
In this paper, a discrete predator-prey model with double Allee effect is discussed. We first simplify the corresponding continuous predator-prey model, and use the semidiscretization method to obtain a new discrete model. Next, the existence and local stability of nonnegative fixed points of the new discrete model are studied by using a key lemma. Then, by using the center manifold theorem and bifurcation theory, the sufficient conditions for the occurrences of transcritical bifurcation and Neimark-Sacker bifurcation and the stability of closed orbit bifurcated are obtained. Finally, the numerical simulations are presented, which not only verify the existence of Neimark-Sacker bifurcation but also reveal some new dynamic phenomena of this model.
Citation: Shaosan Xia, Xianyi Li. Complicate dynamics of a discrete predator-prey model with double Allee effect[J]. Mathematical Modelling and Control, 2022, 2(4): 282-295. doi: 10.3934/mmc.2022027
[1] | Zeyan Yue, Lijuan Dong, Sheng Wang . Stochastic persistence and global attractivity of a two-predator one-prey system with S-type distributed time delays. Mathematical Modelling and Control, 2022, 2(4): 272-281. doi: 10.3934/mmc.2022026 |
[2] | Sheng Wang, Baoli Lei . Dynamics of a stochastic hybrid delay one-predator-two-prey model with harvesting and jumps in a polluted environment. Mathematical Modelling and Control, 2025, 5(1): 85-102. doi: 10.3934/mmc.2025007 |
[3] | Xue Zhang, Bo Sang, Bingxue Li, Jie Liu, Lihua Fan, Ning Wang . Hidden chaotic mechanisms for a family of chameleon systems. Mathematical Modelling and Control, 2023, 3(4): 400-415. doi: 10.3934/mmc.2023032 |
[4] | Hongwei Zheng, Yujuan Tian . Exponential stability of time-delay systems with highly nonlinear impulses involving delays. Mathematical Modelling and Control, 2025, 5(1): 103-120. doi: 10.3934/mmc.2025008 |
[5] | Ankur Jyoti Kashyap, Arnab Jyoti Bordoloi, Fanitsha Mohan, Anuradha Devi . Dynamical analysis of an anthrax disease model in animals with nonlinear transmission rate. Mathematical Modelling and Control, 2023, 3(4): 370-386. doi: 10.3934/mmc.2023030 |
[6] | Zongcheng Li . Exploring complicated behaviors of a delay differential equation. Mathematical Modelling and Control, 2023, 3(1): 1-6. doi: 10.3934/mmc.2023001 |
[7] | Zhaoxia Duan, Jinling Liang, Zhengrong Xiang . Filter design for continuous-discrete Takagi-Sugeno fuzzy system with finite frequency specifications. Mathematical Modelling and Control, 2023, 3(4): 387-399. doi: 10.3934/mmc.2023031 |
[8] | Mrutyunjaya Sahoo, Dhabaleswar Mohapatra, S. Chakraverty . Wave solution for time fractional geophysical KdV equation in uncertain environment. Mathematical Modelling and Control, 2025, 5(1): 61-72. doi: 10.3934/mmc.2025005 |
[9] | Wen Zhang, Jinjun Fan, Yuanyuan Peng . On the discontinuous dynamics of a class of 2-DOF frictional vibration systems with asymmetric elastic constraints. Mathematical Modelling and Control, 2023, 3(4): 278-305. doi: 10.3934/mmc.2023024 |
[10] | Ruxin Zhang, Zhe Yin, Ailing Zhu . Numerical simulations of a mixed finite element method for damped plate vibration problems. Mathematical Modelling and Control, 2023, 3(1): 7-22. doi: 10.3934/mmc.2023002 |
In this paper, a discrete predator-prey model with double Allee effect is discussed. We first simplify the corresponding continuous predator-prey model, and use the semidiscretization method to obtain a new discrete model. Next, the existence and local stability of nonnegative fixed points of the new discrete model are studied by using a key lemma. Then, by using the center manifold theorem and bifurcation theory, the sufficient conditions for the occurrences of transcritical bifurcation and Neimark-Sacker bifurcation and the stability of closed orbit bifurcated are obtained. Finally, the numerical simulations are presented, which not only verify the existence of Neimark-Sacker bifurcation but also reveal some new dynamic phenomena of this model.
The dynamical analysis of complex ecosystem, such as food chain, is based on the interaction among species or between two species, especially the dynamical relationship between predator and prey [1,2,3,4]. The current theory of predator-prey dynamics must depend on the study of nonlinear mathematical model [5]. With the continuous improvement of ecological knowledge like theoretical research, empirical research and observational research, etc., there are more and more basic elements of predation to be considered. Therefore, modelers add some complexities to their abstraction in order to obtain authenticity from the emergence of the far-reaching Lotka-Volterra model[6] and the modifications introduced by Volterra[5], taking into account the self-interference of prey populations.
Allee effect, which affects the number of prey, is one of these factors. It changes the qualitative stability and quantitative aspects of predator-prey model dynamics. Because the interaction between predator and prey is naturally prone to vibration, it is obvious to study this phenomenon as a potential mechanism for the generation of population cycles. Lots of researches about predator-prey models are done with the Allee effect[7,8,9].
The most popular framework for modeling expert predator-prey interaction has the following structure:
{dxdt=xg(x,k)−yh(x),dydt=(ph(x)−c)y, | (1.1) |
where x(t) and y(t) are the prey and predator population sizes in the time t, respectively, p,c>0 indicate the birth rate and background mortality rate, respectively, g(x,k) describes the specific growth rate of the prey in the absence of predator, and h(x) describes the predator functional response.
Any mechanism leading to a positive correlation between the components of individual fitness and the number or density of similar individuals can be regarded as Allee effect; it describes a scenario in which low population size is affected by the positive correlation between population growth rate and density, increasing the possibility of their extinction.
Recent ecological studies have shown that two or more Allee effects can lead to mechanisms acting on a population at the same time. The combined effects of some of these phenomena are called multiple (double) Allee effects. The author's analysis in[10] showed that the results of strong and weak Allee effects on the dynamics of Volterra predator-prey model are similar, which originate from the limit cycle of the model.
In this paper, we continue to consider the following predator-prey model with double Allee effect functional response raised by[10].
{dXdT=rX(1−XK)X−MX+N−qXY,dYdT=(pX−C)Y, | (1.2) |
where r scales the prey growth rate, K is the intrinsic carrying capacity for the environment to the prey in the absence of predator, M is the Allee threshold, and the auxiliary parameter N satisfies N>0, q is the prey captured rate by the predator, p,C>0 indicate the birth rate and background mortality rate, respectively.
One can see that the first equation in the model (1.2) includes double Allee effects, expressed by the first factor m(X)=X−M, and a second term r(X)=rXX+N, This can be interpreted as an approximation of population dynamics, in which the difference between fertile and non-fertile are not clearly modeled.
In order to simplify the analysis of system (1.2), we make a topologically equivalent change of variables and time rescaling as in[11,12,13,14], defining the function ϕ, such that ϕ(x,y)=(Kx,rqy)=(X,Y), rx+ndT=dt. Then, system (1.2) is transformed into
{dxdt=((1−x)(x−m)−(x+n)y)x,dydt=b(x−c)(x+n)y, | (1.3) |
where b=pKr, CpK, m=MK, n=NK, and K>X>M is obtained from equation (1.2), so 1>m>0.
We now use the semidiscretization method, which has been applied in many studies[15,16,17,18], to study the discrete model of system (1.3). For this, suppose that [t] denotes the greatest integer not exceeding t. Consider the following semidiscretization version of (1.3).
{1x(t)dx(t)dt=(1−x([t]))(x([t])−m)−(x([t])+n)y([t]),1y(t)dy(t)dt=b(x([t])−c)(x([t])+n). | (1.4) |
It is easy to see that the system (1.4) has piecewise constant arguments, and that a solution (x(t),y(t)) of the system (1.4) for t∈[0,+∞) possesses the following natures:
1. on the interval [0,+∞), x(t) and y(t) are continuous;
2. when t∈[0,+∞) except for the points t∈{0,1,2,3,⋯}, dx(t)dt and dy(t)dt exist everywhere.
The following system can be obtained by integrating (1.4) over the interval [n,t] for any t∈[n,n+1) and n=0,1,2,⋯
{x(t)=xne(1−xn)(xn−m)−(xn+n)yn(t−n),y(t)=yneb(xn−c)(xn+n)(t−n), | (1.5) |
where xn=x(n) and yn=y(n). Letting t→(n+1)− in (1.5) produces
{xn+1=xne(1−xn)(xn−m)−(xn+n)yn,yn+1=yneb(xn−c)(xn+n), | (1.6) |
where b,c,n>0, 1>m>0 are the same as in (1.3). We mainly study the properties of system (1.6) in the sequel.
The rest of the paper is organized as follows: In Section 2, we investigate the existence and stability of fixed points of the system (1.6). In Section 3, we derive the sufficient conditions for the transcritical bifurcation and the Neimark-Sacker bifurcation of the system (1.6) to occur. In Section 4, numerical simulations are performed to verify the above obtained theoretical results and reveal some new dynamical properties.
Before we analyze the fixed points of the system (1.6), we recall the following lemma see [16,19].
Lemma 1.1. Let F(λ)=λ2+Pλ+Q, where P and Q 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 Q<1;
(ⅰ.2) λ1=−1 and λ2≠−1 if and only if F(−1)=0 and P≠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 Q>1;
(ⅰ.5) λ1 and λ2 are a pair of conjugate complex roots with |λ1|=|λ2|=1 if and only if −2<P<2 and Q=1;
(ⅰ.6) λ1=λ2=−1 if and only if F(−1)=0 and P=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 |Q|=(<,>)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 and then analyze the local stability of each fixed point.
The fixed points of the system (1.6) satisfy
x=xe(1−x)(x−m)−(x+n)y, |
y=yeb(x−c)(x+n). |
Considering the biological meanings of the system (1.6), one only takes into account nonnegative fixed points. Thereout, one finds that the system (1.6) has and only has four nonnegative fixed points E0=(0,0), E1=(1,0), E2=(m,0) and E3=(x0,y0) for m<c<1, where
x0=c,y0=(1−c)(c−m)c+n. |
The Jacobian matrix of the system (1.6) at any fixed point E(x,y) takes the following form
J(E)=([1+x(−2x−y+m+1)]eA−x(x+n)eAb(2x+n−c)yeb(x−c)(x+n)eb(x−c)(x+n)). |
where A=e(1−x)(x−m)−(x+n)y.
The characteristic polynomial of Jacobian matrix J(E) reads
F(λ)=λ2−pλ+q, |
where
p=Tr(J(E))=[1+x(−2x−y+m+1)]eA+eb(x−c)(x+n), q=Det(J(E))=[1+x(−2x−y+m+1)][bx(x+n)(2x+n−c)]eA+b(x−c)(x+n).
For the stability of fixed points E0, E1, E2 and E3, we can easily get the following Theorems 2.1-2.4 respectively.
Theorem 2.1. The fixed point E0=(0,0) of the system (1.6) is a sink.
Proof. The Jacobian matrix J(E0) of the system (1.6) at the fixed point E0=(0,0) is given by
J(E0)=(e−m00e−bcn). |
Obviously, |λ1|=e−m<1 and |λ2|=e−bcn<1, so E0=(0,0) is a sink.
Theorem 2.2. The following statements about the fixed point E1=(1,0) of the system (1.6) are true.
1. If c<1, then E1 is a saddle.
2. If c=1, then E1 is non-hyperbolic.
3. If c>1, then E1 is a stable node.
Proof. The Jacobian matrix of the system (1.6) at E1=(1,0) is
J(E1)=(m−(1+n)0eb(1−c)(1+n)). |
Obviously, λ1=m and λ2=eb(1−c)(1+n).
Note |λ1|<1 is always true. If c<1, then |λ2|>1, so E1 is a saddle; if c=1, then |λ2|=1, therefore E1 is non-hyperbolic; if c>1, implying |λ2|<1, then E1 is a stable node, namely, a sink. The proof is complete.
Theorem 2.3. The following statements about the fixed point E2=(m,0) of the system (1.6) are true.
1. If c<m, then E2 is a source.
2. If c=m, then E2 is non-hyperbolic.
3. If c>m, then E2 is a saddle.
Proof. The Jacobian matrix of the system (1.6) at E2=(m,0) is
J(E2)=(−m2+m+1−m(m+n)0eb(m−c)(m+n)). |
Obviously, λ1=−m2+m+1 and λ2=eb(m−c)(m+n).
Note 0<m<1, so |λ1|>1 is always true. If c<m, then |λ2|>1, so E2 is a source; if c=m, then |λ2|=1, therefore E2 is non-hyperbolic; if c>m, implying |λ2|<1, then E2 is a saddle. The proof is finished.
Theorem 2.4. When (1−c)(c−m)>0, namely, 0<m<c<1, the fixed point E3=(c,(1−c)(c−m)c+n) is a positive fixed point of the system (1.6). Let b0=c2+2cn−mn−m−n(c+n)2(1−c)(c−m). Then the following statements are true about the positive fixed point E3.
Ⅰ. If m<c2+2cn−nn+1, then,
1. for 0<b<b0, E3 is a stable node;
2. for b=b0, E3 is non-hyperbolic;
3. for b>b0, E3 is an unstable node.
Ⅱ. If m≥c2+2cn−nn+1, then, E3 is an unstable node.
Proof. The Jacobian matrix of the system (1.6) at E3 can be simplified as follows
J(E3)=(1+−c(c2+2cn−mn−m−n)c+n−c(c+n)b(1−c)(c−m)1). |
The characteristic polynomial of Jacobian matrix J(E3) reads as
F(λ)=λ2−pλ+q, | (2.1) |
where
p=Tr(J(E3))=2−b0c(1−c)(c−m)(c+n),
q=Det(J(E3))=1+(b−b0)c(1−c)(c−m)(c+n). By calculating we get
F(1)=bc(1−c)(c−m)(c+n)>0,
and
F(−1)=2(2−c2)+2c(n(1−c)+mn+n)c+n+bc(1−c)(c−m)(c+n)>0. |
Ⅰ. If m<c2+2cn−nn+1, then b0>0. So, when 0<b<b0, q<1. By Lemma 1.1 (i.1), |λ1|<1 and |λ2|<1, therefore E3 is a stable node, i.e., a sink. When b=b0, q=1, −2<p<2. By Lemma 1.1 (i.5), Eq. (2.1) has a pair of conjugate complex roots λ1 and λ2 with |λ1|=|λ2|=1, implying E3 is non-hyperbolic. When b>b0<, q>1. Lemma 1.1 (i.4) tells us that |λ1|>1 and |λ2|>1, so E3 is an unstable node, i.e., a source.
Ⅱ. If m≥c2+2cn−nn+1, then b0≤0. So, b>0≥b0. Hence q>1. Lemma 1.1 (i.4) reads that E3 is an unstable node. The proof is complete.
In this section, we are in a position to use the Center Manifold Theorem and bifurcation theorem to analyze the local bifurcation problems of the fixed points E1, E2 and E3. For related work, refer to[20,21,22,23,24,25].
Theorem 2.2 shows that a bifurcation of E1 may occur in the space of parameters (b,c,m,n)∈SE+={(b,c,m,n)∈R4+|b>0,c>0,1>m>0,n>0}.
Theorem 3.1. Set the parameters (b,c,m,n)∈SE+={(b,c,m,n)∈R4+|b>0,c>0,1>m>0,n>0}. Let c0=1, then the system (1.6) undergoes a transcritical bifurcation at E1 when the parameter c varies in a small neighborhood of c0.
Proof. In order to show the detailed process, we proceed according to the following steps.
The first step. Let un=xn−1,vn=yn−0, which transforms the fixed point E1=(1,0) to the origin O(0,0), and the system (1.6) to
{un+1=(un+1)e−un(un−m+1)−(un+n+1)vn−1,vn+1=vneb(un−c+1)(un+n+1). | (3.1) |
The second step. Giving a small perturbation c∗ of the parameter c, i.e., c∗=c−c0, with 0<|c∗|≪1, the system (3.1) is perturbed into
{un+1=(un+1)e−un(un−m+1)−(un+n+1)vn−1,vn+1=vneb(un−c∗)(un+n+1). | (3.2) |
Letting c∗n+1=c∗n=c∗, the system (3.2) can be written as
{un+1=(un+1)e−un(un−m+1)−(un+n+1)vn−1,vn+1=vneb(un−c∗n)(un+n+1),c∗n+1=c∗n. | (3.3) |
The third step. Taylor expanding of the system (3.3) at (un,vn,c∗n)=(0,0,0) takes the form
{un+1=a100un+a010vn+a200u2n+a020v2n+a110unvn+a300u3n+a030v3n+a210u2nvn+a120unv2n+o(ρ31),vn+1=b100un+b010vn+b001c∗n+b200u2n+b020v2n+b002c∗n2+b110unvn+b101unc∗n+b011vnc∗n+b300u3n+b030v3n+b003c∗n3+b210u2nvn+b120unv2n+b201u2nc∗n+b102unc∗n2+b021v2nc∗n+b012vnc∗n2+b111unvnc∗n+o(ρ31),c∗n+1=c∗n, | (3.4) |
where ρ1=√u2n+v2n+(c∗n)2,
a100=m,a010=−(n+1),a200=12(m−1)2+m−2,a020=12(n+1)2,a110=−mn−m−1,a300=16(m−1)3+12(m−1)2−m,a030=−16(n+1)3,a120=12m(n+1)2+n+1,a210=−12(m−1)2(n+1)−mn−2m−2n+3, |
b100=b001=b200=b020=b002=b101=b300=b030=b003=b120=b201=b102=b021=0,b010=1,b110=b(n+1),b011=−b(n+1),b210=b+12b2(n+1)2,b012=12b2(n+1)2,b111=−b−b2(n+2)2. |
Let
J(E1)=(a100a0100b100b0100001) |
i.e.,J(E1)=(m−(n+1)0010001). Then, we derive the three eigenvalues of J(E1) to be
λ1=m,λ2,3=1, |
and the corresponding eigenvectors
(ξ1,η1,φ1)T=(1,0,0)T,(ξ2,η2,φ2)T=(n+1,m−1,0)T,(ξ3,η3,φ3)T=(0,0,1)T. |
The fourth step. Let T=(ξ1ξ2ξ3η1η2η3φ1φ2φ3), namely,
T=(1n+100m−10001), |
then T−1=(11+n1−m001m−10001). |
Taking the following transformation
(un,vn,c∗n)T=T(Xn,Yn,δn)T, |
the system (3.4) is changed into the following form
{Xn+1=mXn+F(Xn,Yn,δn)+o(ρ32),Yn+1=Yn+G(Xn,Yn,δn)+o(ρ32),δn+1=δn, | (3.5) |
where ρ2=√X2n+Y2n+δ2n,
F(Xn,Yn,δn)=m200X2n+m020Y2n+m002δn2+m110XnYn+m101Xnδn+m011Ynδn+m300X3n+m030Y3n+m003δn3+m210X2nYn+m120XnY2n+m201X2nδn+m102Xnδn2+m021Y2nδn+m012Ynδn2+m111XnYnδn, |
G(Xn,Yn,δn)=l200X2n+l020Y2n+l002δn2+l110XnYn+l101Xnδn+l011Ynδn+l300X3n+l030Y3n+l003δn3+l210X2nYn+l120XnY2n+l201X2nδn+l102Xnδn2+l021Y2nδn+l012Ynδn2+l111XnYnδn, |
m200=a200,m300=a300,m002=m101=m003=m201=m102=0,m020=(a200−b110)(1+n)2+a020(m−1)2+a110(m−1)(n+1),m110=(2a200−b110)(1+n)+a110(m−1),m011=−b011(1+n),m030=(a300−b210)(1+n)3+a030(m−1)3+a210(1+n)2(m−1)+a120(1+n)(m−1)2,m210=(3a300−b210)(1+n)+a210(m−1),m120=(3a300−2b210)(1+n)2+2a210(1+n)(m−1)+a120(m−1)2,m021=−b111(1+n)2,m012=−b012(1+n),m111=−b111(1+n),l200=l002=0,l020=b110(1+n),l110=b110,l101=0,l011=b011,l300=0,l030=b210(1+n)2,l003=0,l210=b210,l120=2b210(1+n),l201=l102=0,l021=b111(1+n),l012=b012,l111=b111. |
The fifth step. Suppose on the center manifold
Xn=h(Yn,δn)=h20Y2n+h11Ynδn+h02δ2n+o(ρ23), |
where ρ3=√Y2n+δ2n, then, according to
Xn+1=mh(Yn,δn)+F(h(Yn,δn),Yn,δn)+o(ρ23), |
h(Yn+1,δn+1)=h20Y2n+1+h11Yn+1δn+1+h02δ2n+1+o(ρ23)=h20(Yn+G(h(Yn,δn),Yn,δn))2+h11(Yn+G(h(Yn,δn),Yn,δn))δn+h02δ2n+o(ρ23). |
and Xn+1=h(Yn+1,δn+1), we obtain the center manifold equation to satisfy the following relation
mh(Yn,δn)+F(h(Yn,δn),Yn,δn)=h20(Yn+G(h(Yn,δn),Yn,δn))2+h11(Yn+G(h(Yn,δn),Yn,δn))δn+h02δ2n+o(ρ23).
Comparing the corresponding coefficients of terms with the same orders in the above center manifold equation, we get
h20=(m−2)(1+n)2−(1+n)(2+n)(m−1)−b(1+n)31−m,
h11=b(1+n)21−m,h02=0.
So, the system (3.5) restricted to the center manifold takes as
Yn+1=f(Yn,δn):=Yn+G(h(Yn,δn),Yn,δn)+o(ρ23)=Yn+b(1+n)2Y2n−b(1+n)Ynδn+b(1+n)21−m(1−2m−n−12b(1+n)2(1+m))Y3n+(mb2(1+n)31−m−b(1+n))Yn2δn+12b2(1+n)2Ynδ2n+o(ρ33).
Therefore one has
f(Yn,δn)|(0,0)=0,∂f∂Yn|(0,0)=1,∂f∂δn|(0,0)=0,∂2f∂Yn∂δn|(0,0)=−b(1+n)≠0,∂2f∂Y2n|(0,0)=2b(1+n)2≠0. |
According to (21.1.42)-(21.1.46) in the literature ([26], pp. 507), all the conditions for the occurrence of the transcritical bifurcation are established, hence, it is valid for the {occurrence} of transcritical bifurcation in the fixed point E1. The proof is over.
According to Theorem 2.3, the fixed point E2(m,0) is non-hyperbolic, the system (1.6) may undergo a bifurcation (the correspond eigenvalue are λ1=−m2+m+1, λ2=1). By using the same method as that in Section 3.2, we get the following result.
Theorem 3.2. Set the parameters (b,c,m,n)∈SE+={(b,c,m,n)∈R4+|b>0,c>0,1>m>0,n>0}. Let c1=m, then the system (1.6) undergoes a transcritical bifurcation at E2 when the parameter c varies in a small neighborhood of c1.
When m<c2+2cn−nn+1, b=b0=c2+2cn−mn−n−m(c+n)2(1−c)(c−m), Theorem 2.4 with Lemma 1.2 (i.5) shows that F(1)>0, F(−1)>0, −2<p<2 and q=1, so λ1 and λ2 are a pair of conjugate complex roots with |λ1|=|λ2|=1. At this time we derive that the system (1.6) at the fixed point E3 can undergo a Neimark-Sacker bifurcation in the space of parameters (b,c,m,n)∈SE+={(b,c,m,n)∈R4+|b>0,1>c>m,0<m<c2+2cn−nn+1}.
In order to show the process clearly, we carry out the following steps.
The first step. Take the changes of variables un=xn−x0,vn=yn−y0, which transform fixed point E3=(x0,y0) to the origin O(0,0), and the system (1.6) into
{un+1=(un+x0)e(−un−x0+1)(un+x0−m)−(un+x0+n)vnyn−x0,vn+1=(vn+y0)eb(un+x0−c)(un+x0+n)−y0. | (3.6) |
The second step. Give a small perturbation b∗ of the parameter b, i.e., b∗=b−b0, then the perturbation of the system (3.6) can be regarded as follows
{un+1=(un+x0)e(−un−x0+1)(un+x0−m)−(un+x0+n)vnyn−x0,vn+1=(vn+y0)e(b∗+b0)(un+x0−c)(un+x0+n)−y0. | (3.7) |
The corresponding characteristic equation of the linearized equation of the system (3.7) at the equilibrium point (0, 0) can be expressed as
F(λ)=λ2−p(b∗)λ+q(b∗)=0, |
where
p(b∗)=2+c(−c2−2cn+mn+m+n)c+n, |
and
q(b∗)=c[(c2+2cn−mn−m−n)−(b∗+b0)(c+n)2(1−c)(c−m)]−(c+n)+1.
It is easy to derive p2(b∗)−4q(b∗)<0 when b∗=0, and 0<p(b∗)<2, then the two roots of F(λ)=0 are
λ1,2(b∗)=p(b∗)±√p2(b∗)−4q(b∗)2=p(b∗)±i√4q(b∗)−p2(b∗)2, |
which implies
(|λ1,2(b∗)|)|b∗=0=√q(b∗)|b∗=0=1, |
and
(d|λ1,2(b∗)|db∗)|b∗=0=12c(c+n)(1−c)(c−m)>0. |
The occurrence of Neimark-Sacker bifurcation requires the following conditions to be satisfied
(H.1)(d|λ1,2(b∗)|db∗)|b∗=0≠0; |
(H.2)λi1,2(0)≠1,i=1,2,3,4. |
Since p(b∗)|b∗=0=2+c(−c2−2cn+mn+m+n)c+n and q(b∗)|b∗=0=1, we have λ1,2(0)=2(c+n)+c(−c2−2cn+mn+m+n)±i√4(c+n)2−[2(c+n)+c(−c2−2cn+mn+m+n)]22(c+n), then it is easy to derive λi1,2(0)≠1 for all i=1,2,3,4. According to ([2], pp517-522), they satisfy all of the conditions for Neimark-Sacker bifurcation to occur.
The third step. In order to derive the normal form of the system (3.7), we expand the system (3.7) into power series up to the following third-order form around the origin
{un+1=c10un+c01vn+c20u2n+c11unvn+c02v2n+c30u3n+c21u2nvn+c12unv2n+c03v3n+o(ρ34),vn+1=d10un+d01vn+d20u2n+d11unvn+d02v2n+d30u3n+d21u2nvn+d12unv2n+d03v3n+o(ρ34), | (3.8) |
where ρ4=√u2n+v2n,
c10=1+c(−c2−2cn+mn+m+n)c+n,c01=−c(c+n),c20=−c+−c2−2cn+mn+m+nc+n+c(−c2−2cn+mn+m+n)22(c+n)2,c02=c(c+n)22,c11=−[2c+n+c(−c2−2cn+mn+m+n)],c30=−1−c(−c2−2cn+mn+m+n)c+n+(−c2−2cn+mn+m+n)2c(c+n)2+c(−c2−2cn+mn+m+n)36(c+n)3,c03=−c(c+n)36,c21=−1−(−c2−2cn+mn+m+n)+c[c+n−−c2−2cn+mn+m+nc+n−(−c2−2cn+mn+m+n)22(c+n)],c12=(c+n)22+c(c+n)[1+−c2−2cn+mn+m+n2],d10=−−c2−2cn+mn+m+n(c+n)2,d01=1,d20=(−c2−2cn+mn+m+n)22(c+n)3(1−c)(c−m)−2(−c2−2cn+mn+m+n)(1−c)(c−m)2(c+n)3(1−c)(c−m),d11=−−c2−2cn+mn+m+n(c+n)(1−c)(c−m),d02=d03=d12=0,d30=(−c2−2cn+mn+m+n)2(c+n)4(1−c)(c−m)[1−−c2−2cn+mn+m+n6(1−c)(c−m)],d21=(−c2−2cn+mn+m+n)(c+n)2(1−c)(c−m)[1−−c2−2cn+mn+m+n2(1−c)(c−m)]. |
Let
J(E3)=(c10c01d10d01),namely, |
J(E3)=(1+K−c(c+n)−Kc(c+n)1). |
It is easy to derive the two eigenvalues of the matrix J(E3) are
λ1,2=(1+12K)±βi, |
where K=c(−c2−2cn+mn+m+n)c+n,
β=√−c(−c2−2cn+mn+m+n)[4(c+n)+c(−c2−2cn+mn+m+n)]2(c+n),
with the corresponding eigenvectors
v1,2=(−c(c+n)−12K)±i(0β).
Let
T=(0−c(c+n)β−12K),then, |
T−1=(−K2c(c+n)β1β−1c(c+n)0). |
Make a change of variables
(u,v)T=T(X,Y)T, |
then, the system (3.8) is transformed into the following form
{X→(1+12K)X−βY+¯F(X,Y)+o(ρ35),Y→βX+(1+12K)Y+¯G(X,Y)+o(ρ35), | (3.9) |
where ρ5=√X2+Y2,
¯F(X,Y)=e20X2+e11XY+e02Y2+e30X3+e21X2Y+e12XY2+e03Y3,¯G(X,Y)=f20X2+f11XY+f02Y2+f30X3+f21X2Y+f12XY2+f03Y3, |
e20=c02βK2c01,e11=c01c11K+2c201d11−c02K22c01,e02=4c201(c20K+2c01d20−d11K)8c01β+K2(c02K−2c01c11)8c01β,e30=c03β2K2c01,e21=(2c01c12−3c03K)βK4c01,e12=c01c21K−c12K2+2d21c2012+3c03K38c01,e03=8c301(c30K+2c01d30−d21K)16c01β−K2(4c201c21+c03K2−2c01c12K)16c01β,f20=c02c01β2,f11=c11β−c02c01βK,f02=c01c20−12c11K+c024c01K2,f30=c03c01β3,f21=c12β2−3c032c01β2K,f12=c01c21β−c12βK+3c034c01βK2,f03=c30c201−12c01c21K+14c12K2−c038c01K3. |
Furthermore,
¯FXX=c02βKc01,¯FXY=c01c11K+2c201d11−c02K22c01,¯FXXX=3c03β2Kc01,¯FYY=4c201(c20K+2c01d20−d11K)4c01β+K2(c02K−2c01c11)4c01β,¯FXXY=c12βK−3c03βK22c01,¯FXYY=c01c21K−c12K2+2d21c201+3c03K34c01,¯FYYY=3c301(c30K+2c01d30)−d21Kc01β−3K2(4c201c21+c03K2−2c01c12K)8c01β,¯GXX=2c02β2c01,¯GXY=c11β−c02βKc01,¯GYY=2c01c20−c11K+c02K22c01,¯GXXX=6c03β3c01,¯GXXY=2c12β2−3c03β2Kc01,¯GXYY=2c01c21β−2c12βK+3c03βK22c01,¯GYYY=6c30c201−3c01c21K+32c12K2−3c03K34c01. |
The {fourth} step. In order to ensure that the system (3.9) has a Neimark-Sacker bifurcation occurring, we need to calculate the discriminating quantity
L=−Re((1−2λ1)λ221−λ1ζ20ζ11)−12|ζ11|2−|ζ02|2+Re(λ2ζ21), | (3.10) |
and L is required not to be zero, where
ζ20=18[¯FXX−¯FYY+2¯GXY+i(¯GXX−¯GYY−2¯FXY)],ζ11=14[¯FXX+¯FYY+i(¯GXX+¯GYY)],ζ02=18[¯FXX−¯FYY−2¯GXY+i(¯GXX−¯GYY+2¯FXY)],ζ21=116[¯FXXX+¯FXYY+¯GXXY+¯GYYY+i(¯GXXX+¯GXYY−¯FXXY−¯FYYY)]. |
By calculation we get
ζ20=18(−4c201(c20K+2c01d20−d11K)4c01β−K2(c02K−2c01c11)4c01β+(2c01c11−c02K)βc01)+18(c02(K2+4β2)2c01−2c01(c20+d11))i,ζ11=14(c02βKc01+4c201(c20K+2c01d20−d11K)4c01β+K2(c02K−2c01c11)4c01β)+14(c02(4β2+K2)2c01+2c01c20−c11K)i,ζ02=18(−4c201(c20K+2c01d20−d11K)4c01β−K2(c02K−2c01c11)4c01β+(3c02K−2c01c11)βc01)+14(c02(4β2−3K2)4c01+c11K+c01(d11−c20))i,ζ21=116(2c01(3c30c01−c21K+c01d21)+c12(12K2+2β2))+116(3c03β(K2+2β2)c01+3K2(4c201c21+c03K2−2c01c12K)8c01β+β(2c01c21−3c12K)−3c301(c30K+2c01d30)−d21Kc01β)i. |
Theorem 3.3. Assume the parameters b, c, m, n in the space SE+={(b,c,m,n)∈R4+|b>0,1>c>m,0<m<c2+2cn−nn+1}. Let b0=c2+2cn−mn−m−n(c+n)2(1−c)(c−m) and L be defined as above (3.10). If L≠0 holds and the parameter a varies in the small neighborhood of b0, then the system (1.6) at the fixed point E3 undergoes a Neimark-Sacker bifurcation. In addition, if L<(or>)0, then an attracting (or repelling) invariant closed curve bifurcates from the fixed point E3 for b<(or>)b0.
In this section, we use the bifurcation diagrams, phase portraits and Lyapunov exponents of the system (1.6) to verify our theoretical results and further reveal some new dynamical behaviors to occur as the parameters vary by Matlab software.
Fix the parameter values c=0.8,m=0.3,n=0.8, let b∈(1.5,3.0) and take the initial values (x0,y0)=(0.55,0.25),(0.80,0.05) in Fig. 2 and Fig. 3 respectively. Figure 1(a) shows the {bifurcation} diagram of (b,x)-plane, from which the fixed point E3 is stable when b<b0=2.266. Moreover, the fixed point E3 is unstable when b>b0. Hence, the Neimark-Sacker bifurcation occurs at the fixed point E3=(0.800,0.625) when b=b0, whose multipliers are λ1,2=0.855±0.519i with |λ1,2|=1.
The corresponding maximum Lyapunov exponent diagram of the system (1.6) is plotted in Figure 1(b). Figures 2(a)-2(f) and Figures 3(a)-3(d) show that the dynamical properties of the fixed point E3 change from stable to unstable as the value of the parameter b decreases and there is an occurrence of invariant closed curve around E3 when b=b0, which agrees to the result of Theorem 3.3.
From the phase portraits in Figs 2 and 3, we infer the stability of E3. Figures 2(d)-2(f) show that the closed curve is stable outside, while Figures 3(a)-3(d) indicate that the closed curve is stable inside for the fixed point E3 as long as the assumptions of Theorem 3.2 hold.
In this paper, we discuss the dynamical behaviors of a predator-prey model (1.6) of Gause-type with double Allee effect affecting the prey population. Under the given parametric conditions, we completely show the existence and stability of four nonnegative equilibria E0=(0,0), E1=(1,0), E2=(m,0) and E3=(c,(1−c)(c−m)c+n). Then we derive the sufficient conditions for its transcritical bifurcation and Neimark-Sacker bifurcation to occur. Meanwhile, it is clear that the positive equilibrium E3=(x0,y0) is asymptotically stable when b<b0=c2+2cn−mn−m−n(c+n)2(1−c)(c−m) and unstable when b>b0 under the condition m<c2+2cn−nn+1. Hence, the system (1.6) undergoes a bifurcation which has been shown to be a Neimark-Sacker bifurcation when the parameter b goes through the critical value b0. Finally, numerical simulations illustrate the theoretical analysis results of the system (1.6).
The perturbations of different parameters in this system may lead to different bifurcations. This demonstrates that this system is sensitive to its parameters. Especially, the occurrence of Neimark-Sacker bifurcation implies that the predator and the prey can coexist under such parametric conditions.
All authors contributed equally and significantly in writing this paper. All authors read and approved the final manuscript.
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 National Natural Science Foundation of Zhejiang University of Science and Technology (F701108G14).
The authors declare that they have no competing interests.
[1] |
S. Md, S. Rana, Dynamics and chaos control in a discrete-time ratio-dependent Holling-Tanner model, J. Egypt. Math. Soc., 48 (2019), 1–16. http://doi.org/10.1186/s42787-019-0055-4 doi: 10.1186/s42787-019-0055-4
![]() |
[2] |
A. Q. Khan, Neimark-Sacker bifurcation of a two-dimensional discrete-time predator-prey model, SpringerPlus, 5 (2016), 1–10. https://doi.org/10.1186/s40064-015-1618-y doi: 10.1186/s40064-015-1618-y
![]() |
[3] |
C. Rodrigo, S. Willy, S. Eduardo, Bifurcations in a predator-prey model with general logistic growth and exponential fading memory, Appl. Math. Model., 45 (2017), 134–147. http://doi.org/10.1016/j.apm.2016.12.003 doi: 10.1016/j.apm.2016.12.003
![]() |
[4] | C. Holling, The functional response of predator to prey density and its role in mimicry and population regulation, Mem. Entomol. Soc. Can., 91 (1959), 385–398. |
[5] |
M. C. Saunders, Complex Population Dynamics: A Theoretical/Empirical Synthesis, Entomological Society of America, 35 (2006), 1139–1139. http://doi.org/10.1603/0046-225X-35.4.1139a doi: 10.1603/0046-225X-35.4.1139a
![]() |
[6] |
A. A. Berryman, A. P. Gutierrez, R. Arditi, Credible, , Entomological Society of America, 76 (1995), 1980–1985. http://doi.org/10.2307/1940728 doi: 10.2307/1940728
![]() |
[7] |
Q. Lin, Allee effect increasing the final density of the species subject to the Allee effect in a Lotka–Volterra commensal symbiosis model, Adv. Differ. Equ., 196 (2018), 1–9. http://doi.org/10.1186/s13662-018-1646-3 doi: 10.1186/s13662-018-1646-3
![]() |
[8] |
J. Wang, J. Shi, J. Wei, Predator–prey system with strong Allee effect in prey, J. Math. Biol., 62 (2011), 291–331. http://doi.org/10.1007/s00285-010-0332-1 doi: 10.1007/s00285-010-0332-1
![]() |
[9] |
Y. Cai, C. Zhao, W. Wang, J. Wang, Dynamics of a Leslie–Gower predator–prey model with additive Allee effect, Appl. Math. Model., 39 (2015), 2092–2106. http://doi.org/10.1016/j.apm.2014.09.038 doi: 10.1016/j.apm.2014.09.038
![]() |
[10] |
E. González-Olivares, B. González-Yañez, J. Mena-Lorca, A. Rojas-Palma, J. D. Flores, Consequences of double Allee effect on the number of limit cycles in a predator–prey model, Comput. Math. Appl., 62 (2011), 3449–3463. http://doi.org/10.1016/j.camwa.2011.08.061 doi: 10.1016/j.camwa.2011.08.061
![]() |
[11] | J. D. Flores, J. Mena-Lorca, B. Gonzalez-Yanez, E. Gonzalez-Olivares, Consequences of depensation in a Smith's bioeconomic model for open-access fishery, Proceedings of the 2006 International Symposium on Mathematical and Computational Biology, 27 (2006), 219–232. |
[12] |
E. González-Olivares, A. Rojas-Palma, Multiple limit cycles in a Gause type predator–prey model with holling type Ⅲ functional response and Allee effect on prey, B. Math. Biol., 73 (2011), 1378–1397. http://doi.org/10.1007/s11538-010-9577-5 doi: 10.1007/s11538-010-9577-5
![]() |
[13] |
E. González-Olivares, J. Mena-Lorca, A. Rojas-Palma, J. D. Flores, Dynamical complexities in the Leslie–Gower predator–prey model considering a simple form to the Allee effect on prey, Appl. Math. Model., 35 (2011), 366–381. https://doi.org/10.1016/j.apm.2010.07.001 doi: 10.1016/j.apm.2010.07.001
![]() |
[14] | B. González-Yañez, E. González-Olivares, Consequences of Allee effect on a Gause type predator–prey model with nonmonotonic functional response, Proceedings of the Third Brazilian Symposium on Mathematical and Computational Biology, 2 (2004), 358–373. |
[15] |
Q. Din, Complexity and chaos control in a discrete-time prey-predator model, Commun. Nonliner.Sci. Numer. Simul., 49 (2017), 113–134. https://doi.org/10.1016/j.cnsns.2017.01.025 doi: 10.1016/j.cnsns.2017.01.025
![]() |
[16] | W. Li, X. Li, Neimark–Sacker bifurcation of a semi-discrete hematopoiesis model, J. Appl. Anal. Comput., 8 (2018), 1679–1693. |
[17] |
Z. Hu, Z. Teng, L. Zhang, Stability and bifurcation analysis of a discrete predator-prey model with nonmonotonic functional response, Nonlinear Anal. Real. World. Appl., 12 (2011), 2356–2377. http://doi.org/10.1016/j.nonrwa.2011.02.009 doi: 10.1016/j.nonrwa.2011.02.009
![]() |
[18] |
C. Wang, X. Li, Further investigations into the stability and bifurcation of a discrete predator-prey model, J. Math. Anal. Appl., 422 (2015), 920–939. http://doi.org/10.1016/j.jmaa.2014.08.058 doi: 10.1016/j.jmaa.2014.08.058
![]() |
[19] | C. Wang, X. Li, Stability and Neimark–Sacker bifurcation of a semi-discrete population model, J. Appl. Anal. Comput., 4 (2014), 419–435. |
[20] |
M. Li, X. Zhou, J. Xu, Dynamic properties of a discrete population model with diffusion, Adv. Differ. Equ., 580 (2020), 1–20. http://doi.org/10.1186/s13662-020-03033-w doi: 10.1186/s13662-020-03033-w
![]() |
[21] |
T. Gallay, A center-stable manifold theorem for differential equations in Banach spaces, Commun. Math. Phys., 152 (1993), 249–268. http://doi.org/10.1007/BF02098299 doi: 10.1007/BF02098299
![]() |
[22] |
W. Liu, D. Cai, J. Shi, Dynamic behaviors of a discrete-time predator–prey bioeconomic system, Adv. Differ. Equ., 133 (2018), 1–22. http://doi.org/10.1186/s13662-018-1592-0 doi: 10.1186/s13662-018-1592-0
![]() |
[23] |
A. Jorba, J. Masdemont, Dynamics in the center manifold of the collinear points of the restricted three body problem, Physica D: Nonlinear Phenomena, 132 (1999), 189–213. http://doi.org/10.1016/S0167-2789(99)00042-1 doi: 10.1016/S0167-2789(99)00042-1
![]() |
[24] |
Ü. Ufuktepe, S. Kapcak, O. Akman, Stability and invariant manifold for a predator-prey model with Allee effect, Adv. Differ. Equ., 348 (2013), 1–8. http://doi.org/10.1186/1687-1847-2013-348 doi: 10.1186/1687-1847-2013-348
![]() |
[25] |
M. Zhao, Z. Xuan, C. Li, Dynamics of a discrete-time predator-prey system, Adv. Differ. Equ., 191 (2016), 1–18. https://doi.org/10.1186/s13662-016-0903-6 doi: 10.1186/s13662-016-0903-6
![]() |
[26] | S. Winggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, New York: Springer-Verlag, 2003. |
1. | 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 | |
2. | Chirodeep Mondal, Ritwika Mondal, Dipak Kesh, Debasis Mukherjee, Dynamics of predator-prey system with the consequences of double Allee effect in prey population, 2025, 51, 0092-0606, 10.1007/s10867-025-09670-0 | |
3. | Debjit Pal, Ritwika Mondal, Dipak Kesh, Debasis Mukherjee, Non-spatial Dynamics and Spatiotemporal Patterns Formation in a Predator–Prey Model with Double Allee and Dome-shaped Response Function, 2025, 87, 0092-8240, 10.1007/s11538-025-01411-7 |