
In the paper, stability and bifurcation behaviors of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response are studied theoretically and numerically. Mathematical theory works mainly give some critical threshold conditions to guarantee the existence and stability of all possible equilibrium points, and the occurrence of Hopf bifurcation and Bogdanov-Takens bifurcation. Numerical simulation works mainly display that the Bazykin's predator-prey ecosystem has complex dynamic behaviors, which also directly proves that the theoretical results are effective and feasible. Furthermore, it is easy to see from numerical simulation results that some key parameters can seriously affect the dynamic behavior evolution process of the Bazykin's predator-prey ecosystem. Moreover, limit cycle is proposed in view of the supercritical Hopf bifurcation. Finally, it is expected that these results will contribute to the dynamical behaviors of predator-prey ecosystem.
Citation: Shuangte Wang, Hengguo Yu. Stability and bifurcation analysis of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response[J]. Mathematical Biosciences and Engineering, 2021, 18(6): 7877-7918. doi: 10.3934/mbe.2021391
[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] | Saheb Pal, Nikhil Pal, Sudip Samanta, Joydev Chattopadhyay . Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model. Mathematical Biosciences and Engineering, 2019, 16(5): 5146-5179. doi: 10.3934/mbe.2019258 |
[3] | Mengyun Xing, Mengxin He, Zhong Li . Dynamics of a modified Leslie-Gower predator-prey model with double Allee effects. Mathematical Biosciences and Engineering, 2024, 21(1): 792-831. doi: 10.3934/mbe.2024034 |
[4] | Kunlun Huang, Xintian Jia, Cuiping Li . Analysis of modified Holling-Tanner model with strong Allee effect. Mathematical Biosciences and Engineering, 2023, 20(8): 15524-15543. doi: 10.3934/mbe.2023693 |
[5] | Manoj K. Singh, Brajesh K. Singh, Poonam, Carlo Cattani . Under nonlinear prey-harvesting, effect of strong Allee effect on the dynamics of a modified Leslie-Gower predator-prey model. Mathematical Biosciences and Engineering, 2023, 20(6): 9625-9644. doi: 10.3934/mbe.2023422 |
[6] | Sangeeta Kumari, Sidharth Menon, Abhirami K . Dynamical system of quokka population depicting Fennecaphobia by Vulpes vulpes. Mathematical Biosciences and Engineering, 2025, 22(6): 1342-1363. doi: 10.3934/mbe.2025050 |
[7] | Eric M. Takyi, Charles Ohanian, Margaret Cathcart, Nihal Kumar . Dynamical analysis of a predator-prey system with prey vigilance and hunting cooperation in predators. Mathematical Biosciences and Engineering, 2024, 21(2): 2768-2786. doi: 10.3934/mbe.2024123 |
[8] | Eduardo González-Olivares, Claudio Arancibia-Ibarra, Alejandro Rojas-Palma, Betsabé González-Yañez . Bifurcations and multistability on the May-Holling-Tanner predation model considering alternative food for the predators. Mathematical Biosciences and Engineering, 2019, 16(5): 4274-4298. doi: 10.3934/mbe.2019213 |
[9] | Yuhong Huo, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Renji Han . Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor. Mathematical Biosciences and Engineering, 2023, 20(10): 18820-18860. doi: 10.3934/mbe.2023834 |
[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. Mathematical Biosciences and Engineering, 2022, 19(4): 3402-3426. doi: 10.3934/mbe.2022157 |
In the paper, stability and bifurcation behaviors of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response are studied theoretically and numerically. Mathematical theory works mainly give some critical threshold conditions to guarantee the existence and stability of all possible equilibrium points, and the occurrence of Hopf bifurcation and Bogdanov-Takens bifurcation. Numerical simulation works mainly display that the Bazykin's predator-prey ecosystem has complex dynamic behaviors, which also directly proves that the theoretical results are effective and feasible. Furthermore, it is easy to see from numerical simulation results that some key parameters can seriously affect the dynamic behavior evolution process of the Bazykin's predator-prey ecosystem. Moreover, limit cycle is proposed in view of the supercritical Hopf bifurcation. Finally, it is expected that these results will contribute to the dynamical behaviors of predator-prey ecosystem.
As we all know, the classic Lotka-Volterra predator-prey system has been used to simulate predatory phenomena in nature, and its impact on the researches of mathematical biology and ecology will be roughly equivalent to the atomic bomb effect. In 1965, Holling proposed several different functional responses to characterize the dynamic predator-prey relationship between populations, which could describe the predator population how to transform the captured prey population into its own growth ability. Generally speaking, functional responses in predator-prey systems mainly depend on many internal and external key factors, such as the densities of prey and predator. On the other hand, prey-dependent functional responses are an important role in mathematical ecology, especially the dynamics of predator-prey systems. In recent decades, many scholars have done a lot of research on the predator-prey function response, and have made some excellent research results. Now, it is briefly summarized to enrich the population dynamics modeling system.
(ⅰ) The Holling type Ⅰ functional response [1] is
Φ(x)={cxx⩽x0,cx0x>x0, |
where c is a positive constant.
(ⅱ) The Holling type Ⅱ or the Michaelis-Menten functional response [1,2,3,4,5] is
Φ(x)=αxa+xorΦ(x)=αx1+wx, |
where α, a and w are positive constants.
(ⅲ) The Holling type Ⅲ or the p=2 S-type functional response [1,4,5] is
Φ(x)=αx2β2+x2, |
where α and β are positive constants. The generalized Holling type Ⅲ or sigmoidal is Φ(x)=mx2ax2+bx+1 with b<−2√a, where m and a are positive constants, b is a constant. When b=0, the function Φ(x) can reduce to above Holling type Ⅲ functional response. The S-type functional response with index p is Φ(x)=αxpβ+xp, where α, β and p are positive constants.
(ⅳ) The generalized Holling Ⅳ or the Monod-Haldane functional response [1,4,5,6,7] is
Φ(x)=mxa+bx+x2orΦ(x)=mxax2+bx+1, |
where m and a are all positive constants, but b is a constant. When b=0, it is called the Holling type Ⅳ functional response or the simplified Holling type Ⅳ functional response [8,9].
(ⅴ) The Beddington-DeAngelis type functional response [4,10,11,12,13] is
Φ(x,y)=αxa+bx+cy, |
where α, a, b and c are positive constants, which is originally and independently introduced by Beddington and DeAngelis [10,11]. At the same time, it is similar to the Holling type Ⅱ functional response incorporating an extra term cy in denominator, which can describe mutual interference among predators [12,13]. This functional response has some same qualitative features as the ratio-dependent form, but can avoid some singular behaviors of ratio-dependent models at low densities [12].
(ⅵ) The Hassell-Varley type functional response [14,15,16,17,18] is
Φ(x,y)=Axx+myγ,γ∈(0,1), |
where A and m are positive constants, γ is the Hassell-Varley constant. When γ=0 or γ=1, it can be viewed as limiting cases mathematically. Specially, when γ=0, it is the Holling type Ⅱ functional response regardless of constants.
(ⅶ) The Crowley-Martin type functional response [19] is
Φ(x,y)=αxa+bx+cy+exy, |
where α, a, b, c and e are positive constants. It involves the interference among individual predator engaged in searching and handling the preys. While in [20,21,22], the authors particularly take the Crowley-Martin functional response in the new type of Φ(x,y)=αx(1+ax)(1+by), which is proposed by Bazykin and is an immense breakthrough of the Holling type Ⅱ and Beddington-DeAngelis functional responses.
In this paper, we will continually concentrate on a detailed discussion in the well-known Bazykin's predator-prey ecosystem with the Holling type Ⅱ functional response and interspecific density-restricted effect on the predator, which is also a variation of Volterra's classical predator-prey model and is expressed in the form of following ordinary differential equations (ODEs) [23,24]:
˙x=r1x(1−xK1)−αxya+x−m1x, | (1.1a) |
˙y=αexya+x−m2y−dy2. | (1.1b) |
Here the functions x=x(t) and y=y(t) represent densities of the prey population and predator population at time t, respectively. All positive parameters have practically biological meanings: r1 denotes the intrinsic growth rate of the prey population, K1 represents the carrying capacity of the environment, a is the half-saturation constant; α is the search efficiency of predator for prey, m1 and m2 are the mortality rate of the prey and predator population, e is the biomass conversion, d is the intra-specific competition coefficient. The specific growth term r1x(1−xK1) governs the increase of the prey in the lack of predator. The square non-linear term term dy2, denotes intrinsic decrease of the predator, and represents interspecific density-restricted effect on the predator. Excluding the dy2, the system (1.1) is based on the classical Gause type predator-prey system, which expresses the following form [25]:
˙x=xg(x,k)−yp(x), | (1.2a) |
˙y=y(−d+cq(x)), | (1.2b) |
where g(x,k) is a continues function for x>0, p(x) is a functional response of predators, which is called Holling-type-Ⅱ predator-prey model as well. Such ordinary differential system of predator-prey populations is familiar to the Lotka-Volterra system, in which populations have the addition of damping terms(or self inhabit). The positivity of solutions with respect to initial condition x(0)>0 and y(0)>0 is easy to prove and we omit its proof. This system with positive and bounded solutions is also well behaved as we intuits from the biological significance.
Bazykin [24] wholly discussed the stability of equilibria, global existence of limit cycles, global attractivity of such equilibria, Hopf and codimension 2 bifurcations. In [26], analytical description and alteration of local stability were given. Here the authors investigated a familiar Lotka-Volterra system, in which the populations have self-inhibit(the addition of damping term) for global stability and existence of limit cycles [27]:
˙x=x(1−k1x−k2x2)−xy1+ax, | (1.3a) |
˙y=rxy1+ax−y(δ0+δ1y), | (1.3b) |
where specific growth rate governs the growth of the prey in the absence of predator and it has an increase (or decrease) intrinsic rate on the predator. They already proved the existence of two limit cycles with the help of idea from the Poincare-Bendixson theory. Obviously, at special case k2=0, we note that the system (1.3) can reduce to above system (1.1), which was analyzed by [28] and [29] for stability of equilibria, Hopf bifurcations, global attraction and codimension two bifurcations as well. While in the respect of global behavior, the system (1.1) was investigated by [30]. In [31], for a particular form of the system (1.1) with a modified Holling type Ⅱ functional response β(x−m)1+α(x−m) incorporating a constant prey refuge m, the authors therein gave sufficient conditions to guarantee the global stability of the positive equilibrium and uniqueness of a stable limit cycle. In [32], the authors revealed a rich bifurcation structure, including supercritical and subcritical Bogdanov-Takens bifurcation.
Based on classical biological manipulation theory, the Bazykin's predator-prey ecosystem can be used to explore the dynamic relationship between Microcystis aeruginosa and filter-feeding fish from the perspective of population dynamics, where x(t) and y(t) represent respectively the density of Microcystis aeruginosa and filter-feeding fish (bighead carp and silver carp), the growth kinetics function of Microcystis aeruginosa x is r1x(1−xK1) with intrinsic growth rate r1 and maximum environmental capacity K1. The grazing function of filter-feeding fish y is αxy1+ax with capture coefficient α and density restriction coefficient a. Furthermore, the parameter m1 and m2 are natural mortality of Microcystis aeruginosa and filter-feeding fish, the parameter d and e are internal competition coefficient and energy conversion rate of filter feeding fish. In order to deeply explore the dynamic relationship between Microcystis aeruginosa and filter-feeding fish, it is necessary to investigate some bifurcation dynamic behaviors of the Bazykin's predator-prey ecosystem, thus we mainly focus on the stability and bifurcation of the Bazykin's predator-prey ecosystem in this paper. Firstly, we investigate the existence and stability of hyperbolic equilibrium point and non-hyperbolic equilibrium point, and conveniently study the cusp of condimension 3. Secondly, we explore Hopf bifurcation and Bogdanov-Takens bifurcation in detail, and give some sufficient threshold conditions. Finally, for Hopf bifurcation dynamics, we especially analyze the limit cycle via a perturbation procedure and canonical transformation. Moreover, we think that these mathematical analysis results can provide a theoretical basis for numerical simulation, which can give some biological interpretation for Hopf bifurcation and Bogdanov-Takens bifurcation, hence, we mainly study stability and bifurcation dynamics of the Bazykin's predator-prey ecosystem from the perspective of mathematical analysis, other biological significance issues will be completed in the follow-up work.
All solutions of the system (1.1) are non-negative and bounded with initial conditions x(0)⩾0, y(0)⩾0, thus it is namely dissipative in the first quadrant R+2 of 2 dimensional space R2 and well-defined on the closed domain R2+=¯R+2. Furthermore, the system is uniformly bounded with lim supt→+∞x(t)⩽M1, lim supt→+∞y(t)⩽M2, in which two positive constants M1 and M2 only depend on parameters r1, K1, α, a, m1, e, m2 and d [33]. In other words, the system (1.1) is confined in the domain
{(x,y)|0⩽z⩽M+ϵ,foranyϵ>0}, | (2.1) |
with z=ex+y and a constant M>0. Moreover, the system (1.1) is permanent if the value of all parameters can satisfy
ω1=K1r1(r1−m1−αM2a)>0,αe(1−λ)ω1a+e(1−λ)ω1−m2>0, | (2.2) |
with λ∈(0,1) [34].
It is clear that the system (1.1) admits two biological boundary equilibria E0:=(0,0) and E2:=(x2,0)(if r1>m1), where x2=K1(1−m1r1). For the Jacobian matrix J(E0)=diag{r1−m1,−m2} at point E0, it shows that E0 is a hyperbolic asymptotically stable node(unstable node) when r1<m1(r1>m1), while it is a stable node(non-hyperbolic attractor) with only one zero eigenvalue if r1=m1 (see the Theorem 7.1 in [35]). The Jacobian matrix at point E2 is
J(E2)=[m1−r1−αx2a+x20αex2a+x2−m2], | (2.3) |
then, the hyperbolic point E2 is an asymptotically stable node(unstable node) when αex2a+x2<m2(αex2a+x2>m2). When αex2a+x2=m2, E2 is also a stable node(non-hyperbolic attractor) with characteristic direction tan(θ)=(m1−r1)em2 under the polar coordinate transformation.
Now, we use an example to verify the stability of the equilibrium point E0 and E2 with r1=0.6, a=1.5, α=0.5, e=0.6, K1=20 and d=0.1. It is easy to find from Figure 1(a) that E0 is a stable node with characteristic direction θ=0 when r1=m1 and m2=0.06. Furthermore, it is obvious to see from Figure 1(b) that E2 is a stable node with characteristic direction tan(θ)=(m1−r1)em2 when αex2a+x2=m2 and m1=0.3.
At first, an interior equilibrium E∗=(x∗,y∗) of the system (1.1) always satisfies following algebraic equations
r1(1−xK1)−αya+x−m1=0, | (2.4a) |
αexa+x−m2−dy=0. | (2.4b) |
Furthermore, from Sylvester's resultants in polynomial Equation (2.4), components x∗ and y∗ must be positive roots of third-order polynomial equations(cubic equations) p(x)=3∑i=0aixi=0 and q(y)=3∑i=0biyi=0, respectively. Here the coefficients are listed as follows:
a3=dr1,a2=d(m1−r1)K1+2adr1,a1=[2a(m1−r1)d+α(αe−m2)]K1+a2dr1,a0=[a(m1−r1)d−m2α]aK1,b3=K1d2,b2=−2K1d(αe−m2),b1=[(αe−m2)2+ade(r1−m1)]K1+a2der1,b0=e[(m1−r1)(αe−m2)K1+am2r1]a. |
Then, we define these complicated expressions
px=a1a3−13(a2a3)2,qx=227(a2a3)3−13a2a3a1a3+a0a3,Δx=(qx2)2+(px3)3 |
and
py=b1b3−13(b2b3)2,qy=227(b2b3)3−13b2b3b1b3+b0b3,Δy=(qy2)2+(py3)3 |
which is a discriminant of above cubic equations p(x)=0 and q(y)=0 for later use respectively [26]. The Eq (2.4) also implies that such interior equilibrium E∗ does not exist when one of conditions holds: (ⅰ) r1⩽m1; (ⅱ) αe⩽m2. The rest of our paper always assume the necessary existence condition r1>m1 and αe>m2. If condition 0<am2αe−m2<x2 holds, it is cleat that E∗ always exists.
Thus we define the Jacobian matrix of the system (1.1) at an interior equilibrium E∗=(x∗,y∗) as
J(E∗)=(Jij(E∗))2×2=[αx∗y∗(a+x∗)2−r1x∗K1−αx∗a+x∗αeay∗(a+x∗)2−dy∗]. | (2.5) |
The trace, determinant and discriminant of the matrix J(E∗) are denoted as A1=A1(E∗):=tr[J(E∗)], A2=A2(E∗):=det[J(E∗)] and Δ∗=Δ∗(E∗):=A21−4A2, respectively. Then, by using the Perron's theorems and the Routh-Hurwitz's criteria, we have following local stability of a hyperbolic equilibrium E∗ in general cases:
(a) If A1<0 and (a1) A2>0,Δ∗⩾0, then E∗ is an asymptotically stable node; (a2) A2>0,Δ∗<0, then E∗ is an asymptotically stable focus; (a3) A2<0, then E∗ is a saddle point;
(b) If A1=0 and (b1) A2>0, then E∗ is a center or a focus; (b2) A2<0, then E∗ is a saddle point;
(c) If A1>0, then E∗ is unstable and (c1) Δ∗=0, then E∗ is a node; (c2) Δ∗<0, then E∗ is a focus; (c3) Δ∗>0 and A2>0, then E∗ is a node; (c4) Δ∗>0 and A2<0, then E∗ is a saddle point.
A non-hyperbolic point E∗ is a stable(unstable) node if A1<0(A1>0) and A2=0. The nilpotent E∗ is probable a cusp of codimension at least 2, which can ensure potential Bogdanov-Takens bifurcation when A1=A2=0.
Obviously, when r1⩽m1, the equilibrium point E0 is globally asymptotically stable, which can be proved by using a Lyapunov function V=ex+y. Similarly, when r1>m1 and αe<m2, the point E0 is unstable and E∗ does not exist, thus E2 is globally asymptotically stable. For equilibrium point E∗, we define a positive definite Lyapunov function
V=V(x,y)=(x−x∗−x∗lnxx∗)+a+x∗ae(y−y∗−y∗lnyy∗). | (2.6) |
Now, along solutions of the system (1.1), differentiate V with regard to time t to obtain
dVdt=x−x∗xdxdt+a+x∗aey−y∗ydydt. |
Substituting the value of dxdt and dydt from the system (1.1), we can get
dVdt=(αy∗(a+x)(a+x∗)−r1K1)(x−x∗)2−a+x∗aed(y−y∗)2 |
⩽(αy∗a(a+x∗)−r1K1)(x−x∗)2−a+x∗aed(y−y∗)2. |
Thus, it is obvious that if αy∗a(a+x∗)<r1K1, then dVdt⩽0. This equality holds if and only if (x,y)=E∗, i.e., the equilibrium point E∗ is globally asymptotically stable.
Under a generalized condition αy∗a(a+x∗)⩽r1K1, the hyperbolic point E∗ is a locally asymptotically stable focus or node since A1<0 and A2>0. Hence we assume y∗=a(a+x∗)r1ραK1 with the introducing of a control variable ρ∈(0,1], components of the corresponding equilibrium point E∗=(x∗,y∗) and restricted parameter d are
x∗=(−aρ+K1)r1−m1K1r1,y∗=[(r1−m1)K1+ar1(1−ρ)]ρaαK1,d=−αK1[((−αe+m2)K1+(eρα−m2(ρ−1))a)r1+K1m1(αe−m2)]aρ[(−K1+(ρ−1)a)r1+m1K1]2. | (2.7) |
Thus, we can obtain Theorem 2.1, which can guarantee that the equilibrium point E∗ is globally asymptotically stable.
Theorem 2.1. If the system (1.1) has a unique interior equilibrium point E∗ with αy∗a(a+x∗)<r1K1, then the equilibrium point E∗ is globally asymptotically stable (hyperbolic focus or node).
At the same time, by using the Bendixson-Dulac criteria and a proper Dulac function B(x,y)=1xy or B(x,y)=1(x+c)y with a positive constant c>0, then theorem 2.2 can also guarantee that a unique equilibrium point E∗ is globally asymptotically stable on account of the non-existence of closed orbits and limit cycles.
Theorem 2.2. If the system (1.1) has a unique interior equilibrium E∗, and 0<am2αe−m2<x2<a, then the equilibrium E∗ is globally asymptotically stable.
In order to verify feasibility of Theorem 2.1 and 2.2, we will give some numerical simulations. If we take r1=0.6, m1=0.2, m2=0.1, α=0.5, a=1.5, K1=20, e=0.6 and ρ=0.9, the calculation shows that the values of these parameters can satisfy with the condition of Theorem 2.1, the equilibrium point E∗ is a globally asymptotically stable node, which can be seen in Figure 2(a). If we take r1=0.6, m1=0.2, m2=0.1, α=0.5, a=1.5, e=0.6, d=0.05 and x2=aρ(ρ=0.9), then Theorem 2.2 is true, thus the equilibrium point E∗ is a globally asymptotically stable point, which can be found from Figure 2(b). In a word, the equilibrium point E∗ is globally asymptotically stable under certain conditions.
In this section, we mainly consider existence and stability of interior equilibrium in special cases, which can ensure the occurrence of Hopf bifurcation and Bogdanov-Takens bifurcation. In order to simplify A1 as zero or get potential Hopf bifurcation, we take some parameters in a special case as
m1=αe(λ−1)+r1,m2=λαe,K1=ˉμar1αe, | (2.8) |
where λ∈(0,1) and μ=√1+8λˉμ>1 are two control variables for later use to scale parameters. Then the parameter d is constrained by d=4α(μ+1)φd(λ,μ)a(μ+3)2ψd(λ,μ), where auxiliary functions are
φd(λ,μ)=(μ+3)λ−μ+1,ψd(λ,μ)=(μ+3)λ−μ−1. |
It is quite clear that we can derive a little complicated expressions of a required interior equilibrium point E(2)∗=(x(2)∗,y(2)∗) with x(2)∗=14a(μ−1), y(2)∗=−ae(μ+3)4(μ+1)ψd(λ,μ), A1(E(2)∗)=0 and A2(E(2)∗)=−α2e2(μ+3)2(μ+1)φA2(λ,μ), in which φA2(λ,μ)=(λ−1)2μ3+(λ−1)(7λ+5)μ2+(15λ2+10λ−1)μ+9λ2−6λ+5 is also an auxiliary function.
Obviously, the expressions of y(2)∗ and positive parameter d indicate that φd(λ,μ)<0 and ψd(λ,μ)<0, i.e., μ>μm:=3λ+11−λ. The unique positive root(stationary point) of the following equation
∂∂μφA2(λ,μ)=3(1−λ)2μ2+2(λ−1)(7λ+5)μ+15λ2+10λ−1=0, |
which can satisfy above condition, is μ(2)∗=7λ+5+2√λ2+10λ+73(1−λ). Substituting it into φA2(λ,μ), we have φA2(λ,μ(2)∗)=1627(λ−1)φφA2μ(2)∗(λ), where φφA2μ(2)∗(λ)=(λ2+10λ+7)32−λ3−15λ2+60λ+10 is an auxiliary function of λ. Letting ddλφφA2μ(2)∗(λ)=0, we have a unique negative root λ=−5−154√2. Combining φφA2μ(2)∗(0)=10+7√7 and φφA2μ(2)∗(12)=79, we know that the function φφA2μ(2)∗(λ) must be monotonically increasing and has a positive minimum φφA2μ(2)∗(0) on the interval [0,1], or this function is always positive on (0,1), while function φA2(λ,μ(2)∗) is always negative on (0,1). The second order partial derivative with respect to μ at the point μ(2)∗ is
∂2∂μ2φA2(λ,μ(2)∗)=4(1−λ)√λ2+10λ+7>0, |
which can ensure that the function φA2(λ,μ) owns a negative local minimum at the point μ(2)∗.
At this time, combining φA2(λ,μm)=32λλ−1<0, we have a rough estimation of this function φA2(λ,μ) when μ>μm and λ∈(0,1):
(ⅰ) When μm<μ<μ(2)∗, the function φA2(λ,μ) is monotonically decreasing and negative with respect to variable μ.
(ⅱ) When μ>μ(2)∗, the function φA2(λ,μ) is monotonically increasing with respect to variable μ.
(ⅲ) When μ=μ(2)∗, the function φA2(λ,μ) has a negative (local) minimum.
With the positive coefficient of leading order term μ3 in polynomial φA2(λ,μ) at hand, we have a positive value of variable μ, which is also sufficiently large and is larger than μ(2)∗, such that φA2(λ,μ)>0. The zero theorem implies that the equation φA2(λ,μ)=0 must have a unique positive root, which is denoted as μ1 and on the right hand side of the point μ(2)∗.
All in all, we can confirm the classification of the interior equilibrium point E(2)∗:
(ⅰ) When μm<μ<μ1, we have φA2(λ,μ)<0 and A2(E(2)∗)>0, thus the equilibrium point E(2)∗ is a focus or center;
(ⅱ) When μ=μ1, namely φA2(λ,μ)=0 or μ satisfies the following cubic equation
(λ−1)2μ3+(λ−1)(7λ+5)μ2+(15λ2+10λ−1)μ+9λ2−6λ+5=0, | (2.9) |
we have a degenerate interior equilibrium E(2)∗ with A1(E(2)∗)=A2(E(2)∗)=0;
(ⅲ) When μ>μ1, we have φA2(λ,μ)>0 and A2(E(2)∗)<0, thus the equilibrium point E(2)∗ is a saddle point.
Based on these analysis, we summarize two cases for consideration of stability and type of the interior equilibrium point E(2)∗.
Case Ⅰ: μm<μ<μ1
When μ∈(μm,μ1), making a linear transformation (Ⅰ): x=u+x(2)∗, y=v+y(2)∗, we can transfer the equilibrium point E(2)∗ to the origin O. Then we can construct a transformation (Ⅱ): u=−dy(2)∗X+βY, v=−αeay(2)∗(a+x(2)∗)2X to obtain the Jordan form of the system (1.1), thus the new system is
˙X=−βY+3∑i+j=2aijXiYj+O(|X,Y|4), | (2.10a) |
˙Y=βX+3∑i+j=2bijXiYj+O(|X,Y|4), | (2.10b) |
where β=√A2(E(2)∗)>0. Following Perko's book [36] or [37], the first Lyapunov number of the system (2.10) under the condition (2.8) is
σ=3π2β{3(a30+b03)+(a12+b21)−1β[2(a20b20−a02b02)−a11(a02+a20)+b11(b02+b20)]}=−384πλα3e3(μ−1)ψd(λ,μ)φσ(λ,μ)a2β(μ+1)2(μ+3)4φA2(λ,μ), | (2.11) |
where the auxiliary function is φσ(λ,μ)=(λ−1)2μ3+(λ−1)(9λ−1)μ2+(27λ2−18λ+15)(μ+1). The partial derivative of function φσ(λ,μ) with respect to variable μ is
∂∂μφσ(λ,μ)=3(λ−1)2μ2+2(λ−1)(9λ−1)μ+27λ2−18λ+15, |
which is always positive since its discriminant Δ(∂∂μφσ)=144(λ−1)2(λ−119)<0. The special value φσ(λ,μm)=321−λ>0 can ensure that φσ(λ,μ) is a positive function in this case, i.e., σ<0 or the equilibrium point E(2)∗ is a stable multiple focus with multiplicity one.
Case Ⅱ: μ=μ1
When μ=μ1, we will show that the nilpotent(double-zero eigenvalue) E(2)∗ is a BT cusp of codimension 2. Firstly, by using transformations (Ⅰ): x=X+x(2)∗, y=Y+y(2)∗ and
(II):X=α(μ1+1)φ2d(λ,μ1)u4(μ1+3)ψd(λ,μ1),Y=αeφd(λ,μ1)u+(μ1+3)vμ1+3 |
in original system
˙x=r1x[1−8λαexar1(μ21−1)]−[e(λ−1)α+r1]x−αxya+x, | (2.12a) |
˙y=αexya+x−λαey−4α[(λ−1)μ1+3λ+1](μ1+1)y2a(μ1+3)2[(λ−1)μ1+3λ−1], | (2.12b) |
we can derive a new system
˙u=v+3∑i+j=2aijuivj+O(|u,v|4), | (2.13a) |
˙v=3∑i+j=2bijuivj+O(|u,v|4). | (2.13b) |
With the help of the Lemma 1 in [38], such system (2.13) is equivalent to system
˙x=y, | (2.14a) |
˙y=d1x2+d2xy+O(|x,y|3), | (2.14b) |
where the discriminants are
d1=b20=−4α3e2[(λ−4)μ1+(3λ−4)]φd(λ,μ1)2φd(λ2,μ1)a(λ−1)(μ1−1)(μ1+3)4ψd(λ,μ1)<0,d2=b11+2a20=8eλα2φd(λ,μ1)φd2(λ,μ1)a(λ−1)(μ1−1)(μ1+3)3ψd(λ,μ1), |
and φd2(λ,μ)=(λ−1)(λ+7)μ2+(6λ2+16λ+2)μ+(9λ2−6λ+5) is an auxiliary function.
From the equations φA2(λ,μ)=0 and φd2(λ,μ)=0, the Sylvester's resultant with respect to variable μ is
Rμ(φA2,φd2)=|(λ−1)2(λ−1)(7λ+5)15λ2+10λ−19λ2−6λ+500(λ−1)2(λ−1)(7λ+5)15λ2+10λ−19λ2−6λ+5(λ−1)(λ+7)6λ2+16λ+29λ2−6λ+5000(λ−1)(λ+7)6λ2+16λ+29λ2−6λ+5000(λ−1)(λ+7)6λ2+16λ+29λ2−6λ+5|=2048λ(1−λ)3(9λ2−6λ+5)>0, |
which implies d2≠0. On the other hand, for the quadratic function φd2(λ,μ), which is also an downward opening parabola since the negative coefficient (λ−1)(λ+7) is in the highest order term, its discriminant is Δ(φd2)=656λ2−224λ+144>0, and the symmetry axis μ=6λ2+16λ+22(1−λ)(λ+7) is between the longitudinal axis μ=0 and the vertical line μ=μm. By using the special values φd2(λ,0)>0 and φd2(λ,μm)=32λλ−1<0, we have φd2(λ,μ1)<0 and d2>0. Therefore it completes the non-degeneracy condition d1d2≠0 (actually d1d2<0) and the classification work of codimension 2 cusps in this paper, which is meaningful to Bogdanov-Takens bifurcation analysis of codimension 2.
Now, we will continually use transformations (Ⅲ): u=p+a02pq, v=q−a20p2; (Ⅳ): p=w, q=z−c11wz and (Ⅴ): w=x1+12f02x21, z=y1+f02x1y1 to derive a standard form with discriminants d1 and d2:
˙x1=y1+O(|x1,y1|3),˙y1=d1x21+d2x1y1+O(|x1,y1|3), | (2.15) |
hence it also support above conclusions and we can obtain the Theorem 2.3.
Theorem 2.3. As we take the value of parameters under the condition (2.8), the system (1.1) admits an interior equilibrium point E(2)∗ with zero trace.
(ⅰ) When μm<μ<μ1, the equilibrium point E(2)∗ is a stable multiple focus with multiplicity one.
(ⅱ) When μ=μ1, the equilibrium point E(2)∗ is a cusp of codimension 2 (BT bifurcation point).
(ⅲ) When μ>μ1, the equilibrium point E(2)∗ is a saddle point.
Here we can take the value λ=23 to get some interesting result, the first positive(meaningful) root of Eq (2.9) is μ1=μ1,23:=12+3√17≈24.369317 and μm=9 by using identities
cos[13arctan(1171162√51)]=1508√127(9√17+7),sin[13arctan(1171162√51)]=1508√127(3√51−7√3). |
Solving out the Eq (2.4), we have three possible interior equilibrium point E(i)4=(x(i)4,y(i)4) (i=1,2,3), where some components are
x(1,3)4=a32(μ−9)[(μ−9)(μ2−4μ−29)±Φs(μ)],y(1,3)4=ae(μ+3)24(μ2−1)(μ−9)[(μ−1)(μ−3)(μ+15)±Φs(μ)], |
and auxiliary functions are Φs(μ)=√(μ−1)(μ−3)(μ−9)φs(μ) and φs(μ)=μ3−13μ2−153μ−603. With the techniques in Calculus, we know:
(ⅰ) when μ∈(μm,μ+s), μ+s=13+2√1573≈12.687, the function φs(μ) is negative and monotonically decreasing.
(ⅱ) when μ=μ+s, the function φs(μ) owns a negative (local) minimum.
(ⅲ) when μ>μ+s, the function φs(μ) is monotonically increasing and has a unique positive root, which is denoted as μs, where
μs=173947(4822−36√5997)(2411+18√5997)2/3+23(2411+18√5997)13+133≈21.445494∈(μ+s,μ1). |
Case 1. When μs<μ<μ1 or μ>μ1, that is Φs(μ)>0, the system (1.1) has three interior equilibrium point E(i)4=(x(i)4,y(i)4) (i=1,2,3) due to the following inequalities
(μ−9)2(μ2−4μ−29)2−Φs(μ)2=128(μ−9)(μ3+5μ2−25μ−45)>0,(μ−1)2(μ−3)2(μ+15)2−Φs(μ)2=48(μ−1)(μ+3)(μ−3)(μ2−33)>0. |
Here the equilibrium point E(2)4 is a stable multiple focus with multiplicity one when μ<μ1, while it becomes a saddle point when μ>μ1.
Case 2. When μ=μ1, there exist two interior equilibrium points, including an asymptotically stable node E(1)4=(x(1)4,y(1)4)=(a(11+3√17),3ae) since
A1(E(1)4)=(5√17−21)αe<0,A2(E(1)4)=19(895−217√17))α2e2>0,Δ∗(E(1)4)=149(301−73√17))α2e2>0. |
The second degenerate interior equilibrium point E(2)4=(x(2)4,y(2)4)=(14x(1)4,38(1+√17)ae) is a cusp of codimension 2. Furthermore, it is worth noting that the two equilibrium point E(2)4 and E(3)4 can coincide with each other.
Case 3. When μ=μs or Φs(μ)=0, there exist two interior equilibria, that is to say, above two interior equilibrium point E(1)4 and E(3)4 can coincide with each other (but we still denote it as E(1)4), where
x(1)4=(5μs+27)aμs−9,y(1)4=(μs+3)(μ2s+4μs+27)ae(μs−9)(μ2s−1);x(2)4=14a(μs−1),y(2)4=ae(μ2s−9)12(μs+1). |
It is quite clear that this point E(1)4 is a stable node when A1(E(1)4)=−128αe(55μ2s+324μs+981)9(μs−9)(μ2s−1)(μ2s−9)<0 and A2(E(1)4)=0. Seeing [35] in detail, the interior equilibrium E(2)4 is still a stable multiple focus with multiplicity one.
Case 4. When μm<μ<μs or Φs(μ)<0, there exists a unique stable multiple focus E(2)4 with multiplicity one.
In order to verify the feasibility of theoretical derivation, we take r1=0.6 α=0.5, a=1.5 and e=0.6, then some numerical simulations are implemented. Figure 3 depicts the curves of functions φA2(λ,μ), φd(λ,μ) and ψd(λ,μ), which can show the existence of key values. When μ=20<μ1, the unique equilibrium point E(2)4 is a stable multiple focus with multiplicity one(see Figure 4(a)). When μ=25>μ1, there exist three interior equilibrium points including a stable node E(1)4, a saddle point E(2)4 and an unstable node E(3)4 (see Figure 4(b)). When μ=μ1, a cusp of codimension 2 E(2)4 and a stable node E(1)4 will occur (see Figure 4(c)). When μ=μs, there exist two equilibrium points including a stable multiple focus E(2)4 with multiplicity one and a stable node E(1)4 (see Figure 5(a)). When μ=21.5, there exist three equilibrium points including a stable multiple focus E(2)4 with multiplicity one, a saddle point E(3)4 and a stable node E(1)4 (see Figure 5(b)). In a word, there are several kinds of internal equilibrium points with different characteristics in the system (1.1).
This subsection will show the existence and stability of interior equilibrium point in another special case, which can also ensure potential Hopf bifurcation and Bogdanov-Takens bifurcation. As we take m1=r1−αe−3m2, m2=λαe and K1=ar1s4λαe with dimensionless "control variables" λ∈(0,1), μ>1 and s=√μ2−1>0 for later use, then an interior equilibrium point E∗, the determinant A2(E∗) and the confined positive parameter d are listed as follows:
E∗=(x∗,y∗),x∗=14a(s−1+μ),y∗=ae4sφy∗(λ,μ),A2=−α2e2(s−1+μ)s(s+μ+3)2φA2(λ,μ),d=−2αsφd(λ,μ)aψd(λ,μ), | (2.16) |
where the mentioned auxiliary functions φy∗=φy∗(λ,μ), φd=φd(λ,μ), ψd=ψd(λ,μ), φA2=φA2(λ,μ) are:
φy∗=[(μ+7)λ+μ+3]s+[(μ−1)λ+μ+1](μ−1)>0,φd=(λ−1)s+(μ+3)λ−μ+1,ψd=[(λ+1)μ2+(4λ+3)μ+11λ+4]s+[(λ+1)μ2+(5λ+4)μ+2λ+3](μ−1)>0,φA2=[(μ−1)λ2+(−2μ−14)λ+μ−5]s+(μ2+8μ+7)λ2−2(μ−1)2λ+μ2−1. |
The inequality d>0 or φd<0 deduces a lower bound μm=5λ2+2λ+1(1−λ)(3λ+1). The generalized expression of the first positive root μ1=μ1(λ) in equation φA2(λ,μ)=0 is
μ1(λ)=φμ1(λ)3(3λ+5)(3λ+1)(λ−1)2M(λ)13, | (2.17) |
where
φμ1(λ)=(−39λ4+82λ2+72λ+13)M(λ)13+18λ8−1464λ6−3360λ5+2708λ4+5568λ3+3544λ2+1056λ+2M(λ)23+122,M(λ)=108(λ+53)√3(λ−1)2(λ+13)(λ4+8λ3+392λ2+10λ+32)√8λ6−96λ5+431λ4−168λ3−334λ2−104λ−9+27λ12+1890λ10+13176λ9+423λ8−98784λ7−144532λ6+117360λ5+230793λ4+116640λ3+23874λ2+1368λ−91. |
To reveal complexity in this special case, taking a fascinating value λ=59 with identities
sin[13arctan(3346155947278522√51√113)]=126105404(61√1921−2159)√17√39√29531,cos[13arctan(3346155947278522√51√113)]=126105404(183√1921+2159)√17√13√29531, | (2.18) |
and denoting this case as (C2), we have a threshold μ1=μ1,59=2419240+61240√1921 from equation φA2(59,μ)=0. Therefore, we conversely discuss following two cases and the above mentioned interior equilibrium point E∗(denoted as E(2)5) could be a multiple focus or center with multiplicity one when μ∈(μm,μ1). Following the Eq (2.11) and steps in above subsection, the first Lyapunov coefficient is
σ=−81920π(μ−1)4e3α3φσ(μ)27a2β(μ+1)(s−1+μ)[4μ2+(4s+95)μ−265s+1](s+μ+3)8ψσ(μ)7, | (2.19) |
where ψσ(μ)=14μ3+(14s+47)μ2+(47s−24)μ+91s−37>0, and all coefficients in auxiliary function φσ(μ) is listed in the Appendix for completeness. If μ<μσ(>μσ) or σ<0(>0), the equilibrium point E(2)5 is a stable(unstable) multiple focus with multiplicity one, the system (1.1) undergoes a non-degenerate supercritical(subcritical) Hopf bifurcation around E(2)5, and limit cycles generated by this critical point are stable(unstable). On occasion, there may exist some parameter values such that σ=0 or the system (1.1) may undergoes a degenerate Hopf bifurcation for some values of parameters [36]. Accompanying with the Calculus, Figure 6(a) is the curve of function φσ(μ), which is used to guess the unique positive root of equation φσ(μ)=0:
μσ=2√3θ3/4+√6√−2θ3/2+27814578√θ+12362089428√3+6474θ1/42880θ1/4≈9.276513,θ=80√2√2565425987cos[13arctan(782076303168750157010324√6√1441915345)]+4635763. | (2.20) |
At μ=μσ, by using successor function method, the second focal(Lyapunov) quantity g5≈−0.00003569α4e4a4<0 ensures that the equilibrium point E(2)5 is a stable weak focus of order 2 [36]. The Bautin bifurcation(generalized Hopf bifurcation) may occur.
When μ=μ1,59, the system (1.1) owns two separate interior equilibrium points E(1)5=((151+9√1921)a200,4(56+√1921)ae45) and E(2)5=((39+√1921)a8,(709+17√1921)ae81). The equilibrium point E(1)5 is an unstable node since
A1(E(1)∗)=(3481√1921−143239)αe7290>0,A2(E(1)5)=2(2524559−57521√1921)α2e232805>0,Δ∗(E(1)5)=(5538288481−125878879√1921)α2e226572050>0. |
Now we construct a linear transformation (Ⅱ): u=−36X(√1921−11)α, v= −√1921+31√1921−11eX+Y, then the system (1.1) is apparently equivalent to a new system in the standard form of Eq (2.14) with discriminants
d1=b20=−(12471√1921−547081)e2α323328a≠0,d2=b11+2a20=5eα2(−14825+343√1921)5832a≠0. | (2.21) |
The equilibrium point E(2)5 is just a cusp of codimension 2 due to the non-degeneracy condition d1d2≠0. Hence we can obtain Theorem 2.4.
Theorem 2.4. Under the conditions of case (C2), (ⅰ) when μ∈(μm,μ1,59)∖{μσ}, the equilibrium point E(2)5 is a multiply stable(unstable) focus with multiplicity one if μ<μσ(>μσ); (ⅱ) when μ=μσ, the equilibrium point E(2)5 is a stable weak focus of order 2; (ⅲ) when μ=μ1,59, the equilibrium point E(2)5 is a cusp of codimension 2, and the point E(1)5 is an unstable node; (iv) when μ>μ1,59, the equilibrium point E(2)5 is a saddle point; (v) the equilibrium point E(1)5 is just an unstable node.
In order to verify the feasibility of theoretical derivation, we will give some numerical simulations. Figure 6(b) is the curves of functions φA2(red) and φd(blue) with λ=59, which mainly displays the threshold value of control parameters. And then, we take r1=1, a=1.5, α=0.5 and e=0.6, the system (1.1) exists a multiple stable focus with multiplicity one (see Figure 7(a)), a unique multiple unstable focus with multiplicity one and a limit cycle (see Figure 7(b)), a cusp of codimension 2 and an unstable node (see Figure 7(c)), a unique stable weak focus of order 2 (see Figure 7(d)). In a word, the system (1.1) has different internal equilibrium points with the value change of key parameters.
In the following, we will investigate a cusp of codimension 3 in the system (1.1). First of all, translating the equilibrium point E∗=(x∗,y∗) to the origin O via a transformation (Ⅰ): x=X+x∗, y=Y+y∗, we obtain
˙X=F1(X,Y)=f(X+x∗,Y+y∗),˙Y=G1(X,Y)=g(X+x∗,Y+y∗). | (2.22) |
Next, following the technique in above subsection and making a linear transformation (Ⅱ):
u=Y,v=[(m1−r1)+aαy∗(a+x∗)2+2r1x∗K1]Y+aαey∗(a+x∗)2X, | (2.23) |
the above system can be rewritten as the form
˙u=F2(u,v)=v+∑i+j=2aijuivj+O(|u,v|3),˙v=G2(u,v)=∑i+j=2bijuivj+O(|u,v|3). | (2.24) |
From the Lemma 1 in [38], the system (2.24) is equivalent to the system in standard form
˙x=y,˙y=d1x2+d2xy+O(|x,y|3) | (2.25) |
after some nonsingular transformations in the neighborhood of O, where d1=b20, d2=b11+2a20 are discriminates. Solving out an degenerate equilibrium E3=(a(2αe+m2)αe−m2,12e2αa2αe+m2) from the condition d1=0, which also satisfies A1=0, A2=0 and px=Δx=py=Δy=0, thus we have parameters r1, K1 and d (suppose they are positive) with restrictions
r1=2(αe−m2)(8αe+m2)+3m1(2αe+m2)3(2αe+m2),K1=a[2(αe−m2)(8αe+m2)+3m1(2αe+m2)]2(αe−m2)2,d=(2αe+m2)(αe−m2)18e2αa. | (2.26) |
These restrictions (2.26) can deduce another discriminate d2=(8αe+7m2)(m2−αe)18aαe2≠0, i.e., the degeneracy condition d1d2=0, which also yields that the equilibrium point E3 is a cusp of codimension at least 3. Indeed, the degenerate equilibrium point E3 is a codimension 3 Bogdanov-Takens singularity(focus or center) after some nonsingular transformations. Finally, the existence of the equilibrium point E3 can be seen from Figure 8 in details with parameters a=1.5, α=0.5, m1=0.6, m2=0.06 and e=0.6. At the same time, we can obtain the Theorem 2.5.
Theorem 2.5. The the degenerate equilibrium point E3 with conditions 2.26 is a codimension 3 Bogdanov-Takens singularity(focus or center).
In this section, for the interior equilibrium point E(2)4 and E(2)5, we mainly concentrate on Hopf bifurcation curve when μ∈(μm,μ1) and codimension 2 Bogdanov-Takens bifurcation when μ=μ1, respectively.
In the case of (C1), we firstly discuss the existence of Hopf bifurcation curve when μm<μ<μ1,23. We will choose m2 and d as Hopf bifurcation controlling parameters and consider the following perturbed system
˙x=r1x[1−16xαe3ar1(μ2−1)]−(r1−13αe)x−αxya+x, | (3.1a) |
˙y=αexya+x−(23αe+δ1)y−[4α(μ−9)(μ+1)a(μ−3)(μ+3)2+δ2]y2, | (3.1b) |
where δ=(δ1,δ2) is a sufficiently small parameter vector in a neighbourhood of the origin O=(0,0) in the parameter plane. Letting δ≠(0,0), we suppose the equilibrium point E∗ as (x∗,y∗), where x∗=x(2)4+w, w is a sufficiently small variable and component
y∗=e(aμ2−a−16x∗)(a+x∗)3a(μ2+1). |
Substituting it into A1 and A2, we have the solution δ1=δ1(w), δ2=δ2(w), where
δ1=64weα(aμ+2w)3a(μ2−1)[(μ+3)a+4w)],δ2=−16αwa[(μ+3)a+4w]2[(μ2−4μ+3)a−16w](μ+3)2(μ−3)[(μ5−9μ4+46μ3+258μ2+81μ+135)a2+4w(μ4−12μ3+90μ2+204μ−27)a−64w2(μ+1)(μ−9)]. | (3.2) |
At this time, the approximation of the required Hopf bifurcation curve Hp in a small neighbourhood of the origin is a straight line with slope
kHp=limw→0δ2(w)δ1(w)=−3(μ+1)φk(μ)4aeμ(μ2−9)2<0, | (3.3) |
where φk(μ)=μ4−12μ3+82μ2+12μ+45. Noticing that
dφk(μ)dμ=4μ2(μ−9)+164μ+12>0, |
so the function φk(μ) is monotonically increasing and positive when μ>μm, or kHp<0. Hence we can rewrite the determinant as A2=−α2e2[a(μ−1)+4w]9a2(μ2−1)2(aμ+3a+4w)2ϕA2(μ,w), where an auxiliary function is
ϕA2(μ,w)=(μ5−29μ4+110μ3+74μ2−111μ−45)a3+4w(μ+1)(μ3−33μ2+271μ−111)a2−256(μ2−12μ−5)aw2+4096w3. |
Thus, the bifurcation curve Hp of the system (3.1) at the equilibrium point E(2)4 is analytically defined by the solution (3.2), the variables μ and w can ensure the existence of the equilibrium point E(2)4 and A2>0, and we can obtain the Theorem 3.1.
Theorem 3.1. (Hopf bifurcation curve) For the equilibrium point E(2)4 with μm<μ<μ1,23(A1=0 and A2>0), when parameter δ varies in a small neighbourhood of the origin in parameter plane, the Hopf bifurcation curve of the system (3.1) is defined by (3.2) (notice the range of parameter w) and the approximation is a straight line with slope k in a small neighbourhood of the origin. Furthermore, the curve divides a small neighbourhood of the origin in the parameter plane into two regions Ⅰ and Ⅱ, in which dynamical behaviors of the system (3.1) can be exhibited.
For (C2), starting from a perturbed system (3.1) with μ∈(μm,μ1,59)∖{μσ} and y∗=4e(6as−5x∗)(a+x∗)9as. Solution δ1=δ1(w), δ2=δ2(w) are
δ1=80sαw(aμ+2w)e9a(μ2−1)(aμ+as+3a+4w),δ2=[16(84μ3−84μ2s+251μ2−467μs+676μ−313s+509)wα(444a2μ4+444a2μ3s+263a2μ3+2423a2μ2s+480aμ3w+1056aμ2sw+3155a2μ2−619a2μs+2660aμ2w+4220aμsw−480μ2w2+4537a2μ−2248a2s+9120aμw+2464asw+1000μw2+1201a2+6940aw+1480w2)]/[(μ+1)a(14a3μ3+14a3μ2s+47a3μ2+47a3μs+36a2μ2w+36a2μsw−24a3μ+91a3s−100a2μw+188a2sw−120aμw2+72asw2−37a3−96a2w−200aw2−160w3)(12μ−37)(168μ+193)(3μ+5)2]. | (3.4) |
Therefore, the slope kHp(μ) of the approximate straight line of the Hopf bifurcation curve Hp at O is
kHp(μ)=−18(888sμ2+888μ3−1277sμ+3763μ2−2515s+3797μ+672)(μ+1)5ae(14sμ2+14μ3+61sμ+61μ2+37s+138μ+91)(168μ+193)μ. | (3.5) |
At the same time, we can obtain the Theorem 3.2.
Theorem 3.2. (Hopf bifurcation curve) For the equilibrium point E(2)5 with μ∈(μm,μ1,59)∖{μσ}, when parameter δ varies in a small neighbourhood of the origin in parameter plane, the Hopf bifurcation curve of the system (3.1) is defined by (3.4) (notice the range of parameter w) and the approximation is a straight line with slope k in a small neighbourhood of the origin. Furthermore, the curve divides a small neighbourhood of the origin in the parameter plane into two regions Ⅰ and Ⅱ, in which dynamical behaviors of the system (3.1) can be exhibited.
In order to verify the feasibility of the Theorems 3.1 and 3.2, we will give some numerical simulations. For the equilibrium point E(2)4 with r1=0.6 α=0.5, a=1.5, e=0.6 and μ=10, the Hopf bifurcation curve with w in parameter plane is
Hp={δ∣δ1≈0.172391w(w+7.5)39+8w,δ2≈−0.012398w(w+3.300890)(w−80.116231)(w+4.875)2(w−5.90625)}, |
and the straight line of the approximate representation of bifurcation curve Hp is δ2≈−0.704575δ1. Figure 9(a) depicts the Hopf bifurcation curve, Figure 10(a), (b) depict phase diagrams of asymptotically stable focus and unstable focus(with a limit cycle) corresponding to δ1=δ2=0.001 (in region Ⅰ) and δ1=δ2=−0.001(in region Ⅱ) respectively. For the equilibrium point E(2)5 with r1=0.6 α=0.5, a=1.5, e=0.6 and μ=8 in the case (C2), the bifurcation curve Hp is analytically formulated by such solutions of the system (3.1), i.e., Hp={δ∣δsatisfies(3.1)}, which can be seen from Figure 9(b), and a Hopf bifurcation curve Hp(red) is defined by
δ1(w)≈28.221347(w+6)w1789.570497+252w,δ2(w)≈−(0.010129(w+5.546478))(w−98.756256)w(w+7.101470)(w+7.101470)(w−8.685587) | (3.6) |
and accompanied by its corresponding slope kHp(dashed blue line) at O. Furthermore, we will note
A2(w)≈1(w+7.101470)5(w−8.685587)(−3045.413083−2842.498683w+15.609444w2+561.932800w3+169.930141w4+14.098895w5−1.303023w6−0.278897w7−0.012542w8), | (3.7) |
then the curve Hp divides a parameter plane into separate two regions. Figure 11(a), (b) present phase diagrams of a stable node and an unstable focus with respect to δ1=0.0001, δ2=0.00001 and δ1=−0.0001, δ2=−0.00001 when μ=8, respectively. Figure 11(c), (d) present phase diagrams with same values of parameters δ1 and δ2 when μ=10. In a word, it is obvious to see from the numerical simulation works that the Hopf bifurcation can occur for the equilibrium point E(2)4 and E(2)5, which also indirectly proves the validity of the theoretical derivation.
We firstly recall the system (1.1) with a cusp E(2)4 of codimension 2 when parameter conditions λ=23 and μ=μ1,23 hold, in other words, we can start with an unfolding system
˙x=r1x[1−2xαe3ar1(37+9√17)]−(r1−13αe)x−αxya+x, | (3.8a) |
˙y=α(e+δ1)xya+x−(23αe+δ2)y−(5−√17)α9ay2 | (3.8b) |
by introducing bifurcation parameters e and m2. Naturally, a parameter vector δ=(δ1,δ2) is in a sufficiently small neighbourhood of the origin O in the parameter plane. By using the transformation (Ⅰ): x=X+x(2)4, y=Y+y(2)4 and expanding such system in a power series around the origin, it can be rewritten as
˙X=F1(X,Y)=2∑i+j=1aijXiYj+O(|X,Y|3), | (3.9a) |
˙Y=G1(X,Y)=2∑i+j=0bij(δ)XiYj+O(|X,Y|3), | (3.9b) |
where b00(0,0)=0. Secondly, we will use a transformation (Ⅱ): u=X, v=F1(X,Y), the system (3.9) can be reduced to a new system
˙u=F2(u,v)=v, | (3.10a) |
˙v=G2(u,v)=2∑i+j=0dij(δ)uivj+O(|u,v|3), | (3.10b) |
where dij(δ) can be expressed by coefficients aij and bij(δ), dij(0,0)=0 (i+j⩽1). Thirdly, making a transformation (Ⅲ): p=u+d01(δ)d11(δ), q=v since d11(0,0)=(5√17−21)αe12a≠0, we have
˙p=F3(p,q)=q, | (3.11a) |
˙q=G3(p,q)=2∑i+j=0fij(δ)piqj+O(|p,q|3), | (3.11b) |
where fij(δ) can be expressed by dij(δ), but we will omit them here. At the same tome, there exist f01(δ)=0 and f20(0,0)=(161−39√17)α2e254a>0. Then we will construct a transformation (Ⅳ): w=p, z=(1−f02(δ)p)q, dt=(1−f02(δ)p)dτ, one can rewrite above system as(the symbol τ is denoted as t)
˙w=F4(w,z)=z, | (3.12a) |
˙z=G4(w,z)=h00(δ)+h10(δ)w+h20(δ)w2+h11(δ)wz+O(|w,z|3). | (3.12b) |
Similarly, we omit the expressions of coefficients hij(δ) although they are expressed iteratively by fij(δ), and h20(0,0)=f20(0,0)>0, h11(0,0)=d11(0,0)<0. That is to say, there is a small neighbourhood of the origin such that h20(δ) is positive and h11(δ) is negative when δ falls in this neighbourhood. Finally, the transformation (Ⅴ): m=h11(δ)2h20(δ)w, n=h11(δ)3h20(δ)2z, dt=h11(δ)h20(δ)dτ converts from above system to a generic normal form of Bogdanov-Takens bifurcation
˙m=F5(m,n)=n, | (3.13a) |
˙n=G5(m,n)=l00(δ)+l10(δ)m+m2+mn+O(|m,n|3), | (3.13b) |
where the symbol τ is still denoted as t, and two discriminants are
d1=d1(δ)=l00(δ)=−81(1483+365√17)8192eδ1+243(2361+559√17)32768αeδ2+O(|δ1,δ2|2),d2=d2(δ)=l10(δ)=3(1133+283√17)256eδ1−9(3927+961√17)2048αeδ2+O(|δ1,δ2|2). |
This system (3.8) is indeed a generic family unfolding at the codimension 2 cusp E(2)4 according to the rank of a Jacobian matrix or the nonzero Jacobian determinant
∂(d1,d2)∂(δ1,δ2)|δ=0=2187(176337+42583√17)8388608αe2≠0. |
Therefore we obtain local approximated representations of saddle-node (SN), Hopf (H) and homoclinic (HL) bifurcation curves up to second-order with slope kBT=(1+√17)α6≈0.853851α>0 around the origin for the system (3.8) [39]. These bifurcation curves can divide the parameter plane into several regions, which can exhibit separately dynamical behaviors.
(ⅰ) The saddle-node bifurcation curve is formulated by
SN={δ∣d1=14d22}={δ∣−81(1483+365√17)8192eδ1+243(2361+559√17)32768αeδ2−27(1292261+318563√17)131072e2δ21+81(2696527+641161√17)262144αe2δ1δ2−243(20428527+5081065√17)8388608α2e2δ22+O(|δ1,δ2|3)=0}. | (3.14) |
(ⅱ) The Hopf bifurcation curve is formulated by
H={δ∣d1=0,d2<0}={δ∣−81(1483+365√17)8192eδ1+243(2361+559√17)32768αeδ2−9(1277091+317525√17)65536e2δ21+27(11642831+2746889√17)524288αe2δ1δ2−81(11431247+2867337√17)2097152α2e2δ22+O(|δ1,δ2|3)=0}. | (3.15) |
(ⅲ) The homoclinic bifurcation curve is formulated by
HL={δ∣d1=−625d22,d2<0}={δ∣−81(1483+365√17)8192eδ1+243(2361+559√17)32768αeδ2−9(16056063+4090457√17)1638400e2δ21+27(182198831+42270377√17)13107200αe2δ1δ2−81(192417617+49040343√17)52428800α2e2δ22+O(|δ1,δ2|3)=0}. | (3.16) |
Thus, we can obtain the Theorem 3.3.
Theorem 3.3. (Bogdanov-Takens bifurcation of codimension 2). For the unfolding system (3.8) with bifurcation parameters e and m2, in a small neighbourhood of the equilibrium point E(2)4, the system undergoes an attracting Bogdanov-Takens bifurcation of codimension 2 when the value of parameter δ varies in such sufficiently small neighbourhood of the origin. Furthermore, this system is a generic family unfolding at the cusp E(2)4 of codimension 2.
Here we will take r1=0.6, a=1.5, α=0.5 and e=0.6, then Figure 12 depicts the saddle-node, Hopf and homoclinic bifurcation curves in different colors, which can show the existence of critical thresholds.
(ⅰ) When δ1=δ2=0, it is evident from the Theorem 2.3 that there exist two interior equilibrium points, including an asymptotically stable node E(1)4 and a Bogdanov-Takens cusp of codimension 2 E(2)4.
(ⅱ) When δ1=0.01, δ2=0 (δ lies in positive δ1 axis) or δ falls in region Ⅰ (the region between saddle-node bifurcation curve SN2 and homoclinic bifurcation curve), there exist three interior equilibrium points, where two interior equilibrium points are bifurcated from a stable node in (viii), which can be seen from Figure 13(a), (b).
(ⅲ) When (δ1,δ2)≈(0.01,4.265486×10−3) or δ lies in homoclinic bifurcation curve, there exist three interior equilibrium points and a homoclinic loop.
(ⅳ) When (δ1,δ2)≈(0.01,4.267370×10−3) or δ falls in region Ⅱ(the region between homoclinic bifurcation curve and Hopf bifurcation curve), there exist three interior equilibrium points, including an unstable focus, a saddle point and a stable node, which can be seen from Figure 14(a), (b).
(ⅴ) When (δ1,δ2)≈(0.01,4.269255×10−3) or δ lies in Hopf bifurcation curve, there exist three interior equilibrium points, including a saddle point and a stable node.
(ⅵ) When (δ1,δ2)≈(0.01,4.271301×10−3) or δ falls in region Ⅲ(the region between Hopf bifurcation curve and saddle-node bifurcation curve SN1), there exist three interior equilibrium points, including a stable focus which is unstable in case (ⅳ), which can be seen from Figure 15(a). However, by combining the case (ⅳ), it can ensure the potential Hopf bifurcation, but the homoclinic loop is broken.
(ⅶ) When (δ1,δ2) lies in saddle-node bifurcation curve, there exist two interior equilibrium points.
(ⅷ) When (δ1,δ2)≈(0.01,8.546695×10−3) or δ falls in region Ⅳ(the region on the left hand side of saddle-node bifurcation curve), there exists a unique stable node, which can be seen from Figure 15(b).
However, if we choose m2 and d as Bogdanov-Takens bifurcation parameters and rewrite the original system in the unfolding form (3.1) with μ=μ1,23:
˙x=r1x[1−2xαe3ar1(37+9√17)]−(r1−13αe)x−αxya+x, | (3.17a) |
˙y=αexya+x−(23αe+δ1)y−[(5−√17)α9a+δ2]y2, | (3.17b) |
where δ1 and δ2 are sufficiently small parameters and the vector δ=(δ1,δ2) is in a small neighbourhood of the origin O as well. Following the procedures above and the values of parameters, the unfolding system (3.1) is also a generic family unfolding at the codimension 2 Bogdanov-Takens cusp E(2)4 according to the nonzero Jacobian:
−1∂d1∂δ2⋅∂(d1,d2)∂(δ1,δ2)|δ=0=9(273+23√17)1024αe>0. |
and we have representations of saddle-node, Hopf and homoclinic bifurcation curves with slope kBT=1−√176ae≈−0.520518ae<0 around the origin. Furthermore, it should be noticed that the slope k can be viewed as the limiting case of the slope (3.3) when μ→μ1,23. At the same time, when δ lies on the Hopf bifurcation curve H, for instance, (δ1,δ2)≈(1×10−5,−5.783264×10−6), the Hopf bifurcation can not undergo; when δ lies on the homoclinic bifurcation curve HL, for instance, (δ1,δ2)≈(1×10−5,−5.783278×10−6), the homoclinic loop does not exist. Moreover, it should be noticed more that these two cases both occur owing to d2>0 or the minus of (3.2). While the saddle-node bifurcation curve up to second-order can be formulated by
SN={δ∣243(2361+559√17)32768αeδ1+729(1483+365√17)a32768αδ2−243(20428527+5081065√17)8388608α2e2δ21−729(13636927+3239449√17)a4194304α2eδ1δ2−2187(9087263+2237753√17)a28388608α2δ22+O(|δ1,δ2|3)=0}. | (3.18) |
For the saddle-node bifurcation curve, when δ lies on the region Ⅰ (the left hande side of the SN curve), there exist three interior equilibrium points. When δ lies on the region Ⅱ (the right hand side of the SN curve), there exists a unique interior equilibrium point. When δ lies on the saddle-node bifurcation curve SN1, there exist three interior equilibrium points. When δ lies on the saddle-node bifurcation curve SN2, there exists a unique interior equilibrium point. All the detailed results can be seen in the Figure 16 for saddle-node bifurcation curve in this novel phenomenon with r1=0.6, a=1.5, α=0.5 and e=0.6.
Similarly, for the case (C2) with λ=59 and Bogdanov-Takens bifurcation parameters m2 and d, according to a Jacobian matrix with rank 2 and
−1∂d1∂δ2⋅∂(d1,d2)∂(δ1,δ2)|δ=0=−33692028871+832074361√19211622400000e<0, | (3.19) |
local approximated representations of saddle-node (SN), Hopf (H) and homoclinic (HL) bifurcation curves up to second order with slope kBT=709−17√1921648ae<0 at O are obtained rapidly. It is also the limitation limμ→μ1,59kHp(μ).
(ⅰ) The saddle-node bifurcation curve is formulated by
SN={δ∣d1(δ)=14d2(δ)2}={δ∣−63629496759076709+1454440861891419√192163273600000000αeδ1−(23152747107243364241+528225003996330031√1921)a1281290400000000αδ2+2134120285730918153666651+48639881902310601690341√1921131609088000000000000α2eδ21+(63753096761031669666816023+1454616053502361391140393√1921)a74030112000000000000α2eδ1δ2+(7652528762384382093375762493+174598709679460634819116163√1921)a2749554884000000000000α2δ22+O(|δ1,δ2|3)=0}, | (3.20) |
where the half curves SN1(SN2) is the "right" ("left") part of curve SN in the forth(second) quadrant, respectively.
(ⅱ) The Hopf bifurcation curve is formulated by
H={δ∣d1(δ)=0,d2(δ)<0}={δ∣−63629496759076709+1454440861891419√192163273600000000αeδ1−(23152747107243364241+528225003996330031√1921)a1281290400000000αδ2+178349298417656661730271+4064858239867885041761√192110967424000000000000α2e2δ21+(21199900878564120150636217+483705350254750228637447√1921)a24676704000000000000α2eδ1δ2+(852219601234168435593786269+19444087865263472653017379√1921)a283283876000000000000α2δ22+O(|δ1,δ2|3)=0}. | (3.21) |
(ⅲ) The homoclinic bifurcation curve is formulated by
HL={δ∣d1(δ)=−625d2(δ)2,d2(δ)<0}={δ∣−63629496759076709+1454440861891419√192163273600000000αeδ1−(23152747107243364241+528225003996330031√1921)a1281290400000000αδ2+5329(838970735785952358313+19121465556184117783√1921)274185600000000000000α2e2δ21+(176256789653796176682215483+4021544578154623358037253√1921)a205639200000000000000α2eδ1δ2+(64056051282347703285280481599+1461491238758045401440625409√1921)a26246290700000000000000α2δ22+O(|δ1,δ2|3)=0}. | (3.22) |
Thus, we can obtain the Theorem 3.4.
Theorem 3.4. (Bogdanov-Takens bifurcation of codimension 2) From the unfolding system (3.17) with bifurcation parameters m2 and d in the case (C2), this system undergoes an Bogdanov-Takens bifurcation of codimension 2 when δ varies in a sufficiently small neighbourhood of the origin. Furthermore, the system is a generic family unfolding at the cusp E(2)5 of codimension 2 as well.
For numerical simulation, we take r1=1, α=0.5, a=1.5 and e=0.6, then Figure 17 depicts saddle-node (red), Hopf (green) and homoclinic (blue) bifurcation curves in different colors, which can show the existence of critical thresholds..
(ⅰ) When δ1=δ2=0, it is evident that there exist two interior equilibrium points, including an unstable node E(1)5 and a codimension 2 cusp E(2)5.
(ⅱ) When (δ1,δ2)≈(0.001,−3.095598×10−5) or δ falls in region Ⅰ (the region below saddle-node bifurcation curve SN), there exists a unique unstable node.
(ⅲ) When (δ1,δ2)≈(0.001,−6.188787×10−5) or δ falls in region Ⅱ (the region between Hopf bifurcation curve and saddle-node bifurcation curve SN1), there exist three interior equilibrium points, including an unstable node, a saddle point and an unstable focus.
(ⅳ) When δ lies in Hopf bifurcation curve, there exist three interior equilibrium points, including an unstable node, a saddle point, and a non-hyperbolic equilibrium (focus or center) with zero trace and positive determinant, which can ensure potential Hopf bifurcation.
(ⅴ) When (δ1,δ2)≈(0.001,−6.184072×10−5) or δ falls in region Ⅲ (the region between homoclinic bifurcation curve and Hopf bifurcation curve), there exist three interior equilibrium points, including a stable focus which is unstable in case (ⅲ).
(ⅵ) When δ lies in homoclinic bifurcation curve, there exist three interior equilibrium points and a homoclinic loop, including an unstable node, a saddle point and a stable focus.
(ⅶ) When (δ1,δ2)=(0.001,−3.090882×10−5) or δ falls in region Ⅳ (the region between saddle-node bifurcation curve SN2 and homoclinic bifurcation curve), there still exist three interior equilibrium points, including an unstable node, a saddle point and a stable focus.
(ⅷ) When (δ1,δ2) lies in saddle-node bifurcation curve SN1, there exists a unique unstable node.
(ⅸ) When (δ1,δ2) lies in saddle-node bifurcation curve SN2, there exist three interior equilibrium points.
Now we especially study the limit cycle generated by Hopf bifurcation, they showed the efficiency of the perturbation method by using a topological polynomial version of the classical Rosenzweig-MacArthur (R-M) predator-prey model in the paper [40]. In this section, we focus on the approximate calculation of limit cycles in the original predator-prey system (3.1) via a perturbation procedure and canonical transformation, which can be used to determine the limit cycles and their associated frequencies in general two-dimensional systems. Comparing it with the Lindstedt-Poincare (LP) method, the method can give accurate results, while the LP method is limited to weakly nonlinear systems, although it is simple and is frequently used as an algorithm to approximate steady-state periodic solutions in nonlinear oscillators [41]. Recalling an unfolding system mentioned in Subsection 3.1 with μ∈(μm,μ1,59), σ<0 and a sufficiently small parameter vector (δ1,δ2) in a neighbourhood of the origin O in the parameter plane:
˙x=r1x(1−xK1)−αxya+x−m1x:=P(x,y),˙y=αexya+x−(m2+δ1)y−(d+δ2)y2:=Q(x,y), | (4.1) |
when μ=8, the equilibrium points E∗ is a multiple stable focus with multiplicity one, and the corresponding non-degenerate Hopf bifurcation is supercritical. According to the Figure 10(b), a limit cycle exists when sufficiently small (δ1,δ2) falls in region Ⅱ. In this perturbation procedure, we firstly transfer the equilibrium point E∗ to the origin O by using a linear transformation (Ⅰ): x=X+x∗, y=Y+y∗ and obtain a new system. Secondly, we construct a nonsingular transformation (Ⅱ): X=η, Y=√−J211+2J11J22−4J12J21−J222ξ−(J11−J22)η2J12 such that the system (3.1) has its real Jordan's form:
˙ξ=Aξ−Bη+P1(ξ,η),˙η=Bξ+Aη+Q1(ξ,η), | (4.2) |
where A=12A1, B=12√−Δ∗, P1(ξ,η)=∞∑i+j=2˜aijξiηj and Q1(ξ,η)=∞∑i+j=2˜bijξiηj are analytical functions. Obviously, the limit cycle is enclosing an unstable hyperbolic focus or non-hyperbolic node. Introducing a dimensionless time scale transformation τ=ωt with frequency ω of a limit cycle, above system can be written as a canonical system:
Ωdξdτ=δξ−η+P2(ξ,η),Ωdηdτ=ξ+δη+Q2(ξ,η), | (4.3) |
where Ω=ωB, δ=AB, P2=P1B and Q2=Q1B. Now we suppose that there exist series
Ω=1+∞∑n=1δnΩn=11−ϵ2(1+∞∑n=2ϵnγn),ξ=∞∑n=1ϵnξn,η=∞∑n=1ϵnηn, |
in which ϵ=√δΩ11+δΩ1. Substituting them into Eq (4.3) and noticing the series expansions of functions in the right hand side, we recursively derive following coupled first-order differential equations of ξn and ηn in all orders of ϵ:
ϵ1:dξ1dτ+η1=0,dη1dτ−ξ1=0; | (4.4) |
ϵ2:{dξ2dτ+η2−a02η21−a11ξ1η1−a20ξ21=0,dη2dτ−ξ2−b02η21−b11η1ξ1−b20ξ21=0; | (4.5) |
ϵ3:{dξ3dτ+γ2dξ1dτ−ξ1Ω1+η3−η1−a03η31−a12η21ξ1−a21η1ξ21−a30ξ31−2a02η1η2−a11η1ξ2−a11η2ξ1−2a20ξ1ξ2=0,dη3dτ+γ2dη1dτ−ξ3+ξ1−η1Ω1−b03η31−b12η21ξ1−b21η1ξ21−b30ξ31−2b02η1η2−b11η1ξ2−b11η2ξ1−2b20ξ1ξ2=0;⋯, | (4.6) |
and so on, where all involved coefficients are defined by aij=1B˜aij, bij=1B˜bij.
To illustrate the procedure process, we mainly concentrate on the values of parameters from Figure 10 with small parameters δ1=−0.0001 and δ2=0. The steady-state solutions in order ϵ1 are ξ1=−Asin(τ), η1=Acos(τ), where A is a constant to be determined. Then the straightforward solutions of ξ2 and η2 are
ξ2=A23[(a11+12b02−12b20)cos(2τ)+(a02−a20−12b11)sin(2τ)−32(b02+b20)]≈−A23000000000000[100897370720cos(2τ)+53739942597sin(2τ)−475084764300],η2=A26[(−a02+a20+2b11)cos(2τ)+(a11+2b02−2b20)sin(2τ)+3(a02+a20)]≈−A26000000000000[−110557181337cos(2τ)+575982135020sin(2τ)−49774994529]. | (4.7) |
From the Eq (4.4), we have a second-order differential equation of η3:
d2η3dτ2+η3=C31cos(τ)+S31sin(τ)+C33cos(3τ)+S33sin(3τ), | (4.8) |
where C31, S31, C33, S33 are some constants. Letting C31=S31=0 or eliminating the secular terms in this equation, we know γ2≈−0.011293A2−1, Ω1≈274.507342A2. Hence we accordingly derive solutions
ξ3≈−A33000000000000[33323257806cos(τ)3+2832036878cos(τ)2sin(τ)−52858138173cos(τ)+59719753369sin(τ)],η3≈−A3800000000000[8217691556cos(3τ)+3753250989sin(3τ)]. | (4.9) |
Similarly, based on the previous process, we obtain a second-order differential equation of η4 in the form of
d2η4dτ2+η4=C41cos(τ)+C42cos(2τ)+S42sin(2τ)+C44cos(4τ)+S44sin(4τ). | (4.10) |
Here the coefficient C41 yields γ3=0, thus the solutions of η4 and ξ4 read
ξ4≈A2400000000000000[299979339450A2sin(2τ)cos(2τ)−250523414758A2cos(2τ)2+46357860860A2sin(2τ)−383390712843A2cos(2τ)−1000sin(2τ)+2000cos(2τ)],η4≈−A21500000000000[10532770568A2cos(τ)4−13522443784A2cos(τ)3sin(τ)−7811903244A2cos(τ)2+6097955141A2cos(τ)sin(τ)−2135810211A2−10cos(τ)2+35]. | (4.11) |
Repeating above mentioned steps, we iteratively and formally derive required constants A≈4.202827, γ4≈0.206668, γ5=0, γ6≈−0.014240, γ7=0, ⋯ from coefficients C51, S51, C61, C71, C81, ⋯ and solutions ξn(τ), ηn(τ)(n=5,6,7,8), which are listed in the Appendix. A generalized second-order ODE of ηn reads
d2ηndτ2+ηn=n∑k=0[Cn,k(τ)cos(kτ)+Sn,k(τ)sin(kτ)], | (4.12) |
and the corresponding formal solutions via Leonhard Euler's method is
ξn(τ)=n∑k=0[˜Cn,k(τ)cos(kτ)+˜Sn,k(τ)sin(kτ)],ηn(τ)=n∑k=0[¯Cn,k(τ)cos(kτ)+¯Sn,k(τ)sin(kτ)]. | (4.13) |
Here Cn,k(τ), Sn,k(τ), ˜Cn,k(τ), ˜Sn,k(τ), ¯Cn,k(τ) and ¯Sn,k(τ) are undetermined polynomials of variable τ with Cn,k(0):=Cn,k, Sn,k(0):=Sn,k, Sn,0(τ)=˜Sn,0(τ)=¯Sn,0(τ):=0; n=0,1,2,⋯, k=0,1,2,⋯,n. Finally, a N-order approximate solution to the limit cycle of the Eq (3.1) is
x(t)≈x(N)(t)=x∗+η(N)(t),y(t)≈y(N)(t)=y∗+√−J211+2J11J22−4J12J21−J2222J12ξ(N)(t)−J11−J222J12η(N)(t) | (4.14) |
where
ξ(N)(t)=N∑n=1ϵnξn(BΩt),η(N)(t)=N∑n=1ϵnηn(BΩt). |
Furthermore, up to eight-order approximation of solutions ξ(t) and η(t), for comparison, Figure 18(a), (b) depict limit cycles via the Runge-Kutta 45 method(red) and perturbation procedure(blue) with values μ=8, δ1=−0.0001, δ2=0 and μ=5, δ1=−0.001, δ2=0, respectively. In the first figure, the red and blue curves almost coincide with each other. For the latter option, some main invariants are calculated as A≈10.010405, Ω1≈0.5067015052, γ2≈−3.031960310, γ3=0, γ4≈−6.244350476, γ5=0, γ6≈−29.58780360, γ7≈183.5476191, γ8≈159.9270824.
Within the framework of predator-prey ecological dynamics, this paper mainly discusses the dynamic properties of the Bazykin's predator-prey ecosystem with Holling type Ⅱ functional response and interspecific density-restricted effects on the predators, including the existence and stability of all possible equilibrium points, Hopf bifurcation and Bogdanov-Takens bifurcation.
Aiming at all possible equilibrium points of the Bazykin's predator-prey ecosystem, mathematical theory works mainly investigate the existence and stability of boundary equilibrium point, hyperbolic equilibrium point, non-hyperbolic equilibrium point and cusp of condimension 3, and then give some corresponding threshold conditions of some key parameters, for example, by introducing control variables λ and μ, the stability analysis and type of the interior equilibrium point E(2)∗(∗=4,5) are ascertainable and clear in detail, and this equilibrium point E(2)∗ is a multiple focus with multiplicity one(or cusp of codimension 2 or saddle point) when μm<μ<μ1(or μ=μ1 or μ>μ1). At the same time, it is easy to find from numerical simulation works that all the equilibrium points derived from theoretical derivation always exist and have corresponding stability states, which indirectly prove that the Bazykin's predator-prey ecosystem has different intrinsic dynamic properties with the value change of critical parameters, which also represents that predator population and prey population have different coexistence modes.
For Hopf bifurcation and Bogdanov-Takens bifurcation, mathematical theory works mainly investigate Hopf bifurcation and Bogdanov-Takens bifurcation about the equilibrium point E(2)4 and E(2)5, and give some threshold conditions of some key parameters to ensure the occurrence of Hopf bifurcation and Bogdanov-Takens bifurcation. With the help of numerical simulation, the formulated Hopf bifurcation curve when μm<μ<μ1 and Bogdanov-Takens bifurcation of codimension 2 when μ=μ1 are both presented, which can directly verify the validity and feasibility of theoretical derivation, and indirectly explain that the Bazykin's predator-prey ecosystem has complex bifurcation dynamic evolution process with the value change of critical parameters, such as saddle-node, Hopf and homoclinic bifurcation. Furthermore, it is obvious to find that the Bazykin's predator-prey ecosystem can exhibit different dynamical behaviours when parameter vector (δ1,δ2) varies in different regions in a small neighbourhood of the origin O in the parameter plane. However, it is also worth noting that the Hopf and homoclinic bifurcations do not exist if we choose m2 and d as bifurcation parameters in (C1). Moreover, we specifically investigate the limit problem by comparing the Runge-Kutta 45 method with perturbation procedure.
In the follow-up research works, based on the research results of this paper and biological manipulation theory, we will further explore the dynamic relationship between Microcystis aeruginosa and filter-feeding fish by using bifurcation dynamic analysis and explain the biological significance of the Bazykin's predator-prey ecosystem. At the same time, due to the difference of nutrient load, biological composition and hydrodynamic conditions in different water bodies, the dynamic relationship between filter feeding fish and Microcystis aeruginosa on the basic of the Bazykin's predator-prey ecosystem still needs to be further studied in the experiment.
In summary, all results in this paper further show that the system (1.1) proposed by the paper [24] has more abundant dynamic behavior, and vigorously develop the dynamic properties of the Bazykin's predator-prey ecosystem. Furthermore, the results of bifurcation and stability can be helpful to better understand the interaction mechanism between prey population and predator population in natural real ecosystem.
This work was supported by the National Natural Science Foundation of China (Grant no.31570364 and no.61871293) and the National Key Research and Development Program of China (Grant no. 2018YFE0103700). We thank teachers Min Zhao and Chuanjun Dai in Wenzhou University. Author S. T. Wang thank teachers Maotong Zhang and Shujing Zhao in Liushi No.3 Middle School. We also thank the Editor and Referee(s).
The authors declare that there is no conflict of interest regarding the publication of this paper.
All coefficients in function φσ(μ)=26∑i=0aiμi are:
a0=−901216986993425953s−690735397062449054,a1=−7430425914068984147s−6884009442270443996,a2=−25039469856001943868s−33960715029169842122,a3−36327625126218808192s−106006380987337770808,a4=23813243876181697922s−219774736835924252504,a5=231701500305529576998s−283770759189235628296,a6=562654839813783725532s−133880117821863988900,a7=821527882365809257148s+298839475213006129920,a8=758628501725909295927s+819668812399890357890,a9=307714350506545891573s+1059484475867702695300,a10=−313891529367115361240s+801581088827681437262,a11=−775475240462416420740s+195928837969850670808,a12=−894079031015008419680s−382000374175148440524,a13=−726734929150839367680s−654885975454974356416,a14=−454979042514059201280s−610000421153971812768,a15=−226162103453499020928s−409066946025551638656,a16=−90030584187038049792s−211714542563663790336,a17=−28563209827168347648s−86464340655657245952,a18=−7090481890045417472s−27900627105898709248,a19=−1322666437445936128s−7006358180280438784,a20=−168472574582366208s−1317635181526709248,a21=−10128944271982592s−168919737461608448,a22=888031307333632s−10261691075915776,a23=265257481617408s+875442405031936,a24=25177804603392s+264785229119488,a25=944504995840s+25177804603392,a26=944504995840. |
The required approximate expression of solutions ξn(τ), ηn(τ)(n=5,6,7,8) in Section 4 are:
ξ5(τ)≈8133176134962500000000sin(τ)cos(τ)4−8384602893071500000000000cos(τ)2sin(τ)−5817046626591500000000000sin(τ)+1142118194312500000000cos(τ)5−295303209583300000000000cos(τ)3+27444566793125000000cos(τ),η5(τ)≈37694475731500000000cos(τ)5+561294899187500000sin(τ)cos(τ)4−374071366567120000000000cos(τ)3−519202543400000000cos(τ)2sin(τ)+368324585101480000000000cos(τ)−298170059960000000000sin(τ);ξ6(τ)≈−822047803780000000000sin(2τ)cos(2τ)2−14232508519800000000000sin(2τ)cos(2τ)−163127922259427446320000000000000000000sin(2τ)+671224798731250000000cos(2τ)3+308503714132000000000cos(2τ)2−4636311713284232532000000000000000000cos(2τ)−308503714164000000000,η6(τ)≈12801105171875000000cos(τ)2−41067986718750000000sin(τ)cos(τ)−41746419793750000cos(τ)4+3086299153313125000000cos(τ)3sin(τ)−461024269218750000sin(τ)cos(τ)5+590986371156250000cos(τ)6+277713630715000000000;ξ7(τ)≈−361599192063175000000000sin(τ)cos(τ)6+1874076938759875000000000sin(τ)cos(τ)4−353627379445110500000000000cos(τ)2sin(τ)−3888329829715250000000000sin(τ)−133764712573175000000000cos(τ)7+1946738427715625000000cos(τ)5−78535734073782509100000000000000000cos(τ)3+2220868583350000000000cos(τ);η7(τ)≈−301975471187500000cos(τ)7−426272819375000sin(τ)cos(τ)6+850629111250000000cos(τ)5+63377482731500000000sin(τ)cos(τ)4−21680778471000000000cos(τ)3−687528331250000cos(τ)2sin(τ)+45629687777120000000000cos(τ)−23030459198000000000sin(τ),ξ8(τ)≈−37555522949250000000000cos(2τ)4+1480000000000000000000[25608968881440000000sin(2τ)−36609693930465966080]cos(2τ)3+1480000000000000000000[16558606704080000000sin(2τ)+72468445012762533198]cos(2τ)2+1480000000000000000000[35806451743685700900sin(2τ)+25512157348987112880]cos(2τ)−312205824434843172400000000000000000sin(2τ)−3064748661033755533160000000000000000000,η8(τ)≈−37823585363196875000000cos(τ)2+8200139303113150000000000−978700018145000000000sin(τ)cos(τ)−1413892642759393750000000cos(τ)4+40384900212250000000cos(τ)3sin(τ)+52263332094921875000sin(τ)cos(τ)7−594330709234375000sin(τ)cos(τ)5−37234450370312500cos(τ)8+438072199134921875000cos(τ)6. |
[1] |
Y. Z. Pei, L. S. Chen, Q. R. Zhang, C. G. Li, Extinction and performance of one-prey multi-predators of Holling type Ⅱ function response system with impulsive biological control, J. Theor. Biol., 235 (2005), 495–503. doi: 10.1016/j.jtbi.2005.02.003
![]() |
[2] |
C. S. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, Mem. Entomol. Soc. Canada, 97 (1965), 3–60. doi: 10.4039/entm9741fv
![]() |
[3] |
S. B. Hsu, T. Wei, Y. Kuang, Global analysis of the Michaelis-Menten-type ratio-dependent predator-prey system, J. Math. Biol., 42 (2001), 489–506. doi: 10.1007/s002850100079
![]() |
[4] |
P. Misha, S. N. Raw, Dynamical complexities in a predator-prey system involving teams of two prey and one predator, J. Appl. Math. Comput., 61 (2019), 1–24. doi: 10.1007/s12190-018-01236-9
![]() |
[5] |
J. C. Huang, S. G. Ruan, J. Song, Bifurcation in a predator-prey system of Leslie type with generalized Holling type Ⅲ functional response, J. Differ. Equations, 257 (2014), 1721–1752. doi: 10.1016/j.jde.2014.04.024
![]() |
[6] |
J. Andrews, A mathematical model for the continuous culture of microorganisms utilizing inhibitory substrates, Biotechnol. Bioeng., 10 (1968), 707–723. doi: 10.1002/bit.260100602
![]() |
[7] |
S. N. Raw, P. Mishra, Modeling and analysis of inhibitory effect in plankton-fish model: application to the hypertrophic Swarzedzkie lake in eestern Poland, Nonlinear Anal. Real World Appl., 46 (2019), 465–492. doi: 10.1016/j.nonrwa.2018.09.026
![]() |
[8] |
Y. L. Li, D. M. Xiao, Bifurcations of a predator-prey system of Holling and Leslie types, Chaos Solitons Fractals, 34 (2007), 606–620. doi: 10.1016/j.chaos.2006.03.068
![]() |
[9] |
W. Sokol, J. A. Howell, Kinetics of phenol oxidation by washed cells, Biotechnol. Bioeng., 23 (1981), 2039–2049. doi: 10.1002/bit.260230909
![]() |
[10] |
D. L. DeAngelis, R. A. Goldstein, R. V. O'Neill, A model for trophic interaction, Ecology, 56 (1975), 881–892. doi: 10.2307/1936298
![]() |
[11] |
J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 44 (1975), 331–340. doi: 10.2307/3866
![]() |
[12] |
M. Fan, Y. Kuang, Dynamics of a nonautonomous predator-prey system with the Beddington-DeAngelis functional response, J. Math. Anal. Appl., 295 (2004), 15–39. doi: 10.1016/j.jmaa.2004.02.038
![]() |
[13] |
P. J. Pal, P. K. Mandal, Birfucation analysis of a modified Leslie-Gower predator-prey model with Beddington-DeAngelis functional response and strong Allee-effect, Math. Comput. Simul., 97 (2014), 123–146. doi: 10.1016/j.matcom.2013.08.007
![]() |
[14] |
Y. Zhang, S. J. Gao, K. G. Fan, Q. Y. Wang, Asymptotic behavior of a non-autonomous predator-prey model with Hassell-Varley type functional response and random perturbation, J. Appl. Math. Comput., 49 (2015), 573–594. doi: 10.1007/s12190-014-0854-6
![]() |
[15] |
K. H. Kyung, B. Hunki, The dynamical complexity of a predator-prey system with Hassell-Varley functional response and impulsive effect, Math. Comput. Simul., 94 (2013), 1–14. doi: 10.1016/j.matcom.2013.05.011
![]() |
[16] |
S. B. Hsu, T. W. Hwang, Y. Kuang, Global dynamics of a predator-prey model with Hassell-Varley type functional response, Discrete Contin. Dyn. Syst. B, 10 (2008), 857–871. doi: 10.3934/dcdsb.2008.10.857
![]() |
[17] | K. Wang, Periodic solutions to a delayed predator-prey model with Hassell-Varley type functional response, J. Comput. Appl. Math., 12 (2011), 137–145. |
[18] | F. Rao, S. J. Jiang, Y. Q. Li, H. Liu, Stochastic analysis of a Hassell-Varley type predation model, Abstr. Appl. Anal., 2013 (2013), 1–10. |
[19] | J. P. Tripathi, V. Tiwari, A delayed non-autonomous predator-prey model with Crowley-Martin functional response, International Conference on Mathematics and Computing, 2018. Available from: https://link.springer.com/chapter/10.1007/978-981-13-0023-3_16. |
[20] |
J. L. Ren, L. P. Yu, S. F. Siegmund, Bifurcations and chaos in a discrete predator-prey model with Crowley-Martin functional response, Nonlinear Dyn., 90 (2017), 19–41. doi: 10.1007/s11071-017-3643-6
![]() |
[21] |
B. Dubey, S. H. Agarwal, A. Kumar, Optimal harvesting policy of a prey-predator model with Crowley-Martin-type functional response and stage structure in the predator, Nonlinear Anal. Modell. Control, 23 (2018), 493–514. doi: 10.15388/NA.2018.4.3
![]() |
[22] |
S. B. Li, J. H. Wu, Y. Y. Dong, Uniqueness and stability of a predator-prey model with C-M functional response, Comput. Math. Appl., 69 (2015), 1080–1095. doi: 10.1016/j.camwa.2015.03.007
![]() |
[23] | A. D. Bazykin, Nonlinear Dynamics of Interacting Populations, Singapore World Scientific, 1998. |
[24] | A. D. Bazykin, Structural and Dynamic Stability of Model Predator-Prey Systems, International Institute for Applied Systems Analysis, 1976. |
[25] | H. I. Freedman, Deterministic Mathematical Models in Population Ecology, Marcel Dekker, 1980. |
[26] |
W. Metzler, W. Wischniewsky, Bifurcations of equalibria in Bazykin's predator-prey model, Math. Modell., 6 (1985), 111–123. doi: 10.1016/0270-0255(85)90003-X
![]() |
[27] |
Y. Q. Wang, Z. J. Jing, K. Y. Chan, Multiple limit cycles and global stability in predator prey model, Acta Math. Appl. Sin., 15 (1999), 206–219. doi: 10.1007/BF02720497
![]() |
[28] |
H. I. Freedman, Stability analysis of a predator prey system with mutual interference and density dependent death rate, Bull. Math. Biol., 41 (1979), 67–78. doi: 10.1016/S0092-8240(79)80054-3
![]() |
[29] |
J. Hainzl, Stability and Hopf bifurcation in a predator-prey system with several parameters, SIAM J. Appl. Math., 48 (1988), 170–190. doi: 10.1137/0148008
![]() |
[30] |
N. D. Kazarinoff, P. Van Den Driessche, A model predator-prey system with functional response, Math. Biosci., 39 (1978), 125–134. doi: 10.1016/0025-5564(78)90031-7
![]() |
[31] |
X. X. Qiu, H. B. Xiao, Qualitative analysis of Holling type Ⅱ predator-prey systems with prey refuges and predator restricts, Nonlinear Anal. Real World Appl., 14 (2013), 1896–1906. doi: 10.1016/j.nonrwa.2013.01.001
![]() |
[32] |
M. Lu, J. C. Huang, Global analysis in Bazykin's model with Holling Ⅱ functional response and predator competition, J. Differ. Equations, 280 (2021), 99–138. doi: 10.1016/j.jde.2021.01.025
![]() |
[33] | G. Birkhoff, G. C. Rota, Ordinary Differential Equations Introductions to Higher Mathematics, Ginn and Company, 1962. |
[34] |
F. D. Chen, On a Nonlinear Nonautonomous predator-prey model with diffusion and distributed delay, J. Comput. Appl. Math., 180 (2005), 433–495. doi: 10.1016/j.cam.2004.11.011
![]() |
[35] | Z. F. Zhang, T. R. Ding, W. Z. Huang, Z. X. Dong, Qualitative Theory of Differential Equations, Science Press, 1992. |
[36] | L. Perko, Differential Equations and Dynamical Systems, Springer-Verlag, 2001. |
[37] | S. T. Wang, H. G. Yu, Complexity analysis of a modified predator-prey System with Beddington-DeAngelis functional response and Allee-like effect on predator, Discrete Dyn. Nature Soc., 2021 (2021), 1–18. |
[38] | J. C. Huang, Y. J. Gong, J. Chen, Multiple bifurcations in a predator-prey system of Holling and Leslie type with constant-yield prey harvesting, Int. J. Bifurcation Chaos, 23 (2013), 1–24. |
[39] |
B. Tang, Y. N. Xiao, Bifurcation analysis of a predator-prey model with anti-predator behavior, Chaos Solitons Fractals, 70 (2015), 58–68. doi: 10.1016/j.chaos.2014.11.008
![]() |
[40] |
J. H. Shen, H. X. Chen, Z. Y. Zhou, S. H. Chen, Approximation of limit cycles in two-dimensional nonlinear systems near a Hopf bifurcation by canonical transformations, J. Eng. Math., 92 (2015), 185–202. doi: 10.1007/s10665-014-9762-x
![]() |
[41] |
D. Viswanath, The Lindstedt-Poincare technique as an algorithm for computing periodic orbits, SIAM Rev., 43 (2001), 478–495. doi: 10.1137/S0036144500375292
![]() |
1. | Danyang Li, Hua Liu, Xiaotao Han, Xiaofen Lin, Yumei Wei, Stability and Bifurcation Analysis of Bazykin’s Model with Holling I Functional Response and Allee Effect, 2022, 32, 0218-1274, 10.1142/S0218127422502480 | |
2. | Amina Hammoum, Tewfik Sari, Karim Yadi, The Rosenzweig–MacArthur Graphical Criterion for a Predator-Prey Model with Variable Mortality Rate, 2023, 22, 1575-5460, 10.1007/s12346-023-00739-6 | |
3. | Zina Kh. Alabacy, Azhar A. Majeed, The local bifurcation analysis of two preys stage-structured predator model with anti-predator behavior, 2022, 2322, 1742-6588, 012061, 10.1088/1742-6596/2322/1/012061 | |
4. | Shuangte Wang, Hengguo Yu, Chunrui Zhang, Equilibria and Bogdanov-Takens Bifurcation Analysis in the Bazykin’s Predator-Prey System, 2022, 2022, 1607-887X, 1, 10.1155/2022/4844228 | |
5. | R. P. Gupta, Shristi Tiwari, Arun Kumar, A Comprehensive Study of Bifurcations in an Interactive Population Model with Food-Limited Growth, 2024, 0971-3514, 10.1007/s12591-023-00672-9 | |
6. | Abdul Qadeer Khan, Discrete Bazykin’s Prey–Predator Model with Stability, Control and Bifurcation, 2023, 47, 2731-8095, 1191, 10.1007/s40995-023-01472-0 | |
7. | Soumik Pandey, Debashis Das, Uttam Ghosh, Sarbani Chakraborty, Bifurcation and onset of chaos in an eco-epidemiological system with the influence of time delay, 2024, 34, 1054-1500, 10.1063/5.0177410 | |
8. | Muhammad Aqib Abbasi, Periodic behavior and dynamical analysis of a prey–predator model incorporating the Allee effect and fear effect, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-04909-6 | |
9. | Prajjwal Gupta, Satyabhan Singh, Anupam Priyadarshi, Decoding the role of prey-refuge in food-web systems as stabilizing or destabilizing factor through the analysis of higher-dimensional food-web model, 2025, 18, 1874-1738, 10.1007/s12080-025-00608-9 | |
10. | Chen Wang, Ruizhi Yang, Hopf bifurcation analysis of a pine wilt disease model with both time delay and an alternative food source, 2025, 33, 2688-1594, 2815, 10.3934/era.2025124 | |
11. | HASAN S. PANIGORO, EMLI RAHMI, EBENEZER BONYAH, ALI AKGÜL, SAYOOJ ABY JOSE, DYNAMICAL BEHAVIORS OF A FRACTIONAL-ORDER PREDATOR–PREY MODEL: INSIGHTS INTO MULTIPLE PREDATORS COMPETING FOR A SINGLE PREY, 2025, 33, 0218-348X, 10.1142/S0218348X25400778 | |
12. | Gang Wang, Ming Yi, Zaiyun Zhang, Global Dynamics of a Predator–Prey System with Variation Multiple Pulse Intervention Effects, 2025, 13, 2227-7390, 1597, 10.3390/math13101597 |