
In this paper, a delayed predator-prey system with additional food and asymmetric functional response is investigated. We discuss the local stability of equilibria and the existence of local Hopf bifurcation under the influence of the time delay. By using the normal form theory and center manifold theorem, the explicit formulas which determine the properties of bifurcating periodic solutions are obtained. Further, we prove that global periodic solutions exist after the second critical value of delay via Wu's theory. Finally, the correctness of the previous theoretical analysis is demonstrated by some numerical cases.
Citation: Luoyi Wu, Hang Zheng. Hopf bifurcation in a delayed predator-prey system with asymmetric functional response and additional food[J]. AIMS Mathematics, 2021, 6(11): 12225-12244. doi: 10.3934/math.2021708
[1] | Chenxuan Nie, Dan Jin, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and generalist predator. AIMS Mathematics, 2022, 7(7): 13344-13360. doi: 10.3934/math.2022737 |
[2] | 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 |
[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] | San-Xing Wu, Xin-You Meng . Dynamics of a delayed predator-prey system with fear effect, herd behavior and disease in the susceptible prey. AIMS Mathematics, 2021, 6(4): 3654-3685. doi: 10.3934/math.2021218 |
[5] | 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 |
[6] | 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 |
[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] | 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 |
[9] | 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 |
[10] | Ming Wu, Hongxing Yao . Stability and bifurcation of a delayed diffusive predator-prey model affected by toxins. AIMS Mathematics, 2023, 8(9): 21943-21967. doi: 10.3934/math.20231119 |
In this paper, a delayed predator-prey system with additional food and asymmetric functional response is investigated. We discuss the local stability of equilibria and the existence of local Hopf bifurcation under the influence of the time delay. By using the normal form theory and center manifold theorem, the explicit formulas which determine the properties of bifurcating periodic solutions are obtained. Further, we prove that global periodic solutions exist after the second critical value of delay via Wu's theory. Finally, the correctness of the previous theoretical analysis is demonstrated by some numerical cases.
In the real world, we often need to control or eliminate the target prey in a prey-predator system. Biological control and chemical control are the two essential methods. The main means of chemical control is to spray pesticides on pests at different fixed times (see e.g. [1]). However, chemical control is said to be environmentally detrimental. There are various forms of biological control, which is the focus of current research, such as immunocontraceptive technology, introducing other predators, offering the predator additional food and so on. Saunders et al. [2] used the approaches of immunocontraception and daughterless genes to control the growth of the target pest species. Ghosh et al. [3] investigated the effectiveness of periodic impulsive releases of natural enemies into a two-patch environment.
Numerous studies had shown that the additional food (to predator) can be helpful to increase predators and the prey can be controlled. Tena et al. [4] showed that the melinus females were better able to forage and oviposit by providing sugar to them. Srinivasu et al. [5] studied a standard predator-prey model with additional food and presented the evidence that how to eliminate target pests. Sahoo et al. [6] investigated that the chaotic population dynamics can be controlled to obtain regular population dynamics only by supplying additional food to top predator. Research on the models with additional food, one can refer to the literature [7,8,9,10,11].
Recently, Basheer et al. [12] believed that the additional food can increase the carrying capacity and birth rate of the predators, so they studied the following Holling-Tanner model with additional food:
{dudt=u(t)(1−u(t))−su(t)v(t)mv(t)+αβ+u(t),dvdt=δv(t)(n+β−v(t)αβ+u(t)), | (1.1) |
where u(t) and v(t) denote the prey and predator density. α, β∈R+, and 1α and β represent the quality and quantity of the additional food, respectively. The biological significance of other parameters refer to [12]. Their research indicated that a conditional stable prey-extinction equilibrium could be obtained in system (1.1), but it was nonexistent in the absence of additional food.
On the other hand, if the population maturation time is considered, the corresponding model should be delayed. Therefore, lots of delayed predator-prey systems were studied (see, e.g. [13,14,15,16,17,18,19,20,21,22,23,24]). It is interesting that Jiang et al. [25] applied a delay differential equation (DDE) to two-enterprise interaction mechanism. The dynamical properties of DDE are far more complex than those of ordinary differential equation (ODE). A time delay can cause an equilibrium undergoes from stable state to unstable, which leads to the complicated dynamics of DDE. Arising periodic solution through the Hopf bifurcation is one of the hot topics (see, e.g. [18,19,20,21,22,23,24,26]). However, these bifurcating periodic solutions from Hopf bifurcation are generally local. Whether these local periodic solutions exist globally is an interesting subject. Erbe et al. [27] established the global Hopf bifurcation theorem and Wu [28] applied it to a neural networks with memory. Thereafter, some researchers employed the theorem in [28] to study the global existence of periodic solutions for DDE (see, e.g. [29,30,31,32,33,34]).
Inspiration based on model (1.1), we consider the maturation time of predator and standard Holling type II functional response in the model, then we improve model (1.1) as follows:
{dudt=u(t)(1−u(t))−su(t)v(t)m+αβ+u(t),dvdt=δv(t)(n+β−v(t−τ)αβ+u(t−τ)), | (1.2) |
where τ represent the maturation time of predator. The initial conditions are chosen as:
u(θ)=φ1(θ)≥0,v(θ)=φ2(θ)≥0,θ∈[−τ,0),φ1(0)>0,φ2(0)>0 |
where (φ1(θ),φ2(θ))∈C{[−τ,0],R2+},R2+={u,v:u≥0,v≥0}.
The main purpose of our work are concluded as follows:
(I) We consider the feedback time delay in the maturation time of predator, which is more general than the works in [12]. We investigate the effects of feedback time on the stability of the equilibria and the conditions on occurring Hopf bifurcations.
(II) The delayed feedback is designed for studying bifurcating periodic solutions. We analyze the direction and stability of bifurcating periodic solutions on the center manifold.
(III) The global existence of bifurcating periodic solutions are studied mathematically.
The rest of this paper is organized as follows. We first discuss some properties of system (1.2) to prepare for the next section. In Section 3, we study the local stability of each feasible equilibrium of system (1.2) with the effect of the time delay τ. The formulas determining the direction and stability of bifurcating periodic solutions are obtained via the theory in Hassard et al. [35] in Section 4. In Section 5, the global existence of periodic solutions under the second critical value is proved by using the theory in Wu [28]. Finally, some examples are utilized to demonstrate the validity of the previous results.
Clearly, system (1.2) can be calculated by
{ u(t)=u(0) exp{∫t0(1−u(ξ)−sv(ξ)m+αβ+u(ξ))dξ}, v(t)=v(0) exp{∫t0(δ(n+β−v(ξ−τ)αβ+u(ξ−τ)))dξ}. | (2.1) |
According to positive initial values of system (1.2), it is not difficult to obtain the following lemma.
Lemma 2.1 Any of the solutions of system (1.2) are positive for t≥0 with positive initial values.
Theorem 2.2 System (1.2) is ultimately bounded when τ is bounded.
Proof. From the first equation of system (1.2) and by using the comparison theorem, it is easy to obtain
limt→+∞sup u(t)≤1. |
Namely, there exists a time T1 such that u(t)≤1+ε for arbitrary ε>0 and t>T1. By the second equation of (1.2), we have
dvdt≤σ(δ+1α)v. |
Integrating both sides on the interval [t−τ,t], it produces
v(t)≤v(t−τ)exp{σ(δ+1α)τ}, |
which implies
v(t−τ)≥v(t)exp{−σ(δ+1α)τ}. |
Meanwhile, it follows from the second equation of (1.2) that
dvdt≤δv(n+1α−exp{−σ(δ+1α)τ}αβ+1+εv), =δexp{−σ(δ+1α)τ}αβ+1+εv((αβ+1+ε)(n+1α)exp{σ(n+1α)τ}−v). |
Using the comparison theorem, we obtain
limt→+∞sup v(t)≤(αβ+1)(n+1α)exp{σ(n+1α)τ}. |
Therefore, Theorem 2.2 is confirmed.
Theorem 2.3 In the absence of delay, if s<δ, then there is no closed loop in the first quadrant of system (1.2).
Proof. Let B(u,v)=1uv, f1=u(1−u)−suvm+αβ+u, f2=δv(n+β−vαβ+u)), then
∂(Bf1)∂u+∂(Bf2)∂v=−1v+s(m+αβ+u)2−δu(αβ+u),≤−1v+su(αβ+u)−δu(αβ+u),<0 (provided s<δ). |
By using the Bendixson-Dulac criterion, the proof is completed.
In the paper, we use the following representations for the sake of simplicity:
A≡m+αβ+sn−1,B≡snαβ+sβ−m−αβ,C≡ˉum+αβ+ˉu,D≡δˉvαβ+ˉu, G≡ˉu−C(1−ˉu), F≡G+snC, |
where (ˉu,ˉv) stands for the positive interior equilibrium and is defined in this section.
In order to obtain the equilibria, we discuss the algebraic equations:
{ u(1−u)−suvm+αβ+u=0, δv(n+β−vαβ+u)=0. | (3.1) |
Let (ˉu,ˉv) stands for the interior equilibrium, where ˉu is the positive root of the equation u2+Au+B=0. Thus, we have ˉu±=−A±√A2−4B2 and ˉv=n(αβ+ˉu)+β. The equilibria are as follows:
(i) Trivial equilibrium E0=(0,0).
(ii) Predator-extinction equilibrium E1=(1,0).
(iii) Prey-extinction equilibrium E2=(0,nαβ+β).
(iv) A unique coexisting equilibrium E+=(ˉu+,ˉv+) when B<0; two coexisting equilibria E±=(ˉu±,ˉv±) when B>0, A<0 and A2−4B>0.
Let E=(u∗,v∗) be arbitrary equilibrium. We use linearization technique to analyze the local stability of system (1.2). The Jacobian matrix of system (1.2) at E=(u∗,v∗) is given by
J(u∗,v∗)=(1−2u∗−sv(m+αβ)(m+αβ+u∗)2−sum+αβ+u∗−δ(β−v∗)v∗(αβ+u∗)2e−λτδ(n+β−v∗αβ+u∗)−δv∗αβ+u∗e−λτ). |
It is easy to confirm that E0=(0,0) is an unstable node and E1=(1,0) is a saddle.
At Prey-extinction equilibrium E2=(0,nαβ+β), the corresponding Jacobian matrix is
J(E2)=(1−s(αβn+β)m+αβ0nδ(αβn)αβe−λτ−δ(n+1α)e−λτ) |
and the characteristic equation becomes
(λ−P1)(λ+P2e−λτ)=0, | (3.2) |
where P1=1−s(αβn+β)αβ, P2=δ(n+1α).
Case 1. τ=0.
It follows from (3.2) that E2=(0,nαβ+β) is locally asymptotically stable when P1<0 (equivalent to B>0) and unstable when P1>0.
Case 2. τ>0.
Let λ=iω∗(ω∗>0) be a root of the equation λ+P2e−λτ=0, then iω+P2(cosωτ−isinωτ)=0. By a direct calculation, we get ω∗0=P2 and τ∗k=1P2(2kπ+π2),k=0,1,2,⋯. Differentiating the both sides of λ+P2e−λτ=0 with respect to τ, we have
dλdτ+P2(−τdλdτ−λ)e−λτ=0, |
that is
(dλdτ)−1=1λP2e−λτ−τλ=1ω2−τλ, |
which implies that
sgn{dReλdτ}λ=iω∗0=sgn{Re(dλdτ)−1}λ=iω∗0=1ω∗20>0. |
Lemma 3.1 If B>0, then all roots of the characteristic equation (3.2) have negative real part when 0≤τ<π2P2 and at least one positive real part when τ>π2P2.
Therefore, we have the following conclusions for the boundary equilibria.
Theorem 3.2 (i) The trivial equilibrium E0=(0,0) and predator-extinction equilibrium E1=(1,0) are always unstable for all τ≥0.
(ii) When B>0, the prey-extinction equilibrium E2=(0,nαβ+β) is asymptotically stable for all 0≤τ<π2P2 and unstable for all τ>π2P2. The system (1.2) undergoes a Hopf bifurcation at E2 for τ=π2P2.
Now we investigate the stability of the coexisting equilibrium E=(ˉu,ˉv). The corresponding Jacobian matrix is
J(E)=(−ˉu+C(1−ˉu)−sCnDe−λτ−De−λτ) |
and the corresponding characteristic equation becomes
λ2+Gλ+Dλe−λτ+DFe−λτ=0. | (3.3) |
Case 1. τ=0.
Equation (3.3) becomes
λ2+(G+D)λ+DF=0. |
We calculate F as follows:
F=ˉu−C(1−ˉu)+snC=C(m+αβ+sn+2ˉu−1)=C(A+2ˉu), |
which implies that F<0 when ˉu=ˉu− and F>0 when ˉu=ˉu+. Thus, the following conclusions are obvious.
Lemma 3.3 If the coexisting equilibria exist and τ=0, then E+=(ˉu+,ˉv+) is locally asymptotically stable when G+D>0 and E−=(ˉu−,ˉv−) is always unstable.
Case 2. τ>0.
Let λ=iω(ω>0) be a root of the equation (3.3), then
−ω2+iωG+iωD(cosωτ−isinωτ)+DF(cosωτ−isinωτ)=0. | (3.4) |
We obtain
−ω2+ωDsinωτ+DFcosωτ=0,ωG+ωDcosωτ−DFsinωτ=0, | (3.5) |
that is
cosωτ=ω2(F−G)ω2D+DF2,sinωτ=ω3+ωGFω2D+DF2. | (3.6) |
From (3.6), we have
ω4+(G2−D2)ω2−D2F2=0. | (3.7) |
Let z=ω2, (3.7) turns to
z2+(G2−D2)z−D2F2=0. | (3.8) |
Clearly, (3.8) has a unique positive root z0=−(G2−D2)+√(G2−D2)2+4D2F22. So (3.8) has a pair of purely imaginary roots ±iω0(ω0=√z0). From the first equation of (3.6), we have
τk=1ω0(2kπ+arccosω2(F−G)ω2D+DF2),k=0,1,2,3,⋯. |
Differentiating both sides of (3.3) with respect to τ, we have
2λdλdτ+Gdλdτ+Ddλdτe−λτ+Dλe−λτ(−τdλdτ−λ)+DFe−λτ(−τdλdτ−λ)=0, |
namely,
(dλdτ)−1=(2λ+G)eλτDλ(λ+F)+1λ(λ+F)−τλ. |
Then
Re(dλdτ)−1λ=iω0=ω0Gcosω0τ0−2ω0sinω0τ0−2ω0Fcosω0τ0+Gsinω0τ0+Dω0−ω0D(ω20+F2)=2ω20+(G2−D2)D2(ω20+F2)=√(G2−D2)2+4D2F2D2(ω20+F2), |
which implies that
sgn{dReλdτ}λ=iω0=sgn{Re(dλdτ)−1λ=iω0}>0. |
Lemma 3.4 If condition G+D>0 holds, then all roots of the characteristic equation (3.3) at E+=(ˉu+,ˉv+) have negative real part when 0<τ<τ0 and at least one positive real part when τ>τ0.
From lemma 3.3 and 3.4, we confirm the following conclusions about the interior equilibrium.
Theorem 3.5 (i) If E−=(ˉu−,ˉv−) exists, then it is unstable for all τ≥0.
(ii) If E+=(ˉu+,ˉv+) exists and the condition G+D>0 holds, then E+ is locally asymptotically stable for all 0≤τ<τ0 and unstable for all τ>τ0. The system (1.2) undergoes a Hopf bifurcation at E+ for τ=τ0.
Remark 3.6 If the prey is pest, we just need to control the quantity of additional food satisfy the following conditions:
β>1−m−snα(i.e.A>0) and β>msnα+s−α(i.e.B>0), |
then additional food can induce pest eradication. Meanwhile, the density of predators eventually goes to nαβ+β when the maturation time of predator species is less than π2P2.
Remark 3.7 If the conditions B>0, A<0, A2−4B>0 and G+D>0 hold simultaneously, the system (1.2) is bistable when τ<min{τ∗0,τ0}.
We know from the literature [35] that the properties of Hopf bifurcation are determined by the following three quantities, namely
{ μ2=−Re(c1(0))λ′(˜τ), β2=2Re(c1(0)), T2=−lm(c1(0))+μ2lmλ′(˜τ)˜ω, | (4.1) |
where c1(0)=i2˜ω˜τ[g11g20−2|g11|2−∣g02∣23]+g212, ˜τ is the critical value. μ2, β2 and T2 determine direction, stability and the period of the bifurcating periodic solutions, respectively. So we need to figure out the value of gij in c1(0). We will use the normal form theory and the center manifold theorem to obtain the expression of gij in this section.
It has known that system (1.2) undergoes Hopf bifurcation at coexisting equilibrium E+ when τ=τ0. We denote the critical values τk and E+=(ˉu+,ˉv+) as ˜τ and E∗=(u∗,v∗), respectively. Let x(t)=u(τt)−u∗ and y(t)=v(τt)−v∗, using Taylor expansion, (1.2) can be rewritten as
(˙x(t)˙y(t))=τA1(x(t)y(t))+τB1(x(t−1)y(t−1))+F(xt,yt,τ), | (4.2) |
where
A1=(a1a200), B1=(00a3a4), |
F(xt,yt,τ)=τ(a5x2(t)+a6x(t)y(t)+⋯a7x2(t−1)+a8x(t−1)y(t)+a9x(t−1)y(t−1)+a10y(t)y(t−1)+⋯), |
a1=1−2u∗−s(m+αβ)v∗(m+αβ+u∗)2, a2=−su∗m+αβ+u∗, a3=−δν∗(β−v∗)(αβ+u∗)2=δnv∗αβ+u∗,
a4=−δv∗αβ+u∗, a5=−2+2s(m+αβ)v∗(m+αβ+u∗)3, a6=−2s(m+αβ)(m+αβ+u∗)2,
a7=2δν∗(β−v∗)(αβ+u∗)3, a8=−2δ(β−v∗)(αβ+u∗)2, a9=2δv∗(αβ+u∗)2, a10=−2δαβ+u∗.
Let τ=ˉτ+h, then h=0 is a Hopf bifurcation value of the system (1.2). Choose the phase space C=C([−1,0],R2), ϕ(θ)=(ϕ1(θ),ϕ2(θ))T∈C,θ∈[−1,0], define L(h)
L(h)ϕ=(˜τ+h)A1ϕ(0)+(˜τ+h)B1ϕ(−1). |
By the Riesz representation theorem, we choose the bounded variation function
η(h,θ)=(˜τ+h)A1δ(θ)−(˜τ+h)B1δ(θ+1) |
such that
L(h)ϕ=∫0−1dη(h,θ)ϕ(θ), |
where δ(θ) is delta function.
For ϕ∈C1([−1,0],R2), define
A(h)ϕ(θ)=˙ϕ(θ)+T0(θ)[L(h)(ϕ)−˙ϕ(0)], |
and
R(h)ϕ(θ)=T0(θ)F(ϕ,τ+h), |
where T0(θ)={I, θ=0,0, θ∈[−1,0).
Then (4.2) is written as
˙ut=A(h)ut+R(h)ut, | (4.3) |
where ut=u(t+θ) and ut=(xt,yt)T.
For ϕ∈C1([−1,0],C2) and ψ∈C1([0,1],(C2)∗), define a adjoint operator of A(0)
A∗ψ(s)=−˙ψ(s)+T0(−s)[∫0−1dη(0,t)ψ(−t)+˙ψ(0)] |
and the bilinear form
⟨ψ,ϕ⟩=ˉψ(0)ϕ(0)−∫0−1∫θ0ˉψ(ξ−θ)dη(θ,0)ϕ(ξ)dξ. |
From the previous discussion, we know ±i˜ω˜τ are the eigenvalues of A(0) and A∗. Let q(θ)=(1,α1)Tei˜ω˜τθ be the eigenvector of A(0) corresponding to i˜ω˜τ and q∗(s)=M(1,α2)ei˜ω˜τs be the eigenvector of A∗ corresponding to −i˜ω˜τ, and they satisfy the conditions ⟨q∗,q⟩=1 and ⟨q∗,q¯⟩=0. Therefore, we have
˜τ(a1a2a3e−i˜ω˜τa4e−i˜ω˜τ)(1α1)=i˜ω˜τ(1α1) |
and
˜τM(1α2)(a1a2a3ei˜ω˜τa4ei˜ω˜τ)=−i˜ω˜τM(1α2), |
then α1=i˜ω−a1a2 and α2=−i˜ω−a1a3ei˜ω˜τ.
By the bilinear form, we have
⟨q∗(s),q(θ)⟩=ˉM(1,ˉα2)(1,α1)T−∫0−1∫θξ=0ˉM(1,ˉα2)e−i˜ω˜τ(ξ−θ)dη(θ,0)(1,α1)Tei˜ω˜τξdξ,=ˉM(1+α1ˉα2)−∫0−1ˉM(1,ˉα2)θei˜ω˜τθdη(θ,0)(1,α1)T,=ˉM((1+α1ˉα2)+˜τˉα2(a3+a4α1)e−i˜ω˜τ). | (4.4) |
Thus M=1(1+ˉα1α2)+˜τα2(a3+a4ˉα1)ei˜ω˜τ.
We need the coordinates to describe the center manifold C0 near h=0. Let z and ˉz be local coordinates for C0 in the directions of q∗ and ˉq∗. Assume that ut is a solution of (4.3) at h=0, define
z(t)=⟨q∗,ut⟩, W(t,θ)=ut(θ)−z(t)q(θ)−ˉz(t)ˉq(θ). |
On C0, we have W(t,θ)=W(z(t),ˉz(t),θ), where
W(z,ˉz,θ)=W20(θ)z22+W11zˉz+W02(θ)ˉz22+W30(θ)z32+⋅⋅⋅. | (4.5) |
The manifold of (4.2) on the center manifold is determined by the following equation
˙z(t)=i˜τ˜ωz(t)+¯q∗(0)F(zq(0)+ˉzˉq(0)+W(z,ˉz,0))△=i˜ω˜τz(t)+¯q∗(0)F0 |
which is abbreviated to
˙z(t)=i˜τ˜ωz(t)+g(z,ˉz), | (4.6) |
where the power series form of g(z,ˉz) is
g(z,ˉz)=g20z22+g11zˉz+g02ˉz22+g21z2ˉz2+⋅⋅⋅ | (4.7) |
and
F0=˜τ(a5x2(t)+a6x(t)y(t)a7x2(t−1)+a8x(t−1)y(t)+a9x(t−1)y(t−1)+a10y(t)y(t−1)). |
By a direct calculation, we have
x(t)=z+ˉz+W(1)20(0)z22+W(1)11(0)zˉz+W(1)02(0)ˉz22+W(1)30(0)z32+⋅⋅⋅,y(t)=α1z+ˉα1ˉz+W(2)20(0)z22+W(2)11(0)zˉz+W(2)02(0)ˉz22,x(t−1)=e−i˜ω˜τz+ei˜ω˜τˉz+W(1)20(−1)z22+W(1)11(−1)zˉz+W(1)02(−1)ˉz22+W(1)30(−1)z32+⋅⋅⋅,y(t−1)=α1e−i˜ω˜τz+ˉα1ei˜ω˜τˉz+W(2)20(−1)z22+W(2)11(−1)zˉz+W(2)02(−1)ˉz22+W(2)30(−1)z32+⋅⋅⋅. | (4.8) |
Substituting (4.8) into F0 and comparing with (4.7), we obtain
g20=2˜τˉM(a5+a6α1+ˉα2(a7e−2i˜ω˜τ+a8α1e−i˜ω˜τ+a9α1e−2i˜ω˜τ+a10α21e−i˜ω˜τ)),
g02=2˜τˉM(a5+a6ˉα1+ˉα2(a7e2i˜ω˜τ+a8ˉα1ei˜ω˜τ+a9ˉα1e2i˜ω˜τ+a10ˉα21ei˜ω˜τ)),
g11=2˜τˉM(a5+a6Re{α1}+ˉα2(a7+a8Re{α1ei˜ω˜τ}+a9Re{α1}+a10Re{α1ˉα1ei˜ω˜τ})),
g21=2˜τˉM(a5k5+a6k6+ˉα2(a7k7+a8k8+a9k9+a10k10)),
where
k5=W(1)20(0)+2W(1)11(0),
k6=12ˉα1W(1)20(0)+α1W(1)11(0)+12W(2)20(0)+W(2)11(0),
k7=ei˜ω˜τW(1)20(−1)+2e−i˜ω˜τW(1)11(−1),
k8=12ˉα1W(1)20(−1)+α1W(1)11(−1)+12ei˜ω˜τW(2)20(0)+e−i˜ω˜τW(2)11(0),
k9=12ei˜ω˜τW(2)20(−1)+e−i˜ω˜τW(2)11(−1)+12ˉα1ei˜ω˜τW(1)20(−1)+α1e−i˜ω˜τW(1)11(−1),
k10=12ˉα1W(2)20(−1)+α1W(2)11(−1)+12ˉα1ei˜ω˜τW(2)20(0)+α1e−i˜ω˜τW(2)11(0).
In order to obtain the normal form of (4.6) confined to the center manifold, we need compute W20(θ) and W11(θ).
Using ˙W=˙ut−˙zq−˙ˉzˉq which combines with (4.3) and (4.6), we obtain
˙W={AW−gq(θ)−ˉgˉq(θ),θ∈[−1,0),AW−gq(θ)−ˉgˉq(θ)+F0,θ=0. | (4.9) |
On the other hand, from (4.5) and (4.6), we have
˙W=Wz˙z+Wˉz˙ˉz=[W20(θ)z+W11(θ)ˉz](i˜τ˜ωz(t)+g(z,ˉz))+[W11(θ)z+W02(θ)ˉz](i˜τ˜ωˉz(t)+ˉg(z,ˉz))+⋯. | (4.10) |
We substitute (4.6) into (4.9) and compare the coefficients of z22 and zˉz with (4.10), respectively. It gives
(2˜ω˜τI−A)W20(θ)={−g20q(θ)−ˉg02ˉq(θ),θ∈[−1,0),−g20q(θ)−ˉg02ˉq(θ)+Fz2,θ=0. | (4.11) |
and
−AW11(θ)={−g11q(θ)−ˉg11ˉq(θ),θ∈[−1,0),−g11q(θ)−ˉg11ˉq(θ)+Fzˉz,θ=0. | (4.12) |
According to (4.11) and (4.12), by a direct calculation for θ∈[−1,0), we obatin
W20(θ)=ig20˜ω˜τq(θ)+iˉg023˜ω˜τˉq(θ)+E1e2iω0θ |
and
W11(θ)=−ig11˜ω˜τq(θ)+iˉg11˜ω˜τˉq(θ)+E2, |
where E1 and E2 hold the following equations
(2i˜ω˜τ−a1−a2−a3e−2i˜ω˜τ2i˜ω˜τ−a4e−2i˜ω˜τ)E1=2(a5+a6α1a7e−2i˜ω˜τ+a8α1e−i˜ω˜τ+a9α1e−2i˜ω˜τ+a10α21e−2i˜ω˜τ), |
(−a1−a2−a3−a4)E2=2(a5+a6Re{α1}a7+a8Re{α1ei˜ω˜τ}+a9Re{α1}+a10α1¯α1Re{ei˜ω˜τ}). |
From the above discussions, we have the following conclusions on the center manifold.
Theorem 4.1 (i) If μ2>0 (μ2<0), then the Hopf bifurcation is supercritical (subcritical).
(ii) If β2<0 (β2>0), the bifurcating periodic solution is stable (unstable).
(iii) If T2>0 (T2<0), the period increases (decreases).
Next, we investigate the global continuation of bifurcating periodic solutions from the positive equilibrium (E+,τk). We employ the global Hopf bifurcation theorem and follow closely the notations in Wu [28].
Denote the conditions for the existence of E+ as
(H) Either B<0 or B>0,A<0 and A2−4B>0.
Let R+={(u,v)∈R2,u>0,v>0}, X=C([−τ,0],R2), zt=(ut,vt), the system (1.2) is rewritten as
˙z(t)=F(zt,τ,p), | (5.1) |
where zt(θ)=z(t+θ)∈X and (τ,p)∈R+×R+. Clearly, the mapping F:X×R+×R+⟶R2 is completely continuous. If we take R2 for the subspace of constant functions of X, we obtain ˆF∣R2×R+×R+:R2×R+×R+⟶R2. It is easy to know (E0,τ,p), (E1,τ,p), (E2,τ,p), (E−,τ,p) and (E+,τ,p) are all stationary solutions of (5.1). Now, we verify that (E+,τ,p) holds the conditions (A1), (A2), (A3) and (A4) in [28].
From (1.2), We know easily that ˆF∈C2(R2+×R+×R+,R2+) and F(ϕ,τ,p) is differential with respect to ϕ. That is to say, the conditions (A1) and (A3) are satisfied. We also obtain
DˆF(E+,τ,p)=(−ˉu++C(1−ˉu+)−sCnD−D). |
Directly calculate, we have
DetDzˆF(E+,τ,p)=−D(ˉu+−C(1−ˉu+)+snD)=−DF<0. |
Then DzˆF(E+,τ,p) is a homeomorphism on R2 at E+, which satisfies the condition (A2).
The characteristic matrix of (5.1) at (E+,τ,p) is taken as:
Δ(E+,τ,p)(λ)=λId−DF(E+,τ,p)(eλ.Id). | (5.2) |
A stationary solution (E+,τ,p)(λ) of (5.1) is called a center if detΔ(E+,τ,p)(λ)=0 has purely imaginary characteristic roots of the form im2πp0 for some positive integer m. It follows from (5.2) that
detΔ(E+,τ,p)(λ)=λ2+Eλ+Dλe−λτ+DFe−λτ=0, | (5.3) |
which is the same as (3.3). Taking p0=2πω0, we know i2πp0 is a root of (5.3), namely, i2πp0 is an eigenvalue of (E+,τk,2πω0). Thus, (E+,τk,2πω0) is a center, where τk and ω0 are defined in Section 3. Furthermore, the center (E+,τk,2πω0) is an isolated center, because it satisfies the following two conclusions:
(i) J(E+,τk,2πω0)=1, where J(E+,τk,2πω0) is a positive integer set with respect to m such that im2πp0 are eigenvalues of (E+,τk,2πω0).
(ii) For arbitrary k≥0, there exist εk>0,δk>0 and a smooth function λ:(τk−σk,τk+σk→C)(C is complex field) such that detΔ(E+,τk,2πω0)(λ(τ))=0, |λ(τ)−iω0|<εk for arbitrary τ∈(τk−σk,τk+σk) and λ(τk)=iω0, dλ(τ)dτ|τ=τk>0. That is, (E+,τk,2πω0) is the only center in certain neighborhood of (E+,τk,2πω0).
Let
Ωεk,p0={(r,p)|0<r<εk,p0−εk<p<p0+εk}. |
Clearly, for τ∈(τk−σk,τk+σk) and (r,P)∈∂Ωεk,p0 such that detΔ(E+,τ,p)(r+i2πp)=0 if and only if τ=τk,p=p0,r=0. This is the condition (A4).
So far, we have verified the conditions (A1)−(A4) in [28]. Define
H±m(E+,τk,2πω0)(r,p)=det△(E+,τk±δk,p)(r+im2πp). |
The condition (A4) and J(E+,τk,2πω0)=1 imply H±1(E+,τk,2πω0)(r,p)≠0 for (r,p)∈∂Ωεk,p0. Therefore, the first crossing number γ1(E+,τk,2πω0) is calculated as
γ1(E+,τk,2πω0)=deg(H−1)(E+,τk,2πω0)(r,p),Ωεk,p0)−deg(H+1)(E+,τk,2πω0)(r,p),Ωεk,p0)=−1. | (5.4) |
By the similar arguments, we can know (E0,τ,p) and (E1,τ,p) are not the centers, but (E2,τ∗k,2πω∗0) and (E−,τk,2πω0) are isolated centers. we may also obtain γ1(E2,τ∗k,2πω∗0)=−1 and γ1(E−,τk,2πω0)=−1.
In what follows we define
Σ=Cl{(z,τ,p)|z is p−periodic solution of (5.1)},
N={(ˉz,τ,p)|F(ˉz,τ,p)=0}.
Let C(E+,τk,2πω0) denote the connected component through (E+,τk,2πω0) in Σ.
By Theorem 3.2 in [28], there exists an isolated center (E+,τk,2πω0) which satisfies J(E+,τk,2πω0)=1 and γ1(E+,τk,2πω0)≠0 such that C(E+,τk,2πω0) through (E+,τk,2πω0) in Σ is nonempty. In addition, all the centers of (5.1) are isolated centers and satisfy
∑(ˉz,τ,p)∈(E+,τk,2πω0)∩N(F)γm(ˉz,τ,p)<0. |
By Theorem 3.3 in [28], we obtain the following Lemma.
Lemma 5.1 C(E+,τk,2πω0) is unbounded.
On the other hand, from Theorem 2.2 in Section 2, it is easy to obtain the following lemma.
Lemma 5.2 If τ is bounded, then any nontrivial periodic solution of system (1.2) is uniformly bounded.
Lemma 5.3 When the conditions (H) and s<δ hold, (1.2) has no any nontrivial τ−periodic solutions.
Proof. If (u∗(t),v∗(t)) is a nontrivial τ−periodic solution of (1.2), then it is also a nontrivial periodic solution of the following (5.5).
{dudt=u(1−u)−suvm+αβ+u,dvdt=δv(n+β−vαβ+u). | (5.5) |
When the condition (H) hold, (5.5) has at most five equilibria, namely, E0=(0,0), E1=(1,0), E2=(0,nαβ+β) and E±=(ˉu±,ˉv±). Note that E0, E1, and E2 are located in u-axis and v-axis, they can not produce any nontrivial periodic solution. On the other hand, E± also can not produce any nontrivial periodic solution due to s<δ. Thus, there is no any nontrivial periodic solution in (5.5). The proof is complete.
Theorem 5.4 Suppose the conditions (H), G+D>0 and s<δ hold, then for each τ>τk,k=1,2,3⋯, system (1.2) has at least k−1 periodic solutions.
Proof. Let ProjΘC(E+,τk,2πω0) be the projection of C(E+,τk,2πω0) onto Θ-space. From Lemma 5.2, it is easy to know that ProjzC(E+,τk,2πω0) is bounded. It follows from the proof of Lemma 5.3 that (1.2) with τ=0 has no nontrivial periodic solution. Thus, ProjτC(E+,τk,2πω0) is away from zero.
We suppose that ProjτC(E+,τk,2πω0) is bounded. Then there exist τ∗ such that ProjτC(E+,τk,2πω0) is a subset of interval (0,τ∗). From the definition of τk and ω0 in section 3, we have 2πω0<τk,k=1,2,3⋯. Applying Lemma 5.3, we have p∈(0,τ∗) when (z,τ,p)∈C(E+,τk,2πω0).
The above discussion shows that C(E+,τk,2πω0) is bounded, which contradicts Lemma 5.1. Therefore, ProjτC(E+,τk,2πω0) contains at least an interval (τk,+∞). The proof is complete.
Based on the previous discussion, three numerical results of system (1.2) are presented.
Case 1. We consider system (1.2) with s=0.8, m=0.2, α=1, β=0.6, δ=0.4, n=0.8, that is
{dudt=u(t)(1−u(t))−0.8u(t)v(t)0.2+0.6+u(t),dvdt=0.4v(t)(0.8+0.6−v(t−τ)0.6+u(t−τ)). | (6.1) |
We have A=0.4400, B=0.0640 and √A2−4B=−0.624. It implies system (6.1) has no positive equilibrium and satisfies the condition (ii) in Theorem 3.2. Therefore, the prey-extinction equilibrium E2=(0,1.08) is stable when τ<τ∗0, where τ∗0=2.1817. When τ pass through the critical value τ∗0, E2 loses its stability(see Figure 1).
Case 2. we discuss the following system
{dudt=u(t)(1−u(t))−0.2u(t)v(t)0.1+0.6+u(t),dvdt=0.3v(t)(0.4+0.6−v(t−τ)0.6+u(t−τ)), | (6.2) |
where s=0.2, m=0.1, α=1, β=0.6, δ=0.3, n=0.4. Calculate directly, we have B=−0.5320, G=0.7641, D=0.2444, then there is a unique coexisting equilibrium E+=(0.8476,1.1791). From the condition (ii) in Theorem 3.5, there exists the critical value τ0=6.0476 such that E+ is stable when τ<τ0 and unstable when τ>τ0(see Figure 2).
On the other hand, we obtain the following values by using Matlab
α1=−6.9759−2.3462i, α2=2.7585+7.7731i, M=−0.0090−0.0008i,
g20=−14.4293−14.6239i, g02=19.8351−5.9477i, g11=0.6641−2.1798i,
g21=−41.2805+50.5284i, c1(0)=−27.6348−37.3972i, λ′(τ0)=0.0192−0.0298i.
It follows that μ2=1440.4>0 and β2=−55.2696<0 and T2=312.4076, which, together with Theorem 4.1, implies that the bifurcating periodic solution exists when τ>τ0 and the bifurcating periodic solution is stable on the center manifold and the period increases.
When τ=τ0, all roots of characteristic equation (3.3) have negative real parts except ±iω0. Since the periodic solution on the center manifold is stabile, the periodic solution in the whole phase space is stable(see Figure 2 (c) and (d)).
Case 3. the simulation of the system (1.2) with s=0.6, m=0.2, α=0.3, β=0.5, δ=0.65, n=0.6 is given by
{dudt=u(t)(1−u(t))−0.6u(t)v(t)0.2+0.15+u(t),dvdt=0.65v(t)(0.6+0.5−v(t−τ)0.15+u(t−τ)), | (6.3) |
which has two positive equilibriums E−=(0.0145,0.5987) and E+=(0.2755,0.7553) due to A=−0.2900, B=0.0040 and √A2−4B=0.2610. By the condition (i) and (ii) in Theorem 3.5, we can know that E−=(0.0145,0.5987) is always unstable for any τ>0 and there exists the critical value τ0=1.2379 such that E+ is stable for any τ∈[0,τ0). When τ crosses τ0, E+ is unstable and a Hop bifurcation occurs. We obtain c1(0)=−23.7642+34.8292i and λ′(τ0)=0.5009−0.6450i by using Matlab, then μ2=47.4457>0 and β2=−47.5284<0 and T2=−3.6496, which implies that the bifurcating periodic solution is stable on the center manifold and the period increases. As discussed in system (6.2), for τ>τ0, the periodic solution of system (6.3) in the whole phase space is stable. The corresponding simulation results are shown in Figure 3.
In addition, system (6.3) also has a prey-extinction equilibrium E2=(0,0.59) corresponding to the critical value τ∗0=0.6144. Thus, E2=(0,0.59) and E+=(0.2755,0.7553) are the stability equilibria of system (6.3) when τ<0.6144, which is depicted in Figure 4.
In the paper, we investigate a delayed predator-prey system with additional food and asymmetric functional response. The local stability of all possible equilibria are studied. It shows that we can exterminate the prey by adjusting the quality and quantity of additional food when the prey density and time delay are relatively small. We know that the number of positive equilibria is determined by the value of B. For B<0, there is only one conditionally stable or unstable coexisting equilibrium E+, which depends on the delay. However, there exist an absolutely unstable coexisting equilibriumE− and a conditionally stable or unstable coexisting equilibrium E+ for B>0 in system (1.2). We also find that the model is bistable when B>0, A<0 and A2−4B>0 (see Figure 4).
Our investigation shows that coexisting equilibrium E+ is always unstable after τ passes through the first critical value τ0. That is to say, there does not exist any stability switching. However, stability switching can occurs in some predator-prey systems (see e.g. [26,29,33]). Moreover, the formulas determining the direction (μ2) and stability (β2) of Hopf bifurcation are given. We also show that the local Hopf bifurcation implies the global Hopf bifurcation of positive equilibrium after the second critical value of delay. Finally, we give three examples to illustrate the stability of the system (1.2) near the first critical value, and the simulation results are consistent with our conclusions.
This paper mainly considers the effects of providing additional food and designing a delayed feedback on Holling-Tanner model theoretically. We cannot claim that our method always holds for Holling-Tanner model due to the variety of ways to provide additional food and design delayed feedback. However, our investigation has potential significance for biological control. Therefore, the future works may consider how does the different ways of additional food and delayed feedback affect a predator-prey system, and cover the effects of additional food on an ecoepidemic model and so forth.
The authors thank the editor and referees for their valuable suggestions and comments, which improved the presentation of this manuscript.
For the publication of this article, no conflict of interests among the authors is disclosed.
[1] |
S. Tang, L. Chen, Global attractivity in a "food-limited" population model with impulsive effects, J. Math. Anal. Appl., 292 (2004), 211–221. doi: 10.1016/j.jmaa.2003.11.061
![]() |
[2] |
G. Saunders, B. Cooke, K. McColl, R. Shine, T. Peacock, Modern approaches for the biological control of vertebrate pests: an Australian perspective, Biol. Control, 52 (2010), 288–295. doi: 10.1016/j.biocontrol.2009.06.014
![]() |
[3] | B. Ghosh, F. Grognard, L. Mailleret, Natural enemies deployment in patchy environments for augmentative biological control, Appl. Math. Comput., 266 (2015), 982–999. |
[4] |
A. Tena, A. Pekas, D. Cano, F. Wackers, A. Urbaneja, Sugar provisioning maximizes the biocontrol service of parasitoids, J. Appl. Ecol., 52 (2015), 795–804. doi: 10.1111/1365-2664.12426
![]() |
[5] |
P. Srinivasu, B. Prasad, M. Venkatesulu, Biological control through provision of additional food to predators: a theoretical study, Theor. Popul. Biol., 72 (2007), 111–120. doi: 10.1016/j.tpb.2007.03.011
![]() |
[6] |
B. Sahoo, S. Poria, The chaos and control of a food chain model supplying additional food to top-predator, Chaos Soliton. Fract., 58 (2014), 52–64. doi: 10.1016/j.chaos.2013.11.008
![]() |
[7] |
T. Kar, B. Ghosh, Sustainability and optimal control of an exploited prey predator system through provision of alternative food to predator, Biosystems, 109 (2012), 220–232. doi: 10.1016/j.biosystems.2012.02.003
![]() |
[8] |
P. Srinivasu, B. Prasad, Role of quantity of additional food to predators as a control in predator prey systems with relevance to pest management and biological conservation, Bull Math. Biol., 73 (2011), 2249–2276. doi: 10.1007/s11538-010-9601-9
![]() |
[9] | J. Song, M. Hu, Y. Bai, Y. Xia, Dynamic analysis of a non-autonomous ratio-dependent predator-prey model with additional food, J. Appl. Anal. Comput., 8 (2018), 1893–1909. |
[10] | B. Sahoo, S. Poria, Effects of additional food on an ecoepidemic model with time delay on infection, Appl. Math. Comput., 245 (2014), 17–35. |
[11] | B. Sahoo, Role of additional food in eco-epidemiological system with disease in the prey, Appl. Math. Comput., 259 (2015), 61–79. |
[12] |
A. Basheer, E. Quansah, R. Parshad, The effect of additional food in Holling Tanner type models, Int. J. Dyn. Control, 7 (2019), 1195–1212. doi: 10.1007/s40435-019-00580-3
![]() |
[13] |
Y. Xia, Global analysis of an impulsive delayed Lotka-Volterra competition system, Commun. Nonlinear Sci. Numer. Simul., 16 (2011), 1597–1616. doi: 10.1016/j.cnsns.2010.07.014
![]() |
[14] | X. Meng, L. Chen, Permanence and global stability in an impulsive lotka-volterra n-species competitive system with both discrete delays and continuous delays, Int. J. Biomath., 1 (2012), 179–196. |
[15] |
X. Meng, L. Chen, Almost periodic solution of non-autonomous Lotka-Volterra predator-prey dispersal system with delays, J. Theoret. Biol., 243 (2006), 562–574. doi: 10.1016/j.jtbi.2006.07.010
![]() |
[16] |
Z. Wei, Y. Xia, T. Zhang, Stability and bifurcation analysis of a amensalism model with weak Allee effect, Qual. Theory Dyn. Syst., 19 (2020), 23. doi: 10.1007/s12346-020-00341-0
![]() |
[17] |
K. Kou, Y. Lou, Y. Xia, Zeros of a Class of Transcendental Equation with Application to Bifurcation of DDE, Int. J. Bifurcat. Chaos, 26 (2016), 1650062. doi: 10.1142/S0218127416500620
![]() |
[18] |
R. Xu, Z. Ma, Stability and Hopf bifurcation in a ratio-dependent predator-prey system with stage structure, Chaos Soliton. Fract., 38 (2008), 669–684. doi: 10.1016/j.chaos.2007.01.019
![]() |
[19] |
A. Maiti, B. Dubey, A. Chakraborty, Global analysis of a delayed stage structure prey-predator model with Crowley-Martin type functional response, Math. Comput. Simulat., 162 (2019), 58–84. doi: 10.1016/j.matcom.2019.01.009
![]() |
[20] |
Y. Chen, F. Zhang, Dynamics of a delayed predator-prey model with predator migration, Appl. Math. Model., 37 (2013), 1400–1412. doi: 10.1016/j.apm.2012.04.012
![]() |
[21] | T. Kar, U. Das, Bifurcation analysis of a delayed predator-prey model with Holling type III functional response and predator harvesting, J. Nonlinear Dyn., 2014 (2014), 1–10. |
[22] | J. Jiao, Y. Song, Stability and bifurcation analysis of a delayed Leslie-Gower predator-prey system with nonmonotonic functional response, Abstr. Appl. Anal., 2013 (2013), 152459. |
[23] |
Y. Song, Y. Tao, H. Shu, Dynamics of a ratio-dependent stage-structured predator-prey model with delay, Math. Methods Appl. Sci., 40 (2017), 6451–6467. doi: 10.1002/mma.4467
![]() |
[24] | X. Jiang, X. Chen, T. Huang, H. Yan, Bifurcation and Control for a Predator-Prey System With Two Delays, IEEE T. Circuits-II, 68 (2021), 376–380. |
[25] | X. Jiang, X. Chen, M. Chi, J. Chen, On Hopf bifurcation and control for a delay systems, Appl. Math. Comput., 370 (2019), 124906. |
[26] |
B. Barman, B. Ghosh, Explicit impacts of harvesting in delayed predator-prey models, Chaos Soliton. Fract., 122 (2019), 213–228. doi: 10.1016/j.chaos.2019.03.002
![]() |
[27] |
L. Erbe, K. Geba, W. Krawcewicz, J. Wu, S1-degree and global Hopf bifurcations, J. Differ. Equations, 98 (1992), 277–298. doi: 10.1016/0022-0396(92)90094-4
![]() |
[28] |
J. Wu, Symmetric functional differential equations and neural networks with memory, Trans. Amer. Math. Soc., 350 (1998), 4799–4838. doi: 10.1090/S0002-9947-98-02083-2
![]() |
[29] |
Y. Song, J. Wei, Local Hopf bifurcation and global periodic solutions in a delayed predator-prey system, J. Math. Anal. Appl., 301 (2005), 1–21. doi: 10.1016/j.jmaa.2004.06.056
![]() |
[30] |
S. Yuan, F. Zhang, Stability and global Hopf bifurcation in a delayed predator-prey system, Nonlinear Anal. Real World Appl., 11 (2010), 959–977. doi: 10.1016/j.nonrwa.2009.01.038
![]() |
[31] |
S. Yuan, Y. Song, Stability and Hopf bifurcations in a delayed Leslie-Gower predator-prey system, J. Math. Anal. Appl., 355 (2009), 82–100. doi: 10.1016/j.jmaa.2009.01.052
![]() |
[32] | P. Hao, X. Wang, J. Wei, Global Hopf bifurcation of a population model with stage structure and strong Allee effect, Discrete Contin. Dyn. Syst. Ser. S, 10 (2017), 973–993. |
[33] | X. Yan, W. Li, Hopf bifurcation and global periodic solutions in a delayed predator-prey system, Appl. Math. Comput., 177 (2006), 427–445. |
[34] |
X. Liu, T. Zhang, X. Meng, T. Zhang, Turing-Hopf bifurcations in a predator-prey model with herd behavior, quadratic mortality and prey-taxis, Phys. A, 496 (2018), 446–460. doi: 10.1016/j.physa.2018.01.006
![]() |
[35] | B. Hassard, N. Kazarinoff, Y. Wan, Theory and Applications of Hopf Bifurcation, Cambridge: Cambridge Univ. Press, 1981. |
1. | Osmin Ferrer, José Guerra, Alberto Reyes, Transcritical bifurcation in a multiparametric nonlinear system, 2022, 7, 2473-6988, 13803, 10.3934/math.2022761 | |
2. | Ting Yu, Qinglong Wang, Shuqi Zhai, Exploration on dynamics in a ratio-dependent predator-prey bioeconomic model with time delay and additional food supply, 2023, 20, 1551-0018, 15094, 10.3934/mbe.2023676 | |
3. | XINYU BO, GUANGYING LV, WENJUN LIU, ALI MOUSSAOUI, OPTIMAL HARVEST CONTROL OF PREDATOR–PREY SYSTEMS IN FISHERIES WITH STAGE STRUCTURE AND ADDITIONAL FOOD SUPPLY, 2025, 33, 0218-3390, 221, 10.1142/S0218339025500032 |