small fields. (see Figures 1, 3, 5, 6). We can also observe the orbits of periods 2, 4, and 8 periodic windows of flip bifurcation.

    (c) When k=0, system (1.2) at the positive equilibrium point will generate the Neimark-Sacker bifurcation, flip bifurcation and chaos as the bifurcation parameters r continue to increase.

    (d) When the fear parameter k is greater, both predators and prey populations decrease. It is important to note that the cost of fear does not lead to the extinction of predators, but rather to the extinction of prey. (see Figure 7).

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

    This work was supported by National Natural Science Foundation of China (No.12161079), the Northwest Normal University Graduate Research Grant Project (No.2022KYZZ-S114), and the Gansu Province Innovation Star Project (No.2023CXZX-325).

    The authors declare that there are no conflicts of interest regarding the publication of this paper.



    [1] A. A. Berryman, The origins and evolution of predator-prey theory, Ecology, 73 (1992), 1530–1535. http://dx.doi.org/10.2307/1940005 doi: 10.2307/1940005
    [2] H. Freedman, Deterministic Mathematical Models in Population Ecology, Edmonton: HIFR Consulting Ltd, 1980. http://dx.doi.org/10.2307/2975858
    [3] M. Sen, M. Banerjee, A. Morozov, Bifurcation analysis of a ratio-dependent prey-predator model with the Allee effect, Ecol. Compl., 11 (2012), 12–27. http://dx.doi.org/10.1016/j.ecocom.2012.01.002 doi: 10.1016/j.ecocom.2012.01.002
    [4] R. E. Kooij, A. Zegeling, A predator-prey model with Ivlev's functional response, J. Math. Anal. Appl., 198 (1996), 473–489. http://dx.doi.org/10.1016/j.chaos.2004.07.018 doi: 10.1016/j.chaos.2004.07.018
    [5] R. Lopez-Ruiz, R. Fournier-Prunaret, Indirect Allee effect, bistability and chaotic oscillations in a predator-prey discrete model of logistic type, Chaos Solitons Fract., 24 (2005), 85–101. http://dx.doi.org/10.1016/j.chaos.2004.07.018 doi: 10.1016/j.chaos.2004.07.018
    [6] J. D. Murray, Mathematical Biology, 2nd edtion, Berlin: Springer-Verlag, 1993. http://dx.doi.org/10.1137/1032093
    [7] W. Ma, Y. Takeuchi, Stability analysis on a predator-prey system with distributed delays, J. Comput. Appl. Math., 88 (1998), 79–94. http://dx.doi.org/10.1016/S0377-0427(97)00203-3 doi: 10.1016/S0377-0427(97)00203-3
    [8] S. Sinha, O. Misra, J. Dhar, Modelling a predator-prey system with infected prey in polluted environment, Appl. Math. Model., 34 (2010), 1861–1872. http://dx.doi.org/10.1016/j.apm.2009.10.003 doi: 10.1016/j.apm.2009.10.003
    [9] M. Fan, K. Wang, Periodic solutions of a discrete time non-autonomous ratio-dependent predator-prey system, Math. Comput. Model., 35 (2002), 951–961. http://dx.doi.org/10.1016/s0895-7177(02)00062-6 doi: 10.1016/s0895-7177(02)00062-6
    [10] D. Summers, J. G. Cranford, B. P. Healey, Chaos in periodically forced discrete-time ecosystem models, Chaos Solitons Fract., 11 (2000), 2331–2342. http://dx.doi.org/10.1016/S0960-0779(99)00154-X doi: 10.1016/S0960-0779(99)00154-X
    [11] L. Yuan, Q. Yang, Bifurcation, invariant curve and hybrid control in a discrete-time predator-prey system, Appl. Math. Model., 39 (2015), 2345–2362. http://dx.doi.org/10.1016/j.apm.2014.10.040 doi: 10.1016/j.apm.2014.10.040
    [12] G. Q. Sun, J. Zhang, L. P. Song, Z. Jin, B. L. Li, Pattern formation of a spatial predator-prey system, Appl. Math. Comput., 218 (2012), 11151–11162. http://dx.doi.org/10.1016/j.amc.2012.04.071 doi: 10.1016/j.amc.2012.04.071
    [13] B. Dubey, B. Das, J. Hussain, A predator-prey interaction model with self and cross-diffusion, Ecol. Model., 141 (2001), 67–76. http://dx.doi.org/10.1016/S0304-3800(01)00255-1 doi: 10.1016/S0304-3800(01)00255-1
    [14] G. Q. Sun, Z. Jin, L. Li, M. Haque, B. L. Li, Spatial patterns of a predator-prey model with cross diffusion, Nonlinear Dynam., 69 (2012), 1631–1638. http://dx.doi.org/10.1007/s11071-012-0374-6 doi: 10.1007/s11071-012-0374-6
    [15] C. Robinson, Dynamical Systems: Stability, Symbolic Dynamics, and Chaos, Florida: CRC Press, 1998.
    [16] K. L. Edelstein, Mathematical Model in Biology, New York: McGraw-Hill, 1988.
    [17] N. Hiroyuki, B. Yoshikazu, Introduction to Chaos, Physics and Mathematics of Chaotic Phenomena, Bristol: Institute of Physics Publishing, 1999. http://dx.doi.org/10.1201/9780429187001
    [18] H. F. Huo, W. T. Li, Existence and global stability of periodic solutions of a discrete predator-prey system with delays, Appl. Math. Comput., 153 (2004), 337–351. http://dx.doi.org/10.1016/S0096-3003(03)00635.0 doi: 10.1016/S0096-3003(03)00635.0
    [19] L. F. Cheng, H. G. Cao, Bifurcation analysis of a discrete-time ratio-dependent predator-prey model with Allee Effect, Commun. Nonlinear Sci. Numer. Simul., 38 (2016), 288–302. http://dx.doi.org/10.1016/j.cnsns.2016.02.038 doi: 10.1016/j.cnsns.2016.02.038
    [20] X. Yang, Uniform persistence and periodic solutions for a discrete predator-prey system with delays, Math. Anal. Appl., 316 (2006), 161–177. http://dx.doi.org/10.1016/j.jmaa.2005.04.036 doi: 10.1016/j.jmaa.2005.04.036
    [21] X. Wang, L. Zanette, X. Zou, Modelling the fear effect in predator-prey interactions, Math. Biol., 73 (2016), 1179–1204. http://dx.doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
    [22] M. X. Chen, R. C. Wu, Steady state bifurcation in Previte-Hoffman model, Int. J. Bifur. Chaos, 33 (2023), 2350020. http://dx.doi.org/10.1142/S0218127423500207 doi: 10.1142/S0218127423500207
    [23] S. Gakkhar, A. Singh, Complex dynamics in a prey-predator system with multiple delays, Commun. Nonlinear Sci. Numer. Simul., 17 (2012), 914–929. http://dx.doi.org/10.1016/j.cnsns.2011.05.047 doi: 10.1016/j.cnsns.2011.05.047
    [24] M. X. Chen, R. C. Wu, X. H. Wang, Non-constant steady states and Hopf bifurcation of a species interaction model, Commun. Nonlinear Sci. Numer. Simul., 116 (2023), 106846. http://dx.doi.org/10.1016/j.cnsns.2022.106846 doi: 10.1016/j.cnsns.2022.106846
    [25] X. L. Han, C. Y. Lei, Bifurcation and turing instability analysis for a space-and time-discrete predator-prey system with Smith growth function, Chaos Solitons Fract., 166 (2023), 112920. http://dx.doi.org/10.1016/j.chaos.2022.112910 doi: 10.1016/j.chaos.2022.112910
    [26] S. Lynch, Dynamical Systems with Applications Using Mathematica, Boston: Birkhauser, 2007. http://dx.doi.org/10.1007/978-0-8176-4586-1
    [27] S. Elaydi, An Introduction to Difference Equations, 3rd edition, New York: Springer-Verlag, 2005. http://dx.doi.org/10.1007/0-387-27602-5
    [28] G. Chen, X. Dong, From Chaos to Order: Perspectives, Methodologies, and Applications, Singapore: World Scientific, 1998. http://dx.doi.org/10.1142/3033
    [29] C. Y. Lei, X. L. Han, W. Wang, Bifurcation analysis and chaos control of a discrete-time prey-predator model with fear factor, Math. Biosci. Eng., 19 (2022), 6659–6679. http://dx.doi.org/10.3934/mbe.2022313 doi: 10.3934/mbe.2022313
    [30] X. X. Du, X. L. Han, C. Y. Lei, Behavior analysis of a class of discrete-time dynamical system with capture rate, Mathematics, 10 (2022), 2410. http://dx.doi.org/10.3390/math10142410 doi: 10.3390/math10142410
    [31] Q. Din, Complexity and chaos control in a discrete-time prey-predator model, Commun. Nonlinear Sci. Numer. Simul., 49 (2017), 113–134. http://dx.doi.org/10.1016/j.cnsns.2017.01.025 doi: 10.1016/j.cnsns.2017.01.025
  • Reader Comments
  • © 2023 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(1438) PDF downloads(86) Cited by(0)

Figures and Tables

Figures(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog