Conditions | Existence of fixed points | |
h>14 | nonexistence | |
h=14 | A(12,0) | |
0<h<14 | a>a0 | B(1−√1−4h2,0),C(1+√1−4h2,0) |
a=a0 | B,C,E1(1+√1−3h3,c(1−√1−3h)h(b−c)) | |
a<a0 | B,C,E2(xA,c(b−c)xA),E3(xB,c(b−c)xB) |
Using the forward Euler method, we derive a discrete predator-prey system of Gause type with constant-yield prey harvesting and a monotonically increasing functional response in this paper. First of all, a detailed study for the existence and local stability of fixed points of the system is obtained by invoking an important lemma. Mainly, by utilizing the center manifold theorem and the bifurcation theory some sufficient conditions are obtained for the saddle-node bifurcation and the flip bifurcation of this system to occur. Finally, with the use of Matlab software, numerical simulations are carried out to illustrate the theoretical results obtained and reveal some new dynamics of the system-chaos occuring.
Citation: Jiange Dong, Xianyi Li. Bifurcation of a discrete predator-prey model with increasing functional response and constant-yield prey harvesting[J]. Electronic Research Archive, 2022, 30(10): 3930-3948. doi: 10.3934/era.2022200
[1] | Xianyi Li, Xingming Shao . Flip bifurcation and Neimark-Sacker bifurcation in a discrete predator-prey model with Michaelis-Menten functional response. Electronic Research Archive, 2023, 31(1): 37-57. doi: 10.3934/era.2023003 |
[2] | Zhuo Ba, Xianyi Li . Period-doubling bifurcation and Neimark-Sacker bifurcation of a discrete predator-prey model with Allee effect and cannibalism. Electronic Research Archive, 2023, 31(3): 1405-1438. doi: 10.3934/era.2023072 |
[3] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[4] | Jie Xia, Xianyi Li . Bifurcation analysis in a discrete predator–prey model with herd behaviour and group defense. Electronic Research Archive, 2023, 31(8): 4484-4506. doi: 10.3934/era.2023229 |
[5] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[6] | Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128 |
[7] | Yuan Tian, Yang Liu, Kaibiao Sun . Complex dynamics of a predator-prey fishery model: The impact of the Allee effect and bilateral intervention. Electronic Research Archive, 2024, 32(11): 6379-6404. doi: 10.3934/era.2024297 |
[8] | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150 |
[9] | Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262 |
[10] | San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045 |
Using the forward Euler method, we derive a discrete predator-prey system of Gause type with constant-yield prey harvesting and a monotonically increasing functional response in this paper. First of all, a detailed study for the existence and local stability of fixed points of the system is obtained by invoking an important lemma. Mainly, by utilizing the center manifold theorem and the bifurcation theory some sufficient conditions are obtained for the saddle-node bifurcation and the flip bifurcation of this system to occur. Finally, with the use of Matlab software, numerical simulations are carried out to illustrate the theoretical results obtained and reveal some new dynamics of the system-chaos occuring.
With the continuous development of human society and the continuous progress of civilization, resource consumption and environmental pollution are being increased day by day, and human beings have also been punished by nature, such as frequent occurrences of natural disasters, viruses wreak havoc, etc. So it is very important to find strategies to deal with environmental problems. Mathematical modelling is a force tool to reveal the changing trend of natural environment. More and more scholars use mathematical methods to study ecological balance problem.
Generally speaking, the classical predator-prey model has the following structure:
{dxdt=f(x)x−g(x,y)y,dydt=ϵg(x,y)y−μy, | (1.1) |
where x(t) and y(t) represent the population densities of prey and predator in time t respectively, f(x) is the net growth rate of prey without predator, g(x,y) is the consumption rate of prey by predator, ϵ and μ are the positive constants respectively representing the conversion rate of captured prey into predator and the mortality of predator. In order to show the crowding effect, when the prey is large, the prey growth rate f(x) in model (1.1) is usually a negative value. The most famous example of xf(x) is the logistic form:
xf(x)=rx(1−xK), | (1.2) |
among them, the positive constants r and K respectively represent the inherent growth rate of the prey and the carrying capacity of environment to the prey without the predator. In this paper, we assume that xf(x) takes the logistic form given by above (1.2). Consequently, model (1.1) reads as
{dxdt=rx(1−xK)−g(x,y)y,dydt=ϵg(x,y)y−μy. | (1.3) |
The behavioral characteristics of the predator species can be reflected by the key element g(x,y), called functional response or nutritional function. Ultimately, the functional response plays an important role in determining different dynamical behaviors, such as steady state, oscillation, bifurcation and chaos phenomenon [1]. The functional response g(x,y) in population dynamics (and other disciplines) has several traditional forms:
(ⅰ) g(x,y) depends on x only (meaning g(x,y)=g(x)).
g(x)=mx;
g(x)=mxa+x;
⋄ Holling type Ⅲ [9,10,11,12]:
g(x)=mx2a+x2;
⋄ Holling type Ⅳ [13,14,15,16]:
g(x)=mxa+x2.
(ⅱ) p(x,y) depends on both x and y.
⋄ Ratio-dependent type [17]:
g(x,y)=mxx+ay;
⋄ Beddington-DeAngelis type [18,19]:
g(x,y)=mxa+bx+cy;
⋄ Hassell-Varley type [20,21]:
g(x,y)=mxyγ+ax,γ=12,13.
The parameters m, a, b and c in the above formulas are all positive constants, and they have different biological meanings in different formulas. In order to propose a functional response to show how a group of predators (for example, a group of tuna) search, contact and then hunt a prey or a group of preys, several biological hypotheses were proposed. Based on these assumptions and the logic of Holling [22], the hunting cooperation proposed by Cosner, DeAngelis, Ault and Olson [23] has a special functional response, as shown below:
g(x,y)=Ce0xy1+hCe0xy. | (1.4) |
here, C is the score of the prey killed in each encountering each predator, e0 is the total encountering coefficient between the predator and the prey, and h is the processing time of each prey. It's monotonous. Ryu, Ko and Haque [24] introduced this reaction into model (1.3) and obtained the following system:
{dxdt=rx(1−xK)−Ce0xy1+hCe0xyy,dydt=ϵCe0xy1+hCe0xyy−μy. | (1.5) |
A common phenomenon in the predator-prey model is called cooperative hunting between predators. This phenomenon makes the encountering rate between the predator and the prey change with the number of predators [25,26,27,28,29,30]. However, when encountering a gathering of prey, there may be extreme phenomena leading to the eventual extinction of the predator. Therefore, Shang, Qiao, Duan and Miao [31] added the constant yield harvest H to the first equation of the model (1.5) to study the arrangement of renewable resources that ensures the coexistence of two species. By the transformations ¯t=rt, ¯x=xK, ¯y=hCe0Ky, a=1Ce0h2K2r, b=ϵrh, c=μr and ¯h=HrK, and dropping the bars in the above alphabets, we get the following predator-prey system:
{dxdt=x(1−x)−axy21+xy−h,dydt=bxy21+xy−cy. | (1.6) |
In the system (1.6), we assume that the initial values (x0,y0) are positive to ensure that its solution is positive. Obviously, it is very difficult and complicated to directly find an exact solution of the system (1.6), so we consider to find its approximate solution. This motivates us to study the dynamical properties for the discretization version of the system (1.6).
For a given system, there are many discretization methods, including the forward Euler method, the backward Euler method, semi-discretization, and so on. In this paper, we use the forward Euler method to derive the discrete form of the system (1.6). Applying the forward Euler method to the system (1.6), one has
{xn+1−xnδ=xn(1−xn)−axny2n1+xnyn−h,yn+1−ynδ=bxny2n1+xnyn−cyn, | (1.7) |
which is written as a map to get the followng system
F:(xy)⟶(x+δx(1−x)−δaxy21+xy−δhy+δy(bxy1+xy−c)), | (1.8) |
where δ is the step size, and a, b, c, h all are positive constants.
The outline of this paper is as follows: In Section 2, we investigate the existence and stability of fixed points of the system (1.8). In Section 3, we use the central manifold theorem and the bifurcation theory to derive some sufficient conditions that ensure the flip bifurcation and saddle-node bifurcation of the system (1.8) to occur. In Section 4, numerical simulation results are provided to not only support theoretical analysis derived but also illustrate new and rich dynamical behaviors of this system.
Before we analyze the fixed points of the system (1.8), we recall the following lemma [32,33].
Lemma 1.1. Let F(λ)=λ2+Bλ+C, where B and C are two real constants. Suppose λ1 and λ2 are two roots of F(λ)=0. Then the following statements hold.
(i) If F(1)>0, then
(i.1) |λ1|<1 and |λ2|<1 if and only if F(−1)>0 and C<1;
(i.2) λ1=−1 and λ2≠−1 if and only if F(−1)=0 and B≠2;
(i.3) |λ1|<1 and |λ2|>1 if and only if F(−1)<0;
(i.4) |λ1|>1 and |λ2|>1 if and only if F(−1)>0 and C>1;
(i.5) λ1 and λ2 are a pair of conjugate complex roots and, |λ1|=|λ2|=1 if and only if −2<B<2 and C=1;
(i.6) λ1=λ2=−1 if and only if F(−1)=0 and B=2.
(ii) If F(1)=0, namely, 1 is one root of F(λ)=0, then the another root
λ satisfies |λ|=(<,>)1 if and only if |C|=(<,>)1.
(iii) If F(1)<0, then F(λ)=0 has one root lying in (1,∞). Moreover,
(iii.1) the other root λ satisfies λ<(=)−1 if and only if F(−1)<(=)0;
(iii.2) the other root −1<λ<1 if and only if F(−1)>0.
In this section, we first consider the existence of fixed points of the system (1.8) and then analyze the local stability of these fixed points.
The fixed points of the system (1.8) satisfy the following equations
{x=x+δx(1−x)−δaxy21+xy−δh,y=y+δy(bxy1+xy−c), | (2.1) |
namely,
{x(1−x)−axy21+xy−h=0,y(bxy1+xy−c)=0. | (2.2) |
Considering the biological meanings of the system (1.8), one only takes into account its nonnegative fixed points. Corresponding analysis is as follows:
(ⅰ) if h>14, then x(1−x)−axy21+xy−h<0 for all nonnegative x and y, hence the system (1.8) has no equilibria;
(ⅱ) if h=14, then x(1−x)−axy21+xy−h=0 if and only if x=12 and y=0, so, the system (1.8) has a unique predator free equilibrium A(12,0);
(ⅲ) if 0<h<14, then the system (1.8) has two boundary equilibria B(1−√1−4h2,0) and C(1+√1−4h2,0), and some positive equilibria may take place. Next, we further analyse this case.
If the system (1.8) has a positive equilibrium, denoted as (x,y), then following (2.2) we have
{x3−x2+hx+ac2b(b−c)=0,y=c(b−c)x, | (2.3) |
where 0<x<1, b>c and 0<h<14.
Let
f(x)=x3−x2+hx+ac2b(b−c), |
then
f′(x)=3x2−2x+h. |
Since 0<h<14, f′(x) has two unequal real roots x1=1−√1−3h3 and x2=1+√1−3h3 in the interval (0,1). On the other hand, we can see that 0<f(0)<f(1) and 0<x1<x2<1, and that f(x) is increasing for x∈(0,x1)∪(x2,1) and decreasing for x∈(x1,x2).
For the sake of convenient discussion later, let
a0=b(b−c)[2−9h+(2−6h)√1−3h]27c2. | (2.4) |
It is easy to compute f(x2)=c2(a−a0)b(b−c). So, we have the following results about the positive real roots x∈(0,1) of f(x)=0:
(ⅰ) if a>a0, then f(x2)>0, hence f(x) has no positive real root in (0,1)⇒ the system (1.8) has no positive equilibria;
(ⅱ) if a=a0, then f(x2)=0, hence f(x) has one real root x2 in (0,1), and it is a double root⇒ the system (1.8) possesses a unique positive equilibrium E1(1+√1−3h3,c(1−√1−3h)h(b−c));
(ⅲ) if a<a0, then f(x2)<0, hence f(x) has two positive roots xA and xB in (0,1), and 0<x1<xA<x2<xB<1⇒ the system (1.8) has two positive equilibria E2(xA,c(b−c)xA) and E3(xB,c(b−c)xB).
Summarizing the above discussions, we obtain the following result.
Theorem 2.1. Consider the system (1.8). Suppose a0 is defined in (2.4). The existence conditions for all nonnegative fixed points ofthe system (1.8) are summarized in the Table 1.
Conditions | Existence of fixed points | |
h>14 | nonexistence | |
h=14 | A(12,0) | |
0<h<14 | a>a0 | B(1−√1−4h2,0),C(1+√1−4h2,0) |
a=a0 | B,C,E1(1+√1−3h3,c(1−√1−3h)h(b−c)) | |
a<a0 | B,C,E2(xA,c(b−c)xA),E3(xB,c(b−c)xB) |
Now we begin to analyze the stability of these fixed points. The Jacobian matrix J of the system (1.8) at a fixed point E(x,y) is presented as follows:
J(E)=(1+δ(1−2x−ay2(1+xy)2)−axyδ(2+xy)(1+xy)2by2δ(1+xy)21+δ(bxy(2+xy)(1+xy)2−c)), | (2.5) |
and the characteristic equation of Jacobian matrix J(E) can be written as
λ2+p(x,y)λ+q(x,y)=0, | (2.6) |
where
p(x,y)=−2−δ(1−c−2x+bxy(2+xy)−ay2(1+xy)2),q(x,y)=1+δ(1−c−2x+bxy(2+xy)−ay2(1+xy)2)+δ2(acy2(1+xy)2+(1−2x)(bxy(2+xy)(1+xy)2−c)). |
For the stability of fixed points A(12,0), B(1−√1−4h2,0) and C(1+√1−4h2,0), we can easily get the following Theorems 2.2–2.4, respectively.
Theorem 2.2. The fixed point A=(12,0) of the system (1.8) is non-hyperbolic.
Theorem 2.3. For 0<h<14, the boundary fixed point B=(1−√1−4h2,0) ofthe system (1.8) occurs. Moreover, the following statements about the fixed point B are true.
1) If 0<δ<2c, B is a saddle;
2) if δ=2c, B is non-hyperbolic;
3) if δ>2c, B is a source.
Theorem 2.4. For 0<h<14, the boundary fixed point C=(1+√1−4h2,0) ofthe system (1.8) occurs. In addition, the following results in the Table 2 arevalid about the fixed point C.
Conditions | Eigenvalues | Properties | |
λ1=1−δ√1−4h, λ2=1−δc | |||
c<√1−4h | 0<δ<2√1−4h | |λ1|<1, |λ2|<1 | sink |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
2√1−4h<δ<2c | |λ1|>1, |λ2|<1 | saddle | |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c=√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|=1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c>√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
2c<δ<2√1−4h | |λ1|<1, |λ2|>1 | saddle | |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
δ>2√1−4h | |λ1|>1, |λ2|>1 | source |
For the stability of the positive equilibrium point E1(1+√1−3h3,c(1−√1−3h)h(b−c)), one will discuss it in the next section.
In this section, we use the central manifold theorem and bifurcation theory to discuss the flip bifurcation and saddle-node bifurcation at the boundary fixed point B and the positive equilibrium point E1 of the system (1.8).
From Theorem (2.3) one can see that, when the parameter δ goes through the critical value δ0=2c, the dimension numbers of stable and unstable manifolds of the system (1.8) at the fixed point B change. A bifurcation will occur. Again, for δ=δ0, one eigenvalue −1 appears. So, at this time, the system may produce a flip bifurcation, which is considered in the following, and δ is chosen as bifurcation parameter. Remember the parameters
(a,b,c,h,δ)∈SE+={(a,b,c,h,δ)∈R5+|0<a,0<h<14,0<c<b,δ>0}. |
Let X=x−xB,Y=y−yB,δ∗=δ−δ0. We transform the fixed point B(xB,yB) to the origin and consider the parameter δ∗ as a new independent variable. Thus, the system (1.8) becomes
(XYδ∗)→(X+(δ∗+δ0)(X+xB)[1−(X+xB)−a(Y+yB)21+(X+xB)(Y+yB)]−(δ∗+δ0)hY+(Y+yB)(δ∗+δ0)(b(X+xB)(Y+yB)1+(X+xB)(Y+yB)−c)δ∗). | (3.1) |
Taylor expanding of the system (3.1) at (X,Y,δ∗)=(0,0,0) takes the form:
{Xn+1=a100Xn+a010Yn+a001δ∗n+a200X2n+a020Y2n+a002δ∗n2+a110XnYn+a101Xnδ∗n+a011Ynδ∗n+a300X3n+a030Y3n+a003δ∗n3+a210X2nYn+a201X2nδ∗n+a102Xnδ∗n2+a120XnY2n+a111XnYnδ∗n+a012XnY2n+a021Y2nδ∗n+O(ρ41),Yn+1=b100Xn+b010Yn+b001δ∗n+b200X2n+b020Y2n+b002δ∗n2+b110XnYn+b101Xnδ∗n+b011Ynδ∗n+b300X3n+b030Y3n+b003δ∗n3+b210X2nYn+b201X2nδ∗n+b102Xnδ∗n2+b120XnY2n+b111XnYnδ∗n+b012XnY2n+b021Y2nδ∗n+O(ρ41),δ∗n+1=δ∗n, | (3.2) |
where ρ1=√X2n+Y2n+δ∗n2,
a100=1+δ0√1−4h,a200=−2δ0,a020=aδ0(√1−4h−1),a101=√1−4h,a030=32aδ0(1−√1−4h),a201=−2,a120=−2aδ0,a021=a(√1−4h−1),b010=1−cδ0,b020=bδ0(1−√1−4h),b011=−c,b030=−32bδ0(1−√1−4h),b120=2bδ0,b021=b(1−√1−4h),a011=a010=a110=a001=a002=a300=a210=a012=a003=a102=a111=b100=b001=b200=b002=b110=b101=b300=b003=b210=b201=b102=b012=b111=0. |
Namely, the system (3.2) is equivalent to the following form:
(XYδ∗)→(1+δ0√1−4h0001−cδ00001)(XYδ∗)+(F1(X,Y,δ∗)F2(X,Y,δ∗)0), | (3.3) |
where
F1(X,Y,δ∗)=−2δ0X2+aδ0(√1−4h−1)Y2+√1−4hXδ∗+32aδ0(1−√1−4h)2Y3−2X2δ∗−2aδ0XY2+a(√1−4h−1)Y2δ∗+O(ρ41),F2(X,a∗,Y)=bδ0(1−√1−4h)Y2−cYδ∗−32bδ0(1−√1−4h)2Y3+2bδ0XY2+b(1−√1−4h)Y2δ∗+O(ρ41). |
By the center manifold theorem, the stability of (X,Y)=(0,0) near δ∗=0 can be determined by studying a one-parameter family of map on a center manifold, which can be written as:
Wc(0)={(X,Y,δ∗)∈R3|X=h∗1(Y,δ∗),h∗1(0,0)=0,Dh∗1(0,0)=0}. |
Assume that h∗1(Y,δ∗) has the following form:
h∗1(Y,δ∗)=b∗20Y2+b∗11Yδ∗+b∗02δ∗2+O(ρ33), |
where ρ3=√Y2+δ∗2. Then the center manifold equation must satisfy
h∗1(−Y+F2(h∗1(Y,δ∗),Y,δ∗),δ∗)=(1+δ0√1−4h)h∗1(Y,δ∗)+F1(h∗1(Y,δ∗),Y,δ∗). |
Comparing the corresponding coefficients of terms with the same orders in the above center manifold equation, we get
b∗20=a(1−√1−4h)√1−4h,b∗11=b∗02=0. |
Thus the system (3.3) restricted to the center manifold is given by
F:Y→−Y+bδ0(1−√1−4h)Y2−cYδ∗−32bδ0(1−√1−4h)2Y3+b(1−√1−4h)Y2δ∗+O(ρ43), |
and
F2:Y→Y+cYδ∗+(3−2bδ0)bδ0(1−√1−4h)2Y3+4b(1−√1−4h)Y2δ∗+O(ρ43). |
Therefore, one has
F(Y,δ∗)|(0,0)=0,∂F∂Y|(0,0)=−1,∂F2∂δ∗|(0,0)=0,∂2F2∂Y∂δ∗|(0,0)=c, |
∂2F2∂Y2|(0,0)=0,∂3F2∂Y3|(0,0)=6(3−2bδ0)bδ0(1−√1−4h)2. |
According to [34], if the nondegenerency conditions ∂3F2∂Y3|(0,0)≠0 and ∂2F2∂Y∂δ∗|(0,0)≠0 hold, then the system (1.8) undergoes a flip bifurcation. Obviously, they hold. Therefore, the following result may be derived.
Theorem 3.1. Assume the parameters (a,b,c,h,δ)∈SE+={(a,b,c,h,δ)∈R5+|0<a,0<h<14,0<c<b,δ>0}. Let δ0=2c, then the system (1.8) undergoes a flip bifurcation at B(1−√1−4h2,0) when the parameter δ varies in a small neighborhood of the critical value δ0.
In the next one considers the saddle-node bifurcation of the system (1.8) at the positive fixed point E1(x0,y0), where a is chosen as bifurcation parameter. The characteristic equation of Jacobian matrix J of the system (1.8) at the positive fixed point E1(x0,y0) is presented as
f(λ)=λ2+p(x0)λ+q(x0)=0, | (3.4) |
where x0=1+√1−3h3, y0=c(b−c)x0, 0<h<14 and b>c, p(x0) and q(x0) are given by
p(x0)=−2−δ(1−2x0−ac2b2x20+c(b−c)b),q(x0)=1+δ[1−2x0+c(b−c)b−ac2b2x20]. |
Notice f(1)=0 always holds. So, λ1=1 is a root of f(λ)=0. If
√1−3h≠δ(3c(b−c)+2b−c)−3nb2δ(2b−c),n=0,−2, | (3.5) |
then another eigenvalue of the fixed point E1(x0,y0) satisfies
λ2=1+δ(1+c−2x0−c2b−a0c2b2x20),and|λ2|≠1. |
As this time, the system may produce a fold bifurcation, which is considered in the following.
Let X=x−x0,Y=y−y0,a∗=a−a0, which transform the fixed point (x0,y0) to the origin. Consider the parameter a∗ as a new independent variable, then the system (1.8) becomes
(Xa∗Y)→(X+δ(X+x0)[1−(X+x0)]−δ(a∗+a0)(X+x0)(Y+y0)21+(X+x0)(Y+y0)−δha∗Y+δ(Y+y0)(b(X+x0)(Y+y0)1+(X+x0)(Y+y0)−c)). | (3.6) |
Taylor expanding of the system (3.6) at (X,a∗,Y)=(0,0,0) obtains
{Xn+1=a100Xn+a010a∗n+a001Yn+a200X2n+a002Y2n+a110Xna∗n+a101XnYn+a011a∗nYn+a300X3n+a210X2na∗n+a201X2nYn+a102XnY2n+a111Xna∗nYn+a012a∗nY2n+a003Y3n+O(ρ41),a∗n+1=a∗n,Yn+1=b100Xn+b001Yn+b200X2n+b101XnYn+b002Y2n+b300X3n+b201X2nYn+b102XnY2n+b003Y3n+O(ρ∗14), | (3.7) |
where ρ∗1=√X2n+a∗2n+Y2n,
a100=1+δ(1−2x0)−δa0y20(1+x0y0)2,a010=−δx0y201+x0y0,a001=−δa0x0y0(2+x0y0)(1+x0y0)2,a200=−δ+δa0y30(1+x0y0)3,a002=−δa0x0(1+x0y0)3,a110=−δy20(1+x0y0)2,a101=−2δa0y0(1+x0y0)3,a011=−δx0y0(2+x0y0)(1+x0y0)2,a300=−δa0y40(1+x0y0)4,a003=δa0x20(1+x0y0)4,a210=δy30(1+x0y0)3,a201=3δa0y20(1+x0y0)4,a102=−δa0(1−2x0y0)(1+x0y0)4,a111=−2δy0(1+x0y0)3,a012=−δx0(1+x0y0)3,b100=δby20(1+x0y0)2,b001=1+δ(bx0y01+x0y0+bx0y0(1+x0y0)2−c),b200=−δby30(1+x0y0)3,b002=δbx0(1+x0y0)3,b101=2δby0(1+x0y0)3,b300=δby40(1+x0y0)4,b003=−δbx20(1+x0y0)4,b201=−3δby20(1+x0y0)4,b102=δb(1−2x0y0)(1+x0y0)4,a020=a030=a021=a120=b010=b020=b110=b011=b030=b210=b120=b111=b012=b021=0. |
Then the system (3.7) is equivalent to the following form:
(Xa∗Y)→(a11a12a13010a310a33)(Xa∗Y)+(F1(X,a∗,Y)0F2(X,a∗,Y)), | (3.8) |
where
a11=a100,a12=a010,a13=a001,a31=b100,a33=b001,F1(X,a∗,Y)=a200X2+a020a∗2+a002Y2+a110Xa∗+a101XY+a011a∗Y+a300X3+a210X2a∗+a201X2Y+a120Xa∗2+a102XY2+a111Xa∗Y+a030a∗3+a021a∗2Y+a003Y3+O(ρ41),F2(X,a∗,Y)=b200X2+b101XY+b002Y2+b300X3+b201X2Y+b102XY2+b003Y3+O(ρ∗14). |
Assume that
(a11−1)2+a13a31≠0. | (3.9) |
Take
T=(a131−a11a31λ2−a330(1−a11)2+a13a31a12a3101−a110a31), |
then T−1 exists.
Under the transformation
(Xa∗Y)=T(Ua∗1V), |
the system (3.8) becomes
(Ua∗1V)→(11001000λ2)(Ua∗1V)+(g1(U,a∗1,V)0g2(U,a∗1,V)), | (3.10) |
where ρ∗2=√X2+a∗2+Y2,
g1(U,a∗1,V)=j200X2+j002Y2+j110Xa∗+j101XY+j011a∗Y+j300X3+j210X2a∗+j201X2Y+j102XY2+j111Xa∗Y+j003Y3+O(ρ∗24),g2(U,a∗1,V)=k200X2+k002Y2+k110Xa∗+k101XY+k011a∗Y+k300X3+k210X2a∗+k201X2Y+k102XY2+k111Xa∗Y+k003Y3+O(ρ∗24),X=a13U+1−a11a31a∗1+(λ2−a33)V,a∗=−(1−a11)2+a13a31a12a31a∗1,Y=a31(1−a11)U+a31V,j200=δ(a33−1)a13(1−λ2)+δy30(1−λ2)(1+x0y0)3[a0(1−a33)a13−b],j002=δx0(1−λ2)(1+x0y0)3[b−a0(1−a33)a13],j110=δy20(a33−1)a13(1−λ2)(1+x0y0)2,j011=−δx0y0(1−a33)(2+x0y0)a13(1−λ2)(1+x0y0)2,j101=2δy0(1−λ2)(1+x0y0)3[b−a0(1−a33)a13],j300=δy40(1−λ2)(1+x0y0)4[b−a0(1−a33)a13],j210=δy30(1−a33)a13(1−λ2)(1+x0y0)4,j111=−2δy0(1−a33)a13(1−λ2)(1+x0y0)3,j201=3δy20(1−λ2)(1+x0y0)4[a0(1−a33)a13−b],j102=δ(1−2x0y0)(1−λ2)(1+x0y0)4[b−a0(1−a33)a13],j003=δx20(1−λ2)(1+x0y0)4[a0(1−a33)a13−b],k200=δ1−λ2+δy30(λ2−1)(1+x0y0)3[a0−a13ba11−1],k002=δx0(λ2−1)(1+x0y0)3[a13ba11−1−a0],k110=−δy20(λ2−1)(1+x0y0)2,k011=−δx0y0(2+x0y0)(λ2−1)(1+x0y0)2,k101=2δy0(λ2−1)(1+x0y0)3[a13ba11−1−a0],k300=δy40(λ2−1)(1+x0y0)4[a13ba11−1−a0],k210=δy30(λ2−1)(1+x0y0)4,k111=−2δy0(λ2−1)(1+x0y0)3,k201=3δy20(λ2−1)(1+x0y0)4[a0−a13ba11−1],k102=δ(1−2x0y0)(λ2−1)(1+x0y0)4[a13ba11−1−a0],k003=δx20(λ2−1)(1+x0y0)4[a0−a13ba11−1]. |
By the center manifold theorem, the stability of (U,V)=(0,0) near a∗1=0 can be determined by studying a one-parameter family of map on a center manifold, which can be written as:
Wc(0)={(U,a∗1,V)∈R3|V=h∗2(U,a∗1),h∗2(0,0)=0,Dh∗2(0,0)=0}. |
Assume that h∗2(U,a∗1) has the following form:
h∗2(U,a∗1)=c∗20U2+c∗11Ua∗1+c∗02a∗21+O(ρ3∗3), |
where ρ∗3=√U2+a∗21. Then the center manifold equation must satisfy
h∗2(U+a∗1+g1(U,a∗1,h∗2(U,a∗1)),a∗1)=λ2h∗2(U,a∗1)+g2(U,a∗1,h∗2(U,a∗1)). |
Comparing the corresponding coefficients of terms with the same orders in the above center manifold equation, we get
c∗20=a213k200+(1−a11)2k002+a13(1−a11)k1011−λ2,c∗11=a12(1−a11)[2a13k200+(1−a11)k101]a12a31(1−λ2)+[a13a31+(1−a11)2][a13k110+(1−a11)k011]a12a31(1−λ2)−2[a213k200+(1−a11)2k002+a13(1−a11)k101](1−λ2)2,c∗02=(1−a11)[a12(1−a11)k200+[a13a31+(1−a11)2]k110]a12a231(1−λ2)−a213k200+(1−a11)2k002+a13(1−a11)k101(1−λ2)2−a12(1−a11)[2a13k200+(1−a11)k101]a12a31(1−λ2)2−[a13a31+(1−a11)2][a13k110+(1−a11)k011]a12a31(1−λ2)2+2[a213k200+(1−a11)2k002+a13(1−a11)k101](1−λ2)3. |
Thus the system (3.10) restricted to the center manifold is given by
G∗:U→U+a∗1+h20U2+h02a∗21+h11Ua∗1+h30U3+h21U2a∗1+h12Ua∗21+h03a∗31+O(ρ∗44), |
where
h20=a213j200+(1−a11)2j002+a13(1−a11)j101,h02=(1−a11)2j200a231+(1−a11)[a13a31+(1−a11)2]j110a12a231,h11=(1−a11)[2a13j200+(1−a11)j101]a13+[a13a31+(1−a11)2][a13j110+(1−a11)j011]a12a13,h30=2a13c∗20(λ2−a33)j200+2a31c∗20(1−a11)j002+c∗20[a13a31+(1−a11)(λ2−a33)]j101,h21=2(λ2−a33)[a13c∗11+2c∗20(1−a11)a13]j200+2a31c∗11(1−a11)j002+c∗20[a13a31+(1−a11)2][(λ2−a33)j110+a31j011]a12a31+[c∗1(1−a11)+c∗2[a13a31+(1−a11)(λ2−a11)]]j101,h12=2c∗11(1−a11)(λ2−a33)j200a31+2a31c∗02(1−a11)j002+[c∗11(1−a11)+c∗20[a13a31+(1−a11)(λ2−a33)]]j101+c∗11[a13a31+(1−a11)2][(λ2−a33)j110+a31j011]a12a31,h03=c∗02(λ2−a33)[2a13+2(1−a11)a13]j200+c∗11(1−a11)j101+c∗02[a13a31+(1−a11)2][(λ2−a33)j110+a31j011]a12a13. |
Therefore, one has
G∗(U,a∗1)|(0,0)=0,∂G∗∂U|(0,0)=1≠0, |
∂G∗∂a∗1|(0,0)=1≠0,∂2G∗∂U2|(0,0)=2h20. |
If the condition ∂2G∗∂U2|(0,0)≠0 is true, then the system (1.8) undergoes a saddle-node bifurcation [34]. Therefore, we need assume
h20≠0. | (3.11) |
And the following result may be derived.
Theorem 3.2. Consider the system (1.8). Let a0 be defined in (2.4). Set the parameters (a,b,c,h,δ)∈SE+={(a,b,c,h,δ)∈R5+|0<h<14,0<c<b,√1−3h≠δ(3c(b−c)+2b−c)−3nb2δ(2b−c),n=0,−2}.
If the conditions (3.9) and (3.11) hold, then the system (1.8) undergoes a saddle-node bifurcation at E1(1+√1−3h3,c(1−√1−3h)h(b−c)) when the parameter a varies in a small neighborhood of the critical value a0.
Remark. Since the characteristic equation corresponding to the system (3.10) contains double roots λ1=λ3=1, the normal form can not be obtained by known routine method. Here we use a special mathematical skill to find the invertible matrix T.
In this section, we give the bifurcation diagrams of the system (1.8) to illustrate the above theoretical analyses and further reveal some new dynamical behaviors to occur by Matlab software.
First fix the parameter values a=3, b=1.1, c=0.6, h=0.16, let δ∈(2,5) and take the initial values (x0,y0)=(0.2,0) in Figure 1. We can see that there is a stable fixed point for δ∈(2,3.35), and a flip bifurcation occurs at δ0=3.35, eventually, period-double bifurcation to chaos. The fixed point E is unstable when δ>δ0. This agrees to the results stated in Theorem 3.1.
Then fix the parameter values b=0.56, c=0.25, δ=0.187, h=0.12, and vary a in the range (0.18,0.23) with the initial value (x0,y0)=(0.6,1.3) in Figure 2. One can see that there is a stable fixed point for a∈(0.195,0.205), and that a saddle-node bifurcation occurs at a0=0.2. When a<a0 and is increasing to a0, the fixed point E1 is gradually stable. When a>a0, the fixed point E1 is unstable. This agrees to the results stated in Theorem 3.2.
In this paper, toward a discrete-time predator-prey system of Gause type with constant-yield prey harvesting and a monotonically increasing functional response in R2, we investigate its flip bifurcation and saddle-node bifurcation problems. By using the center manifold theorem and the bifurcation theory, one shows that the flip bifurcation and saddle-node bifurcation of the discrete-time system take place.
We finally present numerical simulations, which not only illustrate the theoretical analysis results, but also find some new properties of the system (1.8)-chaos occurring.
One of the highlights in this paper is to skillfully find an invertible transform to derive the normal form of the flip (fold) bifurcation of the system (1.8), and determine the stability of the closed orbit bifurcated, while it is impossible for one to use routine methods because its two characteristic roots are double so that corresponding invertible matrix does not exist.
This work is partly supported by the National Natural Science Foundation of China (61473340), the Distinguished Professor Foundation of Qianjiang Scholar in Zhejiang Province, and the Natural Science Foundation of Zhejiang University of Science and Technology (F701108G14).
The authors declare there is no conflicts of interest.
[1] |
M. Haque, A detailed study of the Beddington-DeAngelis predator-prey model, Math. Biosci., 234 (2011), 1–16. https://doi.org/10.1016/j.mbs.2011.07.003 doi: 10.1016/j.mbs.2011.07.003
![]() |
[2] |
A. Zegeling, R. E. Kooij, Singular perturbations of the Holling Ⅰ predator-prey system with a focus, J. Differ. Equation, 269 (2020), 5434–5462. https://doi.org/10.1016/j.jde.2020.04.011 doi: 10.1016/j.jde.2020.04.011
![]() |
[3] |
S. M. Li, X. L. Wang, X. L. Li, K. l. Wu, Relaxation oscillations for Leslie-type predator-prey model wemith Holling Type Ⅰ response functional function, Appl. Math. Lett., 120 (2021), 1–6. https://doi.org/10.1016/j.aml.2021.107328 doi: 10.1016/j.aml.2021.107328
![]() |
[4] |
B. Liu, Y. J. Zhang, L. S. Chen, Dynamic complexities of a Holling Ⅰ predator-prey model concerning periodic biological and chemical control, Chaos Solitons Fractals, 22 (2004), 123–134. https://doi.org/10.1016/j.chaos.2003.12.060 doi: 10.1016/j.chaos.2003.12.060
![]() |
[5] |
M. Liu, K. Wang, Dynamics of a Leslie-Gower Holling-type ii predator-prey system with levy jumps, Nonlinear Anal., 85 (2013), 204–213. https://doi.org/10.1016/j.na.2013.02.018 doi: 10.1016/j.na.2013.02.018
![]() |
[6] |
Y. Xu, M. Liu, Y. Yang, Analysis of a stochastic two-predators one prey system with modified Leslie-Gower and holling-type Ⅱ schemes, J. Appl. Anal. Comput., 7 (2017), 713–727. https://doi.org/10.1016/j.physa.2019.122761 doi: 10.1016/j.physa.2019.122761
![]() |
[7] |
X. L. Zou, Y. T. Zheng, L. R. Zhang, J. L. Lv, Survivability and stochastic bifurcations for a stochastic Holling type Ⅱ predator-prey model, Commun. Nonlinear Sci. Numer. Simul., 83 (2020), 1–20. https://doi.org/10.1016/j.cnsns.2019.105136 doi: 10.1016/j.cnsns.2019.105136
![]() |
[8] |
M. Lu, J. C. Huang, Global analysis in Bazykins model with Holling Ⅱ functional response and predator competition, J. Differ. Equation, 280 (2021), 99–138. https://doi.org/10.1016/j.jde.2021.01.025 doi: 10.1016/j.jde.2021.01.025
![]() |
[9] |
A. K. Misra, Modeling the depletion of dissolved oxygen due to algal bloom in a lake by taking Holling type-Ⅲ interaction, Appl. Math. Comput., 217 (2011), 8367–8376. https://doi.org/10.1016/j.amc.2011.03.034 doi: 10.1016/j.amc.2011.03.034
![]() |
[10] |
R. Banerjee, P. Das, D. Mukherjee, Stability and permanence of a discrete-time two-prey one-predator system with Holling Type-Ⅲ functional response, Chaos Solitons Fractals, 117 (2018), 240–248. https://doi.org/10.1016/j.chaos.2018.10.032 doi: 10.1016/j.chaos.2018.10.032
![]() |
[11] |
C. Wang, X. Zhang, Heteroclinic and homoclinic orbits for a slow-fast predator-prey model of generalized Holling type Ⅲ, J. Differ. Equation, 267 (2019), 3397–3441. https://doi.org/10.1016/j.jde.2019.04.008 doi: 10.1016/j.jde.2019.04.008
![]() |
[12] |
J. C. Huang, S. G. Ruan, J. Song, Bifurcations in a predator-prey system of Leslie type with generalized Holling type Ⅲ functional response, J. Differ. Equation, 257 (2014), 1721–1752. https://doi.org/10.1016/J.JDE.2014.04.024 doi: 10.1016/J.JDE.2014.04.024
![]() |
[13] |
D. Jyotiska, J. Debaldev, U. R. Kumar, Bifurcation and bio-economic analysis of a prey-generalist predator model with Holling type Ⅳ functional response and nonlinear age-selective prey harvesting, Chaos Solitons Fractals, 122 (2019), 229–235. https://doi.org/10.1016/j.chaos.2019.02.010 doi: 10.1016/j.chaos.2019.02.010
![]() |
[14] |
Y. L. Li, D. M. Xiao, Bifurcations of a predator-prey system of Holling and Leslie types, Chaos Solitons Fractals, 34 (2007), 606–620. https://doi.org/10.1016/j.chaos.2006.03.068 doi: 10.1016/j.chaos.2006.03.068
![]() |
[15] |
S. W. Zhang, F. Y. Wang, L. S. Chen, A food chain model with impulsive perturbations and Holling Ⅳ functional response, Chaos Solitons Fractals, 26 (2005), 855–866. https://doi.org/10.1016/j.chaos.2005.01.053 doi: 10.1016/j.chaos.2005.01.053
![]() |
[16] |
S. G. Ruan, D. M. Xiao, Global analysis in a predator-prey system with nonmonotonic functional response, SIAM J. Appl. Math., 61 (2001), 1445–1472. https://doi.org/10.1137/S0036139999361896 doi: 10.1137/S0036139999361896
![]() |
[17] |
C. A. I. Claudio, A. Pablo, F. Jos, V. H. Peter, Bifurcation analysis of a predator-prey model with predator intraspecific interactions and ratio-dependent functional response, Appl. Math. Comput., 402 (2021), 1–20. https://doi.org/10.1016/j.amc.2021.126152 doi: 10.1016/j.amc.2021.126152
![]() |
[18] |
X. Y. Zou, Q. W. Li, J. L. Lv, Stochastic bifurcations, a necessary and sufficient condition for a stochastic Beddington-DeAngelis predator-prey model, Appl. Math. Lett., 117 (2021), 1–7. https://doi.org/10.1016/j.aml.2021.107069 doi: 10.1016/j.aml.2021.107069
![]() |
[19] |
D. M. Luo, Q. R. Wang, Global dynamics of a Beddington-DeAngelis amensalism system with weak Allee effect on the first species, Appl. Math. Comput., 408 (2021), 1–19. https://doi.org/10.1007/s12190-021-01533-w doi: 10.1007/s12190-021-01533-w
![]() |
[20] |
G. D. Zhang, Y. Shen, Periodic solutions for a neutral delay Hassell-Varley type predator-prey system, Appl. Math. Comput., 264 (2015), 443–452. https://doi.org/10.1016/j.amc.2015.04.110 doi: 10.1016/j.amc.2015.04.110
![]() |
[21] |
D. S. Wang, On a non-selective harvesting prey-predator model with Hassell-Varley type functional response, Appl. Math. Comput., 246 (2014), 678–695. https://doi.org/10.1016/j.amc.2014.08.081 doi: 10.1016/j.amc.2014.08.081
![]() |
[22] |
C. S. Holling, Some characteristics of simple types of predation and parasitism, Can. Entomol., 91 (1959), 385–398. https://doi.org/10.4039/Ent91385-7 doi: 10.4039/Ent91385-7
![]() |
[23] |
C. Cosner, D. L. DeAngelis, J. S. Ault, D. B. Olson, Effects of spatial grouping on the functional response of predators, Theor. Popul. Biol., 56 (1999), 65–75. https://doi.org/10.1006/tpbi.1999.1414 doi: 10.1006/tpbi.1999.1414
![]() |
[24] |
K. Ryu, W. Ko, M. Haque, Bifurcation analysis in a predator-prey system with a functional response increasing in both predator and prey densities, Nonlinear Dyn., 94 (2018), 1639–1656. https://doi.org/10.1007/s11071-018-4446-0 doi: 10.1007/s11071-018-4446-0
![]() |
[25] |
T. A. Micka¨el, F. M. Hilker, Hunting cooperation and Allee effects in predators, J. Theoret. Biol., 419 (2017), 13–22. https://doi.org/10.1016/j.jtbi.2017.02.002 doi: 10.1016/j.jtbi.2017.02.002
![]() |
[26] |
F. Capone, M. F. Carfora, R. De Luca, I. Torcicollo, Turing patterns in a reaction-diffusion system modeling hunting cooperation, Math. Comput. Simul., 165 (2019), 172–180. https://doi.org/10.1016/j.matcom.2019.03.010 doi: 10.1016/j.matcom.2019.03.010
![]() |
[27] |
Y. S. Chow, S. R. J. Jang, H. M. Wang, Cooperative hunting in a discrete predator-prey system, J. Biol. Dyn., 13 (2019), 247–264. https://doi.org/10.1080/17513758.2018.1555339 doi: 10.1080/17513758.2018.1555339
![]() |
[28] |
J. Duarte, C. Janurio, N. Martins, J. Sardanys, Chaos and crises in a model for cooperative hunting: a symbolic dynamics approach, Chaos, 19 (2009), 1–12. https://doi.org/10.1063/1.3243924 doi: 10.1063/1.3243924
![]() |
[29] |
S. Pal, N. Pal, S. Samanta, J. Chattopadhyay, Effect of hunting cooperation and fear in a predator-prey model, Ecol. Complex., 39 (2019), 1–18. https://doi.org/10.1016/j.ecocom.2019.100770 doi: 10.1016/j.ecocom.2019.100770
![]() |
[30] |
N. C. Pati, G. C. Layek, N. Pal, Bifurcations and organized structures in a predator-prey model with hunting cooperation, Chaos Solitons Fractals, 140 (2020), 1–11. https://doi.org/10.1016/j.chaos.2020.110184 doi: 10.1016/j.chaos.2020.110184
![]() |
[31] |
Z. C. Shang, Y. H. Qiao, L. J. Duan, J. Miao, Bifurcation analysis in a predator-prey system with an increasing functional response and constant-yield prey harvesting, Math. Comput. Simul., 190 (2021), 976–1002. https://doi.org/10.1016/j.matcom.2021.06.024 doi: 10.1016/j.matcom.2021.06.024
![]() |
[32] |
W. Li, X. Y. Li, Neimark-Sacker bifurcation of a semi-discrete hematopoiesis model, J. Appl. Anal. Comput., 8 (2018), 1679–1693. https://doi.org/10.11948/2018.1679 doi: 10.11948/2018.1679
![]() |
[33] |
C. Wang, X. Y. Li, Stability and Neimark-Sacker bifurcation of a semi-discrete population model, J. Appl. Anal. Comput., 4 (2014), 419–435. https://doi.org/10.11948/2014024 doi: 10.11948/2014024
![]() |
[34] | S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Second edition, Springer-verlag, New York, 2003. |
1. | Danyang Li, Xianyi Li, Transcritical bifurcation and Neimark-Sacker bifurcation of a discrete predator-prey model with herd behaviour and square root functional response, 2024, 30, 1387-3954, 31, 10.1080/13873954.2024.2304798 | |
2. | Karima Mokni, Mohamed Ch-Chaoui, Bapin Mondal, Uttam Ghosh, Rich dynamics of a discrete two dimensional predator–prey model using the NSFD scheme, 2024, 225, 03784754, 992, 10.1016/j.matcom.2023.09.024 | |
3. | Luyao Lv, Xianyi Li, Stability and Bifurcation Analysis in a Discrete Predator–Prey System of Leslie Type with Radio-Dependent Simplified Holling Type IV Functional Response, 2024, 12, 2227-7390, 1803, 10.3390/math12121803 | |
4. | Jie Xia, Xianyi Li, Bifurcation analysis in a discrete predator–prey model with herd behaviour and group defense, 2023, 31, 2688-1594, 4484, 10.3934/era.2023229 | |
5. | Chen Zhang, Xianyi Li, Dynamics of a Discrete Leslie–Gower Model with Harvesting and Holling-II Functional Response, 2023, 11, 2227-7390, 3303, 10.3390/math11153303 | |
6. | Xianyi Li, Jiange Dong, Complicate dynamical properties of a discrete slow-fast predator-prey model with ratio-dependent functional response, 2023, 13, 2045-2322, 10.1038/s41598-023-45861-2 | |
7. | Xianyi Li, Luyao Lv, Global attractivity of a rational difference equation with higher order and its applications, 2024, 4, 2767-8946, 260, 10.3934/mmc.2024021 | |
8. | Dongmei Chen, Xianyi Li, Complex bifurcation phenomena seized in a discrete ratio-dependent Holling-Tanner predator-prey system, 2024, 2024, 2731-4235, 10.1186/s13662-024-03855-y |
Conditions | Existence of fixed points | |
h>14 | nonexistence | |
h=14 | A(12,0) | |
0<h<14 | a>a0 | B(1−√1−4h2,0),C(1+√1−4h2,0) |
a=a0 | B,C,E1(1+√1−3h3,c(1−√1−3h)h(b−c)) | |
a<a0 | B,C,E2(xA,c(b−c)xA),E3(xB,c(b−c)xB) |
Conditions | Eigenvalues | Properties | |
λ1=1−δ√1−4h, λ2=1−δc | |||
c<√1−4h | 0<δ<2√1−4h | |λ1|<1, |λ2|<1 | sink |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
2√1−4h<δ<2c | |λ1|>1, |λ2|<1 | saddle | |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c=√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|=1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c>√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
2c<δ<2√1−4h | |λ1|<1, |λ2|>1 | saddle | |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
δ>2√1−4h | |λ1|>1, |λ2|>1 | source |
Conditions | Existence of fixed points | |
h>14 | nonexistence | |
h=14 | A(12,0) | |
0<h<14 | a>a0 | B(1−√1−4h2,0),C(1+√1−4h2,0) |
a=a0 | B,C,E1(1+√1−3h3,c(1−√1−3h)h(b−c)) | |
a<a0 | B,C,E2(xA,c(b−c)xA),E3(xB,c(b−c)xB) |
Conditions | Eigenvalues | Properties | |
λ1=1−δ√1−4h, λ2=1−δc | |||
c<√1−4h | 0<δ<2√1−4h | |λ1|<1, |λ2|<1 | sink |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
2√1−4h<δ<2c | |λ1|>1, |λ2|<1 | saddle | |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c=√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|=1, |λ2|=1 | non-hyperbolic | |
δ>2c | |λ1|>1, |λ2|>1 | source | |
c>√1−4h | 0<δ<2c | |λ1|<1, |λ2|<1 | sink |
δ=2c | |λ1|≠1, |λ2|=1 | non-hyperbolic | |
2c<δ<2√1−4h | |λ1|<1, |λ2|>1 | saddle | |
δ=2√1−4h | |λ1|=1, |λ2|≠1 | non-hyperbolic | |
δ>2√1−4h | |λ1|>1, |λ2|>1 | source |