
Citation: Xin-You Meng, Yu-Qian Wu. Bifurcation analysis in a singular Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting[J]. Mathematical Biosciences and Engineering, 2019, 16(4): 2668-2696. doi: 10.3934/mbe.2019133
[1] | Hongqiuxue Wu, Zhong Li, Mengxin He . Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting. Mathematical Biosciences and Engineering, 2023, 20(10): 18592-18629. doi: 10.3934/mbe.2023825 |
[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. Mathematical Biosciences and Engineering, 2023, 20(8): 15094-15119. doi: 10.3934/mbe.2023676 |
[3] | Rajalakshmi Manoharan, Reenu Rani, Ali Moussaoui . Predator-prey dynamics with refuge, alternate food, and harvesting strategies in a patchy habitat. Mathematical Biosciences and Engineering, 2025, 22(4): 810-845. doi: 10.3934/mbe.2025029 |
[4] | Renji Han, Binxiang Dai, Lin Wang . Delay induced spatiotemporal patterns in a diffusive intraguild predation model with Beddington-DeAngelis functional response. Mathematical Biosciences and Engineering, 2018, 15(3): 595-627. doi: 10.3934/mbe.2018027 |
[5] | G.A.K. van Voorn, D. Stiefs, T. Gross, B. W. Kooi, Ulrike Feudel, S.A.L.M. Kooijman . Stabilization due to predator interference: comparison of different analysis approaches. Mathematical Biosciences and Engineering, 2008, 5(3): 567-583. doi: 10.3934/mbe.2008.5.567 |
[6] | Peter A. Braza . A dominant predator, a predator, and a prey. Mathematical Biosciences and Engineering, 2008, 5(1): 61-73. doi: 10.3934/mbe.2008.5.61 |
[7] | Xiaoying Wang, Xingfu Zou . Pattern formation of a predator-prey model with the cost of anti-predator behaviors. Mathematical Biosciences and Engineering, 2018, 15(3): 775-805. doi: 10.3934/mbe.2018035 |
[8] | Zhihong Zhao, Yan Li, Zhaosheng Feng . Traveling wave phenomena in a nonlocal dispersal predator-prey system with the Beddington-DeAngelis functional response and harvesting. Mathematical Biosciences and Engineering, 2021, 18(2): 1629-1652. doi: 10.3934/mbe.2021084 |
[9] | Yuanfu Shao . Bifurcations of a delayed predator-prey system with fear, refuge for prey and additional food for predator. Mathematical Biosciences and Engineering, 2023, 20(4): 7429-7452. doi: 10.3934/mbe.2023322 |
[10] | Fang Liu, Yanfei Du . Spatiotemporal dynamics of a diffusive predator-prey model with delay and Allee effect in predator. Mathematical Biosciences and Engineering, 2023, 20(11): 19372-19400. doi: 10.3934/mbe.2023857 |
In recent decades, an increasing number of scholars have paid attention to population dynamics of prey-predator ecosystem [1,2,3]. During investigating such biological phenomena, dynamical behavior of biological and mathematical models are affected by many factors, such as the function response. All kinds of predator-prey models with Holling type, Crowley-Martin type and Leslie-Gower type, etc. have been investigated extensively by the researchers [4,5,6,7,8,9]. However, a few of literatures have studied the predator-prey systems with Beddington-DeAngelis type functional response [10,11,12,13,14,15,16,17].
The Beddington et al. [18,19] gave the follows functional response
g(x1,x2)=l1x1lx1+l2x2+d. | (1.1) |
Here, l1 represents that per predator population per time can eat the maximum number of prey population, l is a positive constant and denotes the effect of handling time for predators, and l2 is positive and measures the magnitude of interference among predators, d represents the prey density where the attack rate is half-saturated. It is obvious that two cases are possible as following. One case is that, if l=1, l2=0 and d>0, then it reduces to a Holling Ⅱ functional response (or Michaelis-Menten functional response) [20]. The other case is that, if l=0, l2=0 and d>0, then it reduces to a linear mass-action functional response (or Holling Ⅰ functional response) [21].
Conforto et al. [16] considered a three-dimensional reaction-diffusion system incorporating the dynamics of handling and searching predators, and showed that its solutions converge when a small parameter tends to 0 towards the solutions of a reaction-cross diffusion system of predator-prey type involving a Holling-type Ⅱ or Beddington-DeAngelis functional response. Employing the upper and lower solution method and comparison theory, Li et al. [12] got the sufficient conditions of the upper ultimate boundedness and permanence of this system which implies that impulse always changes the situation of survival for species, and obtained the conditions for the existence of unique globally stable positive periodic solution.
Generally speaking, the introduction of time lag in the mathematical models [22,23,24,25,26,27,28,29,30,31,32] tends to reflect the interaction and coexistence mechanism of population in the past. System with two delays is discussed by using the equivalent system with a single time delay in reference [30]. Liu et al. [33] used a Markovian switching process to model the telephone noise in the environment, proposed a stochastic regime-switching predator-prey model with harvesting and distributed delays, obtained the sufficient and necessary conditions for the existence of an optimal harvesting policy, and gave the explicit forms of the optimal harvesting effort and the maximum of sustainable yield.
Biological resources in the predator-prey system tend to be harvested and sold in order to obtain economic profit. In general, three types of harvesting function have been studied in the literatures [32,34,35,36,37].
(Ⅰ) Constant harvesting function is
H(x,E)=C, |
where C is a suitable constant.
(Ⅱ) Proportionate harvesting function is
H(x,E)=qEx, |
where q is the catchability coefficient.
(Ⅲ) Nonlinear harvesting function is
H(x,E)=qExm1E+m2x, |
where m1,m2 are suitable positive constants.
It is easy to find that there are several unrealistic features in the proportional harvesting, such as stochastic search for prey and unbounded linear harvesting. However, the nonlinear harvesting function (Ⅲ) eliminates the above unrealistic features and satisfies
limE→∞H(x,E)=qxm1,limx→∞H(x,E)=qEm2, |
that is to say, the nonlinear harvesting function shows saturation effects in terms of harvest effort and inventory abundance [38].
As we known that the harvested prey-predator systems have been focused by both theoretical and mathematical biologists [15,39,40]. Chakraborty et al. [40] studied the global dynamics and bifurcation of the predator-prey with constant harvesting. Liu et al. [15] investigated Hopf bifurcation and center stability for a predator-prey biological economic model with linear prey harvesting. Liu et al. [39] discussed bifurcation in a prey-predator model with nonlinear predator harvesting, and obtained the conditions for Turing and Hopf bifurcation. However, little work has been done on dynamic effects of economic interest on prey-predator system with nonlinear predator harvesting and Beddington-DeAngelis functional response.
In 1954, the common-property resource economic theory was proposed by Gordon in [41], which investigated the dynamic effects of harvesting efforts on the ecosystem from an economic point of view. Thus, in order to study the economic interest of commercial harvesting, an equation is proposed:
NetEconomicRevenue(NER)=Total Revenue (TR)−Total Cost (TC). | (1.2) |
Recently, a number of delay differential algebraic systems were proposed to investigate the impact of commercial harvesting on prey-predator model [42,43]. Liu et al. [44] investigated a delayed differential-algebraic system with double time delays, Holling type Ⅱ functional response and linear commercial harvesting effort on predator population. By jointly using the normal form of differential algebraic system and the bifurcation theory, Li et al. [45] discussed the stability and bifurcations, and obtained richer dynamics of the bioeconomic differential algebraic predator-prey model with nonlinear prey harvesting.
Due to the above analysis and discussion, we will extend the work in [44,45] by considering nonlinear predator harvesting, double delays and Beddington-DeAngelis functional response function into our bioeconomic differential algebraic predator-prey system in this paper. The aim of our work is to reveal the dynamical behavior of Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting, to obtain a reasonable profit and provide some guidance for the harvest of biological economy system.
The rest of the paper is organized as follows. In Section 2, a differential algebraic predator-prey model including two delays, Beddington-DeAngelis functional response and nonlinear predator harvesting is proposed. In Section 3, positive solutions are analyzed when m=0 and m>0, respectively. In Section 4, in the absence of time delay, the singularity induced bifurcation is discussed by using differential-algebraic system theory, what's more, state feedback controllers are designed to remove singular induced bifurcation. In the presence of time delay, by analyzing the corresponding characteristic transcendental equation, the local stability around the interior equilibrium are studied. Furthermore, the existence of Hopf bifurcation is investigated. Directions of Hopf bifurcation and stability of the bifurcating periodic solutions are also discussed by using normal form theory and center manifold theorem. Numerical simulations illustrate the effectiveness of the mathematical conclusions in Section 5. Discussions and conclusions are included in the last section.
Leslie and Gower [46] proposed the classical predator-prey system, which is as follows
{.X(T)=s1X(T)(1−X(T)K)−l1X(T)Y(T)X(T)+l2Y(T),.Y(T)=s2Y(T)(1−l3Y(T)X(T)). | (2.1) |
Here, X(T), Y(T) indicate prey and predator population at time T respectively, s1, s2 and K are intrinsic growth rate of prey, intrinsic growth rate of predator and carrying capacity. l1 is the maximum number of prey each predator can eat each time and l2 is a semi saturation constant, and the predation rate is l22. The carrying capacity of predator is proportional to the size of prey population. l3 is the amount of prey needed to support a predator population in equilibrium.
Based on the fact that Beddington-DeAngelis functional response is more authentic [47], we introduce it in system (2.1). Thus, the new system is
{.X(T)=s1X(T)(1−X(T)K)−l1X(T)Y(T)lX(T)+l2Y(T)+d,.Y(T)=s2Y(T)(1−l3Y(T)X(T)). | (2.2) |
Considering nonlinear predator harvesting, system (2.2) is constructed as follows
{.X(T)=s1X(T)(1−X(T)K)−l1X(T)Y(T)lX(T)+l2Y(T)+d,.Y(T)=s2Y(T)(1−l3Y(T)X(T))−qE(T)Y(T)m1E(T)+m2Y(T). | (2.3) |
By using following transformations
t=s1T,x=XK,y=l1Ys1K,α=Km2s1m1l1,β=s2l3l1,r=l1s1l3,d=K,q=s1m1,l=a,b=l2s1l1, |
system (2.3) is non-dimensionalized as follows
{.x(t)=x(t)(1−x(t))−x(t)y(t)1+ax(t)+by(t),.y(t)=βy(t)(r−y(t)x(t))−E(t)y(t)E(t)+αy(t), | (2.4) |
where a,b,α,β,r,q,p,c,m1 and m2 are all positive constants.
Simultaneously, an algebraic equation is also included due to the economic profit of harvesting. Based on Eq (1.2), we have
m=TR−TC=qE(t)(py(t)−c)m1E(t)+m2y(t), | (2.5) |
here p is the price per unit harvested biomass, c is the cost per unit harvest, while m means the net economic revenue.
In general, the delay differential equation model can produce more efficient and accurate dynamics than the ordinary differential equation model as capturing oscillation dynamics.
Therefore, by combining the above biological economic algebraic equation and time delay, a differential-algebraic system with two time delays is given as follows
{.x(t)=x(t−τ1)(1−x(t−τ1))−x(t−τ1)y(t)1+ax(t−τ1)+by(t),.y(t)=βy(t−τ2)(r−y(t−τ2)x(t−τ2))−E(t)y(t)E(t)+αy(t),0=qE(t)m1E(t)+m2y(t)(py(t)−c)−m, | (2.6) |
where τ1 is maturation time for prey population, τ2 is gestation delay for predator population. The initial conditions of system (2.6) take the following form:
x(θ)>0,y(θ)>0,E(0)>0,θ∈[−τ,0],τ=max{τ1,τ2}. |
System (2.6) can be transformed into matrix form as follows
ˉA.U(t)=f(U), | (2.7) |
where
U(t)=(x(t)y(t)E(t)),ˉA=(100010000), |
f(U)=(f1(U)f2(U)f3(U))=(x(t−τ1)(1−x(t−τ1))−x(t−τ1)y(t)1+ax(t−τ1)+by(t)βy(t−τ2)(r−y(t−τ2)x(t−τ2))qE(t)(py(t)−c)m1E(t)+m2y(t)−m). |
Remark 1. Compared with system proposed in [44], nonlinear predator harvesting and Beddington-DeAngelis functional response are considered in system (2.6).
Remark 2. Compared with system proposed in [45], system (2.6) contains two delays and Beddington-DeAngelis functional response, and focuses on economic interest of commercial harvest effort on predator.
For system (2.6), there is a bioeconomic equilibrium state when m=0. First, we give the following assumptions:
(H1): c>rx0andpx0(βr−1)>βc,
(H2): c<rx0andpx0(βr−1)<βc,
(H3): bcp+1−cp>0.
Therefore, if one of the assumptions (H1) and (H2) holds, then the interior equilibrium P0=(x0,y0,E0)=(x0,cp,pacβ(c−rx0)px0(βr−1)−βc) exists. Here x0 satisfies the following equation
ax02+(bcp+1−a)x0−bcp−1+cp=0. | (3.1) |
Obviously, a simple sufficient condition that Eq (3.1) has at least one positive root is that (H1) and (H3) or (H2) and (H3) hold.
In the case of m>0, we suppose
(H4): pq(1+ax∗−x∗−ax∗2)>(qc+mm1)(1+bx∗−b)>0,
(H5): pq(1+ax∗−x∗−ax∗2)<(1+bx∗−b)(qc+mm1)<0.
If one of the assumptions (H4) and (H5) holds, then the interior equilibrium P∗=(x∗,y∗,E∗) exists, here E∗=mm2(1+ax∗−x∗−ax∗2)pq(1+ax∗−x∗−ax∗2)−(qc+mm1)(1+bx∗−b),y∗=1+ax∗−x∗−ax∗21+bx∗−b and x∗ satisfies the following equation
A1x6+A2x5+A3x4+A4x3+A5x2+A6x+A7=0, | (3.2) |
here
A1=a2αbβpqr+a3αβpq,A2=−aβmrm2b2−(αβr(a−1)b−αβra(−b+1))pqa−αβrab(pq(a−1)−(qc+mm1)b)−a2bβmm2−2αβ(a−1)a2pq−a2αβ(pq(a−1)−(qc+mm1)b)+ab2mm2,A3=βmrm2(a−1)b2−2aβmrm2(−b+1)b+αβr(2ab−b−a)(pq(a−1)−(qc+mm1)b)−(αβrb+αβr(a−1)(−b+1)−a2βmm2(−b+1)+αβ(−2a+(a−1)2))pqa−αβrab(pq−(qc+mm1)(−b+1))+2βmm2(a−1)ab−a2αβ(pq−(qc+mm1)(−b+1))+2αβ(a−1)a(pq(a−1)−(qc+mm1)b)−mm2((a−1)b2+2a(−b+1)b),A4=βmrm2b2+2βmrm2(a−1)(−b+1)b−aβmrm2(−b+1)2−αβr(−b+1)pqa+(αβrb+αβr(a−1)(−b+1))(pq(a−1)−(qc+mm1)b)+(pq−(qc+mm1)(−b+1))(αβr(a−1)b−αβra(−b+1))−βmm2(−2a+(a−1)2)b+2βmm2(a−1)a(−b+1)+αβ(2a−2)pqa−αβ(−2a+(a−1)2)(pq(a−1)−(qc+mm1)b)+2αβ(a−1)a(pq−(qc+mm1)(−b+1))−b2mm2−2mm2(a−1)(−b+1)b+amm2(−b+1)2,A5=αβ(pq−(cq+mm1)(−b+1))(rb+r(a−1)(−b+1)+2a−(a−1)2)+αβ(r(−b+1)−2a+2)pq(a−1)−(cq+mm1)b+aαβpq−mm2(2(−b+1)b−(a−1)(−b+1)2−β(−2a+(a−1)2)(−b+1)−β(2a−2)b)+mm2(2βr(−b+1)b+βr(a−1)(−b+1)2),A6=βmrm2(−b+1)2+(αβr(−b+1)−αβ(2a−2))(pq−(cq+mm1)(−b+1))−bβmm2−βmm2(2a−2)(−b+1)−αβ(pq(a−1)−(cq+mm1)b)−mm2(−b+1)2,A7=−βmm2(−b+1)−αβ(pq−(cq+mm1)(−b+1)). |
Based on Routh-Hurwitz criterion [1], Eq (3.2) has at least one positive root when 1<b and αm1<m2. Suppose (τ1,τ2,m)∈Ω1, here
Ω1={(τ1,τ2,m)|τ1≥0,τ2≥0,0<m<αβ(pq+qc(b−1))β(1−b)(αm1−m2)}. |
Here, we are interested in the stability of system (2.6) at the interior equilibrium P0.
When τ1=0,τ2=0, system (2.6) takes the following form
{.x(t)=x(t)(1−x(t))−x(t)y(t)1+ax(t)+by(t),.y(t)=βy(t)(r−y(t)x(t))−E(t)y(t)E(t)+αy(t),0=qE(t)m1E(t)+m2y(t)(py(t)−c)−m. | (4.1) |
Lemma 4.1. (Singularity induced bifurcation theorem [49,50]). If the differential-algebraic equations satisfy the following conditions at the singular equilibrium:
(I) Dyf3 has a simple zero eigenvalue and
trace((DEf1DEf2)adj(DEf3)(Dxf3Dyf3))p0>0, |
(II)
(Dxf1Dyf1DEf1Dxf2Dyf2DEf2Dxf3Dyf3DEf3) |
is nonsingular,
(III)
(Dxf1Dyf1DEf1Dvf1Dxf2Dyf2DEf2Dvf2Dxf3Dyf3DEf3Dvf3DxΔDyΔDEΔDvΔ) |
is also nonsingular, here Δ=DEf3, then the singularity induced bifurcation occurs at the singular equilibrium.
Theorem 4.2. A singularity induced bifurcation takes place at the interior equilibrium P0 of the differential algebraic system (4.1). When the bifurcation parameter m increases through zero, system (4.1) is unstable at P0.
Proof. We can obtain that the Jacobin matrix of system (4.1) around P0 is
JP0=(J11J120βy02x02J22−αy02(αy0+E0)20pqE0(E0m1+m2y0)20), |
where J11=1−2x0−(b+x0y0)(1−x0)2,J12=−x0(1−x0)2(ax0+1)y02,J22=−E02(αy0+E0)2+βr−2βy0x0.
Let m be bifurcation parameter, D be differential operator. We can obtain the following results.
(Ⅰ)
trace((DEf1DEf2)adj(DEf3)(Dxf3Dyf3))P0=αy02pE0(pm1E0+cm2)(αy0+E0)2(m1E0+m2y0)2>0. |
(Ⅱ) If J11≠0 holds, then it can be obtained that
det(Dxf1Dyf1DEf1Dxf2Dyf2DEf2Dxf3Dyf3DEf3)P0=αy2qE0(pE0m1+cm2)J11(αy0+E0)2(m1E0+m2y0)2≠0. |
(Ⅲ) Defining Δ=DEf3=qm2y(py−c)(m1E+m2y)2, it follows from simple computations that
det(Dxf1Dyf1DEf1Dvf1Dxf2Dyf2DEf2Dvf2Dxf3Dyf3DEf3Dvf3DxΔDyΔDEΔDvΔ)P0=−αy03pqm2J11(αy0+E0)2(m1E0+m2y0)3≠0. |
Hence a singularity induced bifurcation occurs around P0 of system (4.1) and the bifurcation value is m=0.
On the other hand,
C1=trace((DEf1DEf2)adj(DEf3)(Dxf3Dyf3))P0=αy02pE0(pm1E0+cm2)(αy0+E0)2(m1E0+m2y0)2>0,C2=(DvΔ−(DxΔDyΔDEΔ)P−11(Dvf1Dvf2Dvf3))P0=m2y0E0(m1E0+m2y0)2>0, |
where P1=(Dxf1Dyf1DEf1Dxf2Dyf2DEf2Dxf3Dyf3DEf3).
According to Lemma 4.1 and Theorem 3 in [49], when m increases through 0, one eigenvalue of system (4.1) moves from C− to C+ along real axis by diverging through ∞. Hence, the Theorem 4.2 is proved.
By using the similar proof of Theorem 4.2, it is easy to show system (4.1) is unstable around P∗, what's more, state feedback controllers are designed to remove singularity induced bifurcation and stabilize system (4.1) around P0 and P∗, respectively.
In the case of m=0, according to the leading matrix ˉA in (4.1) and JP0, we can obtain rank (JP0,ˉAJP0,ˉA2JP0)=3. It is not hard to find that system (4.1) is locally controllable around P0 based on Theorem 2-2.1 in [48]. Therefore, a feedback controller can be used to stabilize system (4.1) around P0. By using Theorem 3-1.2 in [48], a feedback controller is as follows
v(t)=(001)(00ˆk)(x(t)−x0y(t)−y0E(t)−E0), |
where ˆk is a feedback gain.
Applying the controller into system (4.1), we can obtain that a controlled system is
{.x(t)=x(t)(1−x(t))−x(t)y(t)1+ax(t)+by(t),.y(t)=βy(t)(r−y(t)x(t))−E(t)y(t)E(t)+αy(t),0=qE(t)(py(t)−c)m1E(t)+m2y(t)+ˆk(E(t)−E0). | (4.2) |
Theorem 4.3. If the feedback gain ˆk>max{Q1,Q2,Q3}, here
Q1=αy02pqE0(αy0+E0)(E0m1+m2y0)2,Q2=−αy02pqE0(αy0+E0)(E0m1+m2y0)2J11J22,Q3=−αy02pqE0J11(αy0+E0)(E0m1+m2y0)2(−J11−J22+β(1−x0)2(ax0+1)x0), |
then system (4.2) is stable around P0.
Proof. At first, the Jacobin matrix of system (4.2) around P0 is
ˆJP0=(J11J120βy02x02J22−αy02(αy0+E0)20pqE0(E0m1+m2y0)2ˆk). |
The characteristic equation of system (4.2) around P0 is det(λˉA−ˆJP0)=0 based on ˉA in system (4.2) and ˆJP0, which can be written as follows
λ2+Δ1λ+Δ2=0, |
where
Δ1=−J11−J22−1ˆkB1,Δ2=−J11J22+B2+1ˆkJ11B1,B1=αy02pqE0(αy0+E0)2(E0m1+m2y0),B2=β(1−x0)2(ax0+1)x0. |
By simple computation, if ˆk>max{Q1,Q2,Q3}, then system (4.2) is stable around P0.
Similarly, in the case of m>0, a controlled system is
{.x(t)=x(t)(1−x(t))−x(t)y(t)1+ax(t)+by(t),.y(t)=βy(t)(r−y(t)x(t))−E(t)y(t)E(t)+αy(t),0=qE(t)(py(t)−c)m1E(t)+m2y(t)−m+˜k(E(t)−E∗), | (4.3) |
where ˜k is a feedback gain.
By using the similar analysis in Theorem 4.3, we can obtain the following result.
Theorem 4.4. If controller feedback gain ˜k>max{˜Q1,˜Q2}, where
˜Q1=αy∗qE∗(m2c+m1pE∗)−qm2y∗(py∗−c)(αy∗+E∗)2ξ5(αy∗+E∗)2(m1E∗+m2y∗)2(ξ1+ξ3),˜Q2=αy∗2qE∗(m2c+m1pE∗)ξ1(αy∗+E∗)2(m1E∗+m2y∗)2[ξ1+ξ3+ξ4]−qm2y∗(py∗−c)(m1E∗+m2y∗)2, |
ξ0=E∗2(E+αy∗)2−αy∗E∗(pm1E∗+cm2)m2(py∗−c)(E∗+αy∗)2,ξ2=ξ0ξ1,ξ1=2x∗+x∗(1−x∗)2y∗+b(1−x∗)2−1,ξ5=ξ1ξ3,ξ3=E∗2(E+αy∗)2+2βy∗x∗−βr,ξ4=β(1−x∗)2(1+ax∗)x∗, |
then system (4.3) is stable around P∗.
Remark 3. According to design of feedback controller, we can make interior equilibria be stable, which shows that prey-predator ecosystem can be kept sustainable and economic interest can be kept ideal by controlling commercial harvest effort on predator.
By analyzing system (2.6), a characteristic equation of around P∗ is
λ2+ξ0λ+(ξ1λ+ξ2)e−λτ1+(ξ3λ+ξ4)e−λτ2+ξ5e−λ(τ1+τ2)=0. | (4.4) |
When τ1>0, and τ2=0, Eq (4.4) takes the following form
λ2+(ξ0+ξ3)λ+(ξ1λ+ξ2+ξ5)e−λτ1+ξ4=0. | (4.5) |
By simple computation, we can obtain that λ=0 is not a root of Eq (4.5). We suppose that λ=iβ1 (β1 is a positive real number) is a root of Eq (4.5). By separating real and imaginary parts, we can obtain the following two transcendental equations
{ξ1β1sin(β1τ1)+(ξ2+ξ5)cos(β1τ1)=β21−ξ4,ξ1β1cos(β1τ1)−(ξ2+ξ5)sin(β1τ1)=−β1(ξ0+ξ3). | (4.6) |
By computing two equations in Eq (4.6), it gives that
β41+[(ξ0+ξ3)2−2ξ4−ξ21]β21+ξ24−(ξ2+ξ5)2=0. | (4.7) |
Theorem 4.5. If ξ24−(ξ2+ξ5)2<0 holds,
(i) system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω2;
(ii) system (2.6) undergoes Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω3. Here, Ω2 and Ω3 are defined as follows
Ω2={(τ1,τ2,m)|0<τ1<τ10,τ2=0,m>0},Ω3={(τ1,τ2,m)|τ1=τ10,τ2=0,0<m<−αy∗pqE∗m2(E∗+αy∗)2N1−m2E∗2}, |
where N1=(ξ21+2ξ4)12+E∗αy∗(E+αy∗)2−βy∗x∗ and (E∗+αy∗)2N1<E∗2.
Proof. If ξ24−(ξ2+ξ5)2<0, based on Routh-Hurwitz criterion [1], we can guarantee that Eq (4.7) has at least one positive root. Consequently, Eq (4.5) has a pair of purely imaginary roots λ=±iβ∗1.
Based on Eq (4.6), we can calculate τ1k as follows
τ1k=1β∗1arccos[(ξ2+ξ5)−ξ1(ξ0+ξ3)]β∗21−ξ4(ξ2+ξ5)ξ21β∗21+(ξ2+ξ5)2+2kπβ∗1. | (4.8) |
By using Butlers lemma [55], system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω2.
Next, we will determine
Θ=sign{d(Reλ)dτ1}λ=iβ∗1=sign{Re(dλdτ1)−1}λ=iβ∗1. |
By differentiating Eq (4.5) with respect to τ1, we have
(dλdτ1)−1=2λ+(ξ0+ξ3)−λ(λ2+(ξ0+ξ3)λ+ξ4)+ξ1ξ1λ2+(ξ2+ξ5)λ−τ1λ. |
By virtue of Eq (4.6), we have
Θ=sign{Re(dλdτ1)−1}λ=iβ∗1=sign{2β∗21+(ξ0+ξ3)2−2ξ4−ξ21(β∗21−ξ4)2+(ξ0+ξ3)2β∗21}. |
If 0<m<−αy∗pqE∗m2(E∗+αy∗)2N1−m2E∗2 and (E∗+αy∗)2N1<E∗2, then (ξ0+ξ3)2−2ξ4−ξ21>0 holds, which implies that Θ>0. Hence, when (τ1,τ2,m)∈Ω1∩Ω3, transversality condition holds and Hopf bifurcation occurs.
When τ1=0 and τ2>0, Eq (4.4) can be written the following form
λ2+(ξ0+ξ1)λ+ξ2+(ξ3λ+ξ4+ξ5)e−λτ2=0. | (4.9) |
Similarly, it shows that λ=0 is not a root of Eq (4.9). We suppose that λ=iβ2 (β2>0) is a root of Eq (4.9). We can obtain the following two transcendental equations
{ξ3β2sin(β2τ2)+(ξ4+ξ5)cos(β2τ2)=β2−ξ2,(ξ4+ξ5)sin(β2τ2)−ξ3β2cos(β2τ2)=(ξ0+ξ1)β2, | (4.10) |
which gives that
β42+[(ξ0+ξ1)2−2ξ2−ξ23]β22+ξ22−(ξ4+ξ5)2=0. | (4.11) |
Theorem 4.6. If ξ22−(ξ4+ξ5)2<0 holds,
(i) system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω4;
(ii) system (2.6) undergoes Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω5, where Ω4 and Ω5 are defined as follows
Ω4={(τ1,τ2,m)|τ1=0,0<τ2<τ20,m>0},Ω5={(τ1,τ2,m)|τ1=0,τ2=τ20,0<m<αpqy∗E∗√m2(−ξ22+ξ32)(αy∗+E∗)2−m2E∗}, |
here, √m2(−ξ22+ξ32)(αy∗+E∗)2>m2E∗.
By computing Eq (4.10), we can obtain τ2k is
τ2k=1β∗2arccos[(ξ4+ξ5)−ξ3(ξ0+ξ1)]β∗22−ξ2(ξ4+ξ5)ξ23β∗22+(ξ4+ξ5)2+2kπβ∗2. | (4.12) |
The proof of Theorem 4.6 is similar to that of Theorem 4.5. So, we omit it.
In this part, τ1 is considered as a parameter while τ2 is regarded as a fixed value ˆτ2∈(0,τ20), which is a stable interval calculated in Subsection 4.2.2. Here Ω6 is defined as follows
Ω6={(τ1,τ2,m)|τ1>0,τ2=ˆτ2∈(0,τ20),m>0}. |
Let λ=iα1 (α1 is a positive real number) represent a purely imaginary root of Eq (4.4). By separating real and imaginary parts, we have the following two transcendental equations
{α21−ξ3α1sin(α1ˆτ2)−ξ4cos(α1ˆτ2)=ξ1α1sin(α1τ1)+ξ2cos(α1τ1)+ξ5cos[α1(τ1+ˆτ2)],ξ0α1+ξ3α1cos(α1ˆτ2)−ξ4sin(α1ˆτ2)=ξ2sin(α1τ1)−ξ1α1cos(α1τ1)+ξ5sin[α1(τ1+ˆτ2)]. | (4.13) |
Based on Eq (4.13), it derives that
cos(α1τ1)=l10(α1,ˆτ2)l12(α1,ˆτ2),sin(α1τ1)=l11(α1,ˆτ2)l12(α1,ˆτ2), | (4.14) |
where
l10(α1,ˆτ2)=(ξ2−ξ0ξ1)α21−ξ4ξ5+[(ξ5−ξ1ξ3)α21−ξ2ξ4]cos(α1ˆτ2)+(ξ0ξ5−ξ2ξ3+ξ1ξ4)α1sin(α1ˆτ2),l11(α1,ˆτ2)=ξ1α31+(ξ0ξ2+ξ3ξ5)α1−[(ξ5+ξ1ξ3)α21+ξ2ξ4]sin(α1ˆτ2)+(ξ0ξ5+ξ2ξ3−ξ1ξ4)α1cos(α1ˆτ2),l12(α1,ˆτ2)=[ξ2+ξ5cos(α1ˆτ2)]2+[ξ5sin(α1ˆτ2)−ξ1α1]2. |
Let
L1(α1,ˆτ2)=l210(α1,ˆτ2)+l211(α1,ˆτ2)−l212(α1,ˆτ2). | (4.15) |
Due to its complicated form, it is not easy for us to analyze properties of roots of transcendental equation (4.15). Based on dynamical system theory [1], we know that if and only if every eigenvalue has negative real part, system (2.6) is locally asymptotically stable around P∗. What's more, by analyzing existence of Hopf bifurcation around the corresponding interior equilibrium P∗, the periodic oscillation of system (2.6) is investigated. Hale [51] proposed that when the corresponding eigenvalue has a pair of purely imaginary roots, system usually exhibits Hopf bifurcation. Obviously, if Eq (4.15) has finite positive and simple roots 0<α10<α11<⋯<α1n, Eq (4.4) has a pair of purely imaginary roots.
Without loss of generality, we denote α1c= max {α1k},k=0,1,2,⋅⋅⋅n and regard τ1 as the bifurcation parameter, while we have the following corresponding critical value τ1c
τ1c=min{ω1c+2kπα1c}, | (4.16) |
here, ω1c∈[0,2π] satisfies the following equations
cosω1c=l10(α1c,ˆτ2)l12(α1c,ˆτ2),sinω1c=l11(α1c,ˆτ2)l12(α1c,ˆτ2). | (4.17) |
Furthermore, it is important to check transversal condition. The necessary condition for the existence of Hopf branches is that the eigenvalues passes through the imaginary axis with non-zero speed [51].
By differentiating λ with respect to τ1 in Eq (4.4), it derives that
dλdτ1=λ[(ξ1λ+ξ2)e−λτ1+ξ5e−λ(τ1+ˆτ2)]h1(λ,ˆτ2,τ1), | (4.18) |
where
h1(λ,ˆτ2,τ1)=2λ+ξ0+[ξ1−(ξ1λ+ξ2)τ1]e−λτ1+[ξ3−(ξ3λ+ξ4)ˆτ2]e−λˆτ2−ξ5(τ1+ˆτ2)e−λ(τ1+ˆτ2). |
By virtue of Eq (4.18), it can be obtained that
ˆΘ=sign{Re(dλdτ1)−1}τ1=τ1c=sign{B11B13+B12B14B213+B214}, | (4.19) |
where
B11=[(ξ2τ1c−ξ1)α1c−ξ5(τ1c+ˆτ2)α1csin(α1cˆτ2)]sin(α1cτ1c)+[ξ5(τ1c+ˆτ2)α1ccos(α1cˆτ2)−ξ1τ1cα21c]cos(α1cτ1c),+2α21c−ξ3ˆτ2α21ccos(α1cˆτ2)+(ξ4ˆτ2−ξ3)α1csin(α1cˆτ2)B12=−ξ0α1c+ξ3ˆτ2α1csin(α1cˆτ2)+(ξ4ˆτ2−ξ3)α1ccos(α1cˆτ2)+[(ξ2τ1c−ξ1)α1c+ξ5(τ1c+ˆτ2)α1ccos(α1cˆτ2)]cos(α1cτ1c)+[ξ1τ1cα21c−ξ5(τ1c+ˆτ2)α1csin(α1cˆτ2)]sin(α1cτ1c),B13=α41c−[ξ3α1csin(α1cˆτ2)+ξ4cos(α1cˆτ2)]α21c,B14=−ξ0α31c−[ξ3α1ccos(α1cˆτ2)−ξ4sin(α1cˆτ2)]α21c. |
It is obvious to show ˆΘ>0 when B11B13+B12B14>0. Therefore, we have the following results on stability and bifurcation in system (2.6).
Theorem 4.7. For system (2.6), suppose that B11B13+B12B14>0 holds and (τ1,τ2,m)∈Ω1∩Ω6.
(i) If Eq (4.15) has no positive root, then system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω6.
(ii) If Eq (4.15) has at least one positive and simple root α∗1, there exists a critical delay τ∗1=min{ω∗1+2kπα∗1}>0 such that system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω6∩Ω7, here
Ω7={(τ1,τ2,m)∣0<τ1<τ∗1,τ2=ˆτ2,m>0}. |
(iii) If Eq (4.15) has finite positive and simple roots 0<α10<α11<⋅⋅⋅<α1n, there exists a critical delay τ1c defined in (4.16) such that system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω6∩Ω8, here
Ω8={(τ1,τ2,m)∣0<τ1<τ1c,τ2=ˆτ2,m>0}. |
If B11B13+B12B14>0, then system (2.6) undergoes a Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω6∩Ω9, here
Ω9={(τ1,τ2,m)∣τ1=τ1c,τ2=ˆτ2,m>0}. |
In this part, τ2 is seen as a parameter, while τ1 is regarded as a fixed value ˆτ1∈(0,τ10) that is a stable interval calculated in Subsection 4.2.1. Ω10 is defined as follows:
Ω10={(τ1,τ2,m)|τ1=ˆτ1∈(0,τ10),τ2>0,m>0}. |
Let λ=iα2 (α2 is a positive real number) represent a purely imaginary root of Eq (4.4). We define
L2(α2,ˆτ1)=l220(α2,ˆτ1)+l221(α2,ˆτ1)−l222(α2,ˆτ1)=0, | (4.20) |
where,
l20(α2,ˆτ1)=(ξ4−ξ0ξ3)α22−ξ2ξ5+[(ξ5−ξ1ξ3)α22−ξ2ξ4]cos(α2ˆτ1)+(ξ0ξ5+ξ2ξ3−ξ1ξ4)α2sin(α2ˆτ1),l21(α2,ˆτ1)=ξ3α32+(ξ0ξ4+ξ1ξ5)α2−[(ξ5+ξ1ξ3)α22+ξ2ξ4]sin(α2ˆτ1)+(ξ0ξ5−ξ2ξ3+ξ1ξ4)α2cos(α2ˆτ1),l22(α1,ˆτ2)=[ξ4+ξ5cos(α2ˆτ1)]2+[ξ5sin(α2ˆτ1)−ξ3α2]2. |
Theorem 4.8. For system (2.6), suppose that B21B23+B22B24>0 holds and (τ1,τ2,m)∈Ω1∩Ω10.
(i) If Eq (4.20) has no positive root, then system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω10.
(ii) If Eq (4.20) has at least one positive and simple root α∗2, there exists a critical delay τ∗2=min{ω∗2+2kπα∗2}>0 such that system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω10∩Ω11, here
Ω11={(τ1,τ2,m)∣τ1=ˆτ1,0<τ2<τ∗2,m>0}. |
(iii) If Eq (4.20) has finite positive and simple roots 0<α20<α21<⋅⋅⋅α2n, there exists a critical delay τ2c defined in (4.19) such that system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω10∩Ω12, here
Ω12={(τ1,τ2,m)∣τ1=ˆτ1,0<τ2<τ2c,m>0}. |
If B21B23+B22B24>0 holds, then system (2.6) undergoes a Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω10∩Ω13, here
Ω13={(τ1,τ2,m)∣τ1=ˆτ1,τ2=τ2c,m>0}. |
Where the corresponding critical value τ2c satisfies
τ2c=min{ω2c+2kπα2c}, | (4.21) |
here ω2c∈[0,2π] satisfies
cosω2c=l20(α2c,ˆτ1)l22(α2c,ˆτ1),sinω2c=l21(α2c,ˆτ1)l22(α2c,ˆτ1). |
Similarly, by differentiating λ with respect to τ2 in Eq (4.4) and computing Eq (4.13), we can obtain
ˉΘ=sign{Re(dλdτ2)−1}τ2=τ2c=sign{B21B23+B22B24B223+B224}, |
here
B21=2α22c+(ξ2ˆτ1−ξ1)α2csin(α2cˆτ1)−ξ1α22ccos(α2cˆτ1)+[(ξ4τ2c−ξ3)α2c+ξ5(τ2c+ˆτ1)α2ccos(α2cˆτ1)]sin(α2cτ2c)+[ξ5(τ2c+ˆτ1)α2csin(α2cˆτ1)−ξ3τ2cα22c]cos(α2cτ2c),B22=−ξ0α2c+ξ1α21csin(α2cˆτ1)+(ξ2ˆτ1−ξ1)α2ccos(α2cˆτ1)+[(ξ4τ2c−ξ3)α2c+ξ5(τ2c+ˆτ1)α2ccos(α2cˆτ1)]sin(α2cτ2c)+[ξ5(τ2c+ˆτ1)α2csin(α2cˆτ1)−ξ2τ2cα22c]cos(α2cτ2c),B23=α42c−[ξ1α2csin(α2cˆτ1)+ξ2cos(α2cˆτ1)]α22c,B24=[−ξ0α32c−(ξ1α2ccos(α2cˆτ1)+ξ2sin(α2cˆτ1)]α22c. |
In this part, we shall study the direction of the Hopf bifurcation and the stability of bifurcating periodic solution of system (2.6) when τ1 is regarded as a parameter, τ2=ˆτ2 is a fixed value. Similarly, we can discuss other cases. The approach employed here is the normal form method and center manifold theorem introduced by Hassard et al.[52] and Guckenheimer et al.[53]. It follows from implicit function theorem [53] and the third equation of system (2.6) that E(t)=mm2y(t)q(py(t)−c)−mm1. Hence, system (2.6) can be transformed as follows
{.x(t)=x(t−τ1)(1−x(t−τ1))−x(t−τ1)y(t)1+ax(t−τ1)+by(t),.y(t)=βy(t−τ2)(r−y(t−τ2)x(t−τ2))−mm2y(t)mm2+α[q(py(t)−c)−mm1]. | (4.22) |
Some transformations associated with P∗(x∗,y∗) are given as follows:
u1(t)=x(t)−x∗,u2(t)=y(t)−y∗. |
Here, we define ˉτ1=μ+τ1c, then μ=0 is the Hopf bifurcation value of system (2.6).
By simplifying, system (2.6) can be transformed to the following functional differential equation that is in the Banach space of continuous functions mapping C=C([−˜τ,0],R2)(˜τ=max{τ1c,ˆτ2}), here τ1c,ˆτ2 are defined in (4.16) and (4.21), respectively,
{.u1(t)=a11u1(t−τ1c−μ)+a12u2(t)+a13u21(t−τ1c−μ)+a14u1(t−τ1c−μ)u2(t)+a15u22(t),.u2(t)=a21u2(t)+a22u2(t−ˆτ2)+a23u1(t−ˆτ2)+a24u21(t−ˆτ2)+a25u22(t)+a26u22(t−ˆτ2)+a27u1(t−ˆτ2)u2(t−ˆτ2), | (4.23) |
here,
a12=−x∗β(1−x∗)2(1+ax∗)y∗2,a13=2ay∗(1+by∗)(1+ax∗+by∗)3−2,a14=2by∗(by∗−1)−2ax∗−1(1+ax∗+by∗)3,a15=2b(1+ax∗+by∗)3,a21=mm2(mm2−αmm1−αqc)(mm2+αq(py∗−c)−αmm1)3,a23=βy∗2x∗2,a24=−2βy∗2x∗3,a25=−2αpqmm2(mm2−αmm1−αqc)(mm2+αq(py∗−c)−αmm1)3,a26=−2βx∗,a27=2βy∗x∗2,a11=1−2x∗−x∗(1−x∗)2y∗−b(1−x∗)2,a22=βr−2βy∗x∗. |
Based to Riesz representation theorem [54], there is a 2×2 matrix function η(θ,μ), here, θ∈[−˜τ,0] and
Lμ(ϕ)=∫0−˜τdη(θ,μ)ϕ(θ), | (4.24) |
where ϕ(θ)=(ϕ1(θ),ϕ2(θ)∈C([−˜τ,0],R2), and
η(θ,μ)=(0a120a21)δ(θ)+(00a23a22)δ(θ+ˆτ2)−(a11000)δ(θ+τ1c+μ), | (4.25) |
here, δ denotes the Dirac delta function.
Next, for ϕ∈C1([−˜τ,0],R2), we define the operator A(μ) as
A(μ)ϕ={dϕ(θ)dθ,θ∈[−˜τ,0),∫0−˜τdη(μ,s)ϕ(s),θ=0, |
R(μ)ϕ={0,θ∈[−˜τ,0),F(μ,ϕ),θ=0, |
where F(μ,ϕ)=(F1(μ,ϕ),F2(μ,ϕ))T and
{F1(μ,ϕ)=a13ϕ21(−τ1c−μ)+a14ϕ1(−τ1c)ϕ2(0)+a15ϕ22(0),F2(μ,ϕ)=a24ϕ21(−ˆτ2)+a25ϕ22(0)+a26ϕ22(−ˆτ2)+a27ϕ1(−ˆτ2)ϕ2(−ˆτ2), | (4.26) |
then system (4.23) is equivalent to
.ut=A(μ)ut+R(μ)ut. | (4.27) |
For ψ∈C1([−˜τ,0],(R2)∗), we suppose that
A∗ϕ(s)={−dψ(s)ds,s∈(0,˜τ],∫0−˜τdηT(t,0)ψ(−t),s=0, |
and a bilinear inner product
⟨ψ(s),ϕ(θ)⟩=ˉψ(0)ϕ(0)−∫0−˜τ∫θρ=0ˉψ(ρ−θ)dη(θ)ϕ(ρ)dρ, | (4.28) |
where η(θ)=η(θ,0).
From the above analysis, we know that A and A∗ are adjoint operators. According to discussion in Subsection 4.2, ±iα1c are eigenvalues of both A and A∗.
Suppose q(θ)=(1,χ)Teiα1cτ1cθ,θ is eigenvector of A corresponding to iα1c, which gives that Aq(θ)=iα1cq(θ). Based on definition of A, (4.24), (4.25) and (4.26), we have
∫0−˜τdη(θ)q(θ)=(0a120a21)q(0)+(a11000)q(−τ1c)=Aq(0)=iα1cq(0). |
Hence, we can obtain χ=iχ1c−a11e−iα1cτ1ca12.
Similarly, By simple computation, we can obtain eigenvector of A∗ corresponding to iα1c, i.e., q∗(s)=D(1,χ∗)Teiα1cτ1cs, here χ∗=iα1c−a11e−iα1cτ1ca23.
In order to assume ⟨q∗(s),q(θ)⟩=1, we can determine the value of D. Based on (4.28), we have
⟨q∗(s),q(θ)⟩=ˉD(1,¯χ∗)(1,χ)T−∫0−˜τ∫θρ=0ˉD(1,¯χ∗)e−iα1cτ1c(ρ−θ)dη(θ)q(ρ)dρ=ˉD(1+χ¯χ∗)−ˉD∫0−˜τ(1,¯χ∗)θeiα1cτ1cθdη(θ)(1,χ)T=ˉD(1+χ¯χ∗+a11τ1ce−iα1cτ1c). |
Hence, we can choose ˉD=11+χ¯χ∗+a11τ1ce−iα1cτ1c.
Next, let ut be the solution of Eq (4.27) when μ=0, and
z(t)=⟨q∗,ut⟩,W(t,θ)=ut(θ)−Re{z(t)q(θ)}. | (4.29) |
On the center manifold C0, it gives that
W(t,θ)=W(z(t),ˉz(t),θ)=W20(θ)z22+W11(θ)zˉz+W02(θ)ˉz22+⋅⋅⋅, | (4.30) |
z and ˜z are local coordinates for C0 in the direction of q∗ and ˉq∗.
For solution ut∈C0 of Eq (4.27), when μ=0, we have
.z(t)=iα1cz(t)+g(z,ˉz). |
Where,
g(z,ˉz)=g20z22+g11zˉz+g02ˉz22+g21z2ˉz2+⋅⋅⋅. | (4.31) |
It follows from Eq (4.29) and (4.30) that
ut(θ)=W20(θ)z22+W11(θ)zˉz+W02(θ)ˉz22+(1,χ)Teiα1cτ1cθz+(1,ˉχ∗)Teiα1cτ1cθˉz+⋅⋅⋅. | (4.32) |
By virtue of (4.30), (4.31) and (4.32), we have
g(z,ˉz)=τ1cˉD(1,¯χ∗)(ˆF1ˆF2)=τ1cˉD[a13+a14+a15+¯χ∗(a24+a25+a26+a27)][z2+2zˉz+ˉz2]+τ1cˉDa13[W(2)20(−τ1c)+2W(2)11(−τ1c)]+a15[W(2)20(0)+2W(2)11(0)]z2ˉz+τ1cˉDa14[W(1)11(−τ1c)+W(2)11(0)+W(1)20(−τ1c)+W(2)20(0)2]z2ˉz+τ1cˉD¯χ∗[a24(W(1)20(−ˆτ2)+2W(1)11(−ˆτ2))+a25(W(2)20(0)+2W(2)11(0))]z2ˉz+τ1cˉD¯χ∗a26[W(2)20(−ˆτ2)+2W(2)11(−ˆτ2)]z2ˉz+τ1cˉD¯χ∗a27[W(1)11(−ˆτ2)+W(2)11(−ˆτ2)+W(1)20(−ˆτ2)+W(2)20(−ˆτ2)2]z2ˉz+⋅⋅⋅, | (4.33) |
where
ˆF1=a13u21(−τ1c)+a14u1(τ1c)u2(0)+a15u22(0),ˆF2=a24u21(−ˆτ2)+a25u21(0)+a26u22(−ˆτ2)+a27u1(−ˆτ2)u2(−ˆτ2). |
According to comparing (4.31) with (4.33), we can obtain
g20=2τ1cˉD[a13+a14+a15+¯χ∗(a24+a25+a26+a27)],g11=τ1cˉD[a13+a14+a15+¯χ∗(a24+a25+a26+a27)],g02=2τ1cˉD[a13+a14+a15+¯χ∗(a24+a25+a26+a27)],g21=τ1cˉDa13[W(2)20(−τ1c)+2W(2)11(−τ1c)]+a15[W(2)20(0)+2W(2)11(0)]+τ1cˉDa14[W(1)11(−τ1c)+W(2)11(0)+W(1)20(−τ1c)+W(2)20(0)2]+τ1cˉD¯χ∗[a24(W(1)20(−ˆτ2)+2W(1)11(0ˆτ2))+a25(W(2)20(0)+2W(2)11(0))]+τ1cˉD¯χ∗a26[W(2)20(−ˆτ2)+2W(2)11(−ˆτ2)]+τ1cˉD¯χ∗a27[W(1)11(−ˆτ2)+W(2)11(−ˆτ2)+W(1)(−ˆτ2)20+W(2)20(−ˆτ2)2]+⋅⋅⋅. |
By virtue of (4.30) and (4.32), we have
.W=.ut−.zq−.ˉzˉq={AW−2Re{ˉq∗(0)F0q(θ)},θ∈[−1,0)AW−2Re{ˉq∗(0)F0q(0)},θ=0≜AW+H(z,ˉz,θ). | (4.34) |
Here,
H(z,ˉz,θ)=H20(θ)z22+H11(θ)zˉz+H02(θ)ˉz22+⋅⋅⋅. | (4.35) |
By comparing coefficient and computing, we can obtain
(A−2iα1cτ1c)W20(θ)=−H20(θ),AW11(θ)=−H11(θ),⋅⋅⋅. | (4.36) |
We know that
H(z,ˉz,θ)=−g(z,ˉz)q(θ)−ˉg(z,ˉz)ˉq(θ),θ∈[−˜τ,0). | (4.37) |
Based on (4.35) and (4.37), it derives that
H20(θ)=−g20q(θ)−ˉg02ˉq(θ), | (4.38) |
H11(θ)=−g11q(θ)−ˉg11ˉq(θ). | (4.39) |
Based on (4.36) and (4.38), we can obtain
.W20(θ)=2iα1cτ1cW20(θ)+g20q(θ)+ˉg02ˉq(θ). |
Based on q(θ)=(1,χ)Teiα1cτ1cθ, we have
W20(θ)=ig20α1cτ1cq(0)eiα1cτ1cθ+iˉg203α1cτ1cˉq(0)e−iα1cτ1cθ+E1e2iα1cτ1cθ, |
where E1=(E11,E21) is a constant vector.
Similarly, based on (4.36) and (4.39), we have
W11(θ)=−ig11α1cτ1cq(0)eiα1cτ1cθ+iˉg11α1cτ1cˉq(0)e−iα1cτ1cθ+E2, |
here, E2=(E12,E22) is a constant vector.
By above definitions and conditions, we have
E11=1G1|a13e−2iα1cτ1c+a14e−iα1cτ1c+a15−a12(a24+a26+a27)e−2iα1cˆτ2+a252iα1c−a21−a22|,E21=1G1|2iα1ca13e−2iα1cτ1c+a14e−iα1cτ1c+a15−a23(a24+a26+a27)e−2iα1cˆτ2+a25|,E12=1G2|2a13cos(2α1cτ1c)+2a14cos(α1cτ1c)+2a15a122(a24+a26+a27)cos(2α1cˆτ2)+2a25a21+a22|,E22=1G2|a112a13cos(2α1cτ1c)+2a14cos(α1cτ1c)+2a15a232(a24+a26+a27)cos(2α1cˆτ2)+2a25|, |
where G1=2iα1c(2iα1c−a21−a22)−a12a23, G2=a11(a21+a22)−a12a23.
By above analyses and the results of Kuang[54], the following results can be given:
{c1(0)=i2α1cτ1c(g20g11−2|g11|2−|g02|23)+g212,μ2=−Rec1(0)Reλ′(τ1c),ε2=2Rec1(0),T2=−Imc1(0)+ε2Imλ′(τ1c)α1cτ1c. | (4.40) |
By computing Eq (4.40), we have the following theorem.
Theorem 4.9. In Eq (4.40), the following results are true.
(i) The sign of μ2 determines the direction of the Hopf bifurcation: if μ2>0(μ2<0), then Hopf bifurcation is supercritical (subcritical) and the bifurcating periodic solutions exist for τ>τ1c(τ<τ1c).
(ii) The sign of ε2 determines the stability of the bifurcating periodic solution: the bifurcating periodic solutions are stable (unstable) if ε2<0(ε2>0).
(iii) The sign of T2 determines the period of the bifurcating periodic solutions: the period increases (decreases) if T2>0(T2<0).
In this section, some numerical simulation are presented for supporting the analytic results obtained.
First, in order to verify some results of singularity induced bifurcations, the values of some parameters of system (2.6) are given: a=1,b=2,c=0.01,m1=0.8,α=2,β=0.1,m2=1,p=4, q=0.5,r=0.15. When τ1=τ2=0, singularity induced bifurcation occurs as m=0 (see Figure 1) and m=0.18 (see Figure 2) respectively. According to Theorem 4.3, we obtain that the feedback gain satisfies ˆk=1>0.19. Thus state feedback controller u(t)=(E(t)−0.008) is designed to stabilize system (4.1) as m=0 (see Figure 1.(b)). According to Theorem 4.4, state feedback controller ˜u(t)=(E(t)−0.01) is designed to stabilize system (4.1) as m=0.18 (see Figure 2.(b)).
Next, in order to show the phenomenon of Hopf bifurcation, the values of some parameters of system (2.6) are given: a=2,b=1.02,c=1.86,m1=0.3,α=0.5,β=0.5,m2=1,p=15.2,q=0.05,r=2.85. When m=0.9>0, if (H4) or (H5) holds, then x∗=0.745 is obtained by solving Eq (3.2), that is, the interior equilibrium is P∗=(0.745,0.835,2.760).
Based on Theorem 4.5, by computing ξ24−(ξ2+ξ5)2=−0.609<0, we have that system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω2={(τ1,τ2,m)|0<τ1=1.5<τ10=2.4,τ2=0,0<m<22.4}. However, system (2.6) undergoes Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω3={(τ1,τ2,m)|τ1=2.5>τ10=2.4,τ2=0,0<m<22.4} (see Figures 3 and 4).
Similarly, system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩ Ω4={(τ1,τ2,m)|τ1=0,τ2=1.6<τ20=1.65,0<m<22.4}. However, system (2.6) undergoes Hopf bifurcation around P∗ when (τ1,τ2,m)∈Ω1∩Ω5={(τ1,τ2,m)|τ1=0, τ2=1.8>τ20=1.65, 0<m<22.4} (see Figure 5).
When we consider τ1 as a parameter and τ2=2.3 as a fixed value based on Theorem 4.7, we know that Eq (4.15) has finite positive roots, and obtain the critical value of delay is τ1c=2.12. Therefore, system (2.6) is locally asymptotically stable around P∗ when (τ1,τ2,m)∈Ω1∩Ω6∩Ω8={(τ1,τ2,m)|0<τ1=2<τ1c=2.12,τ2=ˆτ2=2.3, 0<m<22.4} (see Figure 6). However, when (τ1,τ2,m)∈Ω1∩Ω6∩Ω9={(τ1, τ2,m)|τ1=2.2>τ1c=2.12,τ2=ˆτ2=2.3,0<m<22.4}, system (2.6) undergoes Hopf bifurcation around P∗ (see Figure 7).
Similarly, we regard τ2 as a parameter and τ1=2 as a fixed value. When (τ1,τ2,m)∈Ω1∩Ω10∩Ω12={(τ1,τ2,m)|τ1=ˆτ1=2,0<τ2=6<τ2c,0<m<22.4}, system (2.6) is locally asymptotically stable around P∗ (see Figure 8). However, when (τ1,τ2,m)∈Ω1∩Ω10∩Ω13={(τ1,τ2,m)|τ1=ˆτ1=2,τ2=90>τ2c,0<m<22.4}, system (2.6) undergoes Hopf bifurcation (see Figure 9).
Based on Theorem 4.9, we can obtain μ2=−0.06<0,ε2=−30.57<0 and T2=−146020<0. Thus, Hopf bifurcation is subcritical and the bifurcating periodic solutions exist for τ1<τ1c, bifurcating periodic solutions are stable and the period decreases because of ε2<0,T2<0.
It is well known that commercial harvesting and economic benefits have a strong impact on the dynamical behavior. Liu et al. [44] investigated a differential-algebraic prey-predator system with linear harvesting on predator and Holling-II. Li et al. [45] analyzed a differential-algebraic prey-predator system without time delay. However, nonlinear harvesting and Beddington-DeAngelis functional response more realistic. Therefore, this paper proposed a singular Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting by extending the work of references [44] and [45], and obtained some results. The existence of equilibria was analyzed. Without considering time delay, the existence of singularity induced bifurcation by regarding economic interest as bifurcation parameter was discussed. In order to remove singularity induced bifurcation and stabilize system (4.1), state feedback controllers u(t)=(E(t)−0.008) (see Figure 1b) and ˜u(t)=(E(t)−0.01) (see Figure 2b) were designed, which shows that the system can be kept in a stable state with benefits by capturing predators. While considering time delay, stability of system were discussed by analyzing the corresponding characteristic transcendental equation. When τ2=0, the critical value of time delay is τ10=2.4; when τ1=0, the critical value of time delay is τ20=1.65; when τ1 is regarded as a parameter and τ2 as a fixed value, the critical value of time delay is τ1c=2.12. At the same time, when τ2 is regarded as a parameter and τ1 as a fixed value, the critical value of time delay is τ2c=81. It was obvious that system lost local stability around it's corresponding interior equilibrium when time delays crossed corresponding the critical values, and Hopf bifurcations occurred (see Figures 4, 5b, 7, and 9). Finally, by using normal form theory and center manifold theorem, Hopf bifurcation is subcritical and the bifurcating periodic solutions exist for τ1<τ1c, bifurcating periodic solutions are unstable and the period decreases because of ε2<0,T2<0, which could be found in Theorem 4.9.
In fact, the prey and predator may be captured simultaneously in real world. Thus, in order to make this model more practical, nonlinear predator harvesting and nonlinear prey harvesting can be introduced into our predator-prey system, which is
{.x(t)=αx(t−τ1)(1−x(t−τ1)k)−x(t−τ1)y(t)1+ax(t−τ1)+by(t)−q1E(t)x(t)m1E(t)+m2x(t),.y(t)=βy(t−τ2)(1−y(t−τ2)x(t−τ2))−q2E(t)y(t)m3E(t)+m4y(t),0=q1E(t)m1E(t)+m2y(t)(px(t)−c)+q2E(t)m3E(t)+m4y(t)(p1y(t)−c1)−m. |
We leave this work in the future.
The authors are grateful to the anonymous reviewers for their constructive comments. This work is supported by the National Natural Science Foundation of China (Grant Nos. 11661050 and 11861044), and the HongLiu first-class disciplines Development Program of Lanzhou University of Technology.
All authors declare no conflicts of interest in this paper.
[1] | M. Kot, Elements of Mathematical Biology, Cambridge University Press, Cambridge, 2001. |
[2] | S. Levin, T. Hallam and J. Cross, Applied Mathematical Ecology, Springer, New York, 1990. |
[3] | R. May, Stability and Complexity in Model Ecosystems, Princeton University Press, Princeton, 1993. |
[4] | C. Ji, D. Jing and N. Shi, A note on a predator-prey model with modified Leslie-Gower and Holling-type II schemes with stochastic perturbatio, J. Math. Anal. Appl., 377 (2011), 435–440. |
[5] | T. Kar and H. Matsuda, Global dynamics and controllability of a harvested prey-predator system with Holling type III functional response, Nonlinear Anal. Hybrid Syst., 1 (2007), 59–67. |
[6] | X. Meng, J. Wang, H. Huo, Dynamical behaviour of a nutrient-plankton model with Holling type IV, delay, and harvesting, Discrete Dyn. Nat. Soc., 2018 (2018), Article ID 9232590. |
[7] | X. Meng, H. Huo, H. Xiang, et al., Stability in a predator-prey model with Crowley-Martin function and stage structure for prey, Appl. Math. Comput., 232 (2014), 810–819. |
[8] | W. Yang, Diffusion has no influence on the global asymptotical stability of the Lotka-Volterra pre-predator model incorporating a constant number of prey refuges, Appl. Math. Comput., 223 (2013), 278–280. |
[9] | Y. Zhu and K.Wang, Existence and global attractivity of positive periodic solutions for a predatorprey model with modified Leslie-Gower Holling-type II schemes, J. Math. Anal. Appl., 384 (2011), 400–408. |
[10] | J. Liu, Dynamical analysis of a delayed predator-prey system with modified Leslie-Gower and Beddington-DeAngelis functional response, Adv. Difference Equ., 2014 (2014), 314–343. |
[11] | X. Liu and Y. Wei, Dynamics of a stochastic cooperative predator-prey system with Beddington- DeAngelis functional response, Adv. Difference Equ., 2016 (2016), 21–39. |
[12] | C. Li, X. Guo and D. He, An impulsive diffusion predator-prey system in three-species with Beddington-DeAngelis response, J. Appl. Math. Comput., 43 (2013), 235–248. |
[13] | T. Ivanov and N. Dimitrova, A predator-prey model with generic birth and death rates for the predator and Beddington-DeAngelis functional response, Math. Comput. Simulat., 133 (2017), 111–123. |
[14] | Q. Meng and L. Yang, Steady state in a cross-diffusion predator-prey model with the Beddington- DeAngelis functional response, Nonlinear Anal.: Real World Appl., 45 (2019), 401–413. |
[15] | W. Liu, C. Fu and B. Chen, Hopf bifurcation and center stability for a predator-prey biological economic model with prey harvesting, Commun. Nonlinear Sci. Numer. Simul., 17 (2012), 3989– 3998. |
[16] | F. Conforto, L. Desvillettes and C. Soresina, About reaction-diffusion systems involving the Holling-type II and the Beddington-DeAngelis functional responses for predator-prey models, Nonlinear Differ. Equ. Appl., 25 (2018), 24. |
[17] | X. Sun, R. Yuan and L. Wang, Bifurcations in a diffusive predator-prey model with Beddington- DeAngelis functional response and nonselective harvesting, J. Nonlinear Sci., 29 (2019), 287–318. |
[18] | J. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 44 (1975), 331–340. |
[19] | D. DeAngilis, R. Goldstein and R. Neill, A model for tropic interaction, Ecology, 56 (1975), 881–892. |
[20] | H. Freedman, Deterministic Mathematical Models in Population Ecology, Marcel Dekker, New York, 1980. |
[21] | G. Seifert, Asymptotical behavior in a three-component food chain model, Nonlinear Anal. Theory Methods Appl., 32 (1998), 749–753. |
[22] | C. Liu, Q. Zhang, X. Zhang, et al., Dynamical behavior in a stage-structured differential-algebraic prey-predator model with discrete time delay and harvesting, J. Comput. Appl. Math., 231 (2009), 612–625. |
[23] | X. Meng, H. Huo, X. Zhang, et al., Stability and hopf bifurcation in a three-species system with feedback delays, Nonlinear Dyn., 64 (2011), 349–364. |
[24] | X. Meng and Y. Wu, Bifurcation and control in a singular phytoplankton-zooplankton-fish model with nonlinear fish harvesting and taxation, Int. J. Bifurcat. Chaos, 28 (2018), 1850042(24 pages). |
[25] | H. Xiang, Y. Wang and H. Huo, Analysis of the binge drinking models with demographics and nonlinear infectivity on networks, J. Appl. Anal. Comput., 8 (2018), 1535–1554. |
[26] | K. Chakraborty, M. Chakraboty and T. Kar, Bifurcation and control of a bioeconomic model of a prey-predator system with a time delay, Nonlinear Anal. Hybrid Syst., 5 (2011), 613–625. |
[27] | W. Liu, C. Fu and B. Chen, Hopf birfucation for a predator-prey biological economic system with Holling type II functional response, J. Franklin Inst., 348 (2011), 1114–1127. |
[28] | G. Zhang, B. Chen, L. Zhu, et al., Hopf bifurcation for a differential-algebraic biological economic system with time delay, Appl. Math. Comput, 218 (2012), 7717–7726. |
[29] | T. Faria, Stability and bifurcation for a delayed predator-prey model and the effect of diffusion, J. Math. Anal. Appl., 254 (2001), 433–463. |
[30] | G. Zhang, Y. She and B. Chen, Hopf bifurcation of a predator prey system with predator harvesting and two delays, Nonlinear Dyn., 73 (2013), 2119–2131. |
[31] | J. Zhang, Z. Jin, J. Yan, et al., Stability and Hopf bifurcation in a delayed competition system, Nonlinear Anal. Theory Methods Appl., 70 (2009), 658–670. |
[32] | Z. Lajmiri, R. K. Ghaziani and I. Orak, Bifurcation and stability analysis of a ratio-dependent predator-prey model with predator harvesting rate, Chaos Soliton Fract., 106 (2018), 193–200. |
[33] | M. Liu, X. He and J. Yu, Dynamics of a stochastic regime-switching predator-prey model with harvesting and distributed delays, Nonlinear Anal. Hybrid Syst., 28 (2018), 87–104. |
[34] | T. Das, R. Mukerjee and K. Chaudhuri, Harvesting of a prey-predator fishery in the presence of toxicity, Appl. Math. Model., 33 (2009), 2282–2292. |
[35] | R. Gupta and P. Chandra, Bifurcation analysis of modied Leslie-Gower predator-prey model with Michaelis-Menten type prey harvesting, J. Math. Anal. Appl., 398 (2013), 278–295. |
[36] | P. Srinivasu, Bioeconomics of a renewable resource in presence of a predator, Nonlinear Anal. Real World Appl., 2 (2001), 497–506. |
[37] | G. Lan, Y. Fu, C. Wei, et al., Dynamical analysis of a ratio-dependent predator-prey model with Holling III type functional response and nonlinear harvesting in a random environment, Adv. Differ. Equ., 2018 (2018), 198. |
[38] | R. Gupta and P. Chandra, Dynamical complexity of a prey-predator model with nonlinear predator harvesting, Discrete Contin. Dyn. Syst. Ser. B, 20 (2015), 423–443. |
[39] | J. Liu and L. Zhang, Bifurcation analysis in a prey-predator model with nonlinear predator harvesting, J. Franklin Inst., 353 (2016), 4701–4714. |
[40] | K. Chakraborty, S. Jana and T. Kar, Global dynamics and bifurcation in a stage structured preypredator fishery model with harvesting, Appl. Math. Comput., 218 (2012), 9271–9290. |
[41] | H. Gordon, The economic theory of a common property resource: the fishery, Bull. Math. Biol., 62 (1954), 124–142. |
[42] | C. Liu, Q. Zhang and X. Duan, Dynamical behavior in a harvested differential-algebraic preypredator model with discrete time delay and stage structure, J. Franklin Inst., 346 (2009), 1038– 1059. |
[43] | X. Zhang and Q. Zhang, Bifurcation analysis and control of a class of hybrid biological economic models, Nonlinear Anal. Hybrid Syst., 3 (2009), 578–587. |
[44] | C. Liu, N. Lu, Q. Zhang, et al., Modelling and analysis in a prey-predator system with commercial harvesting and double time delays, J. Appl. Math. Comput., 281 (2016), 77–101. |
[45] | M. Li, B. Chen and H. Ye, A bioeconomic differential algebraic predator-prey model with nonlinear prey harvesting, Appl. Math. Model., 42 (2017), 17–28. |
[46] | P. Leslie and J. Gower, The properties of a stochastic model for the predator prey type of interaction between two species, Biometrika, 47 (1960), 219–234. |
[47] | R. Cantrell and C. Cosner, On the dynamics of predator-prey models with the Beddington- DeAngelis functional response, J. Math. Anal. Appl., 257 (2001), 206–222. |
[48] | L. Dai, Singular Control System, Springer, New York, 1989. |
[49] | V. Venkatasubramanian, H. Schattler and J. Zaborszky, Local bifurcations and feasibility regions in differential-algebraic systems, IEEE Trans. Automat. Control, 40 (1995), 1992–2013. |
[50] | Q. Zhang, C. Liu and X. Zhang, A singular bioeconomic model with diffusion and time delay, J. Syst. Sci. Complex., 24 (2011), 277–190. |
[51] | J. Hale, Theory of Functional Differential Equations, Springer, New York, 1997. |
[52] | B. Hassard, N. Kazarinoff and Y. Wan, Theory and Applications of Hopf Bifurcation, Cambridge University Press, Cambridge, 1981. |
[53] | J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, Springer, New York, 1983. |
[54] | Y. Kuang, Delay Differential Equations with Applications in Population Dynamics, Academic Press, New York, 1993. |
[55] | H. Freedman and V. S. H. Rao, The trade-off between mutual interference and time lags in predator-prey systems, Bull. Math. Biol., 45 (1983), 991–1004. |
1. | Xin-You Meng, Yu-Qian Wu, Dynamical analysis of a fuzzy phytoplankton–zooplankton model with refuge, fishery protection and harvesting, 2020, 63, 1598-5865, 361, 10.1007/s12190-020-01321-y | |
2. | S Toaha, , Stability analysis of prey predator model with Holling II functional response and threshold harvesting for the predator, 2019, 1341, 1742-6588, 062025, 10.1088/1742-6596/1341/6/062025 | |
3. | Xin-You Meng, Jiao-Guo Wang, Dynamical analysis of a delayed diffusive predator–prey model with schooling behaviour and Allee effect, 2020, 14, 1751-3758, 826, 10.1080/17513758.2020.1850892 | |
4. | Yue Song, Qingjie Liu, Yi Zhang, 2022, Variable Structure Control of Singular Biological Economic System with Restrictive Human Behavior, 978-988-75815-3-6, 918, 10.23919/CCC55666.2022.9902557 | |
5. | Vahid Shojaee, Masoud Shafiee, Asier Ibeas, State feedback H ∞ control for a class of affine nonlinear singular systems: Input restricting approach , 2022, 16, 1751-8644, 166, 10.1049/cth2.12213 | |
6. | Lu Lu, Chengdai Huang, Xinyu Song, Bifurcation control of a fractional-order PD control strategy for a delayed fractional-order prey–predator system, 2023, 138, 2190-5444, 10.1140/epjp/s13360-023-03708-9 | |
7. | Feilong Wang, Min Xiao, Zhengxin Wang, Jing Zhao, Gong Chen, Jinde Cao, Sergey Dashkovskiy, Spatiotemporal Evolution Characteristics of Time-Delay Ecological Competition Systems with Food-Limited and Diffusion, 2022, 2022, 1099-0526, 1, 10.1155/2022/2823303 | |
8. | Jialu Tian, Ping Liu, Global Bifurcation in a Modified Leslie–Gower Predator–Prey Model, 2023, 33, 0218-1274, 10.1142/S0218127423500165 | |
9. | Xin-You Meng, Jie Li, Dynamical behavior of a delayed prey-predator-scavenger system with fear effect and linear harvesting, 2021, 14, 1793-5245, 2150024, 10.1142/S1793524521500248 | |
10. | Juan Ye, Yi Wang, Zhan Jin, Chuanjun Dai, Min Zhao, Dynamics of a predator-prey model with strong Allee effect and nonconstant mortality rate, 2022, 19, 1551-0018, 3402, 10.3934/mbe.2022157 | |
11. | Ahmed M. Yousef, Saad Z. Rida, Soheir Arafat, Sophia R.‐J. Jang, Stability, bifurcation analysis, and chaos control of a discrete bioeconomic model, 2023, 46, 0170-4214, 3204, 10.1002/mma.8686 | |
12. | Qamar Din, A. M. Yousef, A. A. Elsadany, Douglas Anderson, Stability and Bifurcation Analysis of a Discrete Singular Bioeconomic System, 2021, 2021, 1607-887X, 1, 10.1155/2021/6679161 | |
13. | Prahlad Majumdar, Uttam Ghosh, Susmita Sarkar, Surajit Debnath, Study of co-dimension two bifurcation of a prey–predator model with prey refuge and non-linear harvesting on both species, 2023, 0009-725X, 10.1007/s12215-023-00881-9 | |
14. | U. Yadav, A. K. Nayak, S. Gakkhar, Mathematical Scrutiny of Singular Predator-Prey Model with Stage-Structure of Prey, 2024, 189, 0167-8019, 10.1007/s10440-023-00630-1 |