Processing math: 91%
Research article Special Issues

Dynamics analysis of a predator-prey model incorporating fear effect in prey species

  • Fear effects, as spontaneous inherent phenomena among species, ubiquitously manifest in natural ecosystems. In this study, we investigate a Leslie-Gower type predator-prey system incorporating a Holling Type Ⅲ functional response and fear effects under no-flux boundary conditions. For temporal dynamical behaviors, we investigate the Hopf bifurcation for the ordinary differential equations (ODEs). It is found that the system will admit the stable or unstable periodic solution due to supercritical or subcritical Hopf bifurcation based on the first Lyapunov coefficient technique. When we add the diffusion into the system, we rigorously establish the stability conditions of the positive equilibrium. In this manner, the precise existence interval of the Turing instability is obtained so that the spatial profiles of the system can be analyzed. Furthermore, we systematically analyze the existence of Hopf bifurcation in the reaction-diffusion system and characterize the stability of bifurcating periodic solutions by employing normal form theory. The theoretical framework is supported by comprehensive numerical simulations. The research highlights of this paper include the following: (1) The existence of the supercritical and subcritical Hopf bifurcation for the temporal and spatiotemporal predator-prey systems are confirmed; (2) the Turing instability of the spatiotemporal system is found via strict theoretical derivations so that the pattern formation can be observed.

    Citation: Xiaoyan Zhao, Liangru Yu, Xue-Zhi Li. Dynamics analysis of a predator-prey model incorporating fear effect in prey species[J]. AIMS Mathematics, 2025, 10(5): 12464-12492. doi: 10.3934/math.2025563

    Related Papers:

    [1] 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
    [2] 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
    [3] 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
    [4] 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
    [5] 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
    [6] Kwadwo Antwi-Fordjour, Rana D. Parshad, Hannah E. Thompson, Stephanie B. Westaway . Fear-driven extinction and (de)stabilization in a predator-prey model incorporating prey herd behavior and mutual interference. AIMS Mathematics, 2023, 8(2): 3353-3377. doi: 10.3934/math.2023173
    [7] 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
    [8] Kimun Ryu, Wonlyul Ko . Stability and bifurcations in a delayed predator-prey system with prey-taxis and hunting cooperation functional response. AIMS Mathematics, 2025, 10(6): 12808-12840. doi: 10.3934/math.2025576
    [9] Reshma K P, Ankit Kumar . Stability and bifurcation in a predator-prey system with effect of fear and additional food. AIMS Mathematics, 2024, 9(2): 4211-4240. doi: 10.3934/math.2024208
    [10] Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045
  • Fear effects, as spontaneous inherent phenomena among species, ubiquitously manifest in natural ecosystems. In this study, we investigate a Leslie-Gower type predator-prey system incorporating a Holling Type Ⅲ functional response and fear effects under no-flux boundary conditions. For temporal dynamical behaviors, we investigate the Hopf bifurcation for the ordinary differential equations (ODEs). It is found that the system will admit the stable or unstable periodic solution due to supercritical or subcritical Hopf bifurcation based on the first Lyapunov coefficient technique. When we add the diffusion into the system, we rigorously establish the stability conditions of the positive equilibrium. In this manner, the precise existence interval of the Turing instability is obtained so that the spatial profiles of the system can be analyzed. Furthermore, we systematically analyze the existence of Hopf bifurcation in the reaction-diffusion system and characterize the stability of bifurcating periodic solutions by employing normal form theory. The theoretical framework is supported by comprehensive numerical simulations. The research highlights of this paper include the following: (1) The existence of the supercritical and subcritical Hopf bifurcation for the temporal and spatiotemporal predator-prey systems are confirmed; (2) the Turing instability of the spatiotemporal system is found via strict theoretical derivations so that the pattern formation can be observed.



    Predator-prey models constitute a fundamental class of ecological systems, extensively employed to characterize the complex interaction dynamics between prey and predator populations [1,2,3,4]. Typically, in a predator-prey system, the prey species will be consumed by the predators; in other words, the growth of the predator will dependent on the density of the prey species. Nowadays, scholars have developed diverse mathematical frameworks capturing the intrinsic behavioral traits of both populations owing to a basic logistic growth equation framework. Of particular interest in our investigation, we consider the following non-dimensional Leslie-Gower type predator-prey system with a Holling Type Ⅲ functional response:

    {dwdt=ww2aw2vw2+sv2,dvdt=τv(1vw), (1.1)

    where w and v represent the densities of the prey and predators, the term aw2w2+sv2 is the well-known ratio-dependent Holling Type Ⅲ functional response [5,6], and vu corresponds to the Leslie-Gower-type trophic interaction [7,8,9], where the parameters a,τ, and s are positive constants. For prior dynamic analyses of system (1.1), we direct the readers to foundational studies [10,11].

    Fear effects, as ubiquitous non consumptive mechanisms in predator-prey interactions, significantly modulate prey species' intrinsic growth rates through predator-induced physiological stress responses. In fact, the integration of fear effects into predator-prey systems has been systematically investigated through various modeling frameworks. According to a predator-prey model with the fear factor, Tripathi et al. [12] obtained the stability of the system's equilibria, the existence of the Hopf bifurcation, and the Turing instability, as well as pattern formation. By incorporating fear effects and prey refuges for a Filippov prey-predator model, Hamdallah and Arafa [13] analyzed the stability and they obtained bifurcation sets of pseudo equilibrium and local/global sliding bifurcations. By utilizing an ecological model with fear, group defense, and the Allee effect, Kumar et al. [14] investigated the complex dynamical properties of the predator-prey system via saddle-node, Hopf, homoclinic, Bautin, and Bogdanov-Takens bifurcation. Cao et al. [15] reported the Hopf bifurcation, steady-state bifurcation, and the stability of the bifurcating periodic solution of a reaction-diffusion predator-prey system with fear effects. For comprehensive analyses of fear effect dynamics in ecological modeling, see the existing studies in [16,17,18,19,20]. Building upon these theoretical foundations, we propose the following modified system through fear effect coupling:

    {dwdt=w1+θvw2aw2vw2+sv2,dvdt=τv(1vw), (1.2)

    where 11+θv models the fear effect with the positive constant θ, which describes the level of fear. It is clear that |w1+θv||w|. This implies that the fear effect directly impacts the prey population's growth rate, i.e., a larger value of the fear effect control parameter θ corresponds to a lower actual growth rate of the prey population. Meanwhile, if the fear factor disappears in the system (1.2), namely, θ=0, then the system (1.2) will reduce to system (1.1).

    Both systems (1.1) and (1.2) are governed by ordinary differential equations (ODEs), which exclusively describe the temporal dynamics of predator-prey interactions. These formulations capture solely temporal population dynamics without accounting for the spatial distribution mechanisms. However, natural population movements necessitate the incorporation of diffusion processes for both species within the spatial domains, thereby extending the framework to reaction-diffusion systems. Such systems exhibit enhanced dynamic complexity characterized by spatiotemporal pattern formation. For foundational studies on ecological reaction-diffusion systems, see [21,22,23]. Consequently, if we consider the diffusion effect in the system (1.2), then we get

    {wt=d1Δw+w1+θvw2aw2vw2+sv2,xΩ,t>0,vt=d2Δv+τv(1vw),xΩ,t>0,wn=vn=0,xΩ,t0,w(x,0)=w0(x)0,v(x,0)=v0(x)0,xΩ, (1.3)

    where d1Δw and d2Δv describe the random diffusion progress of the prey w and predator v with the diffusion rates d1 and d2, respectively; the notation Δ is the Laplacian operator; Ω=(0,π) is a bounded domain with the smooth boundary Ω; n is the outward unit normal vector; and n is the operator of the directional derivative along the direction n. In addition, w0(x) and v0(x) represent the initial densities of the prey w and predator v, respectively. All parameters involved in the system (1.3) are set to be positive constants.

    In this current paper, we want to explore the spatiotemporal dynamics of the modified systems. To be specific, we will investigate the stability of the positive equilibrium, the Hopf bifurcation, and its natural property for the ODE system (1.2) by virtue of the first Lyapunov coefficient technique; for the reaction-diffusion system (1.3), we will examine the precise conditions of the stability of the equilibrium, the existence of the Turing instability, and the Hopf bifurcation. Especially, we will determine the stability of the bifurcating periodic solution of the reaction-diffusion system (1.3) by using normal form theory and center manifold reduction [24,25,26,27]. Collectively, the complex spatiotemporal dynamics of the reaction-diffusion system (1.3) can be presented by incorporating the fear factor in the system. The outline of this paper is as follows. In Section 2, we study the stability of the equilibrium, and the occurrence conditions of the Hopf bifurcation and its direction for the ODE system. In Section 3, we conduct a stability analysis and explore the existence of the Turing instability and the Hopf bifurcation for the reaction-diffusion system. Section 4 displays the computational validation through numerical simulations. In Section 5, we summarize this paper with some conclusions.

    To yield the equilibria of system (1.2), we set

    {f(w,v):=w1+θvw2aw2vw2+sv2,g(w,v):=τv(1vw). (2.1)

    Now, we look for the positive equilibrium, say E=(w,v), of the system (1.2) by setting f(w,v)=g(w,v)=0 as w>0 and v>0. Utilizing (2.1), one has

    L(w):=θw2+(1+θA)w+A1=0, (2.2)

    where A:=a/(1+s)>0. By using (2.2), we can see that L(w)θw2+(1+θA)w>0 as a1+s. This implies that we must ensure that 0<a<1+s for the possibility of L(w)=0. Now, if 0<a<1+s is valid, we have the existence criterion of the roots of (2.2) is Δ(w):=(θA1)2+4θ>0. That is to say, L(w)=0 must have two real roots. Note that 0<a<1+s, so one has 0<A<1. Therefore, L(w)=0 has a unique positive root

    w=(θA1)2+4θθA12θ>0,

    where A=a/(1+s). Hence, there is the unique positive equilibrium E=(w,v)=(w,w) of the system (1.2) when 0<a<1+s.

    We have the following result concerning the system (1.2).

    Theorem 2.1. Suppose that 0<a<1+s is valid.

    (1) The positive equilibrium E is locally asymptotically stable if one of the following conditions is satisfied:

    (C1) s1;

    (C2) 0<s<1,a(1s)(1+s)2w;

    (C3) 0<s<1,a(1s)(1+s)2>w,τ>a(1s)(1+s)2w.

    However, it becomes unstable when

    (C4) 0<s<1,a(1s)(1+s)2>w,0<τ<a(1s)(1+s)2w.

    (2) If 0<s<1,a(1s)(1+s)2>w, the system (1.2) experiences the Hopf bifurcation as τ=τH0, where

    τH0=a(1s)(1+s)2w.

    Proof. At the positive equilibrium E=(w,v), the Jacobian matrix takes the form:

    JE(τ)=(J11(τ)J12(τ)ττ)=(a(1s)(1+s)2wa(s1)(1+s)2θw(1+θw)2ττ).

    We obtain the characteristic equation as follows:

    λ2TE(τ)λ+DE(τ)=0, (2.3)

    where

    {TE(τ)=a(1s)(1+s)2wτ,DE(τ)=wτ+θτw(1+θw)2.

    (1) If s1 is satisfied, then one has TE(τ)<0 and DE(τ)>0, and it is concluded that the positive equilibrium E is locally asymptotically stable. (2) We can infer that if 0<s<1 and a(1s)(1+s)2w, then one has TE(τ)<0 and DE(τ)>0. This means that the positive equilibrium E is locally asymptotically stable. (3) If 0<s<1,a(1s)(1+s)2>w, and τ>a(1s)(1+s)2w, we can see that TE(τ)<0 and DE(τ)>0. So, the positive equilibria E is locally asymptotically stable. (4) If 0<s<1,a(1s)(1+s)2>w, and 0<τ<a(1s)(1+s)2w, this immediately yields TE(τ)>0, and DE(τ)>0. So, the positive equilibrium E is unstable. Finally, solving λ for (2.3) yields

    λ1,2=TE(τ)2±i4DE(τ)T2E(τ)2:=ζ(τ)±iη(τ). (2.4)

    Therefore, if 0<s<1,a(1s)(1+s)2>w, and τ:=τH0=a(1s)(1+s)2w, we get

    ζ(τH0)=0,η(τH0)=τH0(w+θw(1+θw)2)=(a(1s)(1+s)2w)(w+θw(1+θw)2)>0.

    This implies that the characteristic Eq (2.3) has a pair of purely imaginary roots. On the other hand, we can compute

    dRe{λ}dτ|τ=τH0=dζ(τ)dτ|τ=τH0=12<0.

    Consequently, the system (1.2) presents the Hopf bifurcation as τ=τH0. The proof readily follows.

    Utilizing (2) of Theorem 2.1, we know that the system (1.2) experiences the Hopf bifurcation as τ=τH0, where

    τH0=a(1s)(1+s)2w,a(1s)(1+s)2>w.

    Now, our goal is to investigate the direction of this Hopf bifurcation. To this end, consider the translation ˜w=ww, ˜v=vv and still denote ˜w and ˜v as w and v, respectively. As such, the system (1.2) becomes

    {dwdt=(w+w)1+θ(v+v)(w+w)2a(w+w)2(v+v)(w+w)2+s(v+v)2,dvdt=τ(v+v)(1v+vw+w). (2.5)

    Rewrite the system (2.5) as follows:

    (wtvt)=JE(τ)(wv)+(F(τ,w,v)G(τ,w,v)), (2.6)

    where

    F(τ,w,v)=fww2w2+fwvwv+fvv2v2+fwww3!w3+fwwv2w2v+fwvv2wv2+fvvv3!v3+O(4),G(τ,w,v)=gww2w2+gwvwv+gvv2v2+gwww3!w3+gwwv2w2v+gwvv2wv2+gvvv3!v3+O(4),

    with

    fww=2a(s3)w(s+1)32,fwv=2a(13s)w(s+1)3θ(θv+1)2,fwwv=6a(6ss21)w2(1+s)4,fwww=24a(1s)w2(1+s)4,fwvv=24as(s1)w2(1+s)4+2θ2(θv+1)3,fvv=2as(3s)w(1+s)3+2θ2w(θv+1)3,fvvv=6as(s26s+1)w2(1+s)46θ3w(θv+1)4,

    and

    gwv=2τw,gww=2τw,gwww=6τw2,gwwv=4τw2,gwvv=2τw2,gvv=2τw,gvvv=0.

    Define a matrix

    K=(10ζ(τ)J11(τ)J12(τ)η(τ)J12(τ)),

    where

    J11(τ)=a(1s)(1+s)2w,J12(τ)=a(s1)(1+s)2θw(1+θw)2,

    and ζ(τ) and η(τ) can be found in (2.4). Let

    (wv)=K(ˆwˆv).

    Substituting it into (2.6), we have

    (ˆwtˆvt)=(ζ(τ)η(τ)η(τ)ζ(τ))(ˆwˆv)+(ˆF(τ,ˆw,ˆv)ˆG(τ,ˆw,ˆv)), (2.7)

    where

    ˆF(τ,ˆw,ˆv)=F(τ,ˆw,ζ(τ)J11(τ)J12(τ)ˆww(τ)J12(τ)ˆv),ˆG(τ,ˆw,ˆv)=ζ(τ)J11(τ)η(τ)˜F(τ,ˆw,ˆv)J12(τ)η(τ)G(τ,ˆw,ζ(τ)J11(τ)J12(τ)ˆwη(τ)J12(τ)ˆv).

    The Taylor series expansion of ˆF(τ,ˆw,ˆv) and ˆG(τ,ˆw,ˆv) shows that

    ˆF(τ,ˆw,ˆv)=j20ˆw2+j11ˆwˆv+j02ˆv2+j30ˆw3+j21ˆw2ˆv+j12ˆwˆv2+j03ˆv3+O(4),ˆG(τ,ˆw,ˆv)=k20ˆw2+k11ˆwˆv+k02ˆv2+k30ˆw3+k21ˆw2ˆv+k12ˆwˆv2+k03ˆv3+O(4),

    where

    j20=fww2+fwv(ζ(τ)J11(τ))J12(τ)+fvv2(ζ(τ)J11(τ)J12(τ))2,j11=fvvη(τ)(ζ(τ)J11(τ))J212(τ)fwvη(τ)J12(τ),j02=fvvη2(τ)2J212(τ),j12=fvvvη2(τ)(ζ(τ)J11(τ))2J312(τ)+fwvvη2(τ)2J212(τ),j03=fvvvη3(τ)6J312(τ),j21=fwwvη(τ)2J12(τ)fwvvη(τ)(ζ(τ)J11(τ))J212(τ)fvvvη(τ)(ζ(τ)J11(τ))22J312(τ),j30=fwww6+fwwv(ζ(τ)J11(τ))2J12(τ)+fwvv2(ζ(τ)J11(τ)J12(τ))2+fvvv6(ζ(τ)J11(τ)J12(τ))3,

    and kij=ζ(τ)J11(τ)η(τ)jijJ12(τ)η(τ)˜kij for i,j=0,1,2, with

    ˜k20=gww2+gwv(ζ(τ)J11(τ))J12(τ)+gvv2(ζ(τ)J11(τ)J12(τ))2,˜k11=gvvη(τ)(ζ(τ)J11(τ))J212(τ)gwvη(τ)J12(τ),˜k02=gvvη2(τ)2J212(τ),˜k12=gvvvη2(τ)(ζ(τ)J11(τ))2J312(τ)+gwvvη2(τ)2J212(τ),˜k03=gvvvη3(τ)6J312(τ),˜k21=gwwvη(τ)2J12(τ)gwvvη(τ)(ζ(τ)J11(τ))J212(τ)gvvvη(τ)(ζ(τ)J11(τ))22J312(τ),˜k30=gwww6+gwwv(ζ(τ)J11(τ))2J12(τ)+gwvv2(ζ(τ)J11(τ)J12(τ))2+gvvv6(ζ(τ)J11(τ)J12(τ))3.

    When τ=τH0, we define the first Lyapunov coefficient as follows:

    1(τH0)=i2η(τH0)(g20g112|g11|2|g02|23)+g212,

    where

    g11=14(j20+j02+i(k20+k02)),g02=14(j20j022k11+i(k20k02+2j11)),g20=14(j20j02+2k11+i(k20k022j11)),g21=18(j30+j12+k21+k03+i(k30+k12j21j03)),

    and where

    j20=fww2fwvJ11(τH0)J12(τH0)+fvv2(J11(τH0)J12(τH0))2,j11=fvvη(τH0)(J11(τH0)J212(τH0)fwvη(τH0)J12(τH0),j02=fvvη2(τH0)2J212(τH0),j12=fvvvη2(τH0)J11(τH0)2J312(τH0)+fwvvη2(τH0)2J212(τH0),j03=fvvvη3(τH0)6J312(τH0),j21=fwwvη(τH0)2J12(τH0)+fwvvη(τH0)J11(τH0)J212(τH0)fvvvη(τH0)J211(τH0)2J312(τH0),j30=fwww6fwwvJ11(τH0)2J12(τH0)+fwvv2(J11(τH0)J12(τH0))2fvvv6(J11(τH0)J12(τH0))3,

    and kij=J11(τH0)η(τH0)jijJ12(τH0)η(τH0)˜kij for i,j=0,1,2, with

    ˜k20=gww2gwvJ11(τH0)J12(τH0)+gvv2(J11(τH0)J12(τH0))2,˜k11=gvvη(τH0)J11(τH0)J212(τH0)gwvη(τH0)J12(τH0),˜k02=gvvη2(τH0)2J212(τH0),˜k12=gvvvη2(τH0)J11(τH0)2J312(τH0)+gwvvη2(τH0)2J212(τH0),˜k03=gvvvη3(τH0)6J312(τH0),˜k21=gwwvη(τH0)2J12(τH0)+gwvvη(τH0)J11(τH0)J212(τH0)gvvvη(τH0)J211(τH0)2J312(τH0),˜k30=gwww6gwwvJ11(τH0)2J12(τH0)+gwvv2(J11(τH0)J12(τH0))2gvvv6(J11(τH0)J12(τH0))3.

    By some direct calculations, one has

    Re(1(τH0))=18(3j30+j12+k21+3k03)+18η(τH0)(j11(j20+j02)+2j02k02)18η(τH0)(k11(k20+k02)+2k20j20).

    Now we summarize the analysis above and yield the following result.

    Theorem 2.2. Suppose that 0<a<1+s,0<s<1,a(1s)(1+s)2>w are true. Then, the Hopf bifurcation is supercritical (respectively, subcritical) when Re(1(τH0))<0 (respectively, Re(1(τH0))>0) and the bifurcating periodic solution is stable (respectively, unstable).

    In this subsection, we want to give an estimate with respect to the classic solution (w,v) of the reaction-diffusion system (1.3).

    Theorem 3.1. Suppose that 0<a<2s1+θ is fulfilled. Then the solution (w(x,t),v(x,t)) of the system (1.3) admits

    lim suptmaxx¯Ωw(,t)1,lim suptmaxx¯Ωv(,t)1.

    In addition, one obtains

    lim inftminx¯Ωw(,t)11+θa2s,lim inftminx¯Ωv(,t)11+θa2s.

    Proof. By employing the first equation of the system (1.3), we get

    {wtd1Δwww2,xΩ,t>0,wn=0,xΩ,t0,w(x,0)=w0(x)0,xΩ.

    As such, the comparison principle for the parabolic equations show that t11 and 0<ε11 exist such that w(x,t)1+ε1 for x¯Ω,tt1. Keeping this result in mind and utilizing the second equation of the system (1.3), one derives

    {vtd2Δvτv(1v1+ε1),xΩ,tt1,vn=0,xΩ,tt1,v(x,t1)0,xΩ.

    According to the comparison principle of parabolic equations, we get v(x,t)(1+ε1)+ε2 for x¯Ω,tt2. In the sequel, let us explore the lower-bounds of the classic solution (w(x,t),v(x,t)) of the system (1.3). By the way, the lower bounds of the classic solution (w(x,t),v(x,t)) imply that the system (1.3) is persistence. Using the first equation of the system (1.3) again, we obtain

    {wtd1Δww(11+θ((1+ε1)+ε2)wa2s),xΩ,t0,wn=0,xΩ,t0,w(x,0)0,xΩ.

    Thereby, ε3>0 and t3>0 exist such that

    w(x,t)11+θ((1+ε1)+ε2)a2s+ε3

    for x¯Ω,tt3. Finally, one can obtain

    {vtd2Δvrv(1v11+θ((1+ε1)+ε2)a2s+ε3),xΩ,tt3,vn=0,xΩ,tt3,v(x,t3)0,xΩ.

    This means that

    v(x,t)(11+θ((1+ε1)+ε2)a2s+ε3)+ε4,

    for x¯Ω,tt4. The proof is completed.

    Assume that the domain Ω takes the form Ω=(0,π). Let ˜w=ww, ˜v=vv and still denote ˜w and ˜v as w and v, respectively. As such, the system (1.3) has the form

    {wt=d1Δw+w+w1+θ(v+v)(w+w)2a(w+w)2(v+v)(w+w)2+s(v+v)2,xΩ,t>0,vt=d2Δv+τ(v+v)(1v+vw+w),xΩ,t>0,wn=vn=0,xΩ,t0,w(x,0)=w0(x)0,v(x,0)=v0(x)0,xΩ. (3.1)

    If we let N=(ww,vv)T, at the origin, the system (3.1) can be rewritten as

    Nt=L(τ)N+Q(τ,N), (3.2)

    where

    L(τ)=(J11(τ)+d1ΔJ12(τ)ττ+d2Δ)=(a(1s)(1+s)2w+d1Δa(s1)(1+s)2θw(1+θw)2ττ+d2Δ),

    and

    Q(τ,N)=(F(τ,w,v)G(τ,w,v)),

    where F(τ,w,v) and G(τ,w,v) have been defined in (2.6). Obviously, the local linearized system of (3.1) admits the form

    Nt=L(τ)N. (3.3)

    Now, consider the following solution of (3.3):

    N=m=0(ambm)eλtcosmx,

    where mN0={0,1,2,}, λ is the growth rate of perturbation, and am and bm are two nonzero constants. Substituting it into the linear system (3.3) yields

    m=0(AmλI)(ambm)cosmx=0,

    where I is the identity matrix and

    Am=(J11(τ)d1m2J12(τ)ττd2m2)=(a(1s)(1+s)2wd1m2a(s1)(1+s)2θw(1+θw)2ττd2m2).

    Consequently, the characteristic equation can be obtained by setting |λIAm|=0 for mN0. In this manner, we can get

    λ2Tm(τ)λ+Dm(τ)=0,mN0, (3.4)

    where

    {Tm(τ)=(d1+d2)m2+a(1s)(1+s)2wτ,Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2.

    To obtain the emergence conditions of Turing instability, we require one of Assumptions (C1)–(C3) in Theorem 2.1 is satisfied. This means that Tm(τ)<0 must hold for any mN0. Therefore, we only need to discuss the sign of Dm(τ) to explore the occurrence of Turing instability.

    We have the following.

    Theorem 3.2. Suppose that 0<a<1+s and one of Assumptions (C1)–(C3) in Theorem 2.1 is satisfied. The positive equilibrium E is locally asymptotically stable if one of the following conditions is satisfied:

    (C5) s1;

    (C6) 0<s<1,a(1s)(1+s)2w;

    (C7) 0<s<1,a(1s)(1+s)2>w,τ>a(1s)(1+s)2w,d2d1;

    (C8) 0<s<1,a(1s)(1+s)2>w,τ>a(1s)(1+s)2w,d2τa(1s)(1+s)2wd1;

    and

    (C9){0<s<1,a(1s)(1+s)2>w,τ>a(1s)(1+s)2w,d2>τa(1s)(1+s)2wd1,m2wτ+θτw(1+θw)2d1d2,[d2(a(1s)(1+s)2w)d1τ]24d1d2(wτ+θτw(1+θw)2)0.

    However, E is Turing unstable if

    (C10){0<s<1,a(1s)(1+s)2>w,τ>a(1s)(1+s)2w,d2>τa(1s)(1+s)2wd1,m2=wτ+θτw(1+θw)2d1d2,[d2(a(1s)(1+s)2w)d1τ]24d1d2(wτ+θτw(1+θw)2)>0.

    Proof. If s1 is true, then we have Dm(τ)d1d2m4+wτ+θτw(1+θw)2>0. So, all the eigenvalues of (3.4) possess negative real parts. This demonstrates that E is locally asymptotically stable. Similarly, if 0<s<1,a(1s)(1+s)2w, we also obtain Dm(τ)d1d2m4+wτ+θτw(1+θw)2>0. Hence, the equilibrium E is locally asymptotically stable. If (C7) is satisfied, we can deduce

    Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2d1d2m4d1[(a(1s)(1+s)2w)τ]m2+wτ+θτw(1+θw)2>0.

    Accordingly, the equilibrium E is locally asymptotically stable when (C7) is satisfied. If (C8) is satisfied, it is easy to analyze

    Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2d1d2m4+wτ+θτw(1+θw)2>0.

    Henceforth, the equilibrium E is locally asymptotically stable when (C8) is satisfied. By utilizing Condition (C9), we have

    0<d2(a(1s)(1+s)2w)d1τ2d1d2(wτ+θτw(1+θw)2).

    As such, we get

    Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2d1d2m42d1d2(wτ+θτw(1+θw)2)m2+wτ+θτw(1+θw)2=[d1d2m2wτ+θτw(1+θw)2]2>0, (3.5)

    where ">" holds since

    m2wτ+θτw(1+θw)2d1d2.

    Similarly, if Condition (C10) holds, we have

    d2(a(1s)(1+s)2w)d1τ>2d1d2(wτ+θτw(1+θw)2).

    Therefore, one can derive

    Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2<d1d2m42d1d2(wτ+θτw(1+θw)2)m2+wτ+θτw(1+θw)2=[d1d2m2wτ+θτw(1+θw)2]2=0, (3.6)

    where "=" holds owing to

    m2=wτ+θτw(1+θw)2d1d2.

    Benefitting from (3.6), one can see that Dm(τ)<0 is satisfied, whereas we have Tm(τ)<0. Hence, we can claim that there at least one eigenvalue of (3.4) with a positive real part. Consequently, the equilibrium E is unstable in the Turing sense. The proof is completed.

    For the Hopf bifurcation of the reaction-diffusion system (1.3), we can establish the following.

    Theorem 3.3. Suppose that 0<a<1+s is satisfied. In this case, reaction-diffusion system (1.3) undergoes Hopf bifurcation at the threshold τ=τHm as the following condition is true:

    (C11)0<s<1,a(1s)(1+s)2>w,d2τHmτH0d1,

    where

    τH0=a(1s)(1+s)2w,τHm=a(1s)(1+s)2w(d1+d2)m2=τH0(d1+d2)m2,mN0.

    Proof. Recall that

    {Tm(τ)=(d1+d2)m2+a(1s)(1+s)2wτ,Dm(τ)=d1d2m4[d2(a(1s)(1+s)2w)d1τ]m2+wτ+θτw(1+θw)2.

    By using the definition of τH0, we have

    {Tm(τ)=(d1+d2)m2+τH0τ,Dm(τ)=d1d2m4(d2τH0d1τ)m2+wτ+θτw(1+θw)2.

    Therefore, when τHm=τH0(d1+d2)m2 in (C11), one has Tm(τHm)=(d1+d2)m2+τH0τHm=0 for all mN0. Furthermore, if d2τHmτH0d1 holds, we get

    Dm(τHm)=d1d2m4(d2τH0d1τHm)m2+wτHm+θτHmw(1+θw)2d1d2m4+wτHm+θτHmw(1+θw)2>0.

    Hence, we can conclude that (3.4) must have a pair of purely imaginary roots. On the other hand, by a direct computation, one yields

    dRe{λ}dτ|τ=τHm=12<0,mN0.

    Consequently, the reaction-diffusion system (1.3) undergoes Hopf bifurcation when Condition (C11) is satisfied. We end the proof.

    Remark 3.1. (1) If m=0, the Hopf bifurcation is spatially homogeneous; however, if mN0{0}, the Hopf bifurcation will be spatially inhomogeneous. (2) For the onset of the two types Hopf bifurcation τHm and τH0, we have τHm<τHm1<<τH2<τH1<τH0.

    Now our task is to determine the direction of the Hopf bifurcation for the reaction-diffusion system (1.3). For convenience, we consider a Hopf bifurcation point τ=τH0. Rewrite the system (1.3) as follows:

    {Wt=L(τ)W+F(τ,W),Wx(0,t)=Wx(π,t)=(0,0)T,

    where W=(w,v)T and

    L(τ)=(J11(τ)+d1ΔJ12(τ)ττ+d2Δ),

    and the conjugate operator of L(τ) is given by

    L(τ)=(J11(τ)+d1ΔτJ12(τ)τ+d2Δ).

    In addition,

    F(τ,W)=(f(w,v)J11(τ)wJ12(τ)vg(w,v)τw+τv).

    When τ=τH0, we have the following critical system:

    {Wt=L(τH0)W+F(τH0,W),Wx(0,t)=Wx(π,t)=(0,0)T, (3.7)

    where

    L(τH0)=(J11(τH0)+d1ΔJ12(τH0)τH0τH0+d2Δ),

    and

    F(τH0,W)=(f(w,v)J11(τH0)wJ12(τH0)vg(w,v)τH0w+τH0v).

    Define a inner product

    φ,ϕ=1π×π0ˉφTϕdx.

    Let

    y=(y1y2)=(1J11(τ)J12(τ)+η(τH0)J12(τ)i),

    and

    y=(y1y2)=J12(τ)2πη(τH0)(η(τH0)J12(τ)+J11(τ)J12(τ)ii).

    By some direct computations, we can obtain L(τH0)W1,W2=W1,L(τ)W2,L(τH0)y=iη(τH0)y,L(τH0)y=iη(τH0)y,y,y=1, and y,¯y=0. In the following, we set Bc:=BiB={x1+ix2|x1,x2B}. In light of the existing literature [24], B=BcBs holds with Bc:={zy+¯zy|zC} and Bs={qB|y,q=0}, where z=y,W with W=(w,v)T. For any WB, zC and q=(q1,q2)TBs exist such that

    (wv)=zy+ˉzˉy+(q1q2).

    In this manner, we obtain

    {w=z+ˉz+q1,v=(J11(τ)J12(τ)+η(τH0)J12(τ)i)z+(J11(τ)J12(τ)η(τH0)J12(τ)i)ˉz+q2.

    Accordingly, we rewrite (3.7) as follows:

    {dzdt=iη(τH0)z+y,˜f,dqdt=L(τH0)q+T(z,¯z,q), (3.8)

    where

    ˜f=F(zy+ˉzˉy+q,τH0),T(z,¯z,q)=˜fy,˜fyˉy,˜f¯y.

    By employing [24], the nonlinear term F(τH0,W) in (3.7) can be read as:

    F(τH0,W)=12R(W,W)+16C(W,W,W)+O(|W|4),

    where R(W,W) and C(W,W,W) are symmetric multi linear forms and

    R(U,V)=(R1(U,V)R2(U,V)),C(U,V,Y)=(C1(U,V,Y)C2(U,V,Y)),

    where

    R1(U,V)=fwwu1v1+fwv(u1v2+u2v1)+fvvu2v2,R2(U,V)=gwwu1v1+gwv(u1v2+u2v1)+gvvu2v2,C1(U,V,Y)=fwwwu1v1y1+fwwv(u1v1y2+u1v2y1+u2v1y1)+fwvv(u1v2y2+u2v1y2+u2v2y1)+fvvvu2v2y2,C2(U,V,Y)=gwwwu1v1y1+gwwv(u1v1y2+u1v2y1+u2v1y1)+gwvv(u1v2y2+u2v1y2+u2v2y1)+gvvvu2v2y2,

    for U=(u1,u2)T,V=(v1,v2)T,Y=(y1,y2)T, and U,V,Y in H2([0,π])×H2([0,π]) with

    fww=2a(s3)w(s+1)32,fwv=2a(13s)w(s+1)3θ(θv+1)2,fwwv=6a(6ss21)w2(1+s)4,fwww=24a(1s)w2(1+s)4,fwvv=24as(s1)w2(1+s)4+2θ2(θv+1)3,fvv=2as(3s)w(1+s)3+2θ2w(θv+1)3,fvvv=6as(s26s+1)w2(1+s)46θ3w(θv+1)4,

    and

    gwv=2τH0w,gww=2τH0w,gwww=6τH0w2,gwwv=4τH0w2,gwvv=2τH0w2,gvv=2τH0w,gvvv=0.

    By some direct but complex calculations, one obtains

    ˜f=12R(y,y)z2+R(y,¯y)z¯z+12R(¯y,¯y)¯z2+O(|z|3,|z||q|,|q|2),y,˜f=12y,R(y,y)z2+y,R(y,¯y)z¯z+12y,R(¯y,¯y)ˉz2+O(|z|3,|z||q|,|q|2),ˉy,˜f=12ˉy,R(y,y)z2+ˉy,R(y,¯y)z¯z+12ˉy,R(¯y,¯y)ˉz2+O(|z|3,|z||q|,|q|2).

    So, we can obtain

    T(z,¯z,q)=12z2T20+z¯zT11+12¯z2T02+O(|z|3,|z||q|,|q|2),

    where

    T20=R(y,y)y,R(y,y)yˉy,R(y,y)¯y,T11=R(y,ˉy)y,R(y,¯y)yˉy,R(y,¯y)¯y,T02=R(¯y,¯y)ˉy,R(ˉy,¯y)yˉy,R(¯y,¯y)¯y.

    By direct calculations, we get

    T20=T11=T02=(0,0)T.

    This means

    T(z,¯z,q)=O(|z|3,|z||q|,|q|2).

    From [24], the system (3.8) has a center manifold and it could be written as

    q=12z2q20+z¯zq11+12¯z2q02+O(|z|3).

    Owing to

    Lq+T(z,¯z,q)=dqdt=qzdzdt+q¯zd¯zdt,

    it follows that

    q20=[2iη(τH0)L(τH0)]1T20=(0,0)T,q11=L1(τH0)T11=(0,0)T,q02=[2iη(τH0)L(τH0)]1T02=(0,0)T.

    In this fashion, we know that the reaction-diffusion system (1.3) can be restricted to a center manifold

    dzdt=iη(τH0)z+y,˜f=iη(τH0)z+2i+j3ρiji!j!zi¯zj+O(|z|4), (3.9)

    where

    ρ02=y,R(¯q,¯y),ρ20=y,R(y,y),ρ11=y,R(y,¯y),ρ21=2y,R(q11,y)+y,R(q20,¯y)+y,C(y,y,¯y)=y,C(y,y,¯y).

    In what follows, we rewrite system (3.9) in Poincaré normal form

    dzdt=(ζ(τ)+iη(τ))z+zNj=1δj(τ)(z¯z)j, (3.10)

    where ζ(τ) and η(τ) satisfy

    ζ(τH0)=0,η(τH0)=τH0(w+θw(1+θw)2)=(a(1s)(1+s)2w)(w+θw(1+θw)2)>0,

    and δj(τ) represents complex-valued coefficients. Then

    δ1(τ)=ρ20ρ11[3ζ(τ)+iη(τ)]2[ζ2(τ)+η2(τ)]+|ρ11|2ζ(τ)+iη(τ)+ρ212+|ρ02|22[ζ(τ)+3iη(τ)].

    Thereby, when τ=τH0, we have

    δ1(τH0)=ρ20ρ11i2η(τH0)+|ρ11|2iη(τH0)+ρ212+|ρ02|26iη(τH0)=i2η(τH0)(ρ20ρ112|ρ11|213|ρ02|2)+ρ212.

    It then follows that

    Re{δ1(τH0)}=12η(τH0)(Re{ρ20}Im{ρ11}+Im{ρ20}Re{ρ11})+12Re{ρ21}.

    A direct computation gives

    Re{ρ20}=fww2η2(τH0)+J211(τH0)2J212(τH0)fvv+gwvJ11(τH0)J12(τH0)gvv,Im{ρ20}=J11(τH0)2η(τH0)fww+η2(τH0)+J211(τH0)η(τH0)J12(τH0)fwvη2(τH0)J11(τH0)+J311(τH0)2η(τH0)J212(τH0)fvvJ12(τH0)2η(τH0)gww+J11(τH0)η(τH0)gwv+η2(τH0)J211(τH0)2J12(τH0)η(τH0)gvv,Re{ρ11}=fww2J11(τH0)J12(τH0)fwv+J211(τH0)+η2(τH0)2J212(τH0)fvv,
    Im{ρ11}=J11(τH0)2η(τH0)fwwJ11(τH0)η(τH0)J12(τH0)fwvη2(τH0)J11(τH0)+J311(τH0)2η(τH0)J212(τH0)fvvJ12(τH0)2η(τH0)gww+J11(τH0)η(τH0)gwvη(τH0)2+J211(τH0)2J12(τH0)η(τH0)gvv,Re{ρ21}=fwww2J11(τH0)J12(τH0)fwwv+J211(τH0)+η2(τH0)2J212(τH0)fwvv+gwwv2J11(τH0)J12(τH0)gwvv+gvvv2,

    where

    J11(τH0)=J11(τ)=a(1s)(1+s)2w,J11(τH0)=J12(τ)=a(s1)(1+s)2θw(1+θw)2.

    To summarize, we build the following.

    Theorem 3.4. Suppose that 0<a<1+s is satisfied. Then the Hopf bifurcation is supercritical (respectively subcritical) if 1ζ(τH0)Re{δ1(τH0)}<0(respectively>0). Meanwhile, the bifurcating periodic solution is stable (respectively unstable) if Re{δ1(τH0)}>0(respectively<0).

    In this section, we verify our previous theoretical results by using numerical computational experiments.

    First, we verify the theoretical results established in Theorem 2.1. We select the following parameters in the system (1.2): a=0.5,s=1.5,θ=0.25,and τ=1.0. Under these specific values, we can see that the assumption (C1) is satisfied, yielding the positive equilibrium E=(0.6586,0.6586), which is locally asymptotically stable, as shown in Figure 1(a). A similar stable result of the positive equilibrium E could be found in Figure 1(b), where we set a=0.5,s=1.5,θ=0.25,and τ=2.0 such that (C1) is fulfilled and the positive equilibrium E=(0.6586,0.6586) is locally asymptotically stable. In what follows, we choose the parameters a=0.5,s=0.5,θ=0.25,and τ=1.0 in the system (1.2). As a consequence, we obtain

    a(1s)(1+s)2=0.1111,w=0.5465.

    In this fashion, Assumption (C2) is satisfied. Our numerical experiment shows that the positive equilibrium E=(0.5465,0.5465) is locally asymptotically stable, as demonstrated in Figure 1(c). Now, taking a=0.5,s=0.125,θ=6.25,and τ=1.0, we obtain

    a(1s)(1+s)2=0.3457,w=0.1223.

    These mean that Assumption (C3) holds and the positive equilibrium E=(0.1223,0.1223) is locally asymptotically stable (refer to Figure 1(d)). These numerical results validate the theoretical predictions presented in Theorem 2.1.

    Figure 1.  Stability phase diagrams of the positive equilibrium E with different specific parameters. (a) When taking a=0.5,s=1.5,θ=0.25,and τ=1.0, the positive equilibrium E=(0.6586,0.6586) is locally asymptotically stable. (b) When choosing a=0.5,s=1.5,θ=0.25,and τ=2.0, the positive equilibrium E=(0.6586,0.6586) is locally asymptotically stable. (c) When setting a=0.5,s=0.5,θ=0.25,and τ=1.0, the positive equilibrium E=(0.5465,0.5465) is locally asymptotically stable. (d) When taking a=0.5,s=0.125,θ=6.25,and τ=1.0, the positive equilibrium E=(0.1223,0.1223) is locally asymptotically stable.

    To reveal the validity of Theorem 2.2, we set a=0.5,s=0.125,and θ=6.25, then obtain τH0=0.2234,E=(0.1223,0.1223), and Re(1(τH0))=0.1443<0. Our numerical result illustrates that there is a stable periodic solution around the positive equilibrium E=(0.1223,0.1223); see Figure 2(a). We also set the parameters a=0.35,s=0.14,and θ=7.25. As a result, we obtain τH0=0.0732,E=(0.1584,0.1584), and Re(1(τH0))=0.3673<0. This produces the stable bifurcating periodic solution shown in Figure 2(b).

    Figure 2.  Stable bifurcating periodic solutions due to the existence of supercritical Hopf bifurcation. (a) Taking a=0.5,s=0.125,θ=6.25,and τ=0.21; (b) taking a=0.35,s=0.14,θ=7.25,and τ=0.073.

    We now validate the theoretical results presented in Theorem 3.2 through numerical experiments.

    If we take the parameters a=1.0,s=1.5,θ=0.25,τ=1.0,d1=1.0,and d2=0.5, then Assumption (C5) is satisfied. It is found that the positive equilibrium E=(0.4907,0.4907) remains stable, as shown in Figure 3.

    Figure 3.  The positive equilibrium E=(0.4907,0.4907) is stable, where a=1.0,s=1.5,θ=0.25,τ=1.0,d1=1.0,and d2=0.5.

    When setting the parameters a=1.0,s=0.5,θ=0.25,τ=1.0,d1=1.0,and d2=0.5, one has

    a(1s)(1+s)2=0.2222,w=0.2701.

    Hence, Assumption (C6) is satisfied. It is demonstrated that the positive equilibrium E=(0.2701,0.2701) is stable, as shown in Figure 4.

    Figure 4.  The positive equilibrium E=(0.2701,0.2701) is stable, where a=1.0,s=0.5,θ=0.25,τ=1.0,d1=1.0,and d2=0.5.

    We set the parameters a=0.5,s=0.13,θ=6.5,τ=1.0,d1=1.0,and d2=1.0. We obtain

    a(1s)(1+s)2=0.3407,w=0.1198,

    implying that Assumption (C7) is satisfied. Our numerical experiments illustrate that the positive equilibrium E=(0.1198,0.1198) is stable, as depicted in Figure 5. When we set the parameters a=0.5,s=0.13,θ=6.5,and τ=1.0 in Figure 5, but choose d1=5.0 and d2=1.0, we have

    τa(1s)(1+s)2wd1=22.6355.

    All conditions in (C8) are satisfied, and numerical simulations demonstrate that the positive equilibrium E=(0.1198,0.1198) is stable; refer to Figure 6.

    Figure 5.  The positive equilibrium E=(0.1198,0.1198) is stable, where a=0.5,s=0.13,θ=6.5,τ=1.0,d1=1.0,and d2=1.0.
    Figure 6.  The positive equilibrium E=(0.1198,0.1198) is stable, where a=0.5,s=0.13,θ=6.5,τ=1.0,d1=5.0,and d2=1.0.

    If we set the parameters a=0.5,s=0.2,θ=4.25,τ=0.8,d1=0.35,and d2=5.85, we obtain

    a(1s)(1+s)2=0.2778,w=0.1675,τa(1s)(1+s)2wd1=2.5389,

    and

    [d2(a(1s)(1+s)2w)d1τ]24d1d2(wτ+θτw(1+θw)2)=0.13332.6890=2.5557<0.

    That is to say, all conditions in (C9) are satisfied. The numerical result demonstrates that the equilibrium E=(0.0355,0.0355) is stable; refer to Figure 7.

    Figure 7.  The positive equilibrium E=(0.0355,0.0355) is stable, where a=0.5,s=0.2,θ=4.25,τ=0.8,d1=0.35,and d2=5.85.

    When we set the parameters a=1.0,s=0.2,θ=4.25,τ=0.8,d1=0.35,and d2=5.85, we obtain

    a(1s)(1+s)2=0.5556,w=0.0355,τa(1s)(1+s)2wd1=0.5384,

    and

    [d2(a(1s)(1+s)2w)d1τ]24d1d2(wτ+θτw(1+θw)2)=7.62990.9793=6.6506>0.

    Accordingly, all conditions in (C10) are fulfilled. In this fashion, the equilibrium E=(0.0355,0.0355) becomes unstable in the Turing sense; see Figure 8. If we keep the same parameters in Figure 8 but only change the level of fear, i.e., we set θ=1.85, then we get

    a(1s)(1+s)2=0.5556,w=0.0627,τa(1s)(1+s)2wd1=3.8146,

    and

    [d2(a(1s)(1+s)2w)d1τ]24d1d2(wτ+θτw(1+θw)2)=2.13480.9963=1.1385>0.

    These data demonstrate the validity of (C10), leading to Turing instability; refer to Figure 9.

    Figure 8.  The positive equilibrium E=(0.0355,0.0355) has Turing instability, where a=1.0,s=0.2,θ=4.25,τ=0.8,d1=0.35,and d2=5.85.
    Figure 9.  The positive equilibrium E=(0.0627,0.0627) has Turing instability, where a=1.0,s=0.2,θ=1.85,τ=0.8,d1=0.35,and d2=5.85.

    Figures 39 validate the theoretical conclusions of Theorem 3.2.

    Finally, when setting the parameters a=1.0,s=0.3,θ=3.25,d1=2.35,and d2=0.85, one has E=(0.0623,0.0623),τH0=0.3519, and

    Re{ρ20}=18.5767,Im{ρ20}=22.0461,Re{ρ11}=17.2216,Im{ρ11}=13.7006,Re{ρ21}=794.3671.

    Therefore, we have \mbox{Re}\{\delta_{1}(\tau_0^H)\} = 1.5854e+03 > 0 . From Theorem 3.4, we have the supercritical type Hopf bifurcation, and the bifurcating periodic solution is stable. This prediction is validated by employing numerical simulation, as shown in Figure 10.

    Figure 10.  The bifurcating periodic solution is stable, where a = 1.0, \; s = 0.3, \; \theta = 3.25, \; d_1 = 2.35, \; d_2 = 0.85, \; \text{and}\ \tau = 0.3519 .

    Our findings demonstrate that all theoretical results have been rigorously validated through numerical experiments.

    In this paper, we consider a Leslie-Gower type predator-prey system with a Holling Type Ⅲ functional response and fear effects subject to no-flux boundary conditions. First, we analyze the stability of the unique positive equilibrium and the existence of Hopf bifurcation by considering the ranges of two parameters a and \tau for an ODE system (1.2), as formalized in Theorem 2.1. Furthermore, we rigorously demonstrate the occurrence of Hopf bifurcation at the critical threshold \tau = \tau_0^H and yield its direction; see Theorem 2.2. Next, we shift our focus to the reaction-diffusion system (1.3). An estimate of the classical solution and the stability results of positive equilibrium are outlined in Theorems 3.1 and 3.2, respectively. Notably, we obtain the precise occurrence of Turing instability so that we can observe the spatial pattern formation of the system. Subsequently, by designating \tau as the control parameter, we conclude that spatially inhomogeneous Hopf bifurcation will emerge in the system (1.3) when \tau = \tau_m^H for m\in\mathbb{N}_0 ; see Theorem 3.3. Finally, the normal formal theory of differential equations helps us characterize the supercritical and subcritical nature of the Hopf bifurcation, allowing explicit classification of the stability of bifurcating periodic solutions; see Theorem 3.4. Numerical results illustrate that the rightness of the previous theoretical analysis. The stable positive constant steady states (see Figures 17), the unstable positive constant steady states and pattern formations (see Figures 8 and 9), and the stable bifurcating periodic solutions (see Figure 10) are displayed. These results demonstrate that the predator-prey system with Holling Type Ⅲ functional response and fear effects exhibits rich and complex interaction dynamics.

    Xiaoyan Zhao: Formal analysis, writing-original draft, software, numerical computation, review and editing; Liangru Yu: Correspondence, review and editing, project administration; Xue-Zhi Li: Supervision, writing-original draft, methodology, review and editing, and project administration. 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 National Natural Science Foundation of China (No. 12271143).

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.



    [1] Q. Li, J. F. He, Pattern formation in a ratio-dependent predator-prey model with cross diffusion, Electron, Electron. Res. Arch., 31 (2023), 1106–1118. https://doi.org/10.3934/era.2023055 doi: 10.3934/era.2023055
    [2] J. D. Zhao, T. H. Zhang, Dynamics of two predator-prey models with power law relation, J. Appl. Anal. Comput., 13 (2023), 233–248. https://doi.org/10.11948/20220026 doi: 10.11948/20220026
    [3] N. C. Pati, B. Ghosh, Stability scenarios and period-doubling onset of chaos in a population model with delayed harvesting, Math. Methods Appl. Sci., 46 (2023), 12930–12945. https://doi.org/10.1002/mma.9223 doi: 10.1002/mma.9223
    [4] H. P. Jiang, Stable spatially inhomogeneous periodic solutions for a diffusive Leslie-Gower predator-prey model, J. Appl. Math. Comput., 70 (2024), 2541–2567. https://doi.org/10.1007/s12190-024-02018-2 doi: 10.1007/s12190-024-02018-2
    [5] W. J. Zuo, J. J. Wei, Stability and bifurcation in a ratio-dependent Holling-Ⅲ system with diffusion and delay, Nonlinear Anal.-Model., 19 (2014), 132–153. https://doi.org/10.15388/NA.2014.1.9 doi: 10.15388/NA.2014.1.9
    [6] X. X. Liu, L. H. Huang, Permanence and periodic solutions for a diffusive ratio-dependent predator-prey system, Appl. Math. Model., 33 (2009), 683–691. https://doi.org/10.1016/j.apm.2007.12.002 doi: 10.1016/j.apm.2007.12.002
    [7] V. A. Gaiko, C. Vuik, Global dynamics in the Leslie-Gower model with the Allee effect, Int. J. Bifurcat. Chaos, 28 (2018), 1850151. https://doi.org/10.1142/S0218127418501511 doi: 10.1142/S0218127418501511
    [8] M. X. He, Z. Li, Dynamics of a Lesile-Gower predator-prey model with square root response function and generalist predator, Appl. Math. Lett., 157 (2024), 109193. https://doi.org/10.1016/j.aml.2024.109193 doi: 10.1016/j.aml.2024.109193
    [9] Z. L. Li, Y. Zhang, Dynamic analysis of a fast slow modified Leslie-Gower predator-prey model with constant harvest and stochastic factor, Math. Comput. Simulat., 226 (2024), 474–499. https://doi.org/10.1016/j.matcom.2024.07.027 doi: 10.1016/j.matcom.2024.07.027
    [10] H. B. Shi, Y. Li, Global asymptotic stability of a diffusive predator-prey model with ratio-dependent functional response, Appl. Math. Comput., 250 (2015), 71–77. https://doi.org/10.1016/j.amc.2014.10.116 doi: 10.1016/j.amc.2014.10.116
    [11] H. B. Shi, S. G. Ruan, Y. Su, J. F. Zhang, Spatiotemporal dynamics of a diffusive Leslie-Gower predator-prey model with ratio-dependent functional response, Internat, Int. J. Bifurcat. Chaos, 25 (2015), 1530014. https://doi.org/10.1142/S0218127415300141 doi: 10.1142/S0218127415300141
    [12] D. Tripathi, J. P. Tripathi, S. K. Tiwari, D. Jana, L. F. Hou, Y. Shi, et al., Modified Holling Tanner diffusive and non-diffusive predator-prey models: The impact of prey refuge and fear effect, Results Phys., 65 (2024), 107995. https://doi.org/10.1016/j.rinp.2024.107995 doi: 10.1016/j.rinp.2024.107995
    [13] S. A. A. Hamdallah, A. A. Arafa, Stability analysis of Filippov prey-predator model with fear effect and prey refuge, J. Appl. Math. Comput., 70 (2024), 73–102. https://doi.org/10.1007/s12190-023-01934-z doi: 10.1007/s12190-023-01934-z
    [14] A. Kumar, K. P. Reshma, P. S. Harine, Global dynamics of an ecological model in presences of fear and group defense in prey and Allee effect in predator, Nonlinear Dynam., 113 (2025), 7483–7518. https://doi.org/10.1007/s11071-024-10706-8 doi: 10.1007/s11071-024-10706-8
    [15] J. Z. Cao, F. Li, P. M. Hao, Bifurcation analysis of a diffusive predator-prey model with fear effect, Math. Method. Appl. Sci., 47 (2024), 13404–13423. https://doi.org/10.1002/mma.10198 doi: 10.1002/mma.10198
    [16] S. Mishra, R. Upadhyay, Spatial pattern formation and delay induced destabilization in predator-prey model with fear effect, Math. Method. Appl. Sci., 45 (2022), 6801–6823. https://doi.org/10.1002/mma.8207 doi: 10.1002/mma.8207
    [17] J. Liu, Y. Kang, Spatiotemporal dynamics of a diffusive predator-prey model with fear effect, Nonlinear Anal.-Model., 27 (2022), 841–862. https://doi.org/10.15388/namc.2022.27.27535 doi: 10.15388/namc.2022.27.27535
    [18] H. K. Qi, X. Z. Meng, T. Hayat, A. Hobiny, Influence of fear effect on bifurcation dynamics of predator-prey system in a predator-poisoned environment, Qual. Theor. Dyn. Syst., 21 (2022), 27. https://doi.org/10.1007/s12346-021-00555-w doi: 10.1007/s12346-021-00555-w
    [19] A. Mondal, A. K. Pal, Age-selective harvesting in a delayed predator-prey model with fear effect, Z. NATURFORSCH. A., 77 (2022), 229–248. https://doi.org/10.1515/zna-2021-0217 doi: 10.1515/zna-2021-0217
    [20] H. T. Wang, Y. Zhang, L. Ma, Bifurcation and stability of a diffusive predator-prey model with the fear effect and time delay, Chaos, 33 (2023), 073137. https://doi.org/10.1063/5.0157410 doi: 10.1063/5.0157410
    [21] X. W. Ju, Y. Yang, Turing instability of the periodic solution for a generalized diffusive Maginu model, Comput. Appl. Math., 41 (2022), 290. https://doi.org/10.1007/s40314-022-01992-2 doi: 10.1007/s40314-022-01992-2
    [22] X. Y. Meng, N. N. Qin, H. F. Huo, Dynamics analysis of a predator-prey system with harvesting prey and disease in prey species, J. Biol. Dynam., 12 (2018), 342–374. https://doi.org/10.1080/17513758.2018.1454515 doi: 10.1080/17513758.2018.1454515
    [23] D. Jin, R. Z. Yang, Hopf bifurcation of a predator-prey model with memory effect and intra-species compitition in predator, J. Appl. Anal. Comput., 13 (2023), 1321–1335. https://doi.org/10.11948/20220127 doi: 10.11948/20220127
    [24] B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and applications of Hopf bifurcation, New York: Cambridge University Press, 1981. https://doi.org/10.1137/1024123
    [25] Y. T. Cai, C. C. Wang, D. J. Fan, Bifurcation analysis of a predator-prey model with age structure, Int. J. Bifurcat. Chaos, 30 (2020), 2050114. https://doi.org/10.1142/S021812742050114X doi: 10.1142/S021812742050114X
    [26] Q. An, X. Y. Gu, X. B. Zhang, Normal form and Hopf bifurcation for the memory-based reaction-diffusion equation with nonlocal effect, Math. Method. Appl. Sci., 47 (2024), 12883–12904. https://doi.org/10.1002/mma.10185 doi: 10.1002/mma.10185
    [27] Y. H. Qian, M. R. Ren, H. L. Wang, Dynamic behavior of a class of predator-prey model with two time delays, Acta Mech., 235 (2024), 7453–7473. https://doi.org/10.1007/s00707-024-04111-w doi: 10.1007/s00707-024-04111-w
  • Reader Comments
  • © 2025 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(265) PDF downloads(33) Cited by(0)

Figures and Tables

Figures(10)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog