Processing math: 100%
Research article

Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and generalist predator

  • Received: 22 February 2022 Revised: 24 April 2022 Accepted: 04 May 2022 Published: 17 May 2022
  • MSC : 34K18, 35B32

  • A delayed diffusive predator-prey system with nonlocal competition and generalist predators is considered. The local stability of the positive equilibrium and Hopf bifurcation at positive equilibrium is studied by using time delay as a parameter. In addition, the property of Hopf bifurcation is analyzed using the center manifold theorem and normal form method. It is determined that time delays can affect the stability of the positive equilibrium and induce spatial inhomogeneous periodic oscillation of prey and predator population densities.

    Citation: Chenxuan Nie, Dan Jin, Ruizhi Yang. Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and generalist predator[J]. AIMS Mathematics, 2022, 7(7): 13344-13360. doi: 10.3934/math.2022737

    Related Papers:

    [1] Ruizhi Yang, Dan Jin, Wenlong Wang . A diffusive predator-prey model with generalist predator and time delay. AIMS Mathematics, 2022, 7(3): 4574-4591. doi: 10.3934/math.2022255
    [2] Yingyan Zhao, Changjin Xu, Yiya Xu, Jinting Lin, Yicheng Pang, Zixin Liu, Jianwei Shen . Mathematical exploration on control of bifurcation for a 3D predator-prey model with delay. AIMS Mathematics, 2024, 9(11): 29883-29915. doi: 10.3934/math.20241445
    [3] 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
    [4] Xin-You Meng, Fan-Li Meng . Bifurcation analysis of a special delayed predator-prey model with herd behavior and prey harvesting. AIMS Mathematics, 2021, 6(6): 5695-5719. doi: 10.3934/math.2021336
    [5] 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
    [6] Qinghui Liu, Xin Zhang . Chaos detection in predator-prey dynamics with delayed interactions and Ivlev-type functional response. AIMS Mathematics, 2024, 9(9): 24555-24575. doi: 10.3934/math.20241196
    [7] Hairong Li, Yanling Tian, Ting Huang, Pinghua Yang . Hopf bifurcation and hybrid control of a delayed diffusive semi-ratio-dependent predator-prey model. AIMS Mathematics, 2024, 9(10): 29608-29632. doi: 10.3934/math.20241434
    [8] 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
    [9] Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar . Sex-biased predation and predator intraspecific competition effects in a prey mating system. AIMS Mathematics, 2024, 9(1): 2435-2453. doi: 10.3934/math.2024120
    [10] Liye Wang, Wenlong Wang, Ruizhi Yang . Stability switch and Hopf bifurcations for a diffusive plankton system with nonlocal competition and toxic effect. AIMS Mathematics, 2023, 8(4): 9716-9739. doi: 10.3934/math.2023490
  • A delayed diffusive predator-prey system with nonlocal competition and generalist predators is considered. The local stability of the positive equilibrium and Hopf bifurcation at positive equilibrium is studied by using time delay as a parameter. In addition, the property of Hopf bifurcation is analyzed using the center manifold theorem and normal form method. It is determined that time delays can affect the stability of the positive equilibrium and induce spatial inhomogeneous periodic oscillation of prey and predator population densities.



    Dynamic systems are widely used in nature, such as in infectious disease models [1,2], vegetation-water models [3], and population models [4,5]. For example, Khan et al. [6] studied the dynamics of pine wilt disease with variable population sizes and showed control strategies for the possible elimination of the infection in the pine tree population. Khan et al. [7] obtained an efficient iterated homotopy perturbation transform method (IHPTM) to solve a mathematical model of HIV infection of CD4+ T cells. Khan et al. [8] proposed a new method to solve partial differential equations arising in the fields of science and engineering. Among the applications of dynamical systems, the predator-prey model is an important research topic [9,10,11,12]. Generally, predator-prey models assume that the prey is the only food source of the predator [13,14,15].

    In ecosystems, predators are usually generalist predators. They feed on many types of species and can change their diet to another species when a focal prey population begins to run short [16,17,18,19]. In [17], the authors studied a predator-prey model with a generalist predator. They aimed to achieve biological control by generalists [17]. In [18], the authors studied the spatial pattern formation of a predator-prey model with generalist predator and the harvesting, refuge effect. To better understand the relationship between prey and generalist predators, Upadhyay and Agrawal [19] modified the Leslie-Gower model as

    dvdt=cv2ω1v2u+D1,

    to describe the growth law of generalist predators. The term dv2 represents the growth of predators, where d is the mating rate of the predator. e is another source of food for predators. They considered the following model:

    {dudt=ru(1uK)ωuvA+Bu+v,dvdt=v(dvω1v(tτ)u(tτ)+D1). (1.1)

    u(t) and v(t) represent the prey and predator densities, respectively. The term ωuA+Bu+v represents is the Beddington-DeAngelis functional response. τ is the gestation delay of the predator. Upadhyay and Agrawal studied the invariance, boundedness, and local and global stability and Hopf bifurcation.

    In the real world, the spatial distribution of populations is often inhomogeneous, so it is more practical to use a reaction-diffusion model to describe the relationship between predators and prey. In addition, the reaction-diffusion predator-prey model shows more abundant dynamic properties such as spatial patterns and spatial nonhomogeneous periodic solutions. Considering this factor, Liu et al. [20] proposed the following model with the Crowley-Martin type functional response based on the model (1.1).

    {u(x,t)t=d1Δu+ru(1uK)auv(1+bu)(1+cv),v(x,t)t=d2Δv+v(dvsv(tτ)u(tτ)+e),xΩ,t>0u(x,t)ˉν=v(x,t)ˉν=0,xΩ,t>0u(x,θ)=u0(x,θ)0,v(x,θ)=v0(x,θ)0,xˉΩ,θ[τ,0]. (1.2)

    In the model (1.2), Liu et al. [20] used the Crowley-Martin-type functional response to reflect the impact of predators on prey. This considered the effect of interference among predators. a, b, and c represent the capture rate, handling time, and magnitude of interference among predators, respectively. They mainly analyzed the instability and Hopf bifurcation induced by time delay [20]. Although they pointed out that time delay may cause spatially inhomogeneous periodic solutions, the numerical simulations did not show stably spatially inhomogeneous periodic solutions. This is because in the delayed reaction-diffusion predator-prey model, the spatial nonhomogeneous Hopf bifurcation curve is usually above the spatial homogeneous Hopf bifurcation curve. In addition, Turing instability cannot occur for the model (1.2). For model (1.2), it is unfortunate that the inhomogeneous distribution of prey and predator in space is not shown. This may be due to the lack of stable spatially inhomogeneous periodic solutions in the delayed reaction-diffusion predator-prey model.

    In nature, competition within populations exists widely, and this competition is often nonlocal since resources are limited. In [21,22], the authors suggested that the internal competition of the population caused by the natural environment is related not only to the population density at the current location but also to the population density nearby. They measured this effect by weighting and integrating, and modified the uK as 1KΩG(x,y)u(y,t)dy. G(x,y) is a kernel function. Wu and Song studied Hopf and state-Hopf bifurcations in a diffusive predator-prey model with a nonlocal effect and delay [23]. Geng et al. studied a diffusive predator-prey model with nonlocal competition, including Hopf, Turing, double-Hopf, and Turing-Hopf bifurcations [24]. The works in [25,26] show that stable spatially inhomogeneous periodic solutions often exist in predator-prey models with nonlocal competition.

    Inspired by the above work, we assume there is nonlocal competition in prey and modify the model (1.3) as follows.

    {u(x,t)t=d1Δu+ru(11KΩG(x,y)u(y,t)dy)auv(1+bu)(1+cv),v(x,t)t=d2Δv+v(dvsv(tτ)u(tτ)+e),xΩ,t>0u(x,t)ˉν=v(x,t)ˉν=0,xΩ,t>0u(x,θ)=u0(x,θ)0,v(x,θ)=v0(x,θ)0,xˉΩ,θ[τ,0]. (1.3)

    where d1 and d2 represent the diffusion coefficients of prey and predator, respectively. The term au(1+bu)(1+cv) is the Crowley-Martin-type functional response. Ω is the prey and predator's living region. Just for the convenience of calculation, we choose Ω=(0,lπ), where l>0. The boundary condition is the Newman boundary condition, which means that the living region is closed and no prey and predator enter or leave the region. ΩG(x,y)u(y,t)dy represents the nonlocal competition effect, and G(x,y) is the kernel function. We assume that the competition strength among prey individuals in the habitat is the same; we choose G(x,y)=1lπ as in previous works [23,24,25,26].

    To our knowledge, there is no work about the predator-prey model (1.3) with nonlocal competition in prey and generalist predators from the point of Hopf bifurcation. The aim of this paper is to study the combined effect of time delay and nonlocal competition on model (1.3). Compared with the model (1.2), do new dynamics appear, such as stably spatially inhomogeneous periodic solutions and Turing instability?

    This paper is organized as follows: the stability and existence of Hopf bifurcation are studied in Section 2, the property of Hopf bifurcation is analyzed in Section 3, and numerical simulations are given in Section 4. Finally, a short conclusion is presented.

    (0,0) and (K,0) are boundary equilibria of system (1.3). The existence of positive equilibria of system (1.3) was studied in [20], that is,

    Lemma 2.1. [20]If the following condition

    (H0)de<s<d(K+e),a>cr(bs+dbde)(de+dKs)d2K, (2.1)

    holds, then the model (1.3) has a coexisting equilibrium point (u,v), where

    u=sdedandv=r(bu+1)(Ku)aKcr(bu+1)(Ku).

    Next, we just denote E(u,v) as a coexisting equilibrium. Linearize system (1.3) at E(u,v)

    ut(u(x,t)u(x,t))=D(Δu(t)Δv(t))+L1(u(x,t)v(x,t))+L2(u(x,tτ)v(x,tτ))+L3(ˆu(x,t)ˆv(x,t)), (2.2)

    where

    D=(d100d2),L1=(a1a20b1),L2=(00b2b1),L3=(a3000),

    and

    a1=abuv(1+bu)2(1+cv)>0,a2=au(1+bu)(1+cv)2>0,a3=ruK>0,
    b1=dv>0,b2=sv2(e+u)2>0,ˆu=1lπlπ0u(y,t)dy.

    The characteristic equation is

    λ2+Anλ+Bn+(Cn+b1λ)eλτ=0,nN0, (2.3)

    where

    A0=a3a1b1,B0=(a1a3)b1,C0=a1b1+a3b1+a2b2,An=(d1+d2)n2l2a1b1,Bn=d1d2n4l4(a1d2+b1d1)n2l2+a1b1,Cn=b1d1n2l2a1b1+a2b2,nN. (2.4)

    When τ=0, the characteristic equation (2.3) is

    λ2+(An+b1)λ+Bn+Cn=0,nN0, (2.5)

    where

    {A0+b1=a3a1,B0+C0=a2b2,An+b1=(d1+d2)n2l2a1,Bn+Cn=d1d2n4l4a1d2n2l2+a2b2,nN. (2.6)

    Make the following hypothesis:

    (H1)An+b1>0,Bn+Cn>0,fornN0,(H2)a3a1>0Ak+b1<0,(orBk+Ck<0),for somekN.

    Theorem 2.2. For system (1.3), assume τ=0 and (H0) holds. Then, E(u,v) is locally asymptotically stable under (H1) and is Turing unstable under (H2).

    Proof. If (H1) holds, then we can determine that the characteristic roots of (2.5) all have negative real parts. Then, E(u,v) is locally asymptotically stable. If (H1) holds, then the characteristic roots of (2.5) with kN have at least one positive real part, but with n=0 all having a negative real part. This implies that E(u,v) is Turing unstable.

    Lemma 2.3. When τ=0, the Turing instability of E(u,v) cannot occur for the system (1.2), which lacks the nonlocal competition term.

    Proof. When τ=0, for the system (1.3) without nonlocal competition, the characteristic equation (2.3) is

    λ2+Tnλ+Dn=0,nN0, (2.7)

    where

    Tn=(d1+d2)n2l2(a1+ˆa),
    Dn=d1d2n4l4d2(a1+ˆa)n2l2+a2b2.

    Assume E(u,v) is locally asymptotically stable in the absence of spatial diffusion, which means that a1+ˆa<0. However, Tn>0 and Dn>0 when a1+ˆa<0 holds. Then, the Turing instability of E(u,v) cannot occur.

    Let iω (ω>0) be a solution of Eq (2.3); then,

    ω2+iωAn+Bn+(Cn+b1iω)(cosωτisinωτ)=0.

    We can obtain

    cosωτ=ω2(Cnb1An)BnCnC2n+d2ω2,sinωτ=ω(AnCnBnb1+b1ω2)C2n+b21ω2.

    It leads to

    ω4+ω2(A2n2Bnb21)+B2nC2n=0. (2.8)

    Let z=ω2; then, (2.8) becomes

    z2+z(A2n2Bnb21)+B2nC2n=0, (2.9)

    and the roots of (2.9) are

    z±=12[Pn±P2n4QnRn],

    where Pn=A2n2Bnb21, Qn=Bn+Cn, and Rn=BnCn. If (H0) and (H1) hold, Qn>0(nN0). By direct calculation, we have

    P0=(a1a3)2>0,R0=2(a3a1)b1a2b2<0Pk=(a1d1n2l2)2+d2n2l2(d2n2l22b1),Rk=d1d2k4l4(2b1d1+a1d2)k2l2+2a1b1a2b2,forkN.

    Define

    W1={n|Rn<0,nN0},W2={n|Rn>0,Pn<0,P2n4QnRn>0,nN},W3={n|Rn>0,P2n4QnRn<0,nN},

    and

    ω±n=z±n,τj,±n={1ω±narccos(V(n,±)cos)+2jπ,V(n,±)sin0,1ω±n[2πarccos(V(n,±)cos)]+2jπ,V(n,±)sin<0.V(n,±)cos=(ω±n)2(Cnb1An)BnCnC2n+b21(ω±n)2,V(n,±)sin=ω±n(AnCnBnb1+b1(ω±n)2)C2n+b21(ω±n)2. (2.10)

    We have the following lemma.

    Lemma 2.4. Assuming that (H0) and (H1) hold, the following results hold:

    (1) Eq (2.3) has a pair of purely imaginary roots±iω+n at τj,+n for jN0 and nW1.

    (2) Eq (2.3) has two pairs of purely imaginary roots±iω±n at τj,±n for jN0 and nW2.

    (3) Eq (2.3) has no purely imaginary root for nW3.

    Lemma 2.5. Assume (H0) and (H1) hold. Then, Re(dλdτ)|τ=τj,+n>0, Re(dλdτ)|τ=τj,n<0 for nW1W2 and jN0.

    Proof. By (2.3), we have

    (dλdτ)1=2λ+An+b1eλτ(Cn+b1λ)λeλττλ.

    Then

    [Re(dλdτ)1]τ=τj,±n=Re[2λ+An+b1eλτ(Cn+b1λ)λeλττλ]τ=τj,±n=[1C2n+b21ω2(2ω2+A2n2Bnb21)]τ=τj,±n=±[1C2n+b21ω2(A2n2Bnb21)24(B2nC2n)]τ=τj,±n.

    Therefore, Re(dλdτ)|τ=τj,+n>0, Re(dλdτ)|τ=τj,n<0.

    Denote τ=min{τ0n|nW1W2}. We have the following theorem:

    Theorem 2.6. Assume that (H0) and (H1) hold. Then, the following statements are true for system (1.3):

    (1) E(u,v) is locally asymptotically stable for τ>0 when W1W2=.

    (2) E(u,v) is locally asymptotically stable for τ[0,τ) when W1W2.

    (3) E(u,v) is unstable for τ(τ,τ+ε) for some ε>0 when W1W2.

    (4) Hopf bifurcation occurs at(u,v) when τ=τj,+n(τ=τj,n), jN0, nW1W2.

    In [27,28], we studied the property of Hopf bifurcation. For fixed jN0 and nW1W2, we denote ˜τ=τj,±n. Let ˉu(x,t)=u(x,τt)u and ˉv(x,t)=v(x,τt)v. Drop the bar, and (1.3) can be written as

    (3.1)

    We rewrite system (4) as follows:

    {ut=τ[d1Δu+a1ua2va3ˆu+α1u2rKuˆu+α2uv+α3v2+α4u3+α5u2v+α6uv2+α7v3]+h.o.t.,vt=τ[d2Δv+b1v+b2u(t1)b1v(t1)+dv2+β1u2(t1)+β2u(t1)v(t1)+β3u2(t1)+β4u(t1)v+β5v(t1)v+β6u3(t1)+β7u2(t1)v(t1)+β8u(t1)v2(t1)+β9v3(t1)+β10u2(t1)v+β11v2(t1)v]+h.o.t., (3.2)

    where

    α1=abv(1+bu)3(1+cv),α2=a(1+bu)2(1+cv)2,α3=acu(1+bu)(1+cv)3
    α4=ab2v(1+bu)4(1+cv),α5=ab(1+bu)3(1+cv)2,
    α6=ac(1+bu)2(1+cv)3,α7=ac2u(1+bu)(1+cv)4,
    β1=sv2(e+u)3,β2=sv(e+u)2,β3=0,β4=sv(e+u)2,β5=se+u,
    β6=sv2(e+u)4,β7=sv(e+u)3,β8=0,β9=0,β10=sv(e+u)3,β11=0.

    Define the real-valued Sobolev space as

    X:={(u,v)T:u,vH2(0,lπ),(ux,vx)|x=0,lπ=0},

    the complexification of XXC:=XiX={x1+ix2|x1,x2X}. and the inner product

    <˜u,˜v>:=lπ0¯u1v1dx+lπ0¯u2v2dx

    for ˜u=(u1,u2)T, ˜v=(v1,v2)T, ˜u, ˜vXC. The phase space C:=C([1,0],X) has the sup norm. Then, we can write ϕtC, ϕt(θ)=ϕ(t+θ) or 1θ0. Denote β(1)n(x)=(γn(x),0)T, β(2)n(x)=(0,γn(x))T, and βn={β(1)n(x),β(2)n(x)}, where {β(i)n(x)} is an an orthonormal basis of X. We define the subspace of C as

    Bn:=span{<ϕ(),β(j)n>β(j)n|ϕC,j=1,2},nN0.

    There exists a 2×2 matrix function ηn(σ,˜τ)1σ0 such that

    ˜τDn2l2ϕ(0)+˜τL(ϕ)=01dηn(σ,τ)ϕ(σ)

    for ϕC. The bilinear form on C×C is defined by

    (ψ,ϕ)=ψ(0)ϕ(0)01σξ=0ψ(ξσ)dηn(σ,˜τ)ϕ(ξ)dξ, (3.3)

    for ϕC, ψC. Defining τ=˜τ+μ, the system undergoes a Hopf bifurcation at (0,0) when μ=0, with a pair of purely imaginary roots ±iωn0. Let A denote the infinitesimal generators of the semigroup, and A be the formal adjoint of A under the bilinear from (3.3). Define the following function:

    δ(n0)={1n0=0,0n0N. (3.4)

    Choose ηn0(0,˜τ)=˜τ[(n20/l2)D+L1+L3δ(nn0)], ηn0(1,˜τ)=˜τL2, ηn0(σ,˜τ)=0 for 1<σ<0. Let

    p(θ)=p(0)eiωn0˜τθ(θ[1,0]),
    q(ϑ)=q(0)eiωn0˜τϑ(ϑ[0,1])

    be the eigenfunctions of A(˜τ) and A corresponding to iωn0˜τ, respectively. We can choose p(0)=(1,p1)T, q(0)=M(1,q2), where

    p1=1a2(a1d1n2l2a3δn0iωn0),
    q2=a2/(b1b1eiωn0˜τd2n2/l2iωn0),

    and

    M=(1+p1q2+(b2q2b1p1q2)τeiωn0˜τ)1.

    Then, (4) can be rewritten in abstract form as

    dU(t)dt=(˜τ+μ)DΔU(t)+(˜τ+μ)[L1(Ut)+L2U(t1)+L3ˆU(t)]+F(Ut,ˆUt,μ), (3.5)

    where

    F(ϕ,μ)=(˜τ+μ)(α1ϕ1(0)2rKϕ1(0)ˆϕ1(0)+α2ϕ1(0)ϕ2(0)+α3ϕ2(0)2+α4ϕ31(0)+α5ϕ21(0)ϕ2(0)+α6ϕ1(0)ϕ22(0)+α7ϕ32(0)dϕ22(0)+β1ϕ21(1)+β2ϕ1(1)ϕ2(1)+β3ϕ22(1)+β4ϕ1(1)ϕ2(0)+β5ϕ2(1)ϕ2(0)+β6ϕ31(1)+β7ϕ21(1)ϕ2(1)+β8ϕ1(1)ϕ22(1)+β9ϕ32(1)+β10ϕ21(1)ϕ2(0)+β11ϕ22(1)ϕ2(0)) (3.6)

    respectively, for ϕ=(ϕ1,ϕ2)TC and ˆϕ1=1lπlπ0ϕdx. Then, the space C can be decomposed as C=PQ, where

    P={zpγn0(x)+ˉzˉpγn0(x)|zC},
    Q={ϕC|(qγn0(x),ϕ)=0and(ˉqγn0(x),ϕ)=0}.

    Then, system (3.6) can be rewritten as

    Ut=z(t)p()γn0(x)+ˉz(t)ˉp()γn0(x)+ω(t,)

    and ^Ut=1lπlπ0Utdx, where

    z(t)=(qγn0(x),Ut),ω(t,θ)=Ut(θ)2Re{z(t)p(θ)γn0(x)}. (3.7)

    Then, we have

    ˙z(t)=iω)n0˜τz(t)+ˉq(0)<F(0,Ut),βn0>.

    There exists a center manifold C0, and ω can be written as follows near (0,0):

    ω(t,θ)=ω(z(t),ˉz(t),θ)=ω20(θ)z22+ω11(θ)zˉz+ω02(θ)ˉz22+. (3.8)

    Then, restrict the system to the center manifold as ˙z(t)=iωn0˜τz(t)+g(z,ˉz). Denoting

    g(z,ˉz)=g20z22+g11zˉz+g02ˉz22+g21z2ˉz2+.

    By direct computation, we have

    g20=2˜τM(ς1+q2ς2)I3,g11=˜τM(ϱ1+q2ϱ2)I3,g02=ˉg20,
    g21=2˜τM[(κ11+q2κ21)I2+(κ12+q2κ22)I4],

    where I2=lπ0γ2n0(x)dx, I3=lπ0γ3n0(x)dx, I4=lπ0γ4n0(x)dx, and

    ς1=α1rδn0K+ξ(α2+α3ξ),ς2=e2iωn0˜τ(β1+ξ(β2+de2iωn0˜τξ+β3ξ+eiωn0˜τ(β4+β5ξ))),ϱ1=r2Kδn0+14(2α1+2α3ˉξξ+α2(ˉξ+ξ)),ϱ2=14eiωn0˜τ(e2iωn0˜τ(β4+β5ˉξ)ξ+ˉξ(β4+β5ξ)+eiωn0˜τ(2β1+2(d+β3)ˉξξ+β2(ˉξ+ξ))),κ11=2ω(1)11(0)(rK(1+δn0)2α1α2ξ)+2ω(2)11(0)(α2+2α3ξ)+ω(1)20(0)(rK(1+δn0)+(2α1+α2ˉξ))+ω(2)20(0)(α2+2α3ˉξ),κ12=12(3α4+α5(ˉξ+2ξ)+ξ(2α6ˉξ+α6ξ+3α7ˉξξ)),
    κ21=2eiωn0˜τω(2)11(0)(β4+(2deiωn0˜τ+β5)ξ)+ω(2)20(0)(2dˉξ+eiωn0˜τ(β4+β5ˉξ))+2eiωn0˜τω(1)11(1)(2β1+(β2+eiωn0˜τβ4)ξ)+2eiωn0˜τω(2)11(1)(β2+2β3ξ+eiωn0˜τβ5ξ)+ω(1)20(1)(β4ˉξ+eiωn0˜τ(2β1+β2ˉξ))+ω(2)20(1)(β5ˉξ+eiωn0˜τ(β2+2β3ˉξ)),κ22=12e2iωn0˜τ(2e2iωn0˜τξ(β10+β11ˉξξ)+ˉξ(β10+β11ξ2))+12eiωn0˜τ(3β6+β7(ˉξ+2ξ)+ξ(2β8ˉξ+β8ξ+3β9ˉξξ)).

    Now, we compute W20(θ) and W11(θ) for θ[1,0] to give g21. By (3.7), we have

    ˙ω=˙Ut˙zpγn0(x)˙ˉzˉpγn0(x)=Aω+H(z,ˉz,θ), (3.9)

    where

    H(z,¯z,θ)=H20(θ)z22+H11(θ)z¯z+H02(θ)¯z22+. (3.10)

    Comparing the coefficients of (3.8) with (3.9), we have

    (A2iωn0˜τI)ω20=H20(θ),Aω11(θ)=H11(θ). (3.11)

    Then, we have

    ω20(θ)=g20iωn0˜τp(0)eiωn0˜τθˉg023iωn0˜τˉp(0)eiωn0˜τθ+E1e2iωn0˜τθ,ω11(θ)=g11iωn0˜τp(0)eiωn0˜τθˉg11iωn0˜τˉp(0)eiωn0˜τθ+E2, (3.12)

    where E1=n=0E(n)1, E2=n=0E(n)2,

    E(n)1=(2iωn0˜τI01e2iωn0˜τθdηn0(θ,ˉτ))1<˜F20,βn>,E(n)2=(01dηn0(θ,ˉτ))1<˜F11,βn>,nN0,
    <˜F20,βn>={1lπˆF20,n00,n=0,12lπˆF20,n00,n=2n0,1lπˆF20,n0=0,n=0,0,other,
    <˜F11,βn>={1lπˆF11,n00,n=0,12lπˆF11,n00,n=2n0,1lπˆF11,n0=0,n=0,0,other,

    and ˆF20=2(ς1,ς2)T, ˆF11=2(ϱ1,ϱ2)T.

    Thus, we can obtain

    c1(0)=i2ωn˜τ(g20g112|g11|2|g02|23)+12g21,μ2=Re(c1(0))Re(λ(˜τ)),T2=1ωn0˜τ[Im(c1(0))+μ2Im(λ(τjn))],β2=2Re(c1(0)). (3.13)

    Theorem 3.1. For any critical value τjn (nS,jN0), we have the following results:

    (1) When μ2>0 (resp. <0), the Hopf bifurcation is forward (resp. backward).

    (2) When β2<0 (resp. >0), the bifurcating periodic solutions on the center manifold are orbitally asymptotically stable(resp. unstable).

    (3) When T2>0 (resp. <0), the period increases (resp. decreases).

    To verify our theoretical results, we give the following numerical simulations. Fix parameters

    r=2,K=15,b=1,d=2,e=1,a=5,s=20,l=1.5,d1=2,d2=0.2.

    The bifurcation diagram of system (1.3) with the parameter of interference magnitude among predators c is given in Figure 1, where (H0) and (H1) hold. We can see that with an increase in parameter c, the Hopf bifurcation curves decrease, which implies that the stable region of positive equilibrium (u,v) will decrease. This means that increasing the interference magnitude among predators is not conducive to the homogeneous distribution of prey and predators in space. This causes spatial oscillations in prey and predator densities. Since the inhomogeneous Hopf bifurcation curve τ01 is always under the homogeneous Hopf bifurcation curve τ00. This spatial oscillation of prey and predator density is inhomogeneous.

    Figure 1.  Bifurcation diagram of system (1.3) with parameter c.

    Fixing c=0.1, we can determine that (u,v)(9.0000,1.9048) is the unique positive equilibrium, and (H1) holds. By direct calculation, we have τ=τ010.0758<τ000.1002, μ20.0105, β20.0584 and T20.8602. If we choose τ=0.05<τ, then (u,v) is locally asymptotically stable for model (1.3) (shown in Figure 2) and model (1.2) (shown in Figure 3). This means that the nonlocal competition term has no effect on the stability of the model (1.3). Prey and predator will coexist in a spatially homogeneous form, and their densities will converge to the positive equilibrium (u,v).

    Figure 2.  Numerical simulations of system (1.3) with τ=0.05. (u,v) is locally asymptotically stable.
    Figure 3.  Numerical simulations of system (1.2) with τ=0.05. (u,v) is locally asymptotically stable.

    If we choose τ<τ=0.09<τ00, then (u,v) is unstable, and inhomogeneous periodic solutions exist for the model (1.3) (shown in Figure 4). This means that prey and predator will coexist in the form of spatially inhomogeneous oscillations, and their densities will distribute inhomogeneously in space. To compare our result with the work in [20], we give the numerical simulations of model (1.2) under the same parameter τ=0.09 in Figure 5. We can see that (u,v) is locally asymptotically stable (1.2). This means that prey and predators will still coexist in a spatially homogeneous form and that their densities will converge to positive equilibrium (u,v).

    Figure 4.  Numerical simulations of system (1.3) with τ=0.09. (u,v) is unstable, and inhomogeneous periodic solutions exist.
    Figure 5.  Numerical simulations of system (1.2) with τ=0.09. (u,v) is locally asymptotically stable.

    Assuming there is nonlocal competition caused by limited resources in prey, we modified the model (1.2) to the model (1.3) with the nonlocal competition in prey, gestation delay in predator, and generalist predator. We study the local stability of the positive equilibrium and Hopf bifurcation at the positive equilibrium by using the time delay as a parameter. Through the center manifold and normal form method, we give some parameters that can determine the direction of Hopf bifurcation, the stability of bifurcating periodic solutions, and the period of periodic solutions.

    For the time delay when τ=0, we determine that nonlocal competition can induce Turing instability of the coexisting equilibrium point (u,v), but Turing instability cannot occur in the model (1.2) without nonlocal competition. Similar to the results in [20], time delays can affect the stability of the positive equilibrium. The positive equilibrium is stable when the delay is shorter than the critical value but unstable when the delay is longer than the critical value, and prey and predator densities will produce periodic oscillations.

    Through the bifurcation diagram, we find that increasing the interference magnitude among predators is not conducive to the homogeneous distribution of prey and predators in space. This causes spatial oscillations in prey and predator densities. Compared with the work in [20], the numerical simulations show that the inhomogeneous Hopf bifurcation curve τ01 is always under the homogeneous Hopf bifurcation curve τ00 for the model (1.3) with nonlocal competition. When the time delay is longer than the critical value τ01 and shorter than τ00, the prey and predator will coexist in space, but their densities will oscillate periodically in the form of spatially inhomogeneous oscillations for the model (1.3). However, the prey and predator will still coexist in a spatially homogeneous form, and their densities will converge to the positive equilibrium for the model (1.2). The model (1.3) is more realistic than the model (1.2) in the real world since it is difficult for the population density to exist in a completely uniform way. Through our theoretical analysis and numerical simulation, we show that stable inhomogeneous periodic solutions exist when the delay crosses the critical value, which is different from the work in [20].

    In the inhomogeneous periodic oscillations of prey and predator densities (Figure 4), we find that the amplitude near the boundary of the region is larger than that at the center of the region. This is because the prey needs to avoid the hunting of predators. They will escape from the interior of the area and move out to the boundary. However, because the living area is closed, they cannot cross the boundary. Thus, they can only move inward again. After reaching the interior of the area, they have to escape the hunting of predators again. This forms the periodic oscillation of this mode.

    In applying the predator-prey model, we suggest that scholars consider the nonlocal competition of prey since the dynamics of the models are different in the absence and presence of nonlocal competition. We hope that our model can be applied to ecological environment protection and population control. In future work, we will study the predator-prey model with a nonlocal competition term, nonconstant kernel function, and the Allee effect.

    We would like to thank you for following the instructions above very closely in advance. It will definitely save us lot of time and expedite the process of your paper's publication. This research is supported by the Fundamental Research Funds for the Central Universities (Grant No. 2572022BC01), Postdoctoral Program of Heilongjiang Province (No. LBH-Q21060), and College Students Innovations Special Project funded by Northeast Forestry University (No. 202210225157).

    The authors declare that they have no competing interests.



    [1] E. F. D. Goufo, Y. Khan, Q. A. Chaudhry, HIV and shifting epicenters for COVID-19, an alert for some countries, Chaos Solitons Fract., 139 (2020), 110030. http://dx.doi.org/10.1016/j.chaos.2020.110030 doi: 10.1016/j.chaos.2020.110030
    [2] N. Faraz, Y. Khan, E. Goufo, A. Anjum, A. Anjum, Dynamic analysis of the mathematical model of COVID-19 with demographic effects, Z. Naturforsch C. J. Biosci., 75 (2020), 389–396. https://doi.org/10.1515/znc-2020-0121 doi: 10.1515/znc-2020-0121
    [3] Q. Xue, C. Liu, L. Li, G. Q. Sun, Z. Wang, Interactions of diffusion and nonlocal delay give rise to vegetation patterns in semi-arid environments, Appl. Math. Comput., 399 (2021), 126038. http://dx.doi.org/10.1016/j.amc.2021.126038 doi: 10.1016/j.amc.2021.126038
    [4] R. Yang, D. Jin, Dynamics in a predator-prey model with memory effect in predator and fear effect in prey, Electronic Res. Arch., 30 (2022), 1322–1339. http://dx.doi.org/10.3934/era.2022069 doi: 10.3934/era.2022069
    [5] R. Yang, C. Zhang, Dynamics in a diffusive predator-prey system with a constant prey refuge and delay, Nonlinear Anal.: Real World Appl., 31 (2016), 1–22. http://dx.doi.org/10.1016/j.nonrwa.2016.01.005 doi: 10.1016/j.nonrwa.2016.01.005
    [6] M. A. Khan, R. Khan, Y. Khan, S. Islam, A mathematical analysis of Pine Wilt disease with variable population size and optimal control strategies, Chaos Solitons Fract., 108 (2018), 205–217. http://dx.doi.org/10.1016/j.chaos.2018.02.002 doi: 10.1016/j.chaos.2018.02.002
    [7] Y. Khan, H. Vˊazquez-Leal, Q. Wu, An efficient iterated method for mathematical biology model, Neural Comput. Applic., 23 (2013), 677–682. http://dx.doi.org/10.1007/s00521-012-0952-z doi: 10.1007/s00521-012-0952-z
    [8] Y. Khan, N. Faraz, Z. Smarda, An efficient iterated method for mathematical biology model, Neural Comput. Applic., 27 (2016), 671–675. http://dx.doi.org/10.1007/s00521-015-1886-z doi: 10.1007/s00521-015-1886-z
    [9] S. Mishra, R. K. Upadhyay, Exploring the cascading effect of fear on the foraging activities of prey in a three species Agroecosystem, Eur. Phys. J. Plus, 136 (2021), 974. http://dx.doi.org/10.1140/epjp/s13360-021-01936-5 doi: 10.1140/epjp/s13360-021-01936-5
    [10] R. Yang, Q. Song, Yong. An, Spatiotemporal dynamics in a predator-prey model with functional response increasing in both predator and prey densities, Mathematics, 10 (2022), 17. http://dx.doi.org/10.3390/math10010017 doi: 10.3390/math10010017
    [11] R. Yang, X. Zhao, Yong. An, Dynamical analysis of a delayed diffusive predator-prey model with additional food provided and anti-predator behavior, Mathematics, 10 (2022), 469. http://dx.doi.org/10.3390/math10030469 doi: 10.3390/math10030469
    [12] Y. Shao, Dynamics of an impulsive stochastic predator-prey system with the Beddington-DeAngelis functional response, Axioms, 10 (2021), 323. http://dx.doi.org/10.3390/axioms10040323 doi: 10.3390/axioms10040323
    [13] R. Yang, F. Wang, D. Jin, Spatially inhomogeneous bifurcating periodic solutions induced by nonlocal competition in a predator-prey system with additional food, Math. Methods Appl. Sci., 2022. http://dx.doi.org/10.1002/mma.8349
    [14] R. Yang, D. Jin, W. Wang, A diffusive predator-prey model with generalist predator and time delay, AIMS Math., 7 (2022), 4574–4591. http://dx.doi.org/10.3934/math.2022255 doi: 10.3934/math.2022255
    [15] C. Qin, J. Du, Y. Hui, Dynamical behavior of a stochastic predator-prey model with Holling-type III functional response and infectious predator, AIMS Math., 7 (2022), 7403–7418. http://dx.doi.org/10.3934/math.2022413 doi: 10.3934/math.2022413
    [16] Y. Kang, L. Wedekin, Dynamics of a intraguild predation model with generalist or specialist predator, J. Math. Biol., 67 (2013), 1227–1259. http://dx.doi.org/10.1007/s00285-012-0584-z doi: 10.1007/s00285-012-0584-z
    [17] S. Madec, J. Casas, G. Barles, C. Suppo, Bistability induced by generalist natural enemies can reverse pest invasions, J. Math. Biol., 75 (2017), 543–575. http://dx.doi.org/10.1007/s00285-017-1093-x doi: 10.1007/s00285-017-1093-x
    [18] L. N. Guin, S. Acharya, Dynamic behaviour of a reaction-diffusion predator-prey model with both refuge and harvesting, Nonlinear Dyn., 88 (2017), 1501–1533. http://dx.doi.org/10.1007/s11071-016-3326-8 doi: 10.1007/s11071-016-3326-8
    [19] R. K. Upadhyay, R. Agrawal, Dynamics and responses of a predator-prey system with competitive interference and time delay, Nonlinear Dyn., 83 (2016), 821–837. http://dx.doi.org/10.1007/s11071-015-2370-0 doi: 10.1007/s11071-015-2370-0
    [20] F. Liu, R. Yang, L. Tang, Hopf bifurcation in a diffusive predator-prey model with competitive interference, Chaos Solitons Fract., 120 (2019), 250–258. http://dx.doi.org/10.1016/j.chaos.2019.01.029 doi: 10.1016/j.chaos.2019.01.029
    [21] N. F. Britton, Aggregation and the competitive exclusion principle, J. Theor. Biol., 136 (1989), 57–66. https://doi.org/10.1016/S0022-5193(89)80189-4 doi: 10.1016/S0022-5193(89)80189-4
    [22] J. Furter, M. Grinfeld, Local vs. non-local interactions in population dynamics, J. Math. Biol., 27 (1989), 65–80. http://dx.doi.org/10.1007/BF00276081 doi: 10.1007/BF00276081
    [23] S. Wu, Y. Song, Spatiotemporal dynamics of a diffusive predator-prey model with nonlocal effect and delay, Commun. Nonlinear Sci. Numer. Simul., 89 (2020), 105310. http://dx.doi.org/10.1016/j.cnsns.2020.105310 doi: 10.1016/j.cnsns.2020.105310
    [24] D. Geng, W. Jiang, Y. Lou, H. Wang, Spatiotemporal patterns in a diffusive predator-prey system with nonlocal intraspecific prey competition, Stud. Appl. Math., 2021 (2021), 396–432. http://dx.doi.org/10.1111/sapm.12444 doi: 10.1111/sapm.12444
    [25] S. Chen, J. Yu, Stability and bifurcation on predator-prey systems with nonlocal prey competition, Discrete Cont. Dyn. Syst., 38 (2018), 43–62. http://dx.doi.org/10.3934/dcds.2018002 doi: 10.3934/dcds.2018002
    [26] Y. Liu, D. Duan, B. Niu, Spatiotemporal dynamics in a diffusive predator-prey model with group defense and nonlocal competition, Appl. Math. Lett., 103 (2020), 106175. https://doi.org/10.1016/j.aml.2019.106175 doi: 10.1016/j.aml.2019.106175
    [27] J. Wu, Theory and applications of partial functional differential equations, Springer, 1996. https://doi.org/10.1007/978-1-4612-4050-1
    [28] B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and applications of Hopf bifurcation, Cambridge University Press, 1981.
  • Reader Comments
  • © 2022 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(2037) PDF downloads(85) Cited by(0)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog