Plan
Comptes Rendus

Population biology/Biologie des populations
Oscillations for a delayed predator–prey model with Hassell–Varley-type functional response
Comptes Rendus. Biologies, Volume 338 (2015) no. 4, pp. 227-240.

Résumé

In this paper, a delayed predator–prey model with Hassell–Varley-type functional response is investigated. By choosing the delay as a bifurcation parameter and analyzing the locations on the complex plane of the roots of the associated characteristic equation, the existence of a bifurcation parameter point is determined. It is found that a Hopf bifurcation occurs when the parameter τ passes through a series of critical values. The direction and the stability of Hopf bifurcation periodic solutions are determined by using the normal form theory and the center manifold theorem due to Faria and Maglhalaes (1995). In addition, using a global Hopf bifurcation result of Wu (1998) for functional differential equations, we show the global existence of periodic solutions. Some numerical simulations are presented to substantiate the analytical results. Finally, some biological explanations and the main conclusions are included.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2015.01.002
Mots clés : Predator-prey model, Hassell–Varley functional response, Stability, Hopf bifurcation
Changjin Xu 1 ; Peiluan Li 2

1 Guizhou Key Laboratory of Economics System Simulation, School of Mathematics and Statistics, Guizhou College of Finance and Economics, Guiyang 550004, China
2 School of Mathematics and Statistics, Henan University of Science and Technology, Luoyang, Henan 471023, China
@article{CRBIOL_2015__338_4_227_0,
     author = {Changjin Xu and Peiluan Li},
     title = {Oscillations for a delayed predator{\textendash}prey model with {Hassell{\textendash}Varley-type} functional response},
     journal = {Comptes Rendus. Biologies},
     pages = {227--240},
     publisher = {Elsevier},
     volume = {338},
     number = {4},
     year = {2015},
     doi = {10.1016/j.crvi.2015.01.002},
     language = {en},
}
TY  - JOUR
AU  - Changjin Xu
AU  - Peiluan Li
TI  - Oscillations for a delayed predator–prey model with Hassell–Varley-type functional response
JO  - Comptes Rendus. Biologies
PY  - 2015
SP  - 227
EP  - 240
VL  - 338
IS  - 4
PB  - Elsevier
DO  - 10.1016/j.crvi.2015.01.002
LA  - en
ID  - CRBIOL_2015__338_4_227_0
ER  - 
%0 Journal Article
%A Changjin Xu
%A Peiluan Li
%T Oscillations for a delayed predator–prey model with Hassell–Varley-type functional response
%J Comptes Rendus. Biologies
%D 2015
%P 227-240
%V 338
%N 4
%I Elsevier
%R 10.1016/j.crvi.2015.01.002
%G en
%F CRBIOL_2015__338_4_227_0
Changjin Xu; Peiluan Li. Oscillations for a delayed predator–prey model with Hassell–Varley-type functional response. Comptes Rendus. Biologies, Volume 338 (2015) no. 4, pp. 227-240. doi : 10.1016/j.crvi.2015.01.002. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2015.01.002/

Version originale du texte intégral

1 Introduction

Since the pioneering work of Volterra [1] and Lotka [2] in the mid-1920s, there has been increasing interest in investigating the dynamical behaviors of predator–prey models in both ecology and mathematical ecology [3–11]. In particular, one of the important dynamical predator–prey behaviors, such as periodic phenomena and bifurcation has become even more interesting [6–10,12–24]. In 1980, Freeman [25] proposed a most popular predator–prey model with Michaelis–Menten-type functional response:

dx1dt=rx11x1Kcx1x2m+x1,dx2dt=x2d+fx1m+x1,x0>0,y0>0,(1)
where x1, x2 denote the population of preys and predators at time t, respectively. r, K, c, m, d, and f are positive constants that denote the prey's intrinsic growth rate, carrying capacity, capturing rate, half-saturation constant, predator death rate, maximal predator growth rate, respectively. For more details about the model, the reader is referred to [25].

Considering that in many situations, predators must search and share or compete for food, Arditi and Ginzburg [3] introduced and studied the following ratio-dependent-type functional response model:

dx1dt=rx11x1Kcx1x2mx2+x1,dx2dt=x2d+fx1mx2+x1,x0>0,y0>0(1.2)

Since the functional response depends on the predator density in a different way, Hassel and Varley [26] reconstructed the predator–prey model with Hassell–Varly-type functional response, which takes the following form:

dx1dt=rx11x1Kcx1x2mx2γ+x1,dx2dt=x2d+fx1mx2γ+x1,x0>0,y0>0,(1.3)
where γ ∈ (0,1) is called the HV constant. Generally, the consumptions of prey by the predator throughout its past history governs the present birth rate of the predator. Motivated by this point of view, Wang [27] introduced and investigated the periodic solutions to the following delayed predator–prey model:
dx1tdt=x1tatbtx1tτtctx2tmx2γt+x1t,dx2tdt=x2tdt+rtx1tmx2γ+x1,x00,y0>0(1.4)
with the following initial condition:
x1t=φθ,θ[δ,0],φ0=φ0>0,x2t=ψθ,θ[δ,0],ψ0=ψ0>0,(1.5)
where δ=supt[0,ω]τt,φ,ψCδ,0 with the norm x=suptδ,0xt. It is worth pointing out that during the course of the predator–prey interaction when predators do not form groups, one can assume that the HV constant is equal to 1, that is, γ = 1. Moreover, it is more reasonable to incorporate the delay into Hassell–Varly-type functional response. From the point of view of biology, we will consider the following model with delayed Hassell–Varly-type functional response:
dx1dt=x1abx1tτcx2tτmx2tτ+x1tτ,dx2dt=x2d+rx1tτmx2tτ+x1tτ(1.6)

In this paper, we will devote our attention to investigating the properties of a Hopf bifurcation of system (1.6), that is to say, we shall take the delay τ as the bifurcation parameter and show that when τ passes through a certain critical value, the positive equilibrium loses its stability and a Hopf bifurcation will take place. Furthermore, when the delay τ takes a sequence of critical values containing the above critical value, the positive equilibrium of system (1.6) will undergo a Hopf bifurcation. In particular, by using the normal form theory and the center manifold reduction due to Faria and Maglhalaes [28], the formulae for determining the direction of Hopf bifurcations and the stability of bifurcating periodic solutions are obtained. In addition, the existence of periodic solutions for τ far away from the Hopf bifurcation values is also established by means of the global Hopf bifurcation result of Wu [29].

In order to obtain the main results of our paper, throughout this paper, we assume that the coefficients of system (1.6) satisfy the following condition:

H1

am2 + cd − cr > 0, r > d

This paper is organized as follows. In Section 2, the stability of the positive equilibrium and the existence of a Hopf bifurcation at the positive equilibrium are studied. In Section 3, the direction of Hopf bifurcation and the stability of bifurcating periodic solutions on the center manifold are determined. In Section 4, numerical simulations are carried out to illustrate the validity of the main results. In Section 5, some conditions that guarantee the global existence of the bifurcating periodic solutions to the model are given. Biological explanations and some main conclusions are drawn in Section 6.

2 Stability of the equilibrium and existence of the local Hopf bifurcation

In the section, by analyzing the characteristic equation of the linearized system of system (1.6) at the positive equilibrium, we investigate the stability of the positive equilibrium and the existence of the local Hopf bifurcations occurring at the positive equilibrium.

Considering the biological meaning, we only study the property of a unique positive equilibrium (i.e., coexistence equilibrium). It is easy to see that under the hypothesis (H1), system (1.6) has a unique positive equilibrium E*x1*,x2*, where

x1*=am2+cdcrabm,x2*=rdam2+cdcrabdm2

Let u1t=x1tx1*,u2t=x2tx2*, then, system (1.6) takes the following form:

du1dt=m1u1tτ+m2u2tτ+m3u12tτ+m4u22tτ   +m5u1tτu2tτ+m6u1(t)u1tτ+m7u1tu2tτ   +m8u1tu12tτ+m9u1tu22tτ+h.o.t.,du2dt=n1u1tτ+n2u2tτ+n3u12tτ+n4u22tτ      +n5u1tτu2tτ+n6u1tτu2t+n7u2tu2(tτ)      +n8u1tτu22tτ+n9u12tτu2tτ+n10u2tu22tτ      +n11u1tτu2(t)u2tτ+n12u12tτu2t+h.o.t.,(2.1)
where
m1=cx2*mx2*+x1*bx1*,m2=mcx2*(mx2*+x1*)2cmx2*+x1*x1*,m3=cx1*x2*mx2*+x1*3,m4=mcx1*mx2*+x1*2m2cx2*mx2*+x1*3,m5=cmx2*+x1*22mcx2*mx2*+x1*2,m6=cx2*mx2*+x1*b,m7=mcx2*mx2*+x1*2cmx2*+x1*,m8=cx2*mx2*+x1*3,m9=mcmx2*+x1*2m2cx2*mx2*+x1*3,n1=rx2*mx2*+x1*rx1*x2*mx2*+x1*2,n2=mrx1*x2*mx2*+x1*2,n3=rx2*mx2*+x1*2,n4=m2rx1*x2*mx2*+x1*3,n5=2mrx1*x2*mx2*+x1*3mrx2*mx2*+x1*2,n6=rmx2*+x1*rx1*mx2*+x1*2,n7=mrx1*mx2*+x1*2,n8=m2rx2*mx2*+x1*3,n9=2mrx2*mx2*+x1*3,n10=m2rx1*mx2*+x1*3,n11=2mrx1*mx2*+x1*3mrmx2*+x1*2,n12=rmx2*+x1*2

Then, we obtain the linearized system of (2.1)

du1dt=m1u1tτ+m2u2tτ,du2dt=n1u1tτ+n2u2tτ(2.2)
with characteristic equation:
λ2m1+n2λ eλτ+m1n2m2n1 e2λτ=0(2.3)

In order to investigate the distribution of roots of the transcendental equation (2.3), the following Lemma stated in [30] is helpful.

Lemma 2.1.[30]For the transcendental equation

Pλ,eλτ1,,eλτm=λn+p10λn1++pn10λ+pn0+p11λn1++pn11λ+pn1 eλτ1++p1mλn1++pn1mλ+pnm eλτm=0,
as τ1,τ2,τ3,,τm vary, the sum of orders of the zeros of Pλ,eλτ1,,eλτm in the open right half plane can change, and only a zero appears on or crosses the imaginary axis.

Regarding τ as the parameter, we can apply Lemma 2.1 to (2.3), which is a special case of

Pλ,eλτ1,,eλτm

Obviously, if m1n2 ≠ m2n1, then, λ = 0 is not a root of (2.3). For τ = 0 the characteristic equation (2.3) becomes:

λ2m1+n2λ+m1n2m2n1=0(2.4)

It is easy to see that Eq. (2.4) have two negative real roots if the following condition:

H2

m1 + n2 < 0, m1n2 − m2n1 > 0

holds.

Multiplying eλτ on both sides of (2.3), it is obvious to obtain:

λ2eλτm1+n2λ+m1n2m2n1 eλτ=0(2.5)

For ω0 > 0, iω0 is a root of (2.5) if and only if

ω02 eiω0τm1+n2iω+m1n2m2n1 eiω0τ=0(2.6)

Separating the real and imaginary parts of (2.6), we get:

m1n2m2n1ω02cosω0τ=0,ω02+m1n2m2n1sinω0τ=m1+n2ω0(2.7)

If the condition (H2) holds, we can easily check that cosω0τ ≠ 0. Then, it follows from (2.7) that:

m1n2m2n1=ω02
which leads to:
ω0=m1n2m2n1(2.8)

From the second equation of (2.7), we can easily obtain:

τk=1ω0arcsin(m1+n2)m1n2m2n12(m1n2m2n1)+2kπ,k=0,1,2,.(2.9)

From (2.7), we know that (2.3) has a simple pair of purely imaginary roots ±iω0 at τkk=0,1,2,

Let λkτ=αkτ+iωkτ be the root of Eq. (2.3) satisfying αkτk=0,ωkτk=ω0

Assume that

H3

4m1n2m2n12+m1+n24m1n2m2n12m1+n22>2m1+n22

Then, the following transversality condition holds.

Lemma 2.2. If ( H1 ), ( H2 ) and ( H3 ) are satisfied, then d α k τ d τ τ = τ k > 0

Proof. Differentiating the equation (2.3) with respect to τ leads to:

dλdτ1=2λ eλτ+m1+n2τ2m1n2m2n1τ eλτ2m1n2m2n1τ eλτλm1+n2λ

When τ = τk, iω0 is a purely imaginary root of (2.3). We then easily get:

dαkτdτ1|τ=τk=4ω04cos2ω0τk2ω03m1+n2cosω0τk[2ω03sinω0τk]2+[2ω03cosω0τkm1+n2ωk]2(2.10)

From (2.7), we have:

sinω0τk=m1+n22ω0,cosω0τk=±4ω02m1+n222ω0

Hence,

dαkτdτ1|τ=τk=ω024m1n2m2n12m1+n2222ω03sinω0τk2+2ω03cosω0τkm1+n2ωk2±ω024m1n2m2n1m1+n222ω03sinω0τk2+2ω03cosω0τkm1+n2ωk2

Under the condition (H3), we know that

dαkτdττ=τk>0,
completing the proof. Lemma 2.2 implies that the roots λk (τ) of characteristic equation (2.3) near τk crosses the imaginary axis from the left to the right as τ continuously varies from a number less than τk to one greater than τk by Rouché’s theorem [31].

Applying Lemma 2.1, we obtain the following results: Lemma 2.3. If (H1), (H2) and (H3) are satisfied, then

(i) if τ [ 0 , τ 0 , all roots Eq. (2.3) have negative real parts;

(ii) if τ = τ0, all roots of Eq. (2.3) except ±iω0 have negative real parts;

(iii) if τ ∈ [τk, τk+1) for k = 0, 1, 2, ..., Eq. (2.3) have 2 (k + 1) roots with positive real parts.

Spectral properties of Eq. (2.3) immediately lead to the properties of the zero solutions to system (2.2), and equivalently, of the positive equilibrium E for system (1.6).

Theorem 2.4.Suppose that (H1), (H2) and (H3) are satisfied. Then, the positive equilibrium Eof system(1.6)is asymptotically stable whenτ[0,τ0, and unstable when τ > τ0. Moreover, at τ = τk, k = 0, 1, 2, ..., ± 0 are a simple pair imaginary roots of (2.3), and (1.6) undergoes a Hopf bifurcation near E.

3 Direction and stability of the Hopf bifurcation

In the previous section, we have obtained conditions for Hopf bifurcations to occur when τ = τk. In this section, we will employ the algorithm of Faria and Maglhalaes [28] to compute explicitly the normal forms of system (1.6) on the center manifold. After that, we will investigate the direction of the Hopf bifurcation and stability of the bifurcating periodic orbits from the positive equilibrium E of system (1.6) at these critical values of τk. We know that Eq. (2.3) has a pair of purely imaginary roots ±iω0 when τ = τk and system (1.6) undergoes a Hopf bifurcation from E. Let μ = τ–τk, then μ = 0 is the Hopf bifurcation value of system (1.6).

Throughout this section, we refer to Faria and Maglhalaes [28] for the meaning of the notations involved.

Normalizing the delay τ by the time-scaling t → t/τ, then the system (2.1) can be rewritten as a functional differential equation in C[1,0],R2:

du1dt=τ[m1u1t1+m2u2t1+m3u12t1+m4u22t1   +m5u1t1u2t1+m6u1tu1t1+m7u1tu2t1   +m8u1tu12t1+m9u1tu22t1+h.o.t.],   du2dt=τ[n1u1t1+n2u2t1+n3u12t1+n4u22t1      +n5u1t1u2t1+n6u1t1u2t+n7u2tu2t1      +n8u1t1u22t1+n9u12t1u2t1+n10u2tu22t1      +n11u1t1u2(t)u2t1+n12u12t1u2t+h.o.t.](3.1)

Let u = (u1(t), u2(t))T, then, system (3.1) can be rewritten in the following vector form:

dudt=Lτut+Fut,τ,(3.2)
where
Lτϕ=τm1ϕ11+m2ϕ21n1ϕ11+n2ϕ21,
Fϕ,τ=τF1ϕ,τF2ϕ,τ
where ϕ=ϕ1,ϕ2TC1,0,R2, and
F1ϕ,τ=m3ϕ121+m4ϕ221+m5ϕ11u21+m6ϕ10ϕ11+m7ϕ10ϕ21+m8ϕ10ϕ121+m9ϕ10ϕ221+h.o.t.,F2ϕ,τ=n3ϕ121+n4ϕ221+n5ϕ11ϕ21+n6ϕ11ϕ20+n7ϕ20ϕ21+n8ϕ11ϕ221+n9ϕ121ϕ21+n10ϕ20ϕ221+n11ϕ11ϕ20ϕ21+n12ϕ121ϕ20+h.o.t..

Obviously, L(τ) is a continuous linear function mapping C1,0,R2 into R2. By the Riesz representation theorem, there exists a 2 × 2 matrix function ηθ,τ,1θ0, whose elements are of bounded variation such that

ϕΦΨ,ϕ(3.3)

In fact, we can choose

ηθ,τ=τm1m2n1n2δθ+1,(3.4)
where δ denotes Dirac delta function. Then, (3.3) is satisfied. If ϕ is any a given function in C1,0,R2 and u(ϕ) is the unique solution to the linearized equation dudt=Lτut of Eq. (3.2) with the initial function ϕ at zero, then the solution operator Tt:C1,0,R2C1,0,R2 is defined by
Ttϕ=utϕ,t0(3.5)

From Lemma 7.1.1 in Hale [32], we know that Tt,t0, is a strongly continuous semigroup of linear transformation on 0, and the infinitesimal generator A(τ) of T(t), t ≥ 0 is given by

Aτϕθ=ϕ˙θ+X0θLτϕϕ˙0   for   ϕC1,0,R2,(3.6)
where
X0θ=0,1θ<0,I,θ=0(3.7)

For ψC1,0,R2*, define

A*ψs=ψ˙s+X0s10ψtdηt,τk+ψ˙0(3.8)
and a linear inner product
<ψ,ϕ>=ψ¯0ϕ010ξ=0θψTξθdηθϕξdξ,(3.9)
where ηθ=ηθ,0 the A=A0 and A* are adjoint operators. By the discussions in Section 2, we know that ±iω0τk are eigenvalues of A(0), and they are also eigenvalues of A* corresponding to iω0τk and –iω0τk, respectively. Let Λ=iω0τk,iω0τk and denote by P the invariant space of A(τk) associated with Λ, where the dimension of P equals to 2. Now, we can decompose the space CC1,0,R2 as C = PQ by applying the formal adjoint theory for functional differential equations in [32].

Consider the complex coordinates and still denote C1,0,R2 as C. Suppose that Φ = (Φ1, Φ2) is a basis of P and

Φ1(θ)=eiω0τkθ(1,ξ,)T,Φ2(θ)=Φ1(θ)¯,1θ0,
where
ξ=iω0 eiω0τkm1m2.

Also, the two eigenvectors Ψ1, Ψ2 of A* corresponding to the eigenvalues iω0τk, –iω0τk, respectively, construct a basis Ψ = (Ψ1, Ψ2)T of the adjoint space P* of P and

Ψ1θ=Eeiω0τks1,ηT,Ψ2θ=Ψ1θ¯,0s1,
where
η=iω0 eiω0τk+m1n1,
E=11+ξ¯η+m1 eiω0τk+m2η eiω0τk+m2 eiω0τk+n2ξ¯η eiω0τk.

Thus, Ψ,Φ=Ψj,Φi,i,j=1,2=I2, where I2 is a second-order identical matrix. It is known that Φ˙=ΦB, where

B=iω0τk00iω0τk      .

Take the enlarged phase space C by considering the space BCϕ:1,0C2|ϕ   is   continuous   on   1,0   and   limθ0ϕθ   exists. The projection ϕΦΨ,ϕ of C upon P, associated with the decomposition C = PQ, is now replaced by π : BC → P such that πϕ+X0α=ΦΨ,ϕ+Ψ0α.

Thus, we have the decomposition BC = PKerπ. Using the decomposition ut=ϕxt+yt,xtC2,ytKerπC1=Q1, and by Theorem 7.6.1 in [32], we can decompose (3.2) as

dxdt=Bx+ψ0F0Φx+y,μ,dydt=Aτ˜Q1y+IπX0F0Φx+y,μ,(3.10)
where F0ϕ,μ=Lμϕ+Fϕ,τk+μ. In view of the Taylor expansion, we denote, respectively, Ψ0F0Φx+y,μ and IπX0F0Φx+y,μ as:
Ψ0F0Φx+y,μ=12f21x,y,μ+13f31x,y,μ+,   IπX0F0Φx+y,μ=12f22x,y,μ+13f32x,y,μ+,(3.11)
where fj1x,y,μ and fj2x,y,μ are homogeneous polynomials in (x, y, μ) of degree j, j = 2, 3,…, with coefficients in C2 and Kerπ, respectively. The normal form method gives a normal form on the center manifold of the origin for (3.10) as
x˙=Bx+12g21x,0,μ+13g31x,0,μ+,(3.12)
where gj1x,0,μ are homogeneous polynomials in (x, μ) of degree j, j = 2, 3,…

In what follows, we first define the operators Mj1 as

Mj1px,μ=Dxpx,μBxBpx,μ,j2(3.13)

In particular, Mj1μlxqek=iω0τkq1q2+1kμlxqek,   l+q1+q2=j,k=1,2   for   j=2,3,   q=q1,q2N02,   lN0   and   e1,e2 is the canonical basis for C2. Therefore, we have

KerM21=spanx1μ0,0x2μ,KerM31=spanx12x20,x1μ20,0x1x22,0x2μ2

By (3.10), we have

f21x,y,μ=2Ψ0LμΦx+y+F2Φx+y,τk(3.14)

Noting that L(μ) = μ / τkLk), we have

f21x,0,u=2A1x1μ+2A1μx2+a20x12+2a11x1x2+a02x222A¯1x1μ+2A¯2x2μx2+a¯20x12+2a¯11x1x2+a¯02x22(3.15)
where
A1=Em1 eiω0τk+m2 eiω0τkξ+ηn1 eiω0τk+n2 eiω0τkξ,A2=Em1 eiω0τk+m2 eiω0τkξ¯+ηn1 eiω0τk+n2 eiω0τkξ¯,a20=2Eτkm3 e2iω0τk+m4 e2iω0τkξ2+m5 e2iω0τkξ+m6 eiω0τk+m7 eiω0τkξ,a11=2Eτkm3+m4|ξ|2+m5Reξ+m6Reeiω0τk+m7 eiω0τkξ¯,a02=2Eτkm3 e2iω0τk+m4 e2iω0τk+m5 e2iω0τkξ¯+m6 eiω0τk+m7 eiω0τkξ¯(3.16)

Since the second-order terms in (μ, x) on the center manifold are given by

12g21x,0,μ=12ProjKerM21f21x,0,μ,(3.17)
it follows that
12g21x,0,μ=A1x1μA¯1x2μ,(3.18)
where A1 defined by (3.16).

In the following, we shall compute the cubic term g31x,0,μ. First, we note that

g31x,0,μKerM31=Spanx12x20,x1μ20,0x1x22,0x2μ2

However, the terms O0xμ2 are irrelevant to determine the generic Hopf bifurcation. Hence, it is only needed to compute the coefficients of x12x20 and 0x1x22. Let

sspanx1μ0,0x2μ,
then we have
13g31x,0,μ=13!ProjKerM31f¯31x,0,μ=13!Projsf¯31x,0,0+O0xμ2,
where
f¯31x,0,0=f31x,0,0+32Dxf21U21DxU21g21x,0,0+32Dyf21hx,0,0
is the third-order term of the equation, which is obtained after computing the second-order terms of the normal form, U21x,0 is the solution to equation M21U21x,0=f21x,0,0, and h = (h1, h2)T is a second homogeneous polynomial in (x1, x1, μ) with coefficients in Q1.

From (3.1) and (3.2) and from (3.18) we can easily see that g21x,0,0=0 and

Projsf31x,0,0=3a21x12x23a¯21x1x22,
where
a21=E2m8+2m9|ξ|2+Eη2n8|ξ|2 eiω0τk+2n9ξ eiω0τk+2n10|ξ|2ξ+n11|ξ|2 eiω0τk+ξ2+ξ¯+2n12ξ

Therefore, we have DxU21g21x,0,0=0. In the sequel, we only need to compute U21x,0 and h(x)(θ).

From (3.15), we have:

f21x,0,0=a20x12+2a11x1x2+a02x22a¯02x12+2a¯11x1x2+a¯20x22

In view of the definition of M21, the equation M21U21x,0=f21x,0,0 can be written as the following partial differential equations:

x1u1x1x2u1x2u1=1iω0τka20x12+2a11x1x2+a02x22,x1u2x1x2u2x2+u2=1iω0τka¯02x12+2a¯11x1x2+a¯20x22(3.19)

From (3.19), we can easily obtain:

U21x,0=1iω0τka20x122a11x1x213a02x221iω0τk13a¯02x12+2a¯11x1x2a¯20x22

Thus, we obtain:

ProjsDxf21U21x,0,0=2iω0τka20a112|a11|213|a02|2x12x22iω0τka¯20a¯112|a11|213|a02|2x12x2

In what follows, we shall compute ProjsDxf21hx,0,0. From (3.14), we know

f21x,y,0=2Ψ0FΦx+y,τk   =2τkEm3e12+m4e22+m5e1e2+m6e1e3+m7e2e3+Eηn3e12+n5e22+n6e1e4+n7e2e4E¯m3e12+m4e22+m5e1e2+m6e1e3+m7e2e3+E¯η¯n3e12+n5e22+n6e1e4+n7e2e4,(3.20)
where
e1=eiω0τkx1+eiω0τkx2+y11,e2=eiω0τkξx1+eiω0τkξ¯x2+y21,e3=x1+x2+y10,e4=x1+x2+y20
since h = (h(1), h(2))T is a second-order homogeneous polynomial in (x1, x2, μ) with coefficients in Q1. Hence, we can let:
h=h110x1x2+h101x1μ+h011x2μ+h200x12+h020x22+h002μ2(3.21)

Thus, from (3.20), we obtain

Dyf21hx,0,0,=2τkK1K2,
whereK1=E2m3e10h11+m5e20h11+e10h(1)0+m7e20h10+Eη2n3e10h11+n6e40h11+E2m4e20h21,+m5e10h21+m7e30h21+Eη2n5e20h2(1)+n6e10h20+n7e40h2(1)+e20h20,K2=E¯2m3e10h(1)1+m5e20h(1)1+e10h(1)1+m6e30h11+m7e20h10+E¯η¯2n3e10h1(1)+n6e40h11+E¯2m4e20h21+m5e10h21+m7e30h21+E¯η¯2n5e20h21+n6e10h20+n7(e40h21+e20h20,where
e10=eiω0τkx1+eiω0τkx2,e20=eiω0τkξx1+eiω0τkξ¯x2,e30=x1+x2,e40=x1+x2

Therefore,

Dyf21hx,0,0,=2B3x12x22B¯3x1x22,
where
B3=mEτkξ3h11010+ξ¯3h20010+h11030+h20030

Since h110(θ) and h200(θ) for θ1,0 appear in B3, we still need to compute them.

Following [28], we know that h = h (θ; x1, x2, μ) is the unique solution in the linear space of homogeneous polynomials of degree 2 in 3 real variable (x, μ) = (x1, x2, μ) of the equation

M22hx,μ=2IπX0LμΦx+FΦx,τk,(3.22)
since
M22hx,μ=Dxhx,μBxAτkQ1hx,μ(3.23)

Combining the definition (3.6) of the operator A(τ), we can obtain:

Dxhx,μBxh˙x,μX0θLτhx,μh˙x,μ0=2IπX0LμΦx+FΦx,τk

For the sake of simplicity, let:

h=h110θx1x2+h200θx12+h020θx22(3.24)

Then, h0 can be evaluated by the system

h˙0xDxh0x,μBx=2ΦΨ0L0Φx+FΦx,τk,(3.25)
h˙0xLτkh0x=2L0Φx+FΦx,τk,(3.26)
where h˙0 denote the derivative of h0 with respect to θ.

In view of (3.14), (3.15), (3.25) and (3.26), we know that h110=h1101,h1102T and h200=h2001,h2002T are the solution to the following two equations:

h˙110=2a11Φ1+a¯11Φ2,h˙1100Lτkh0x=4τkp1p2,(3.27)
h˙2002iω0τkh200=a20Φ1+a¯02Φ2,h˙2000Lτkh200=τkq1q2,(3.28)
respectively, where
p1=m3+m4ξ2+m5Reξ+m6Reeiω0τk+m7Reξ eiω0τkp2=n3+n4ξ2+n5Reξ+n6Reξ eiω0τk+n7Reξ eiω0τk,q1=m3 e2iω0τk+m4 e2iω0τkξ2+m5ξ+m6eiω0τk+m7ξ eiω0τk,q2=n3 e2iω0τk+n4 e2iω0τkξ2+n5ξ+n6ξ eiω0τk+n7ξ eiω0τk

Solving the Eq. (3.27) and (3.28), we have

h110θ=2iω0τka11Φ1a¯11Φ2+C,h200θ=a20iω0τkΦ1a¯203iω0τkΦ2+D e2iω0τk,(3.29)
where C = (c1, c2)T, D = (d1, d2)T are the solution to the following linear algebra equations:
m1 eiω0τkm2 eiω0τkn1 eiω0τkn2 eiω0τkc1c2=4p1p2,(3.30)
2iω0m1 e2iω0τkm2 e2iω0τkn1 e2iω0τk2iω0m2 e2iω0τkd1d2=q1q2,(3.31)
respectively. We know that
13g31x,0,0=A3x12x2A¯3x1x22,(3.32)
where
A3=i2ω0τka20a112a11213a02212B3(3.33)

Accordingly, the normal form (3.12) of (3.10) has the following form

=Bx+A1x1μA¯1x2μ+A3x12x2A¯3x1x22+Oxμ2+x4

The normal form (3.12) relative to P can be written in real coordinates (ω1, ω2) through the change of variables x1 = ω1 − iω2, x2 = ω1 + iω2. Setting ω1 = ρ cosν, ω2 = ρ sinν, then this form becomes

ρ˙=k1μρ+k2ρ3+Oμ2ρ+ρ,μ4,ν˙=σk+Oρ,μ,(3.34)
where k1=ReA1,k2=ReA3. Following [33], we know that the sign of k1k2 determines the direction of the Hopf bifurcation and the sign of k2 determines the stability of the nontrivial periodic solution bifurcating from the Hopf bifurcation. The Hopf bifurcation is supercritical when k1k2 < 0 and subcritical if k1k2 > 0. In addition, the bifurcating periodic solution on the center manifold is stable provided that k2 < 0 and unstable if k2 > 0.

Summarizing the above analysis, we have the following result.

Theorem 3.1.The flow of Eq.(3.2)with μ = 0 on the center manifold of the origin is given by (3.34). Hopf bifurcation is supercritical if k1k2 < 0 and subcritical if k1k2 > 0. Moreover, the nontrivial periodic solution is stable if k2 < 0 and unstable if k2 > 0.

4 Numerical examples

In this section, we present some numerical results for some particular values of the parameters associated with the model system (1.6). We consider the system (1.6) with a = 1, b = 2, c = 0.3, m = 0.5, d = 0.8, r = 2. That is,

dx1dt=x112x1tτ0.3x2tτ0.5x2tτ+x1tτ,dx2dt=x20.8+2x1tτ0.5x2tτ+x1tτ,(4.35)
which has a positive equilibrium E=0.32,0.96. By means of software Matlab, we obtain:
ω0=0.9254,τ0=1.940,ξ=10.20325.3126i,η=0.1073+0.2435i,E=0.01070.04131i,A1=0.1912+0.3103i,A2=0.29560.2391i,a20=0.74430.3476i,a11=0.02520.0395i,a02=0.70050.4669i,c1=1.0411+0.0012i,c2=0.35890.8993i,d1=2.1251+0.9696i,d2=0.4998+0.5933i,h11010=0.201310.0882i,h200(1)0=2.3266+7.0852i,B3=0.30920.5112i,A3=0.13771.3126i

Then, the stability determining quantities for Hopf bifurcating periodic solutions are given by k1 = 0.1912, k2 = −0.1377. From Theorem 2.4, we know that the positive equilibrium E0.32,0.96 is asymptotically stable when τ0,1.94 and is unstable when τ > 1.94. The numerical simulations for τ = 1.9 and τ = 2 are shown on Figs. 1–8, respectively. From Theorem 3.1, we know that system (4.35) with τ = 1.94 has a supercritical Hopf bifurcation and the nontrivial periodic solution bifurcating from the Hopf bifurcation of (4.35) with τ = 1.94 is orbitally asymptotically stable on the center. Moreover, all the roots of Eq. (2.3) with τ = 1.94, except ±0.9254i, have negative real parts. Thus, the center manifold theory implies that the stability of the periodic solutions projected on the center manifold coincides with the stability of the periodic solutions in the whole phase space.

Fig. 1

Trajectory portrait and phase portrait of system (4.35) with τ = 1.9 < τ0 ≈ 1.94. The positive equilibrium E0.32,0.96 is asymptotically stable. The initial value is (0.2, 1.5).

Fig. 2

Trajectory portrait and phase portrait of system (4.35) with τ = 1.9 < τ0 ≈ 1.94. The positive equilibrium E0.32,0.96 is asymptotically stable. The initial value is (0.2, 1.5).

Fig. 3

Trajectory portrait and phase portrait of system (4.35) with τ = 1.9 < τ0 ≈ 1.94. The positive equilibrium E0.32,0.96 is asymptotically stable. The initial value is (0.2, 1.5).

Fig. 4

Trajectory portrait and phase portrait of system (4.35) with τ = 1.9 < τ0 ≈ 1.94. The positive equilibrium E0.32,0.96 is asymptotically stable. The initial value is (0.2, 1.5).

Fig. 5

Trajectory portrait and phase portrait of system (4.35) with τ = 2 < τ0 ≈ 1.94. Hopf bifurcation occurs from the positive equilibrium E0.32,0.96. The initial value is (0.2, 1.5).

Fig. 6

Trajectory portrait and phase portrait of system (4.35) with τ = 2 < τ0 ≈ 1.94. Hopf bifurcation occurs from the positive equilibrium E0.32,0.96. The initial value is (0.2, 1.5).

Fig. 7

Trajectory portrait and phase portrait of system (4.35) with τ = 2 < τ0 ≈ 1.94. Hopf bifurcation occurs from the positive equilibrium E0.32,0.96. The initial value is (0.2, 1.5).

Fig. 8

Trajectory portrait and phase portrait of system (4.35) with τ = 2 < τ0 ≈ 1.94. Hopf bifurcation occurs from the positive equilibrium E0.32,0.96. The initial value is (0.2, 1.5).

5 Global continuation of local Hopf bifurcations

In the earlier sections we have established that the model system (1.6) undergoes a Hopf bifurcation at Ex1,x2 for τ = τk and also investigated the stability of bifurcating periodic solutions. We all know that periodic solutions can arise through the Hopf bifurcation in delay differential equations. However, these bifurcating periodic solutions are generally local, i.e., they only exist in a small neighborhood of the critical value. Therefore, we want to know that whether these nonconstant periodic solutions obtained through local Hopf bifurcations exist globally or not. In this section, we will consider the global continuation of periodic solutions bifurcating from the positive equilibrium Ex1,x2 of system (1.6).

Throughout this section, we closely follow the notations in Wu [29]. For the simplification of notations, setting zt=z1t,z2tT=x1t,x2tT, we can rewrite system (1.6) as the following form

z˙t=Fzt,τ,p,(5.1)
where ztθ=z1tθ,z2tθT=z1t+θ,z2t+θτ,0,R2. It is easy to see that if the condition (H1) holds, system (5.1) has a positive equilibrium Ex1,x2.

Following the work of Wu [29], we need to define:

X=Cτ,0,R2,Γ=Clz,τ,pX×R×R+;z   is   a   nonconstant   periodic   solution   to   5.1,N=z¯,τ,p;Fz¯,τ,p=0.

Let z¯ be the equilibrium of system (5.1). Then, the characteristic matrix of (5.1) at the equilibrium z¯ takes the following form

Δz¯,τ,pλ=λIdDFz¯,τ,peλId,
where Id is the identity matrix and DFz¯,τ,p is the Fréchet derivative of F with respect to zt evaluated at z¯,τ,p. From Wu [29], we know that z¯,τ,p is called a center if z¯,τ,pN and Δz¯,τ,pλ=0. A center z¯,τ,p is said to be isolated if it is the only center in some neighborhood of z¯,τ,p.

For the benefit of the readers, we first state the global Hopf bifurcation theory due to Wu [29] for functional differential equations.

Lemma 5.1. Assume that z ¯ , τ , p is an isolated center satisfying the hypotheses (A 1 ) – (A 4 ) in Wu [7] . Denote by

C z ¯ , τ , p the connected component of z ¯ , τ , p in Γ. Then, either

(i) C z ¯ , τ , p is unbounded, or

(ii)Cz¯,τ,pis bounded, Cz¯,τ,pΓis finite and

z¯,τ,pCz¯,τ,pNγmz¯,τ,p=0
for all m = 1,2,…, whereγmz¯,τ,pis the m-th crossing number ofz¯,τ,pifmJz¯,τ,p, or it is zero if otherwise.

Obviously, if (ii) of Lemma 5.1 is not true, then Cz¯,τ,p is unbounded. Thus, if the projections of Cz¯,τ,p onto z-space and onto p-space are bounded, then the projections of Cz¯,τ,p onto τ-space is unbounded. Further, if we can show that the projections of Cz¯,τ,p onto τ-space is away from zero, then the projections of Cz¯,τ,p onto τ-space must include the interval τ,. Based on the idea, we can prove our results on the global continuation of the local Hopf bifurcation.

Lemma 5.2. If τ is bounded, then all periodic solutions to (1.6) is uniformly bounded.

Proof. Let x1t,x2t be a nontrivial solution to (1.6) and define

x1ξ1=minx1t,x1η1=maxx1t,x2(ξ2)=minx2t,x2η2=maxx2t
with initial value
x1t=φ1(t)0,x10=φ10>0;x2t=φ2t0,x20=φ20>0,
where tτ,0. Then, it follows from (1.6) that
x1t=x10exp0tabx1(tτ)cx2(tτ)mx2(tτ)+x1(tτ)ds,x2t=x20exp0td+rx1(tτ)mx2(tτ)+x1(tτ)ds

Thus, x1t>0,x2t>0, which implies that solutions to (1.6) cannot cross the x-axes and y-axes. Thus, the nonconstant periodic orbits must be located in the interior of the first quadrant. If (x1(t), x2(t)) is a solution to (1.6), then it follows from the first equation of (1.6) that

dx1dt<x1abx1tτ(5.2)

Clearly, x1t<x1tτeaτ for t > τ, which implies that x1tτ>x1teaτ. This, together with (5.2), leads to

dx1dt<x1ab eaτx1(t),
for t > τ. Then, we obtain that
limtsupx1ta eaτb

Thus,

x1ta eaτb+εM

From the second equation of (1.6), we have

dx2dt<rdx2,
which leads to
x2t<erdτx2tτ

Then,

x2tτ>edrτx2t

Therefore,

x2η2τ>edrτx2η2

Applying the second equation of (1.6), we get

d+rx1η2τmx2η2τ+x1η2τ=0,

i.e.

rx1η2τmx2η2τ+x1η2τ=d

It follows that

rMm edrτx2η2+M>d,
which leads to
x2η2<rdMdm edrτN

Thus, the possible periodic solutions lying in the first quadrant of (1.6) must be uniformly bounded. This completes the proof of Lemma 5.2.

Lemma 5.3.If a < d, r < c, then system (1.6) has nonconstant periodic solution with period τ.

Proof. Suppose for a contradiction that system (1.6) has a nonconstant periodic solution with period τ. Then, the following ordinary differential equations

dx1dt=x1abx1tcx2tmx2t+x1t,dx2dt=x2d+rx1tmx2t+x1t(5.3)
has nonconstant periodic solution. System (5.3) has a boundary equilibrium E0(0, 0) and a positive equilibrium E*x1*,x2*, where
x1*=am2+cdcrabm,x2*=rdam2+cdcrabdm2

Note that x-axis and y-axis are the invariable manifold of system (5.3) and the orbits of system (5.3) do not intersect each other. Thus, there are no solutions crossing the coordinate axes. On the other hand, consider that if system (5.3) has a periodic solution, then there must be an equilibrium in its interior, and that E0(0, 0) is located on the coordinate axis. Thus, we can conclude that the periodic orbit of system (5.3) must lie in the first quadrant.

In the sequel, we will prove that system (5.3) has no nonconstant periodic solution in the first quadrant.

Define

D=x1,x2R2|0x1M,0x2N.

It is easy to show that D is an ultimately bounded region (or absorbing and positively invariant set) of system (5.3). Let

Px,y=x1abx1tcx2tmx2t+x1t,Qx,y=x2d+rx1tmx2t+x1t

Then, a direct computation show that, for (x, y) ∈ D and a < d, r < c,

Px,yx+Qx,yy=ad2bx1+rcmx22mx2+x12<0

Thus, the Bendixson–Dulac criterion, together with the fact that D is an ultimately bounded region of (5.3), implies that (5.3) has no nontrivial periodic solutions, leading to a contradiction. Thus, the proof is complete.

Theorem 5.4.Assume that a < d, r < c. Let ω0 and τk (k = 0,1,2,…) be defined by (2.8) and (2.9), respectively. Then, for each τ > τk (k ≥ 1), system (1.6) has at least k periodic solutions.

Proof. Obviously, E*,τk,2πω0 is an isolated center of (1.6). Let

CE*,τk,2πω0 denote the connected component passing through E*,τk,2πω0 in Γ. It follows from Theorem 2.4 that CE*,τk,2πω0 is nonempty. It is only necessary to prove that the projection of CE*,τk,2πω0 onto τ-space is τ¯, for each j > 0, where τ¯τk.

By Lemma 2.2 and 2.3, there exists ɛ > 0, δ > 0 and a smooth curve λ: (τk – δ, τk + δ) → C such that detΔλτ=0,|λτiω0|<ε for all ττkδ,τk+δ, and λτk=iω0,ddτReλτ|τ=τk>0

Let Ωε,2πω0=μ,p:0<μ<ε,|p2πω0|<ε. It is easy to verify that on τkδ,τk+δ×Ωε,2πω0, detΔE*,τk,pμ+2πpi=0 if and only if μ = 0, τ = τk and p=2πω0. This verifies assumption (A4) of Wu [29]. Moreover, if we put

H±E*,τk,2πω0=detΔ(E*,τk,p)μ+i2πp,
then, we have the crossing number of isolated centers E*,τk,2πω0 as follows:
γ1E*,τk,2πω0=degBHE*,τk,2πω0,Ωε,2πω0degBH+E*,τk,2πω0,Ωε,2πω0=1

By Theorem 3.3 of Wu [29], we conclude that the connected component CE*,τk,2πω0 through E*,τk,2πω0 in Σ is nonempty, and

z¯,τ,pCz¯,τ,pγz¯,τ,p<0

Thus, CE*,τk,2πω0 is unbounded. From (2.9), one can show that, for k ≥ 1,

τk=1ω0arcsinm1+n2m1n2m2n12m1n2m2n1+2kπ,k=1,2,

Hence,

2πω0<τk(5.4)

Now, we are in position to prove that the projection of CE*,τk,2πω0 onto τ-space is [τ, ∞), where τ ≤ τk. Clearly, It follows from the proof of Lemma 5.3 that system (1.6) with τ = 0 has no nontrivial periodic solution. Hence, the projection of CE*,τk,2πω0 onto τ-space is away from zero.

For the sake of contradiction, we suppose that the projection of CE*,τk,2πω0 onto τ-space is bounded. This means that the projection of CE*,τk,2πω0 onto τ-space is included in an interval (0, τ*). From (5.4) and applying Lemma 5.3, we get 0 < p < τ* for zt,τ,pCE*,τk,2πω0. This implies that the projection of CE*,τk,2πω0 onto p-space is also bounded. Thus, combining this with Lemma 5.2, we can get that the connected component CE*,τk,2πω0 is bounded. This contradiction completes the proof.

6 Conclusions and biological explanations

In this paper, we have investigated the local stability of the positive equilibrium E*x1*,x2* and the local Hopf bifurcation in a delayed predator–prey model with Hassell–Varley-type functional response. We have showed that if conditions (H1)–(H3) hold, the positive equilibrium E*x1*,x2* of system (1.6) is asymptotically stable for all τ ∈ [0, τ0) and unstable for τ > τ0. This shows that, in this case, the population of preys and predators will tend to stabilization and still keep stable whenever the delay parameter lies in the range τ ∈ [0, τ0) and unstable when τ > τ0. We have also showed that, if conditions (H1)–(H3) hold, as the delay τ increases, the equilibrium loses its stability and a sequence of Hopf bifurcations occur at E*x1*,x2*, i.e., a family of periodic orbits bifurcate from the positive equilibrium E*x1*,x2*. This means that the population of preys and predators may coexist and keep in an oscillatory mode. Moreover, the direction of Hopf bifurcation and the stability of the bifurcating periodic orbits are discussed by applying the normal form theory and the center manifold theorem. Numerical simulations supporting our theoretical results are also included. Further, sufficient conditions ensuring the existence of global Hopf bifurcation are given, i.e., if a < d, r < c, then system (1.6) has at least k periodic solutions for τ > τk(k ≥ 1). It is shown that the population of preys and predators still keep in an oscillatory mode near the positive equilibrium E*x1*,x2* for τ > τk(k ≥ 1).


Bibliographie

[1] V. Volterra Variazioni e fluttuazioni del numero di individui in specie animali conviventi, Mem. Accd. Lincei., Volume 2 (1926), pp. 31-113

[2] A. Lotka Elements of Physical Biology, Williams and Wilkins, Baltimore, MD, USA, 1925

[3] R. Arditi; L.R. Ginzburg Coupling in predator–prey dynamics: ratio-dependence, J. Theoret. Biol., Volume 139 (1989), pp. 311-326

[4] H.K. Baek Qualitative analysis of Beddington-DeAngelis type impulsive predator–prey models, Nonlinear Anal.: Real World Appl., Volume 11 (2010), pp. 1312-2132

[5] H. Baek; Y. Lim Dynamics of an impulsively controlled Michaelis–Menten-type predator–prey system, Commun. Nonlinear Sci. Numer. Simul., Volume 16 (2011), pp. 2041-2053

[6] Y.L. Song; S.L. Yuan; J.M. Zhang Bifurcation analysis in the delayed Leslie–Gower predator–prey system, Appl. Math. Model., Volume 33 (2009), pp. 4049-4061

[7] C.J. Xu; X.H. Tang; M.X. Liao Stability and bifurcation analysis of a delayed predator–prey model of prey dispersal in two-patch environments, Appl. Math. Comput., Volume 216 (2010), pp. 2920-2936

[8] C.J. Xu; M.X. Liao; X.F. He Stability and Hopf bifurcation analysis for a Lokta–Volterra predator–prey model with two delays, Int. J. Appl. Math. Comput. Sci., Volume 21 (2011), pp. 97-107

[9] C.J. Xu; X.H. Tang; M.X. Liao; X.F. He Bifurcation analysis in a delayed Lokta–Volterra predator–prey model with two delays, Nonlinear Dynam., Volume 66 (2011), pp. 168-183

[10] J.Z. Zhang; J. Zhen; J. Yan; G.Q. Sun Stability and Hopf bifurcation in a delayed competition system, Nonlinear Anal.: Real World Appl., Volume 70 (2009), pp. 658-670

[11] T. Zhao; Y. Kuang; H.L. Simith Global existence of periodic solutions in a class of delayed Gause-type predator–prey system, Nonlinear Anal.: Real World Appl., Volume 28 (1997), pp. 1373-1394

[12] B. Rai; W. Krawcewicz Hopf bifurcation in symmetric configuration of predator–prey-mutualist systems, Nonlinear Anal.: Real World Appl., Volume 71 (2009), pp. 4279-4296

[13] N. Bairagi; D. Jana On the stability and Hopf bifurcation of a delay-induced predator–prey system with habitat complexity, Appl. Math. Modell., Volume 35 (2011), pp. 3255-3267

[14] M. Baurmann; T. Gross; U. Feudel Instabilities in spatially extended predator–prey systems: spatio-temporal patterns in the neighborhood of Turing-Hopfbifurcations, J. Theor. Biol., Volume 245 (2007), pp. 220-229

[15] S. Kumar; S.K. Srivastava; P. Chingakham Hopf bifurcation and stability analysis in a harvested one-predator-two-prey model, Appl. Math. Comput., Volume 129 (2002), pp. 107-118

[16] T.K. Kar; A. Batabyal Stability and bifurcation of a prey–predator model with time delay, C.R. Biol., Volume 332 (2009), pp. 642-651

[17] C. Çelik Hopf bifurcation of a ratio-dependent predator–prey system with time delay, Chaos, Solitons Fractals, Volume 42 (2009), pp. 1474-1484

[18] C. Çelik The stability and Hopf bifurcation for a predator–prey system with time delay, Chaos, Solitons Fractals, Volume 37 (2008), pp. 87-99

[19] K. Chakraborty; M. Chakraborty; T.K. Kar Optimal control of harvest and bifurcation of a prey–predator model with stage structure, Appl. Math. Comput., Volume 217 (2011), pp. 8778-8792

[20] I. Kusbeyzi; O.O. Aybar; A. Hacinliyan Stability and bifurcation in two species predator–prey models, Nonlinear Anal.: Real World Appl., Volume 12 (2011), pp. 377-387

[21] P. Auger; B.W. Kooi; R.B. de la Parra; J.C. Poggiale Bifurcation analysis of a predator–prey model with predators using hawk and dove tactics, J. Theoret. Biol., Volume 238 (2006), pp. 597-607

[22] S. Kovács; K. Kiss; M. Farkas Qualitative behaviour of a ratio-dependent predator–prey system, Nonlinear Anal.: Real World Appl., Volume 10 (2009), pp. 1627-1642

[23] X.P. Yan; C.H. Zhang Hopf bifurcation in a delayed Lokta–Volterra predator–prey system, Nonlinear Anal.: Real World Appl., Volume 9 (2008), pp. 114-127

[24] S.L. Yuan; Y.L. Song Stability and Hopf bifurcations in a delayed Lesile-Gower predator–prey system, J. Math. Anal. Appl., Volume 355 (2009), pp. 82-100

[25] H. Freedman Deterministic Mathematical Models in Population Ecology, Marcel Dekker, New York, 1980

[26] M. Hassell; G. Varley New inductive population model for insect parasites and its bearing on biological control, Nature, Volume 223 (1969), pp. 1133-1136

[27] K. Wang Periodic solutions to a delayed predator–prey model with Hassell–Varley-type functional response, Nonlinear Anal.: Real World Appl., Volume 12 (2011), pp. 137-145

[28] T. Faria; L.T. Magalhaes Normal forms for retared functional differential equations with parameters and applications to Hopf bifurcations, J. Differ. Equations, Volume 122 (1995), pp. 181-200

[29] J. Wu Symmertic functional differential equations and neural networks with memory, Trans. Am. Math. Soc., Volume 350 (1998), pp. 4799-4838

[30] S. Ruan; J. Wei On the zero of some transcendential functions with applications to stability of delay differential equations with two delays, Dyn. Contin. Discrete Impuls. Syst. Ser. A, Volume 10 (2003), pp. 863-874

[31] X. Dieuonnné Founcations of Modern Analysis, Academic Press, New York, 1960

[32] J.K. Hale Theory of Functional Differential Equations, Springer-Verlag, New York, 1977

[33] S.N. Chow; J.K. Hale Methods of Bifurcation Theory, Springer, New York, 1982


This work is supported by National Natural Science Foundation of China (No. 11261010 and No. 60902044), the Doctoral Foundation of Guizhou College of Finance and Economics (2010), the Science and Technology Program of Hunan Province (No. 2010FJ6021) and the soft Science and Technology Program of Guizhou Province (No. 2011LKC2030).

Commentaires - Politique


Ces articles pourraient vous intéresser

Does mutual interference always stabilize predator–prey dynamics? A comparison of models

Roger Arditi; Jean-Marc Callois; Yuri Tyutyunov; ...

C. R. Biol (2004)