Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access March 26, 2019

Hopf bifurcation and stability in a Beddington-DeAngelis predator-prey model with stage structure for predator and time delay incorporating prey refuge

  • Zaowang Xiao , Zhong Li , Zhenliang Zhu and Fengde Chen EMAIL logo
From the journal Open Mathematics

Abstract

In this paper, we consider a Beddington-DeAngelis predator-prey system with stage structure for predator and time delay incorporating prey refuge. By analyzing the characteristic equations, we study the local stability of the equilibrium of the system. Using the delay as a bifurcation parameter, the model undergoes a Hopf bifurcation at the coexistence equilibrium when the delay crosses some critical values. After that, by constructing a suitable Lyapunov functional, sufficient conditions are derived for the global stability of the system. Finally, the influence of prey refuge on densities of prey species and predator species is discussed.

MSC 2010: 34D23; 92B05; 34D40

1 Introduction

Predator-prey model is one of the basic models between different species in real world. As we all know, there are always many species going through two stages, immature and mature, which reflect the different characteristics of species at each stage. Therefore, to exhibit the real world phenomenon, stage structure population models are more reasonable than other models. In recent years, numerous papers have considered the predator-prey system with stage structure (see [1, 2, 3, 4, 5]).

On the other hand, in general, the consumption of prey by predator throughout its past history governs the present birth rate of the predator, in other words, time delay due to gestation is a common example. Obviously, delay differential equations exhibit much more complicated dynamics than ordinary differential equations (see [6, 7, 8, 9, 10, 11, 12, 13, 14]). For example, Wang and Chen [15] considered the following predator prey system with stage structure for the predator population:

x˙(t)=x(t)(rax(tτ1)μy2(t)),y˙1(t)=ημx(tτ2)y2(tτ2)βy1(t)dy1(t),y˙2(t)=βy1(t)ey2(t). (1.1)

The authors studied the asymptotic behavior of system (1.1). When a time delay due to gestation of the predators and a time delay from a crowding effect of the prey are incorporated, we establish conditions for the permanence of the populations and sufficient conditions for the existence of globally stable positive equilibrium of system (1.1).

Beddington [16] and DeAngelis et al. [17] established a famous B-D functional response that is a predator dependent functional response. In many cases, predators need to search for food and share or compete for food. Therefore, the stage-structured predator-prey models incorporating Beddington-DeAngelis functional response better reflect the ecology. Chen et al. [18] discussed the stability of the boundary solution of a nonautonomous predator-prey model with the Beddington-DeAngelis functional response, which reflects the dynamics of interacting predators and prey in a fluctuating environment. Xia et al. [19] considered stability and traveling waves in a Beddington-DeAngelis type stage-structured predator-prey reaction-diffusion systems with nonlocal delays and harvesting. Chen et al. [20] discussed the extinction of a two species non-autonomous competitive system with Beddington-DeAngelis functional response and the effect of toxic substances. Khajanchi [21] investigated the dynamic behavior of a Beddington-DeAngelis type stage structured predator-prey model:

x˙(t)=x(t)(rrkx(t))μx(t)y2(t)1+bx(t)+cy2(t),y˙1(t)=ημx(t)y2(t)1+bx(t)+cy2(t)βy1(t)dy1(t),y˙2(t)=βy1(t)ey2(t). (1.2)

By analyzing the above system, conditions for positivity, boundedness, uniform persistence, existence of positive equilibria with their local stability have been established. Also, the author showed the existence of Hopf bifurcation when the conversion parameter k1 passes the critical value. Finally, the conditions for the occurrence of global stability for the unique interior equilibrium point were derived.

In the real world, refuge is a strategy to reduce the risk of predation. It is clear that the existence of refuge can have a significant impact on the coexistence of predator species and prey species. In recent years, many papers [22, 23, 24, 25, 26, 27] have proposed and analyzed predator-prey models incorporating prey refuges. Recently, Wei and Fu [28] discussed the Hopf bifurcation and stability for predator-prey systems with Beddington-DeAngelis type functional response and stage structure for prey incorporating refuge

x˙1(t)=ax2(t)rx1(t)bx1(t),x˙2(t)=bx1(t)cx2(t)αx22(t)β(1m)x2(t)y(t)a1+b1(1m)x2(t)+c1y(t),y˙(t)=dβ(1m)x2(tτ)y(tτ)a1+b1(1m)x2(tτ)+c1y(tτ)γy(t). (1.3)

By using the characteristic equations, the local stability of each feasible equilibrium of model (1.3) was discussed, and the existence of a Hopf bifurcation at the coexistence equilibrium was established.

Motivated by the works [21] and [28], a Beddington-DeAngelis predator-prey model with stage structure for predator and time delay incorporating prey refuge is investigated in this paper. The proposed model is as follows:

x˙(t)=x(t)(rax(t))μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t),y˙1(t)=ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)βy1(t)dy1(t),y˙2(t)=βy1(t)ey2(t), (1.4)

where x(t), y1(t) and y2(t) denote the densities of prey species, immature predator species and mature predator species at time t, respectively; r is the intrinsic growth rate of prey species; a is the intraspecific competition rate of prey species; d and e are the death rates of the immature and mature predator species respectively; μ(1 – m) is the capturing rate of the mature predator; η is the conversion rate of nutrients into the production of predator species; τ denotes the time delay due to the gestation of the mature predator; m ∈ [0, 1) is refuge rate to prey; the predator species consumes the prey species with Beddington-DeAngelis functional response incorporating prey refuge μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t),andημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ) denotes the growth rate of predator which are pregnant at time tτ.

The initial conditions for system (1.4) take the form

x(θ)=ϕ(θ),y1(θ)=ψ1(θ),y2(θ)=ψ2(θ),ϕ(θ)0,ψ1(θ)0,ψ2(θ)0,θ[τ,0),ϕ(0)>0,ψ1(0)>0,ψ2(0)>0, (1.5)

where (ϕ(θ), ψ1(θ), ψ2(θ)) ∈ C([–τ, 0], R+3 ), which is the Banach space of continuous functions mapping the interval [–τ, 0] into R+3 , where R+3 = {(x1, x2, x3) : xi ≥ 0, i = 1, 2, 3}.

The rest of this paper is organized as follows. The boundedness and local stability of the equilibrium and the existence of Hopf bifurcation at positive equilibrium of system (1.4) are derived in the next section. In Section 3, we study the permanence of system (1.4). In Section 4, the global stability of system (1.4) are investigated. In Section 5, the influence of refuge rate on the densities to predator species and prey species is discussed. We end this paper with some examples and a briefly discussion.

2 Boundedness, Local stability and Hopf bifurcation

In this section, we study the boundedness and local stability of the equilibrium as well as the existence of Hopf bifurcation at positive equilibrium of system (1.4). It is obvious that solutions of model (1.4) with initial conditions (1.5) are positive for all t ≥ 0. The result is a direct consequence of Nagumo’s theorem [29].

2.1 Boundedness

Theorem 2.1

Every solution of system (1.4) with initial conditions (1.5) is bounded for all t ≥ 0 and all of these solutions are ultimately bounded.

Proof

Let V(t) = ηx(tτ) + y1(t) + y2(t), and calculating the derivative of V(t) with respect to t along the positive solution of system (1.4), we have

V˙(t)=ηx˙(t)+y˙1(t)+y˙2(t)=η[x(tτ)(rax(tτ))μ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)]+ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)βy1(t)dy1(t)+βy1(t)ey2(t)=ηx(tτ)[rax(tτ)]dy1(t)ey2(t). (2.1)

For a small positive constant s ≤ min{d, e},

V˙(t)+sV(t)=(sd)y1(t)+(se)y2(t)+ηx(tτ)[s+rax(tτ)]ηx(tτ)[s+rax(tτ)]. (2.2)

Hence there exists a positive constant M=η(s+r)24a such that

V˙(t)+sV(t)M, (2.3)

that is

V(t)(V(0)Ms)est+Ms. (2.4)

Thus V(t) is ultimately bounded, that is, each solution z(t) = (x(t), y1(t), y2(t)) of system (1.4) is ultimately bounded. The proof is complete.

2.2 Equilibria

Obviously, system (1.4) always has a trivial equilibrium E0(0, 0, 0) and a predator-extinction equilibrium E1(r/a, 0, 0). Further, if the following holds:

μη>be(1+dβ),0m<1ae(1+dβ)r[μηbe(1+dβ)], (H1)

then model (1.4) has a unique coexistence equilibrium E(x,y1,y2), where

x=K+Δ2,y1=eβy2,y2=(1m)[μηbe(1+dβ)]xe(1+dβ)ce(1+dβ) (2.5)

with

K=1a((1m)[μηbe(1+dβ)]ηcr),Δ=K2+4eacη(1+dβ).

Let = (, 1, 2) be any arbitrary equilibrium, then the variational matrix of system (1.4) at is given by

J=r2ax~μ(1m)y~2(1+cy~2)[1+b(1m)x~+cy~2]20μ(1m)x~[1+b(1m)x~][1+b(1m)x~+cy~2]2ημ(1m)y~2(1+cy~2)eλτ[1+b(1m)x~+cy~2]2(β+d)ημ(1m)x~[1+b(1m)x~]eλτ[1+b(1m)x~+cy~2]20βe

and the characteristic equation becomes

(λ+β+d)(λ+e)(λr+2ax~+μ(1m)y~2(1+cy~2)[1+b(1m)x~+cy~2]2)+μ(1m)y~2(1+cy~2)[1+b(1m)x~+cy~2]2×βημ(1m)x~[1+b(1m)x~][1+b(1m)x~+cy~2]2eλτ(λr+2ax~+μ(1m)y~2(1+cy~2)[1+b(1m)x~+cy~2]2)×βημ(1m)x~[1+b(1m)x~][1+b(1m)x~+cy~2]2eλτ=0. (2.6)

2.3 E0 = (0, 0, 0)

First, we analyze the stability of equilibrium E0.

Theorem 2.2

The trivial equilibrium E0(0, 0, 0) of system (1.4) is unstable.

Proof

The characteristic equation (2.6) takes the form at the trivial equilibrium E0(0, 0, 0)

(λr)(λ+β+d)(λ+e)=0. (2.7)

It is readily seen that Eq. (2.7) has a positive root, thus the equilibrium E0 is always unstable. The proof is complete.

2.4 E1 = (r/a, 0, 0)

After that, we consider the stability of equilibrium E1.

Theorem 2.3

If the following holds:

μη>be(1+dβ),1ae(1+dβ)r[μηbe(1+dβ)]<m<1, (H2)

then the predator-extinction equilibrium E1(r/a, 0, 0) of system (1.4) is locally asymptotically stable; if (H1) holds, then E1 is unstable.

Proof

The characteristic equation (2.6) at predator-extinction equilibrium E1 becomes

(λ+r)[(λ+β+d)(λ+e)βημr(1m)a+br(1m)eλτ]=0. (2.8)

Clearly, the equation λ + r = 0 has one negative real root, which implies that all other roots of Eq. (2.8) are determined by

λ2+h1λ+h2+h3eλτ=0, (2.9)

where h1 = e + β + d > 0, h2 = e(β + d), h3 = βημr(1m)a+br(1m).

When τ = 0, Eq. (2.9) turns to

λ2+h1λ+h2+h3=0. (2.10)

According to (H2), we have h2 + h3 > 0. By the Routh-Hurwitz criterion, the boundary equilibrium E1 is locally asymptotically stable. If (H1) holds, then Eq. (2.10) has at least a positive real root, thus the predator-extinction equilibrium E1 is unstable.

For τ > 0, we investigate the existence of purely imaginary roots of (2.9). If 1(ω1 > 0) is a solution of (2.9) if and only if ω1 satisfies

ω12+h1ω1i+h2+h3(cos(τω1)isin(τω1))=0.

Separating the real and imaginary parts, we obtain

h1ω1=h3sin(τω1),ω12h2=h3cos(τω1), (2.11)

which implies

ω14+(h122h2)ω12+h22h32=0. (2.12)

Note that

h122h2=(e+β+d)22e(β+d)=e2+(β+d)2>0,h2h3=e(β+d)+βημr(1m)a+br(1m)>0,

and h2 + h3 > 0, then h22h32>0. Hence (2.9) has no positive real roots. By Theorem 3.4.1 in [30], if (H2) holds, then all the roots of (2.9) have negative real parts for all τ ≥ 0, this implies that the boundary equilibrium E1 is locally asymptotically stable for all τ ≥ 0. The proof is complete.

2.5 E=(x,y1,y2)

Further, we analyze the stability of equilibrium E*.

For the positive equilibrium E=(x,y1,y2), the characteristic equation (2.6) reduces to

λ3+p2λ2+p1λ+p0+(q1λ+q0)eλτ=0, (2.13)

where

p2=β+d+e+A+2axr,p1=Q+(β+d+e)(A+2axr),p0=Q(A+2axr),q1=B,q0=B(2axr),

with

Q=e(β+d),A=μ(1m)y2(1+cy2)[1+b(1m)x+cy2]2,B=βημ(1m)x[1+b(1m)x][1+b(1m)x+cy2]2=Q[e(1+dβ)+be(1+dβ)(1m)x]μη(1m)x.

When τ = 0, Eq. (2.13) turns to

λ3+p2λ2+(p1+q1)λ+p0+q0=0. (2.14)

From (H1), we derive that B=Q[e(1+dβ)+be(1+dβ)(1m)x]μη(1m)x<Q, thus

p2=β+d+e+(A+2axr),p1+q1=QB+(β+d+e)(A+2axr),p0+q0=QA+(QB)2ax(QB)r,p2(p1+q1)(p0+q0)=(β+d+e+A+2axr)(QB)+(A+2axr)×[(β+d)(β+d+A+2axr)+e(e+A+2axr)]+QA+(Q+B)2ax(Q+B)r.

Denote

r1=QAQ+B+2ax,r2=A+2ax,r3=QAQB+2ax,r4=r1+(β+d+e+r2r)(QB)Q+B+(r2r)[(β+d)(β+d+r2r)+e(e+r2r)]Q+B.

We see that if r < min{r2, r4} and (H1) hold, then p2 > 0, p1 + q1 > 0, p0 + q0 > 0 and p2(p1 + q1) – (p0 + q0) > 0. Therefore, by the Routh-Hurwitz criterion, the positive equilibrium E* is locally asymptotically stable.

For τ > 0, if (ω > 0) is a solution of (2.13) if and only if ω satisfies

ω3ip2ω2+p1ωi+p0+(q1ωi+q0)(cos(τω)isin(τω))=0.

Separating the real and imaginary parts, we have

ω3p1ω=q1ωcos(τω)q0sin(τω),p2ω2p0=q1ωsin(τω)+q0cos(τω), (2.15)

which implies

ω6+(p222p1)ω4+(p122p0p2q12)ω2+p02q02=0, (2.16)

where

p222p1=(β+d+e+A+2axr)22Q2(β+d+e)(A+2axr)=(β+d)2+e2+(A+2axr)2>0,p122p0p2q12=[Q+(β+d+e)(A+2axr)]22Q(A+2axr)×(β+d+e+A+2axr)B2>(A+2axr)2[(β+d)2+e2]>0,p02q02=[QA+(QB)2ax(QB)r]×[QA+(Q+B)2ax(Q+B)r].

Obviously, if r < r1 and (H1) hold, then p02q02>0, this implies that (2.16) has no positive real roots. Therefore, by Theorem 3.4.1 in [30], if r < r1 and (H1) are satisfied, then all the roots of (2.16) have negative real parts for all τ ≥ 0. Hence the positive equilibrium E=(x,y1,y2) is locally asymptotically stable for all τ ≥ 0.

If r1 < r < min{r2, r4} holds, which implies p02q02<0, then there exists a unique positive root ω0 satisfying (2.16). From (2.15), we have

cos(τω0)=q1ω04+(p2q0p1q1)ω02p0q0q12ω02+q02. (2.17)

Denote

τ0n=1ω0arccosq1ω04+(p2q0p1q1)ω02p0q0q12ω02+q02+2nπω0,n=0,1,2,. (2.18)

By Theorem 3.4.1 in Kuang [30], we see that if p02q02<0 hold, then E* remains stable for τ < τ0 := τ00.

We now claim that

d(Reλ)dττ=τ0>0.

This shows that there exists at least one eigenvalue with a positive real part for τ > τ0. Moreover, the conditions for the existence of a Hopf bifurcation [31] are then satisfied yielding a periodic solution. To this end, differentiating Eq. (2.13) with respect to τ, it follows that

dλdτ1=3λ2+2p2λ+p1λ(λ3+p2λ2+p1λ+p0)+q1λ(q1λ+q0)τλ.

Hence, a direct calculation shows that

sgnd(Reλ)dτλ=iω0=sgnRedλdτ1λ=iω0=sgn(p13ω02)(ω02p1)+2p2(p0p2ω02)(ω03p1ω0)2+(p0p2ω02)2q12q12ω02+q02.

We derive from (2.15) that

(ω03p1ω0)2+(p0p2ω02)2=q12ω02+q02.

Therefore, we yield

sgnd(Reλ)dτλ=iω0=sgn3ω04+2(p222p1)ω02+p122p0p2q12q12ω02+q02>0.

Thus, the transversal condition holds and a Hopf bifurcation occurs at ω = ω0, τ = τ0. Now, let us summarize our results as follows:

Theorem 2.4

  1. If (H1) and r < r1 hold, then the positive equilibrium E* of system (1.4) is locally asymptotically stable for all τ ≥ 0.

  2. If (H1) and r1 < r < min{r2, r4} hold, then there exists a τ0 > 0 such that E* is locally asymptotically stable when τ ∈ [0, τ0). Furthermore, system (1.4) undergoes a Hopf bifurcation at E* when τ = τ0.

Remark

  1. Let τ = 0 and m = 0 in (1.4), then Theorem 2.3 of this paper is equivalent to Theorem 3 of [21]. Obviously, we generalize the conclusion of boundary equilibrium in [21], and show that prey refuge affect the stability of the boundary equilibrium. Further, the global stability of the boundary equilibrium will be studied in Section 4.

  2. Notice that the conditions of positive equilibrium E* is locally asymptotically stable in [21] is very complicated. Let τ = 0 and m = 0 in model (1.4), compare Theorem 2.4 of this paper with Lemma 4 of [21], we find that the conditions of our positive equilibrium locally stable are more extensive and concise than that of [21].

3 Permanence

Lemma 3.1

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5). Assume that (H1) holds, then

lim supt+x(t)L1,lim supt+y1(t)L2,lim supt+y2(t)L3, (3.1)

where

L1=ra,L2=βημ(1m)L1e(β+d)[1+b(1m)L1]cβ(β+d),L3=βeL2.

Proof

It follows from (H1) that L2=βημ(1m)L1e(β+d)[1+b(1m)L1]cβ(β+d)>0. Hence, there exists an enough small positive constant ε such that

L2ε=defβημ(1m)(L1+ε)e(β+d)[1+b(1m)(L1+ε)]cβ(β+d)>0,L3ε=defβeL2ε>0.

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5). From the first equation of system (1.4), it follows that

x˙(t)=x(t)(rax(t))μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t)x(t)(rax(t)). (3.2)

Applying Lemma 2.3 in [32] to (3.2), it immediately follows that

lim supt+x(t)ra=defL1. (3.3)

Then for above ε > 0 sufficiently small there exists a T1 > 0 such that if t > T1, x(t) ≤ L1 + ε. We derive from the second and the third equations of system (1.4) that for t > T1 + τ,

y˙1(t)ημ(1m)(L1+ε)y2(tτ)1+b(1m)(L1+ε)+cy2(tτ)βy1(t)dy1(t),y˙2(t)=βy1(t)ey2(t). (3.4)

Consider the following auxiliary equations:

u˙1(t)=ημ(1m)(L1+ε)u2(tτ)1+b(1m)(L1+ε)+cu2(tτ)βu1(t)du1(t),u˙2(t)=βu1(t)eu2(t). (3.5)

Using a similar argument as that in the proof of Lemma 2.4 in [33], it follows from (3.5) that

limt+u1(t)=L2ε,limt+u2(t)=L3ε. (3.6)

By comparison, we obtain that

lim supt+y1(t)L2ε,lim supt+y2(t)L3ε. (3.7)

Let ε → 0, it follows that

lim supt+y1(t)L2,lim supt+y2(t)L3. (3.8)

The proof is complete.

Lemma 3.2

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5), if (H1) and

r>μ(1m)L31+cL3,μη>e(β+d)[1+b(1m)l1]β(1m)l1 (3.9)

hold, then

lim inft+x(t)l1,lim inft+y1(t)l2,lim inft+y2(t)l3, (3.10)

where

l1=r(1+cL3)μ(1m)L3a(1+cL3),l2=βημ(1m)l1e(β+d)[1+b(1m)l1]cβ(β+d),l3=βel2.

Proof

It follows from (H1) and condition (3.9) that l1=r(1+cL3)μ(1m)L3a(1+cL3)>0,l2=βημ(1m)l1e(β+d)[1+b(1m)l1]cβ(β+d)> 0. Hence, there exists an enough small positive constant ε such that

l1ε=r(1+c(L3+ε))μ(1m)(L3+ε)a(1+c(L3+ε))>0,l2ε=βημ(1m)(l1ε)e(β+d)[1+b(1m)(l1ε)]cβ(β+d)>0,l3ε=βel2ε>0.

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5). For above ε > 0 sufficiently small, it follows from Lemma 3.1 that there exists a T2 > 0 such that if t > T2, y1(t) ≤ L2 + ε, y2(t) ≤ L3 + ε. Hence, we derive from the first equation of system (1.4) that for t > T2,

x˙(t)x(t)(rax(t)μ(1m)(L3+ε)1+c(L3+ε)).

According to condition (3.9) and similar to the proof of Lemma 3.1, we have

lim inft+x(t)r(1+cL3)μ(1m)L3a(1+cL3)=defl1>0. (3.11)

For above ε > 0 sufficiently small, there exists a T3T2 such that if t > T3, x(t) ≥ l1ε. Therefore, it follows from the second and the third equations of system (1.4) that for t > T3 + τ,

y˙1(t)ημ(1m)(l1ε)y2(tτ)1+b(1m)(l1ε)+cy2(tτ)βy1(t)dy1(t),y˙2(t)=βy1(t)ey2(t). (3.12)

Consider the following auxiliary equations:

u˙1(t)=ημ(1m)(l1ε)u2(tτ)1+b(1m)(l1ε)+cu2(tτ)βu1(t)du1(t),u˙2(t)=βu1(t)eu2(t). (3.13)

Using a similar argument as that in the proof of Lemma 2.4 in [33], it follows from (3.13) that

limt+u1(t)l2ε,limt+u1(t)l3ε. (3.14)

By comparison, we obtain that

lim inft+y1(t)l2ε,lim inft+y2(t)l3ε. (3.15)

Let ε → 0, we conclude that

lim inft+y1(t)l2,lim inft+y2(t)l3. (3.16)

The proof is complete.

As a direct corollary of Lemmas 3.1 and 3.2 we have the following theorem.

Theorem 3.1

Assume that (H1) and (3.9) hold, then system (1.4) is permanent.

4 Global stability

In this section, we study the global stability of the predator-extinction equilibrium E1 and the global attractivity of the coexistence equilibrium E* of system (1.4). The strategy of proofs is to use Lyapunov functionals and the LaSalle invariance principle.

Theorem 4.1

If (H2) holds, then the predator-extinction equilibrium E1(r/a, 0, 0) of system (1.4) is globally asymptotically stable.

Proof

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5). Denote x0 = r/a.

Define

V11(t)=xx0x0lnxx0+ξ1y1+ξ2y2, (4.1)

where ξ1 = 1+b(1m)x0η,ξ2=μ(1m)x0e .

Calculating the derivative of V11 along positive solutions of system (1.4), it follows that

ddtV11(t)=(1x0x)[x(t)(rax(t))μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t)]+ξ1[ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)βy1(t)dy1(t)]+ξ2[βy1(t)ey2(t)]. (4.2)

Substituting r = ax0 into (4.2), we obtain that

ddtV11(t)=rx0(x(t)x0)2[1+b(1m)x0]μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t)+μ(1m)x0y2(t)μ(1m)cx0y22(t)1+b(1m)x(t)+cy2(t)+ξ1ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)ξ1βy1(t)ξ1dy1(t)+ξ2βy1(t)ξ2ey2(t). (4.3)

Define

V1(t)=V11(t)+ξ1ημ(1m)tτtx(s)y2(s)1+b(1m)x(s)+cy2(s)ds. (4.4)

We derive from (4.3) and (4.4) that

ddtV1(t)=rx0(x(t)x0)2μ(1m)cx0y22(t)1+b(1m)x(t)+cy2(t)[(β+d)[a+br(1m)]ηaβμr(1m)ae]y1(t). (4.5)

If (H2) holds, it then follows from (4.5) that 1(t) ≤0. By Theorem 5.3.1 in [34], solutions limit to 𝓜, the largest invariant subset of {1(t) = 0}. Obviously, we see from (4.5) that 1(t) = 0 if and only if x = x0, y2 = 0. Noting that 𝓜 is invariant, for each element in 𝓜, we have x(t) = x0, y2(t) = 0. It therefore follows from the third equation of system (1.4) that 0 = 2(t) = βy1(t), which yields y1(t) = 0. Therefore, 1(t) = 0 if and only if (x, y1, y2) = (x0, 0, 0). Accordingly, the global asymptotic stability of E1 follows from LaSalle invariance principle. This completes the proof.

Theorem 4.2

Assume that (H1) and (3.9) are satisfied. If the following holds:

a(l1+x)rρ¯1,l1x+b(1m)l1x2ρ¯2, (4.6)

where ρ¯1=2μ(1m)cL3y2(1+cy2)+μc(1m)xy2[b(1m)L1+cL3]2[1+b(1m)l1+cl3][1+b(1m)x+cy2],ρ¯2=xy2[b(1m)L1+cL3]2 , and Li, li (i = 1, 2, 3) are defined as those in Theorem 3.1, then the positive equilibrium E(x,y1,y2) of system (1.4) is globally attractive.

Proof

Let (x(t), y1(t), y2(t)) be any positive solution of system (1.4) with initial conditions (1.5). Define

V21(t)=xxxlnxx+k1(y1y1y1lny1y1)+k2(y2y2y2lny2y2), (4.7)

where k1=1+b(1m)x+cy2η,k2=k1(β+d)β .

Calculating the derivative of V11 along positive solutions of system (1.4), it follows that

ddtV21(t)=(1xx)[x(t)(rax(t))μ(1m)x(t)y2(t)1+b(1m)x(t)+cy2(t)]+k1(1y1y1)[ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)βy1(t)dy1(t)]+k2(1y2y2)[βy1(t)ey2(t)]. (4.8)

Substituting r=ax+μ(1m)y21+b(1m)x+cy2 into (4.8), we derive that

ddtV21(t)=(1xx)[x(rax)x(rax)+μ(1m)xy21+b(1m)x+cy2]μ(1m)[[1+b(1m)x+cy2]x[1+b(1m)x+cy2]x+c(xy2xy2)x]×x(t)y2(t)1+b(1m)x(t)+cy2(t)+k1(1y1y1)×[ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)dy1(t)βy1(t)]+k2(1y2y2)[βy1(t)ey2(t)]
=(1xx)[x(rax)x(rax)+μ(1m)xy21+b(1m)x+cy2]μ(1m)[1+b(1m)x+cy2]×x(t)y2(t)1+b(1m)x(t)+cy2(t)μ(1m)cx(t)y2(t)(xy2xy2)x[1+b(1m)x(t)+cy2(t)]+k1ημ(1m)x(tτ)y2(tτ)1+b(1m)x(tτ)+cy2(tτ)k1ημ(1m)y1x(tτ)y2(tτ)y1[1+b(1m)x(tτ)+cy2(tτ)]+k1(β+d)y1k2βy1y2y1y1(t)y2(t)+k2ey2. (4.9)

Define

V2(t)=V21(t)+k1ημ(1m)tτt[x(s)y2(s)1+b(1m)x(s)+cy2(s)xy21+b(1m)x+cy2xy21+b(1m)x+cy2×ln[1+b(1m)x+cy2]x(s)y2(s)xy2[1+b(1m)x(s)+cy2(s)]]ds. (4.10)

It follows from (4.9) and (4.10) that

ddtV2(t)=(1xx)[x(rax)x(rax)+μ(1m)xy21+b(1m)x+cy2]μ(1m)cx(t)y2(t)(xy2xy2)x[1+b(1m)x(t)+cy2(t)]k1ημ(1m)xy21+b(1m)x+cy2×y1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]+k1(β+d)y1k2βy1y2y1y1(t)y2(t)+k2ey2+k1ημ(1m)xy21+b(1m)x+cy2×ln[1+b(1m)x(t)+cy2(t)]x(tτ)y2(tτ)x(t)y2(t)[1+b(1m)x(tτ)+cy2(tτ)]. (4.11)

Noting that

k2ey2=k2βy1=k1(β+d)y1=k1ημ(1m)xy21+b(1m)x+cy2=μ(1m)xy2,

and

(1xx)μ(1m)xy21+b(1m)x+cy2=μ(1m)xy2(1x[1+b(1m)x(t)+cy2(t)]x(t)[1+b(1m)x+cy2])+μ(1m)cxy2(xy2xy2)x(t)[1+b(1m)x+cy2],

we derive from (4.11) that

ddtV2(t)=(xx)2x[ra(x+x)]μ(1m)cx(t)y2(t)(xy2xy2)x(t)[1+b(1m)x(t)+cy2(t)]+μ(1m)cxy2(xy2xy2)x(t)[1+b(1m)x+cy2]μ(1m)xy2×[x[1+b(1m)x+cy2]x[1+b(1m)x+cy2]1lnx[1+b(1m)x+cy2]x[1+b(1m)x+cy2]]μ(1m)xy2×[y2y1(t)y1y2(t)1lny2y1(t)y1y2(t)]μ(1m)xy2×[y1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]1lny1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]]
(xx)2x[a(x+x)rρ1]μ(1m)c(yy2)2[xx+b(1m)xx2ρ2]x(t)[1+b(1m)x(t)+cy2(t)][1+b(1m)x+cy2]μ(1m)xy2×[x[1+b(1m)x+cy2]x[1+b(1m)x+cy2]1lnx[1+b(1m)x+cy2]x[1+b(1m)x+cy2]]μ(1m)xy2×[y2y1(t)y1y2(t)1lny2y1(t)y1y2(t)]μ(1m)xy2×[y1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]1lny1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]], (4.12)

where ρ1=2μ(1m)cy2y2(1+cy2)+μc(1m)xy2[b(1m)x+cy2]2[1+b(1m)x(t)+cy2(t)][1+b(1m)x+cy2],ρ2=xy2[b(1m)x+cy2]2 .

Hence, if (4.6) holds, it then follows from (4.12) that 1(t) ≤0, with equality if and only if x = x*, y1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]=y2y1(t)y1y2(t) = 1. We now look for the invariant subset 𝓜 within the set M = {(x, y1, y2) : x = x*, y1[1+b(1m)x+cy2]x(tτ)y2(tτ)xy2y1[1+b(1m)x(tτ)+cy2(tτ)]=y2y1(t)y1y2(t) = 1}. Since x = x* on 𝓜 and due to 0 = (t) = x*(rax*) − μ(1m)xy2(t)1+b(1m)x+cy2(t) , we obtain y2(t) = y2 . From the third equation of model (1.4) that 0 = 2(t) = βy1(t) − e y2 , which yields y1 = y1 . Therefore, the only invariant set in M is 𝓜 = {(x,y1,y2)} . Using the LaSalle invariance principle, then E* of system (1.4) is globally attractive.

5 The influence of prey refuge

In this section, we investigate the influence of prey refuge. Under the condition (H1), let us compute the derivative along the positive equilibrium E* with respect to m, that is

dxdm=[μηbe(1+dβ)]xaηcΔ>0,dy2dm=[μηbe(1+dβ)]x[(1m)[μηbe(1+dβ)]aηcΔ]aηc2e(1+dβ)Δ,dy1dm=[μηbe(1+dβ)]x[(1m)[μηbe(1+dβ)]aηcΔ]βaηc2(1+dβ)Δ.

Denote

R1=ηcr2+4ae(1+dβ)2r[μηbe(1+dβ)]>ae(1+dβ)r[μηbe(1+dβ)]=R2.

Due to the existence of E*, which implies that x* is a strictly increasing function of m, that is, increasing the constant amount of prey refuge m leads to the increase of prey densities. When 0 < m < 1 − ℜ1, dyidm > 0 (i = 1, 2), it then yields that both y1andy2 are strictly increasing functions on m ∈ (0, 1 − ℜ1); when 1 − ℜ1 < m < 1 − ℜ2, dyidm < 0 (i = 1, 2), which implies that both y1andy2 are strictly decreasing functions on m ∈ (1 − ℜ1, 1 − ℜ2); when m = 1 − ℜ1, the predator species reaches its maximum, and when m = 1 − ℜ2, the predator species goes to extinction.

6 Numerical simulations

Example 6.1

In (1.4), let m = τ = 0, then system (1.4) is degenerated into model (2.1) of [21]. Let r = 6, a = 3, μ = 20, b = d = e = 1, η = β = 2 and c = 4, then system (1.4) has a positive equilibrium E* = (0.5168, 1.5330, 3.0660). By calculation, μη = 40 > 1.5 = be(1 + dβ ), 0 < 0.9805 = 1 − ae(1+dβ)r[μηbe(1+dβ)] , r < 7.3835 = r2 and r < 17.7279 = r4. Thus E* is locally asymptotically stable.

Note that ηβ(1 + bx*) = 6.0671 > 0.4778 = b2xy2(β+d+e)1+cy2 . Therefore, condition (S3): ηβ(1 + bx*) < min e(β+d)(1+bx+cy2)2μx,b2xy2(β+d+e)1+cy2 in [21] which ensures the local stability of the positive equilibrium E* is not satisfied. This implies that our conditions are weaker that those in [21] (see Figs. 1 and 2).

Figure 1 
m = τ = 0, E* = (0.5168, 1.5330, 3.0660) of system (1.4) is locally asymptotically stable.
Figure 1

m = τ = 0, E* = (0.5168, 1.5330, 3.0660) of system (1.4) is locally asymptotically stable.

Figure 2 
m = τ = 0, E* = (0.5168, 1.5330, 3.0660) of system (1.4) with different initial values is locally asymptotically stable.
Figure 2

m = τ = 0, E* = (0.5168, 1.5330, 3.0660) of system (1.4) with different initial values is locally asymptotically stable.

Example 6.2

In (1.4), let r = 6, a = 3, m = 0.2, b = d = e = 1, η = β = 2 and c = 4, then be(1 + dβ ) = 1.5.

  1. If μ = 1, τ = 10, it is easy to show that μη = 2 > be(1 + dβ ), 1 − m = 0.8 < 1.5 = ae(1+dβ)r[μηbe(1+dβ)] and the predator-extinction equilibrium E1 = (2, 0, 0). By Theorem 4.1, E1 is globally asymptotically stable (see Figs. 3 and 4).

    Figure 3 
μ = 1, τ = 10, E1 = (2, 0, 0) of system (1.4) is globally asymptotically stable.
    Figure 3

    μ = 1, τ = 10, E1 = (2, 0, 0) of system (1.4) is globally asymptotically stable.

    Figure 4 
μ = 1, τ = 10, E1 = (2, 0, 0) of system (1.4) with different initial values is globally asymptotically stable.
    Figure 4

    μ = 1, τ = 10, E1 = (2, 0, 0) of system (1.4) with different initial values is globally asymptotically stable.

  2. If μ = 20, τ = 10, it is easy to show that μη = 40 > be(1 + dβ ), m < 0.9805 = 1 − ae(1+dβ)r[μηbe(1+dβ)] , r < 7.9443 = r1 and the positive equilibrium E* = (0.7953, 1.9162, 3.8323). By Theorem 2.4, E* is locally asymptotically stable (see Figs. 5 and 6).

    Figure 5 
μ = 20, τ = 10, E* = (0.7953, 1.9162, 3.8323) of system (1.4) is locally asymptotically stable.
    Figure 5

    μ = 20, τ = 10, E* = (0.7953, 1.9162, 3.8323) of system (1.4) is locally asymptotically stable.

    Figure 6 
μ = 20, τ = 10, E* = (0.7953, 1.9162, 3.8323) of system (1.4) with different initial values is locally asymptotically stable.
    Figure 6

    μ = 20, τ = 10, E* = (0.7953, 1.9162, 3.8323) of system (1.4) with different initial values is locally asymptotically stable.

  3. If μ = 36, it is easy to show that μη = 72 > be(1 + dβ ), m < 0.9894 = 1 − ae(1+dβ)r[μηbe(1+dβ)] , r > 5.2798 = r1, r < 6.2736 = r2, r < 8.8358 = r4, and the positive equilibrium E* = (0.1302, 0.4868, 0.9735). Further, by calculation, we have τ0 = 0.9429. By Theorem 2.4, when τ < τ0, E* is locally asymptotically stable (see Fig. 7); when τ > τ0, the positive equilibrium E* of model (1.4) is unstable, which yields a periodic solution (see Fig. 8).

    Figure 7 
τ = 0.9 < 0.9429 = τ0, μ = 36, E* = (0.1302, 0.4868, 0.9735) of system (1.4) is locally asymptotically stable, system (1.4) undergoes a Hopf bifurcation at E* when τ0.
    Figure 7

    τ = 0.9 < 0.9429 = τ0, μ = 36, E* = (0.1302, 0.4868, 0.9735) of system (1.4) is locally asymptotically stable, system (1.4) undergoes a Hopf bifurcation at E* when τ0.

    Figure 8 
τ = 0.96 > 0.9429 = τ0, μ = 36, E* = (0.1302, 0.4868, 0.9735) of system (1.4) is locally asymptotically stable, system (1.4) undergoes a Hopf bifurcation at E* when τ0.
    Figure 8

    τ = 0.96 > 0.9429 = τ0, μ = 36, E* = (0.1302, 0.4868, 0.9735) of system (1.4) is locally asymptotically stable, system (1.4) undergoes a Hopf bifurcation at E* when τ0.

  4. If μ = 2, τ = 2, it is easy to show that μη = 4 > be(1 + dβ ), m < 0.7 = 1 − ae(1+dβ)r[μηbe(1+dβ)] , a(l1 + x*) − r = 5.5962 > 0.6417 = ρ1, l1 x* + b(1 − m)l1x*2 = 9.5580 > 1.2781 = ρ2, r > 0.25 = μ(1m)L31+cL3 , μη > 2.4783 = e(β+d)[1+b(1m)l1]β(1m)l1 and the positive equilibrium E* = (1.9487, 0.1998, 0.3996). By Theorem 4.2, E* is globally attractive (see Fig. 9).

    Figure 9 
μ = 2, τ = 2, E* = (1.9487, 0.1998, 0.3996) of system (1.4) with different initial values is globally attractive.
    Figure 9

    μ = 2, τ = 2, E* = (1.9487, 0.1998, 0.3996) of system (1.4) with different initial values is globally attractive.

Example 6.3

In (1.4), let r = 10, μ = 6, c = e = 0.5, a = b = d = 1, η = β = 2 and τ = 0. By simple computations, we find μη = 10 > 0.75 = be(1 + dβ ), 1 − ℜ1 = 0.4432, 1 − ℜ2 = 0.9919. Note that, if m ∈ [0, 0.9919), by calculation, we can obtain m < 1 − ae(1+dβ)r[μηbe(1+dβ)] , r < min{r2, r4}, then E* is locally asymptotically stable when τ = 0. Our simulations show that the constant prey refuge m plays an important role on the coexistence of prey-predator population (see Figs. 10 and 11). When m is small and increasing, the predator density increases, due to the fact that predators have sufficient preys available for predation, even though the refuge increases (see Fig. 12). But, as m crosses its threshold value, the predator density decreases with increasing m, the predators was unable to catch preys to sustain themselves and ultimately goes to extinction due to starvation (see Fig. 13).

Figure 10 
Bifurcation diagram of the prey species of model (1.4), m as the bifurcation parameter.
Figure 10

Bifurcation diagram of the prey species of model (1.4), m as the bifurcation parameter.

Figure 11 
Bifurcation diagram of the predator species of model (1.4), m as the bifurcation parameter.
Figure 11

Bifurcation diagram of the predator species of model (1.4), m as the bifurcation parameter.

Figure 12 
τ = 0, E* is locally asymptotically stable. Change the values of the refuge m from zero to 1 − ℜ1 = 0.4432, the predator species will increase.
Figure 12

τ = 0, E* is locally asymptotically stable. Change the values of the refuge m from zero to 1 − ℜ1 = 0.4432, the predator species will increase.

Figure 13 
τ = 0, E* is locally asymptotically stable. Change the values of the refuge m from 1 − ℜ1 = 0.4432 to 0.9919 = 1 − ℜ2, the predator species will decrease.
Figure 13

τ = 0, E* is locally asymptotically stable. Change the values of the refuge m from 1 − ℜ1 = 0.4432 to 0.9919 = 1 − ℜ2, the predator species will decrease.

7 Conclusion

In this paper we investigate the influence of prey refuge on the dynamics of a Beddington-DeAngelis predator-prey system with stage structure for predator and time delay. Sufficient conditions are derived to ensure the predator-extinction and the locally asymptotically stability of positive equilibrium. Compared with [21] we get more precise conditions. Also, we find that time delay can cause a stable equilibrium to become unstable one, even Hopf bifurcation to occur, when time delay passes through some critical values. Furthermore, the persistence is investigated. After that, we study the global stability of the predator-extinction and positive equilibrium by constructing some suitable Lyapunov functionals.

Also, we discuss the influence of prey refuge on the densities of predator species and prey species. When prey refuge m in the interval (0, 1 − ℜ1), the density of predators will increase with prey refuge m, due to predator species having sufficient food for their predation with sufficiently small prey refuge m. Predator population attains its maximum when the prey refuge m = 1 − ℜ1. In case of larger values of m (m > 1 − ℜ1), this implies that predators species are less likely to catch prey, and the predator species deceases with the increasing of prey refuge m. Eventually, the predator population will be extinct when prey refuge m = 1 − ℜ2.

  1. Competing interests The authors declare that there is no conflict of interests.

  2. Authors’ Contributions All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Acknowledgement

This work was supported by the National Natural Science Foundation of China (11601085), the Natural Science Foundation of Fujian Province (2017J01400) and the foundation of Education Department of Fujian Province (JAT160063).

References

[1] Li T.T., Chen F.D., Chen J.H., Lin Q.X., Stability of a stage-structured plant-pollinator mutualism model with the Beddington-DeAngelis functional response, J. Nonlinear Funct. Anal., 2017, 2017, Article ID 5010.23952/jnfa.2017.50Search in Google Scholar

[2] Zhuo X.L., Global attractability and permanence for a new stage-structured delay impulsive ecosystem, J. Appl. Anal. Comput., 2018, 8, 457-47010.11948/2018.457Search in Google Scholar

[3] Li Z., Han M.A., Chen F.D., Global stability of a predator-prey system with stage structure and mutual interference, Discrete Contin. Dyn. Syst. Ser. B, 2014, 19, 173-18710.3934/dcdsb.2014.19.173Search in Google Scholar

[4] Chen F.D., Xie X.D., Chen X.F., Dynamic behaviors of a stage-structured cooperation model, Commun. Math. Biol. Neurosci., 2015, 2015, Article ID 4Search in Google Scholar

[5] Xu R., Global stability and Hopf bifurcation of a predator-prey model with stage structure and delayed predator response, Nonlinear Dynam., 2012, 67, 1683-169310.1007/s11071-011-0096-1Search in Google Scholar

[6] Li L., Wu R.X., Extinction of a delay differential equation model of plankton allelopathy, Commun. Math. Biol. Neurosci., 2015, 2015, Article ID 13Search in Google Scholar

[7] Wu R.X., Li L., Extinction of a reaction-diffusion model of plankton allelopathy with nonlocal delays, Commun. Math. Biol. Neurosci., 2015, 2015, Article ID 8Search in Google Scholar

[8] Zhang T.Q., Meng X.Z., Song Y., Zhang T.H., A stage-structured predator-prey SI model with disease in the prey and impulsive effects, Math. Model. Anal., 2013, 18, 505-52810.3846/13926292.2013.840866Search in Google Scholar

[9] Jiang Z.C., Wang L., Global Hopf bifurcation for a predator-prey system with three delays, Int. J. Bifurc. Chaos, 2017, 27, 175010810.1142/S0218127417501085Search in Google Scholar

[10] Wang Z., Wang X.H., Li Y.X., Huang X., Stability and Hopf bifurcation of fractional-order complex-valued single neuron model with time delay, Int. J. Bifurc. Chaos, 2017, 27, 175020910.1142/S0218127417502091Search in Google Scholar

[11] Li L., Wang Z., Li Y.X., Shen H., Lu J.W., Hopf bifurcation analysis of a complex-valued neural network model with discrete and distributed delays, Appl. Math. Comput., 2018, 330, 152-16910.1016/j.amc.2018.02.029Search in Google Scholar

[12] Li Z., He M.X., Hopf bifurcation in a delayed food-limited model with feedback control, Nonlinear Dynam., 2014, 76, 1215-122410.1007/s11071-013-1205-0Search in Google Scholar

[13] Chen L.J., Chen F.D., Dynamic behaviors of the periodic predator-prey system with distributed time delays and impulsive effect, Nonlinear Anal. RWA, 2011, 12, 2467-247310.1016/j.nonrwa.2011.03.002Search in Google Scholar

[14] Liu Z.J., Chen L.S., Positive periodic solution of a general discrete non-autonomous difference system of plankton allelopathy with delays, J. Comput. Appl. Math., 2006, 197, 446-45610.1016/j.cam.2005.09.023Search in Google Scholar

[15] Wang W.D., Chen L.S., A predator-prey system with stage-structure for predator, Computers Math. Applic., 1997, 33, 83-9110.1016/S0898-1221(97)00056-4Search in Google Scholar

[16] Beddington J.R., Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 1975, 44, 331-34010.2307/3866Search in Google Scholar

[17] DeAngelis D.L., Goldstein R.A., O’Neill R.V., A model for tropic interaction, Ecology, 1975, 56, 881-89210.2307/1936298Search in Google Scholar

[18] Chen F.D., Chen Y.M., Shi J.L., Stability of the boundary solution of a nonautonomous predator-prey system with the Beddington-DeAngel is functional response, J. Math. Anal. Appl., 2008, 344, 1057-106710.1016/j.jmaa.2008.03.050Search in Google Scholar

[19] Xia J., Yu Z.X., Zheng S.W., Stability and traveling waves in a Beddington-DeAngelis type stage-structured predator-prey reaction-diffusion systems with nonlocal delays and harvesting, Adv. Difference Equ., 2017, 2017, 6510.1186/s13662-017-1093-6Search in Google Scholar

[20] Chen F.D., Chen X.X., Huang S.Y., Extinction of a two species non-autonomous competitive system with Beddington-DeAngelis functional response and the effect of toxic substances, Open Math., 2016, 14, 1157-117310.1515/math-2016-0099Search in Google Scholar

[21] Khajanchi S., Dynamic behavior of a Beddington-DeAngelis type stage structured predator-prey model, Appl. Math. Comput., 2014, 244, 344-36010.1016/j.amc.2014.06.109Search in Google Scholar

[22] Chen F.D., Lin Q.X., Xie X.D., Xue Y.L., Dynamic behaviors of a nonautonomous modified Leslie-Gower predator-prey model with Holling-type III schemes and a prey refuge, J. Math. Comput. Sci., 2017, 17, 266-27710.22436/jmcs.017.02.08Search in Google Scholar

[23] Chen L.J., Chen F.D., Chen L.J., Qualitative analysis of a predator-prey model with Holling type II functional response incorporating a constant prey refuge, Nonlinear Anal. RWA, 2010, 11, 246-25210.1016/j.nonrwa.2008.10.056Search in Google Scholar

[24] Chen L.J., Chen F. D., Global stability and bifurcation of a ratio-dependent predator-prey model with prey refuge, Acta Math. Sin. Chin. Ser., 2014, 57, 301-310Search in Google Scholar

[25] Wang Y.Q., Chen L.J., Gao H.Y., Global analysis of a ratio-dependent predator-prey system incorporating a prey refuge, J. Nonlinear Funct. Anal., 2017, 2017, Article ID 5410.23952/jnfa.2017.54Search in Google Scholar

[26] Khajanchi S., Banerjee S., Role of constant prey refuge on stage structure predator-prey model with ratio dependent functional response, Appl. Math. Comput., 2017, 314, 193-19810.1016/j.amc.2017.07.017Search in Google Scholar

[27] Huang Y.J., Chen F.D., Li Z., Stability analysis of a prey-predator model with Holling type III response function incorporating a prey refuge, Appl. Math. Comput., 2006, 182, 2006, 672-68310.1016/j.amc.2006.04.030Search in Google Scholar

[28] Wei F.Y., Fu Q.Y., Hopf bifurcation and stability for predator-prey systems with Beddington-DeAngelis type functional response and stage structure for prey incorporating refuge, Appl. Math. Model., 2016, 40, 126-13410.1016/j.apm.2015.04.042Search in Google Scholar

[29] Nagumo N., Über die lage der integeralkueven gewönlicher differentialgleichunger, Proc. Phys. Math. Soc. Jpn., 1942, 24, 551-559Search in Google Scholar

[30] Kuang Y., Delay Differential Equations with Applications in Population Dynamics, 1993, New York: Academic Press.Search in Google Scholar

[31] Hu H.J., Huang L.H., Stability and Hopf bifurcation in a delayed predator-prey system with stage structure for prey, Nonlinear Anal. RWA, 2010, 11, 2757-276910.1016/j.nonrwa.2009.10.001Search in Google Scholar

[32] Chen F.D., Li Z., Huang Y.J., Note on the permanence of a competitive system with infinite delay and feedback controls, Nonlinear Anal. RWA, 2007, 8, 680-68710.1016/j.nonrwa.2006.02.006Search in Google Scholar

[33] Xu R., Ma Z.E., Stability and hopf bifurcation in a ratio-dependent predator-prey system with stage structure, Chaos Soliton. Fract., 2008, 38, 669-68410.1016/j.chaos.2007.01.019Search in Google Scholar

[34] Hale J.K., Theory of Functional Differential Equations, 1976, New York: Springer.10.1007/978-1-4612-9892-2Search in Google Scholar

Received: 2018-07-03
Accepted: 2018-12-12
Published Online: 2019-03-26

© 2019 Xiao et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Downloaded on 27.5.2024 from https://www.degruyter.com/document/doi/10.1515/math-2019-0014/html
Scroll to top button