Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Dynamics of a predator-prey model with strong Allee effect and nonconstant mortality rate


  • In this paper, dynamics analysis for a predator-prey model with strong Allee effect and nonconstant mortality rate are taken into account. We systematically studied the existence and stability of the equilibria, and detailedly analyzed various bifurcations, including transcritical, saddle-node, Hopf and Bogdanov-Takens bifurcation. In addition, the theoretical results are verified by numerical simulations. The results indicate that when the mortality is large, the nonconstant death rate can be approximated to a constant value. However, it cannot be considered constant under small mortality rate conditions. Unlike the extinction of species for the constant mortality, the nonconstant mortality may result in the coexistence of prey and predator for the predator-prey model with Allee effect.

    Citation: Juan Ye, Yi Wang, Zhan Jin, Chuanjun Dai, Min Zhao. Dynamics of a predator-prey model with strong Allee effect and nonconstant mortality rate[J]. Mathematical Biosciences and Engineering, 2022, 19(4): 3402-3426. doi: 10.3934/mbe.2022157

    Related Papers:

    [1] Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh . Combined impact of fear and Allee effect in predator-prey interaction models on their growth. Mathematical Biosciences and Engineering, 2024, 21(10): 7211-7252. doi: 10.3934/mbe.2024319
    [2] Fang Liu, Yanfei Du . Spatiotemporal dynamics of a diffusive predator-prey model with delay and Allee effect in predator. Mathematical Biosciences and Engineering, 2023, 20(11): 19372-19400. doi: 10.3934/mbe.2023857
    [3] 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
    [4] Moitri Sen, Malay Banerjee, Yasuhiro Takeuchi . Influence of Allee effect in prey populations on the dynamics of two-prey-one-predator model. Mathematical Biosciences and Engineering, 2018, 15(4): 883-904. doi: 10.3934/mbe.2018040
    [5] Yun Kang, Sourav Kumar Sasmal, Amiya Ranjan Bhowmick, Joydev Chattopadhyay . Dynamics of a predator-prey system with prey subject to Allee effects and disease. Mathematical Biosciences and Engineering, 2014, 11(4): 877-918. doi: 10.3934/mbe.2014.11.877
    [6] 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
    [7] 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
    [8] 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
    [9] Eduardo Liz, Alfonso Ruiz-Herrera . Delayed population models with Allee effects and exploitation. Mathematical Biosciences and Engineering, 2015, 12(1): 83-97. doi: 10.3934/mbe.2015.12.83
    [10] Yuxuan Zhang, Xinmiao Rong, Jimin Zhang . A diffusive predator-prey system with prey refuge and predator cannibalism. Mathematical Biosciences and Engineering, 2019, 16(3): 1445-1470. doi: 10.3934/mbe.2019070
  • In this paper, dynamics analysis for a predator-prey model with strong Allee effect and nonconstant mortality rate are taken into account. We systematically studied the existence and stability of the equilibria, and detailedly analyzed various bifurcations, including transcritical, saddle-node, Hopf and Bogdanov-Takens bifurcation. In addition, the theoretical results are verified by numerical simulations. The results indicate that when the mortality is large, the nonconstant death rate can be approximated to a constant value. However, it cannot be considered constant under small mortality rate conditions. Unlike the extinction of species for the constant mortality, the nonconstant mortality may result in the coexistence of prey and predator for the predator-prey model with Allee effect.



    Interactions between a wide variety of species are a feature of ecosystems. Understanding the relationship between predator and prey, which plays a key status in predicting the dynamics of ecosystems. Mathematical models are increasingly influential in theoretical ecology, contributing not only to the quantitative understanding of ecosystems [1,2,3,4,5] but also the development of mathematical capability [6,7,8,9,10,11]. A classical general predator-prey model is shown below [12]:

    {dNdt=f(N)Ng(N,P)P,dPdt=h(g(N,P),P)P, (1)

    where the densities of predator and prey are respectively expressed as P and N, the per capita growth rate of predator and prey are respectively denoted as h(g(N,P),P) and f(N), g(N,P) is the functional response.

    There are many key factors that affect the dynamics in predator-prey models, especially functional response and nonconstant mortality rate of the predator. Functional response is generally of two kinds: ratio-dependent and prey dependent. Holling [13] introduced the concept of functional response and the study of small-mammal predation of the European Pine Sawfly1 revealed the composition of predation. Subsequently, there have been many discussions on the effects of functional response on predator-prey models [14,15,16]. Abrams and Ginzburg [12] studied the nature of predation and compared the effects of different functional responses on population dynamics. Zhang et al. [17] studied Beddington-DeAngelis functional response, discussed the linear stability and obtained the condition for Turing instability.

    Consider h(g(N,P),P)=eg(N,P)α(P), then we get the following equation:

    dPdt=(eg(N,P)α(P))P,

    where e and α(P) stand for the conversion efficiency and the specific mortality rate of predator without prey, respectively. In most researches, the mortality rate of predator is assumed to be constant, i.e., α(P)=u [18,19]. However, Cavani and Farkas [20] introduced a nonconstant mortality rate of predator α(P)=γ+δp1+P, where γ>0 and δ>0 represent mortality rate at the low density and the maximal mortality, respectively(γδ). The nonconstant mortality α(P) is a bounded and increasing function of P. When γ=δ, it is a constant mortality rate [21,22]. About the nonconstant death rate in predator-prey model, many researchers have found that nonconstant mortality has a major impact on the population dynamics [23,24,25,26,27].

    Experiments have shown that due to the factors, such as mate finding, predator satiation, cooperative defense and habitat attention, Allee effect exists during the growth of prey species [28,29]. Allee effect is any mechanism guiding a positive correlation between a component of individual fitness and population density. This means that population at low density will increase the risk of extinction because of positive correlation between density and growth rate [30,31,32]. In the past several years, many researchers have shown a keen interest in population models with Allee effect and proposed several interesting mathematical models [33,34]. Gonzˊalez-Olivares et al. [35] found Allee effect induces the equilibria to loss their stability, and may be a destabilizing force. The species will extinct due to strong Allee effect when the density of prey population is low, while weak Allee effect does not cause species to die out [32]. Huang et al. [30] illustrated that the Allee effect or the fear effect plays a negligible role in the density of prey, but it influences the predator population greatly.

    Besides Allee effect, prey refuge should be considered in predator-prey models. Generally, preys have refuges where they can avoid predators. Therefore, to better describe the interactions between predator and prey, the inclusion of prey refuge is necessary in predator-prey models. As many researchers have made rich achievements in the predator-prey model with prey refuge, it has gradually become the focus of research in this area [36,37,38]. Chen et al. [39] found that prey refuge plays a negligible role in the persistence and extinction of the prey and the predator, but it influences the population density greatly. In [40], the authors concluded that the prey refuge parameter affect the dynamic behavior of the model greatly. As the amount of refuge increases, the prey density rises resulting in an outbreak of prey population.

    In this paper, under the assumption that the prey population growth rate obeys logistic type, we describe the predation by using the prey dependent functional response. Additionally, we consider nonconstant mortality, Allee effect and prey refuge, and propose the following predator-prey model:

    {dxdt=ax(1xK)(xm)b(1θ)xy,dydt=eb(1θ)xyγ+δy1+yy, (2)

    where K and a describe the carrying capacity of prey and the intrinsic growth rate without predators, respectively. m is the Allee effect threshold of the prey species, θ is the refuge protecting of the prey (0<θ<1), e and b respectively represent the conversion efficiency of predator by consuming prey and the predation rate, γ and δ are the same as above.

    The rest of the article is arranged as follows. In Section 2, we discussed the existence of the equilibria and their stability in detail. Subsequently, the bifurcation analysis of model (2) is given in Section 3. Some numerical simulations are presented to verify the theoretical results in Section 4. Finally, conclusions are presented in Section 5.

    Firstly, model (2) can be rewrited as below:

    {dxdt=b(1θ)x(f(x)y),dydt=eb(1θ)y(xg(y)), (3)

    here,

    f(x)=ab(1θ)(1xK)(xm),g(y)=1eb(1θ)γ+δy1+y.

    By direct calculation, we can get

    f(x)=ab(1θ)(1+mK2xK),g(y)=1eb(1θ)δγ(1+y)2.

    According to the biological significance of the model variables, set of definitions for model (3) is

    R+0×R+0={(x,y)R2x0,y0}.

    According to model (3), we can draw the following results.

    About the boundary equilibria of model (3), we have

    (i) The origin E0=(0,0);

    (ii) The predator free equilibria E1=(m,0) and E2=(K,0).

    The positive equilibria should satisfy the following equation

    {f(x)y=0,xg(y)=0. (4)

    Obviously, the above equation is equivalent to y=aKb(1θ)(Kx)(xm) and H(x)=0. Here,

    H(x)=aeb(1θ)x3(Kaeb(1θ)+maeb(1θ)+aδ)x2+(Kmaeb(1θ)+aδK+maδKeb2(1θ)2)x+(Kb(1θ)γmaδK).

    If Δ>0, then the root of dH(x)dx=0 is

    xA=2(Kaeb(1θ)+maeb(1θ)+aδ)Δ6aeb(1θ),
    xB=2(Kaeb(1θ)+maeb(1θ)+aδ)+Δ6aeb(1θ),

    where

    Δ=4(Kaeb(1θ)+maeb(1θ)+aδ)212aeb(1θ)(Kmaeb(1θ)+aδK+maδKeb2(1θ)2).

    By direct calculation, we have

    H(K)=Kb(1θ)(γKeb(1θ)),H(m)=Kb(1θ)(γmeb(1θ)).

    For x=g(y), (γeb(1θ),0) is where it intersects the X-axis, and for y=f(x), (m,0) and (K,0) are where it intersects the X-axis. We need to consider the following situations to find all the positive equilibria:

    (i) γ<eb(1θ)m;

    (ii) γ=eb(1θ)m;

    (iii) eb(1θ)m<γ<eb(1θ)K;

    (iv) γeb(1θ)K.

    When γeb(1θ)K, the model has no positive equilibrium. That is to say, if there exist positive equilibria, then γ<eb(1θ)K, i.e., H(K)<0.

    In conclusion, when m<xA<K, H(xA)=0 and γ<eb(1θ)m, the model has a unique positive equilibrium E=(x,y), where x=xA and y=a(Kx)(xm)Kb(1θ); when m<xA<K, H(xA)>0 and γ<eb(1θ)m, E3=(x3,y3) and E4=(x4,y4) are two positive equilibria of the model, where H(x3)=0, H(x4)=0 and y3,4=a(Kx3,4)(x3,4m)Kb(1θ); when m<xA<K, γ=eb(1θ)m or eb(1θ)m<γ<eb(1θ)K, the model has a unique positive equilibrium E4=(x4,y4).

    About the stability of all the equilibria, we get the following theorems.

    Theorem 2.1. (i) E0=(0,0) is always a stable node.

    (ii) E1 is a saddle if γ>eb(1θ)m; E1 is an unstable node if γ<eb(1θ)m; E1 is a repelling saddle-node if γ=eb(1θ)m.

    (iii) E2 is a stable node if γ>eb(1θ)K; E2 is a saddle if γ<eb(1θ)K; E2 is an attracting saddle-node if γ=eb(1θ)K.

    Proof. The Jacobian matrix of model (3) is

    J(x,y)=(J11J12J21J22),

    here, J11=b(1θ)(f(x)y)+b(1θ)xf(x), J12=b(1θ)x, J21=eb(1θ)y, J22=eb(1θ)(xg(y))eb(1θ)yg(y).

    By direct calculation, we can get

    (i) The eigenvalues of J at E0=(0,0) are λ1=ma<0 and λ2=γ<0. Therefore, E0=(0,0) is a stable node.

    (ii) The eigenvalues of J at E1=(m,0) are λ1=ma(1mK)>0 and λ2=eb(1θ)mγ. If γ>eb(1θ)m, then λ2<0, E1 is a saddle; if γ<eb(1θ)m, then λ2>0, E1 is an unstable node. For the case γ=eb(1θ)m, we need to discuss further.

    Let X=xm and Y=y, the model can be written as

    {dXdt=a10X+a01Y+a20X2+a02Y2+a11XY+P1(X,Y),dYdt=b10X+b01Y+b20X2+b02Y2+b11XY+Q1(X,Y), (5)

    where P1(X,Y) and Q1(X,Y) are C functions at least of third order with respect to (X,Y) and

    a10=ma(1mK),a01=mb(1θ),a20=a2amK,a02=0,a11=b(1θ),
    b10=0,b01=eb(1θ)mγ,b20=0,b02=γδ,b11=eb(1θ).

    Let X=xa01a10y, Y=y and τ=a10t, the model becomes

    {dxdτ=x+a20a10x2+[a01a210(b02a01b11a10)+a20a201a310a11a01a210]y2+(a11a10+a01b11a2102a20a01a210)xy+P2(x,y),dydτ=(b02a10a01b11a210)y2+b11a10xy+Q2(x,y). (6)

    By Theorem 7.1 in [41], since the coefficient of y2 is b02a10a01b11a210=γb(1θ)ma(1mK)(δγ)m2a2(1mK)20 and a10=ma(1mK)>0, the equilibrium E1 is a repelling saddle-node.

    (iii) At E2=(K,0), the eigenvalues of J are λ1=a(Km)<0 and λ2=eb(1θ)Kγ. When γ>eb(1θ)K, E2 is a stable node due to λ2<0. When γ<eb(1θ)K, E2 is a saddle due to λ2>0.

    In the following, we will discuss the case γ=eb(1θ)K.

    According to the translation X=xK and Y=y to transform E2 to (0,0), we have

    {dXdt=a10X+a01Y+a20X2+a02Y2+a11XY+P1(X,Y),dYdt=b10X+b01Y+b20X2+b02Y2+b11XY+Q1(X,Y), (7)

    where P1(X,Y) and Q1(X,Y) are C functions at least of third order with respect to (X,Y), and

    a10=a(Km),a01=Kb(1θ),a20=2a+amK,a02=0,a11=b(1θ),
    b10=0,b01=0,b20=0,b02=γδ,b11=eb(1θ).

    Let X=xa01a10y, Y=y and τ=a10t, the model becomes

    {dxdτ=x+a20a10x2+[a01a102(b02a01b11a10)+a20a012a103a11a01a102]y2+(a11a10+a01b11a1022a20a01a102)xy+P2(x,y),dydτ=(b02a10a01b11a102)y2+b11a10xy+Q2(x,y). (8)

    The coefficient of y2 is b02a10a01b11a102=δγa(Km)+Keb2(1θ)2a2(Km)2>0 and a10=a(Km)<0, so the equilibrium E2 is an attracting saddle-node by Theorem 7.1 in [41].

    That is the end of the proof.

    Theorem 2.2. If the condition m<xA<K,H(xA)=0 and γ<eb(1θ)m holds, the unique positive equilibrium E=(x,y) exists in model (3).

    (i) When c10+d01=0, E is a degenerate critical point (cusp). In addition, if f20(f11+2e20)0, then E is a cusp of codimension 2; else E is a cusp of codimension at least 3.

    (ii) When a10+b010 and h200, if f01>0, then E is a repelling saddle-node; else E is an attracting saddle-node.

    Proof. According to the Jacobian matrix of model (3), we can get

    det(J(E))=eb2(1θ)2xy(1f(x)g(y)),
    tr(J(E))=b(1θ)xf(x)eb(1θ)yg(y).

    Obviously, det(J(E))=0.

    Letting X=xx, Y=yy to transform E to (0,0) for studying the stability of E. We can rewrite model (3) as

    {dXdt=c10X+c01Y+c20X2+c11XY+c02Y2+R1(X,Y),dYdt=d10X+d01Y+d20X2+d11XY+d02Y2+S1(X,Y), (9)

    where R1(X,Y) and S1(X,Y) are C functions at least of third order with respect to (X,Y) and

    c10=x[aK(xm)+a(1xK)],c01=b(1θ)x,c20=maK+a3aKx,c11=b(1θ),
    c02=0,d10=eb(1θ)y,d01=(γδ)y(1+y)2,d20=0,d02=γδ(1+y)3,d11=eb(1θ).

    At E=(x,y), the Jacobian matrix of model (3) is

    J(E)=(c10c01d10d01),

    where det(J(E))=c10d01c01d10=0. Then the eigenvalues of J(E) are λ1=0, λ2=c10+d01.

    Case 1: c10+d01=0.

    Obviously, the eigenvalues of J(E) are λ1=λ2=0. Through the transformation of (XY)=(c010c101)(xy), model (9) becomes

    {˙x=y+e20x2+e11xy+R2(x,y),˙y=f20x2+f02y2+f11xy+S2(x,y). (10)

    By Lemma in [44], we can obtain the equivalent model of model(10) as follows:

    {˙x=y,˙y=f20x2+(f11+2e20)xy+o(|x,y|2). (11)

    If f20(f11+2e20)0, then E is a cusp of codimension 2; else E is a cusp of codimension at least 3.

    Case 2: c10+d010.

    Obviously, the eigenvalues of J(E) are λ1=0 and λ20. Through the transformation of (XY)=(d01c10d10d10)(xy), model (9) becomes

    {˙x=e20x2+e02y2+e11xy+R2(x,y),˙y=f01y+f20x2+f02y2+f11xy+S2(x,y), (12)

    where R2(x,y) and S2(x,y) are C functions at least of third order with respect to (x,y) and

    e20=1c10+d01(c10d11d01+c20d201c11d01d10c10d02d10),
    e02=1c10+d01(c20c210+c11c10d10d11c210c10d02d10),
    e11=1c10+d01(2c10d02d10+c11d01d10+2c20d01c10+d11c210c11c10d10d01d11c10),
    f01=1c10+d01(c210+c01d10+d01c10+d201),
    f20=1c10+d01(c20d201+d01d02d10c11d01d10d11d201),
    f02=1c10+d01(c11c10d10+c20c210+d01d11c10+d01d02d10),
    f11=1c10+d01(c11d01d10+2c20d01c10+d11d201c11c10d10d01d11c102d01d02d10).

    Redefine τ by τ=f01t, we can get

    {˙x=h20x2+h02y2+h11xy+R3(x,y),˙y=y+k20x2+k02y2+k11xy+S3(x,y), (13)

    where R3(x,y) and S3(x,y) are C functions at least of third order with respect to (x,y) and hij=eijf01,kij=fijf01. The coefficient of x2 is

    h20=e20f01=c10d11d01+c20d201c11d01d10c10d02d10c210+c01d10+d01c10+d201.

    If h200, the equilibrium E is a repelling saddle-node if f01>0 by Theorem 7.1 in [41]; otherwise it is an attracting saddle-node.

    That is the end of the proof.

    Theorem 2.3. When m<xA<K, H(xA)>0 and γ<eb(1θ)m, model (3) has two positive equilibria E3=(x3,y3) and E4=(x4,y4).

    (i) E3=(x3,y3) is always a saddle.

    (ii) if x4[m+K2,K), then E4=(x4,y4) is local asymptotically stable; if x4(m,m+K2), then E4=(x4,y4) may be a focus or a center or a node.

    Proof. According to det(J(E)) and tr(J(E)) in Theorem 2.2.

    (i) If x3(m,m+K2), then det(J(E3))=eb2(1θ)2x3y3(1f(x3)g(y3))<0. Therefore, E3=(x3,y3) is a saddle.

    (ii) If x4[m+K2,K), then det(J(E4))>0 and tr(J(E4))=b(1θ)x4f(x4)eb(1θ)y4g(y4)<0 due to f(x4)0,g(y4)>0. Thus, E4=(x4,y4) is local asymptotically stable. If x4(m,m+K2), then det(J(E4))=eb2(1θ)2x4y4(1f(x4)g(y4))>0 since 1g(x4)>f(x4). The sign of tr(J(E4)) is uncertain, so E4=(x4,y4) may be a focus or a center or a node.

    That is the end of the proof.

    For the stability of the unique positive equilibrium E4 in other cases, the analysis is the same as E4 in Theorem 2.3, and we omit it here.

    We are interested in various possible bifurcations of model (3) in this section, including transcritical, saddle-node, Hopf and Bogdanov-Takens bifurcation.

    According to Theorem 2.1, the boundary equilibrium E2 is unstable when γ<eb(1θ)K, while it is stable when γ>eb(1θ)K. Particularly, when γ=eb(1θ)K, E2 coincides with the positive equilibrium E4. Therefore, we consider the existence of transcritical bifurcation around E2 by taking γ as bifurcation parameter.

    Theorem 3.1. A transcritical bifurcation occurs in model (3) when γ=eb(1θ)KγTC.

    Proof. We have det(J(E2))=0 when γ=eb(1θ)K. Hence, there is a zero eigenvalue for J(E2). For zero eigenvalue, the eigenvectors corresponding to matrices J(E2) and JT(E2) are denoted as V and W, respectively. After a simple calculation, we have

    V=(V1V2)=(b(1θ)a(1mK)1),W=(W1W2)=(01).

    Moreover,

    Fγ(E2;γTC)=(0y1+y)(E2;γTC)=(00),
    DFγ(E2;γTC)V=(01(1+y)2)(E2;γTC)=(01),
    D2F(E2;γTC)(V,V)=(2F1x2V12+22F1xyV1V2+2F1y2V222F2x2V12+22F2xyV1V2+2F2y2V22)(E2;γTC)=((2(Km)2a)b2(1θ)2K2a2(Km)2+2b(1θ)b(1θ)Ka(Km)2eb(1θ)b(1θ)Ka(Km)2(δeb(1θ)K)).

    For the transversality conditions, we have

    WTFγ(E2;γTC)=0,WT[DFγ(E2;γTC)V]=10,
    WT[D2F(E2;γTC)(V,V)]=2eb(1θ)b(1θ)Ka(Km)2(δeb(1θ)K)0.

    A transcritical bifurcation occurs at E2 when γ=eb(1θ)K according to Sotomayor's theorem.

    This is the end of the proof.

    From the previous analysis, we can see clearly that under the condition m<xA<K and γ<eb(1θ), model (3) has two positive equilibria E3=(x3,y3) and E4=(x4,y4) if H(xA)>0. The two positive equilibria coincide if H(xA)=0, while they disappear if H(xA)<0. Therefore, a saddle-node bifurcation will occur at E=(x,y). Saddle-node bifurcation at E can be summarized into the following theorem.

    Theorem 3.2. A saddle-node bifurcation occurs in model (3) when γ=γ1γSN, where γ1 satisfies H(xA)=0.

    Proof. According to Theorem 2.2, there is a zero eigenvalue λ1 for J(E). For zero eigenvalue, the eigenvectors corresponding to matrices J(E) and JT(E) are denoted as V and W, respectively.

    V=(V1V2)=(c01c101)=(b(1θ)aK(xm)+a(1xK)1),
    W=(W1W2)=(d10c101)=(eb(1θ)yx(aK(xm)+a(1xK))1).

    Furthermore, we can get

    Fγ(E;γSN)=(0y1+y)(E;γSN)=(0y1+y),
    D2F(E;γSN)(V,V)=(2F1x2V21+22F1xyV1V2+2F1y2V222F2x2V21+22F2xyV1V2+2F2y2V22)(E;γSN)=(2ab2(1θ)2xK(aK(xm)+a(1xK))22eb2(1θ)2aK(xm)+a(1xK)2(δγ)(1+y)3).

    V and W satisfy the transversality conditions

    WTFγ(E;γSN)=y1+y0,
    WT[D2F(E;γSN)(V,V)]=2aeb3(1θ)3xyKx[aK(xm)+a(1xK)]3+2eb2(1θ)2aK(xm)+a(1xK)+2(γδ)(1+y)20.

    A saddle-node bifurcation occurs at E when γ=γSN according to Sotomayor's theorem.

    This is the end of the proof.

    The positive equilibrium E4 may be a focus from Theorem 2.3. In this subsection, the conditions for Hopf bifurcation at E4 will be discussed. Obviously, det(J(E4)|γ=γH)>0, and there exists γγH such that tr(J(E4)|γ=γH)=0.

    Theorem 3.3. A Hopf bifurcation occurs at E4 in model (3) when γ=γH.

    Proof. According to tr(J(E4)), we have

    ddγtr(J(E4)|γ=γH)=y4(1+y4)2|γ=γH0,

    which satisfies the transversality condition for Hopf bifurcation.

    For the stability of the limit cycle, we need to calculate the first Lyapunov number l1 at E4 first.

    Letting X=xx4, Y=yy4 to transform E4 to (0,0) and we rewrite model (3) as

    {dXdt=l10X+l01Y+l11XY+l20X2+l02Y2+l30X3+l21X2Y+l12XY2+l03Y3+P4(X,Y)dYdt=n10X+n01Y+n11XY+n20X2+n02Y2+n30X3+n21X2Y+n12XY2+n03Y3+Q4(X,Y), (14)

    where

    l10=x4[aK(x4m)+a(1x4K)],l01=b(1θ)x4,l11=b(1θ),l20=maK+a3aKx4,
    l30=a2K,l21=l02=l12=l03=0,n10=eb(1θ)y4,n01=(γδ)y4(1+y4)2,
    n11=eb(1θ),n02=γδ(1+y4)3,n03=δγ2(1+y4)4,n20=n21=n12=n30=0,

    where P4(X,Y) and Q4(X,Y) are the power series in (X,Y) with the term XiYj satisfying i+j4.

    For the first Lyapunov number l1, which is given by the formula

    l1=3π2l01Δ32((l10n10(l211+l11n02+l02n11)+l10l01(n211+l20n11+l11n02)+n210(l11l02+2l02n02)2l10n10(n202l20l02)2l10l01(l220n20n02)l201(2l20n20+n11n20)+(l01n102l210)(n11n02l11l20))(l210+l01n10)(3(n10n03l01l30)+2l10(l21+n12)+(n10l12l01n21))),

    where Δ=det(J(E4))=l10n01l01n10.

    The limit cycle is stable via a supercritical Hopf bifurcation if l1<0, and it is unstable via a subcritical Hopf bifurcation if l1>0.

    According to Theorem 3.2, positive equilibrium appears via saddle-node bifurcation. The stability of E4 may change by Hopf bifurcation in Theorem 3.3. When saddle-node bifurcation crosses Hopf bifurcation, a Bogdanov-Takens (BT) bifurcation will occur under a small parameter perturbation in model (3). By choosing γ and m as two bifurcation parameters, γBT and mBT are the threshold value, i.e., det(J(E)|(γ,m)=(γBT,mBT))=0 and tr(J(E)|(γ,m)=(γBT,mBT))=0).

    Theorem 3.4. A Bogdanov-Takens bifurcation of codimension 2 will occur in a small neighborhood of E when (m,γ) varies near (mBT,γBT) with a10+b01=0 and g20g110.

    For the proof of the theorem, please see Appendix 5.

    In this section, several numerical simulations are provided to verify the reliability of the theoretical results under the parameters conditions as a=0.8, b=0.6, e=0.8, θ=0.1, δ=0.9, K=0.9, m=0.2. The parameter γ is chosen as the bifurcation parameter.

    The possible equilibria of model (3) are discussed in the Section 2 and the qualities of them are also given, through which we plot a bifurcation diagram (Figure 1) and the phase portraits (Figures 25). It can be seen from Figure 2(a) that for a small γ, there exist equilibria E0, E1 and E2. As γ increases and reaches the line L1, E is the unique interior equilibrium in model (3) (see Figure 2(b)). With the increase of the parameter γ, the positive equilibria E3 and E4 appear as the taking place of a saddle-node bifurcation in model (3). From Figure 2(c), it is clearly that both E3 and E4 are unstable near the line L1, the former is a saddle and the later is a node.

    Figure 1.  Bifurcation graph in the plane of γx for a=0.8, b=0.6, e=0.8, θ=0.1, δ=0.9, K=0.9, m=0.2. S0, S1 and S2 stand for the three boundary equilibria E0, E1 and E2, respectively. S3 and S4 represent the two positive equilibria E3 and E4. The solid and dotted lines denote stable and unstable equilibria.
    Figure 2.  Phase portraits (a)(c): (a) There exists three boundary equilibria E0, E1 and E2; (b) There exists an interior equilibrium E (E3 coincides with E4) and three boundary equilibria; (c) Three boundary equilibria E0, E1, E2 and two positive equilibria E3 and E4. E3 is a saddle.
    Figure 3.  Phase portraits (a)-(d): (a) E4 is a center when γ=0.05400367556=γH; (b) Local amplification of (a); (c) E4 is an unstable focus when γ=0.0535<γH; (d) E4 is a stable focus when γ=0.056>γH. The equilibrium E3 is always a saddle in all of the above cases.
    Figure 4.  Phase portraits (a)-(c): (a) E2 is an unstable saddle and E4 is a stable node; (b) E2 is an attracting saddle-node; (c) E2 is a stable node.
    Figure 5.  Phase portraits (a)(c): (a) The equilibrium E is a critical cusp of degradation; (b) The equilibrium E is a repelling saddle-node; (c) The equilibrium E is an attracting saddle-node.

    As the parameter γ increases and crosses the line L2 (γH=0.05400367556), a Hopf bifurcation occurs. The first Lyapunov number l1=6.414584574π is greater than zero, which indicates that the direction of Hopf bifurcation is subcritical. Thus, an unstable limit cycle emerges bifurcating from the positive equilibrium E4, shown in Figure 3(a), (b). Additionally, the positive equilibrium E4 recovers its stability owing to the occurrence of Hopf bifurcation. The equilibrium E4 is an unstable focus in region II (γ<γH, see Figure 3(c)), and becomes a stable focus in region III (γ>γH, see Figure 3(d)). In region IV, E4 is the unique positive equilibrium in model (3), which coincides with the boundary equilibrium E2 when γ reaches the line L4. Specially, on the left of line L4 (region IV), the boundary equilibrium E2 is a saddle (see Figure 4(a)), while on the right (region V), E2 is a stable node (see Figure 4(c)). As γ increases and crosses the line L4, E2 changes from unstable to stable. Therefore, we can conclude that transcritical bifurcation occurs.

    Moreover, Theorem 3.4 indicates that a BT bifurcation with codimension 2 occurs around the equilibrium E in model (3). We choose γ and m as the control parameters to numerically study BT bifurcation. Direct numerical computation shows that BT bifurcation may occur when m=mBT=0.1347130155 and γ=γBT=0.003455490134. The equilibrium E is a cusp when m=mBT and γ=γBT, which is dispalyed in Figure 5(a). For m=0.4 and γ=0.1682189296, the positive equilibrium E is a repelling saddle-node (see Figure 5(b)). However, it becomes an attracting saddle-node when m=0.1301 and γ=0.00005717997942 (see Figure 5(c)).

    We present the bifurcation diagram of model (3) in Figure 6 to illustrate all the possible dynamic behaviors in the γm parameter plane, where the curve SN, H and T divide the parameter plane into four regions. In region I, model (3) has no positive equilibrium. However, two positive equilibria E3 and E4 appear through the saddle-node bifurcation in region II. On the curve H, the model undergoes a Hopf bifurcation which leads to the appearance of a limit cycle. In addition, as the parameters go through the transcritical bifurcation curve (curve T) and enter region IV, the positive equilibrium E4 disappears. Particularly, when the curve SN meets the curve H at the BT point, model (3) may undergo a BT bifurcation of codimension 2.

    Figure 6.  Bifurcation diagram of model (3) in the γm plane. The blue, red and black curves represent the curves SN, H and T respectively. The point BT denotes the Bogdanov-Taken bifurcation point.

    For m=mBT and γ=γBT, we can obtain |(ζ1,ζ2)(ϵ1,ϵ2)|ϵ1=ϵ2=00. Thus, the parametric transformation ζ1(ϵ)=γ00(ϵ)γ411(ϵ), ζ2(ϵ)=γ01(ϵ)γ11(ϵ) is nonsingular. For small ϵ1 and ϵ2, θ20>0 and θ110, which means that BT bifurcation with codimension 2 occurs in model (3). Under the second-order approximation, the bifurcation curves can be locally expressed as:

    1) SN: the local expression of the saddle-node bifurcation curve

    {(ϵ1,ϵ2):20.56464756ϵ127.99484334ϵ2+853.8953780ϵ213562.665889ϵ1ϵ2+3243.982017ϵ22,ζ2(ϵ1,ϵ2)0}

    2) H: the local expression of the Hopf bifurcation curve

    {(ϵ1,ϵ2):20.56464756ϵ127.99484334ϵ2+868.9514392ϵ213529.702770ϵ1ϵ2+3262.024034ϵ22,ζ1(ϵ1,ϵ2)<0}

    3) HL: the local expression of the homoclinic bifurcation curve

    {(ϵ1,ϵ2):20.56464756ϵ127.99484334ϵ2+883.4052627ϵ213498.058175ϵ1ϵ2+3279.344371ϵ22,ζ1(ϵ1,ϵ2)<0}

    A small area at the origin in the (ϵ1,ϵ2) parameter plane is divided into four regions by the above three local bifurcation curves, as is shown in Figure 7. When (ϵ1,ϵ2)=(0,0), Figure 8(a) suggests that E is a cusp with codimension 2, and it is the unique positive equilibrium in model (3). In region I of Figure 7, there exists three boundary equilibria, but no positive equilibrium (see Figure 8(b)). There exists one positive saddle-node in model (3) when (ϵ1,ϵ2) is on the saddle-node bifurcation curve SN. Passing through the curve SN from region I to region II, an unstable saddle and an unstable node appear via saddle-node bifurcation (see Figure 8(c)). Furthermore, on the Hopf bifurcation curve H, a limit cycle emerges via Hopf bifurcation. The limit cycle bifurcating from the positive equilibrium is unstable, since there occurs a subcritical Hopf bifurcation (see Figure 8(d)). When (ϵ1,ϵ2) reaches the homoclinic bifurcation curve HL, the limit cycle becomes an unstable homoclinic one shown in Figure 8(e). Finally, the unstable homoclinic cycle disappears in region IV. From Figure 8(f), we can see clearly that a saddle and a stable focus can coexist near the curve HL.

    Figure 7.  Bifurcation diagram of model (3) with m=mBT=0.1347130155 and γ=γBT=0.003455490134. The red, green and blue curves represent the curves SN, HL and H respectively.
    Figure 8.  Phase portraits of model (15) (a)(f): (a) A codimension 2 cusp for (ϵ1,ϵ2)=(0,0); (b) No interior equilibrium for (ϵ1,ϵ2)=(0.02,0.008) in region I; (c) There exists an unstable saddle and an unstable node in region II for (ϵ1,ϵ2)=(0.01096442,0.008); (d) There exists an unstable limit cycle for (ϵ1,ϵ2)=(0.010714,0.008); (e) An unstable homoclinic orbit for (ϵ1,ϵ2)=(0.010478,0.008) on the curve HL; (f) There exists a stable focus for (ϵ1,ϵ2)=(0.005,0.008) in region IV.

    The bifurcation diagram of model (3) according to the parameters γ and δ is shown in Figure 9(a). When γ and δ are small, model (3) only has boundary equilibria (see region I in Figure 9(a)), and the equilibrium (0,0) is a stable node. For a small γ, as δ increases from region I to region II, two positive equilibria coexist (see Figure 9(a)). For a large δ, with the increase of γ from region II to region III, two positive equilibria coincide and become one equilibrium via saddle-node bifurcation (see Figure 9(a)). Due to transcritical bifurcation, the unique positive equilibrium coincides with the boundary equilibrium. In region IV, there are only three boundary equilibria, and (K,0) is a stable node. In addition, Hopf bifurcation may occur on the curve H. Particularly, a Bogdanov-Takens bifurcation of codimension 2 takes place at the BT point in model (3), once the curves SN and H meet.

    Figure 9.  (a) Bifurcation diagram in the γδ plane of model (3) when m=0.1347. The curves SN and L denote the saddle node bifurcation, and T and H represent the transcritical bifurcation and the Hopf bifurcation, respectively. The BT point is the point where occurs the Bogdanov-Takens bifurcation; (b) The critical value of m for saddle-node bifurcation (SN) in the γδ plane (γ×δ=(0,0.058)×(0,1)), where the red line represents γ=δ.

    For γ=δ, when γ is small, (0,0) is a stable node (see the red dashed line in Figure 9(a)). When γ is large, (0,0) and (K,0) are the only two boundary equilibria, and both of them are stable node (see the red dotted line in Figure 9(a)). However, when γ is at the middle region, there exists olny one positive equilibrium in the model (see the red solid line in Figure 9(a)). Unlike γ=δ, the dynamics of predator-prey model with γ<δ are complicated when γ is small (see Figure 9(a)). We also trace the critical value of m for saddle-node bifurcation in Figure 9(b). From Figure 9(b), we can see that for γ×δ=(0,0.058)×(0,1), there is always a m(0,1) satisfying the occurrence conditions of saddle-node bifurcation.

    A predator-prey model with strong Allee effect and nonconstant mortality rate is studied in this paper. The aim is to explore what dynamic behaviors will occurs in the predator-prey model with a strong Allee effect caused by the nonconstant mortality. Our results indicate that there always exists three boundary equilibria in model (3), where (0,0) is always a stable node, (K,0) is stable only if γ>eb(1θ)K and (m,0) is always unstable whenever exists. As for positive equilibria, when γ<eb(1θ)m and m<xA<K, there is a unique positive equilibrium when H(xA)=0 in model (3), and the occurrence of saddle-node bifurcation brings two positive equilibria when H(xA)>0. A limit cycle appears with the positive equilibrium undergoing a Hopf bifurcation at γ=γH, whose stability is determined by the first Lyapunov number l1. Additionally, the positive equilibrium coincides with (K,0) at γ=eb(1θ)K via transcritical bifurcation.

    When the curve H meets the curve SN, a BT bifurcation of codimension 2 occurs. The BT bifurcation point is sensitive to the perturbation intensity. The predator-prey model can generate rich dynamics when we introduce two small perturbation into the BT bifurcation parameters γBT and mBT. Two positive equilibria exist in model (3), one is a saddle, and the other undergoes a Hopf bifurcation then produces limit cycles. A homoclinic cycle emerges bifurcating from the limit cycle via homoclinic bifurcation. While for a large perturbation to γBT and mBT, there are only boundary equilibria.

    For γ=δ in model (3), the predator has a constant mortality rate. There is a γ such that model (3) only has boundary equilibria for 0<γ<γ, and (0,0) is a stable node. However, when γ<δ, the mortality rate of predator is nonlinear. There is a δ such that two positive equilibria exist in model (3) and the model undergoes a saddle-node bifurcation for 0<γ<γ and δ>δ. Moreover, when 0<γ<γ, a m satisfying the occurrence conditions of saddle-node bifurcation always exists. These results suggest that the predator dies out (no positive equilibrium) under constant mortality rate, but nonconstant mortality can result in the coexistence of predator and prey (positive equilibrium exists). Obviously, nonconstant mortality rate is the cause of the emergence of two positive equilibria. Therefore, nonconstant mortality plays an important role in the dynamics of predator-prey model, and we cannot simply approximate nonconstant mortality rate to a constant, especially for small γ.

    From the ecological view point, when the mortality γ at low density and the limiting, maximal mortality rate δ are small, both the prey and the predator will go extinct due to Allee effect. Interestingly, the prey and the predator will reach a coexistence state when the mortality γ is small enough and the limiting, maximal death rate δ is large enough. The reason is that the increase in the mortality rate of predator relieves the feeding pressure on the prey and reduces the competition among predators for prey. However, when γ is large, (K,0) is a asymptotically stable boundary equilibrium. This means that the predator dies out and the prey increases quickly to reach its carrying capacity K when the mortality rate of predator is too high. Therefore, we can approximate the nonconstant mortality rate to a constant value when γ is large. Moreover, prey refuge plays a negligible role in the persistence and extinction of the prey and the predator, but it influences the population density greatly. It is worth noting that the diffusion can also affect the dynamics of predator-prey model with strong Allee effect significantly [42,43]. A detailed analysis on the effect of diffusion is beyond the scope of this work and will be presented in the future.

    In short, the nonconstant mortality of predator do have a significant impact on the dynamics of predator-prey model. We cannot simply approximate the nonconstant death rate to a constant unless the mortality rate is large. We expect that our results provide a new insight into the predator-prey model with nonconstant mortality rate.

    This work was supported by the National Natural Science Foundation of China (Grants No. 61901303 and No. 61871293), the National Key Research and Development Program of China (Grant No. 2018YFE0103700), and Wenzhou basic agricultural science and technology project (Grant No. S20180006).

    The author declares that there are no conflicts of interest.

    The proof of Theorem 3.4

    Proof. To obtain the local critical expressions of the bifurcation curves around the BT point, firstly we need to transform model (3) into a normal form of BT bifurcation.

    Introducing two small perturbations to the parameters γ and m at the critical BT bifurcation value through m=mBT+ϵ1 and γ=γBT+ϵ2, where ϵ(ϵ1,ϵ2) is extremely close to zero vector. Under this operation, model (3) can be rewritten as

    {˙x=x(a(1xK)(xmBTϵ1)b(1θ)y),˙y=y(eb(1θ)xγBT+ϵ2+δy1+y). (A1)

    Then we can introduce the transformations n1=xx and n2=yy to let (0,0) be the bifurcation point, it follows that

    {˙n1=p00(ϵ)+p10(ϵ)n1+p01(ϵ)n2+p20(ϵ)n21+p11(ϵ)n1n2+T1(n1,n2,ϵ),˙n2=q00(ϵ)+q10(ϵ)n1+q01(ϵ)n2+q20(ϵ)n21+q02(ϵ)n22+q11(ϵ)n1n2+R1(n1,n2,ϵ), (A2)

    where

    p00(ϵ)=x(a(1xK)(xmBTϵ1)b(1θ)y),
    p10(ϵ)=a(1xK)(xmBTϵ1)b(1θ)y+x(a(xmBTϵ1)K+a(1xK)),
    p01(ϵ)=b(1θ)x,p20(ϵ)=a(12xK)a(xmBTϵ1)K,
    p11(ϵ)=b(1θ),q00(ϵ)=y(eb(1θ)xγBT+ϵ2+δy1+y),
    q10(ϵ)=eb(1θ)y,q01(ϵ)=eb(1θ)xγBT+ϵ2+δy1+y+y(δ1+y+γBT+ϵ2+δy(1+y)2),
    q20(ϵ)=0,q11(ϵ)=eb(1θ),
    q02(ϵ)=δ1+y+γBT+ϵ2+δy(1+y)2+y(δ(1+y)2γBT+ϵ2+δy(1+y)3),

    and T1(n1,n2,ϵ) and R1(n1,n2,ϵ) are at least of the third order with terms ni1nj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Next, we introduce the following C variable substitutions near the origin

    {z1=n1,z2=p00(ϵ)+p10(ϵ)n1+p01(ϵ)n2+p20(ϵ)n21+p11(ϵ)n1n2+T1(n1,n2,ϵ). (A3)

    Model (A2) becomes

    {˙z1=z2,˙z2=η00(ϵ)+η10(ϵ)z1+η01(ϵ)z2+η20(ϵ)z21+η11(ϵ)z1z2+η02(ϵ)z22+R2(z1,z2,ϵ), (A4)

    where

    η00(ϵ)=p01(ϵ)q00(ϵ)q01(ϵ)p00(ϵ)+q02(ϵ)p200(ϵ)p01(ϵ),
    η10(ϵ)=p01(ϵ)q10(ϵ)+p11(ϵ)q00(ϵ)q01(ϵ)p10(ϵ)+2q02(ϵ)p00(ϵ)p10(ϵ)p01(ϵ)p00(ϵ)q11(ϵ),
    η01(ϵ)=p10(ϵ)+q01(ϵ)2q02(ϵ)p00(ϵ)p01(ϵ)p11(ϵ)p00(ϵ)p01(ϵ),
    η20(ϵ)=p01(ϵ)q20(ϵ)+p11(ϵ)q10(ϵ)q01(ϵ)p20(ϵ)+q02(ϵ)p01(ϵ)p210(ϵ)q11(ϵ)p10(ϵ),
    η02(ϵ)=q02(ϵ)+p11(ϵ)p01(ϵ),η11(ϵ)=2p20(ϵ)2q02(ϵ)p10(ϵ)p01(ϵ)+q11(ϵ)p10(ϵ)p11(ϵ)p01(ϵ),

    and R2(z1,z2,ϵ) is at least of the third order with terms zi1zj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Next we let τ to replace the time variable such that dt=(1η02(ϵ)z1)dτ. Model (A4) can be written as model (A5) when we rewritten t to denote τ.

    {˙z1=z2(1η02(ϵ)z1),˙z2=(1η02(ϵ)z1)(η00(ϵ)+η10(ϵ)z1+η01(ϵ)z2+η20(ϵ)z21+η11(ϵ)z1z2+η02(ϵ)z22+R2(z1,z2,ϵ)). (A5)

    Let υ1=z1,υ2=z2(1η02(ϵ)z1), then model (A5) can be expressed as

    {˙υ1=υ2,˙υ2=θ00(ϵ)+θ10(ϵ)υ1+θ01(ϵ)υ2+θ20(ϵ)υ21+θ11υ1υ2+R3(υ1,υ2,ϵ), (A6)

    where

    θ00(ϵ)=η00(ϵ),θ10(ϵ)=η10(ϵ)2η00(ϵ)η02(ϵ),
    θ01(ϵ)=η01(ϵ),θ20(ϵ)=η20(ϵ)2η02(ϵ)η10(ϵ)+η00(ϵ)η202(ϵ),
    θ11(ϵ)=η01(ϵ)η02(ϵ)+η11(ϵ),

    and R3(v1,v2,ϵ) is at least of the third order with terms υi1υj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    The express of θ20(ϵ) is complex enough, we omit it. And we can not get any information about the sign of θ20(ϵ). Therefore, the following two situations need to be considered to ensure the following process is meaningful.

    Case 1: θ20(ϵ)>0 for small ϵi(i=1,2), making the following variable substitutions

    μ1=υ1,μ2=υ2θ20(ϵ),τ=θ20(ϵ)t.

    Retaining τ as t, model (A1) can be expressed as

    {˙μ1=μ2,˙μ2=s00(ϵ)+s10(ϵ)μ1+s01(ϵ)μ2+μ21+s11μ1μ2+R4(μ1,μ2,ϵ), (A7)

    where

    s00(ϵ)=θ00(ϵ)θ20(ϵ),s10(ϵ)=θ10(ϵ)θ20(ϵ),s01(ϵ)=θ01(ϵ)θ20(ϵ),s11(ϵ)=θ11(ϵ)θ20(ϵ),

    and R4(μ1,μ2,ϵ) is at least of the third order with terms μi1μj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Model (A7) can be expressed as model (A8) under the changes of ω1=μ1+s10(ϵ)2,ω2=μ2.

    {˙ω1=ω2,˙ω2=γ00(ϵ)+γ01(ϵ)ω2+ω21+γ11(ϵ)ω1ω2+R5(ω1,ω2,ϵ), (A8)

    where

    γ00(ϵ)=s00(ϵ)14s210(ϵ),γ01(ϵ)=s01(ϵ)12s10(ϵ)s11(ϵ),γ11(ϵ)=s11(ϵ),

    and R5(ω1,ω2,ϵ) is at least of the third order with terms ωi1ωj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Assuming θ11(ϵ)0, then r11(ϵ)=s11(ϵ)=θ11(ϵ)θ20(ϵ)0. Making the following variable substitutions

    x=γ211(ϵ)ω1,y=γ311(ϵ)ω2,τ=1γ11(ϵ)t,

    then a versal unfolding form of model (A1) can be written as

    {˙x=y,˙y=ζ1(ϵ)+ζ2(ϵ)y+x2+xy+R6(x,y,ϵ), (A9)

    where

    ζ1(ϵ)=γ00(ϵ)γ411(ϵ),ζ2(ϵ)=γ01(ϵ)γ11(ϵ),

    and R6(x,y,ϵ) is at least of the third order with terms xiyj, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Case 2: θ20(ϵ)<0 for small ϵi(i=1,2), making the following variable substitutions

    μ1=υ1,μ2=υ2θ20(ϵ),τ=θ20(ϵ)t.

    Retaining τ as t, model (A1) can be expressed as

    {˙μ1=μ2,˙μ2=s00(ϵ)+s10(ϵ)μ1+s01(ϵ)μ2μ21+s11μ1μ2+R4(μ1,μ2,ϵ), (A10)

    where

    s00(ϵ)=θ00(ϵ)θ20(ϵ),s10(ϵ)=θ10(ϵ)θ20(ϵ),s01(ϵ)=θ01(ϵ)θ20(ϵ),s11(ϵ)=θ11(ϵ)θ20(ϵ),

    and R4(μ1,μ2,ϵ) is at least of the third order with terms μi1μj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Model (A10) can be expressed as model (A11) under the changes of ω1=μ1s10(ϵ)2,ω2=μ2.

    {˙ω1=ω2,˙ω2=γ00(ϵ)+γ01(ϵ)ω2ω21+γ11(ϵ)ω1ω2+R5(ω1,ω2,ϵ), (A11)

    where

    γ00(ϵ)=s00(ϵ)+14s210(ϵ),γ01(ϵ)=s01(ϵ)+12s10(ϵ)s11(ϵ),γ11(ϵ)=s11(ϵ),

    and R5(ω1,ω2,ϵ) is at least of the third order with terms ωi1ωj2, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Assuming θ11(ϵ)0, then r11(ϵ)=s11(ϵ)=θ11(ϵ)θ20(ϵ)0. Making the following variable substitutions

    x=γ211(ϵ)ω1,y=γ311(ϵ)ω2,τ=1γ11(ϵ)t.

    Then a versal unfolding form of model (A1) can be written as

    {˙x=y,˙y=ζ1(ϵ)+ζ2(ϵ)y+x2+xy+R6(x,y,ϵ), (A12)

    where ζ1(ϵ)=γ00(ϵ)γ411(ϵ),ζ2(ϵ)=γ01(ϵ)γ11(ϵ). R6(x,y,ϵ) is at least of the third order with terms xiyj, whose coefficients are smooth functions of ϵ1 and ϵ2.

    Retain ζ1(ϵ) and ζ2(ϵ) to denote ζ1(ϵ) and ζ1(ϵ) to combine the above two situations. In a small area near the origin, the transformations of ζ1(ϵ) and ζ2(ϵ) are homeomorphisms when |(ζ1,ζ2)(ϵ1,ϵ2)|ϵ1=ϵ2=0 is a nonsingular matrix, in the mean while ζ1(ϵ) and ζ2(ϵ) are independent parameters. Then we know there takes place a BT bifurcation when ϵ=(ϵ1,ϵ2) around the origin. Under the second-order approximation, the bifurcation curves can be locally expressed as (``+" for θ20(ϵ)>0, ``-" for θ20(ϵ)<0):

    1) SN: the local expression of the saddle-node bifurcation curve

    SN={(ϵ1,ϵ2):ζ1(ϵ1,ϵ2)=0,ζ2(ϵ1,ϵ2)0}

    2) H: the local expression of the Hopf bifurcation curve

    H={(ϵ1,ϵ2):ζ2(ϵ1,ϵ2)=±ζ1(ϵ1,ϵ2),ζ1(ϵ1,ϵ2)<0}

    3) HL: the local expression of the homoclinic bifurcation curve

    HL={(ϵ1,ϵ2):ζ2(ϵ1,ϵ2)=±57ζ1(ϵ1,ϵ2),ζ1(ϵ1,ϵ2)<0}


    [1] E. Sˊaez, E. Gonzˊalez-Olivares, Dynamics of a predator-prey model, SIAM J. Appl. Math., 59 (1999), 1867–1878. https://doi.org/10.1137/S0036139997318457 doi: 10.1137/S0036139997318457
    [2] R. P. Gupta, M. Banerjee, P. Chandra, Bifurcation analysis and control of Leslie-Gower predator-prey model with Michaelis–Menten type prey-harvesting, Differ. Equations Dyn. Syst., 20 (2012), 339–366. https://doi.org/10.1007/s12591-012-0142-6 doi: 10.1007/s12591-012-0142-6
    [3] D. Hu, H. Cao, Stability and bifurcation analysis in a predator-prey system with Michaelis-Menten type predator harvesting, Nonlinear Anal.: Real World Appl., 33 (2017), 58–82. https://doi.org/10.1016/j.nonrwa.2016.05.010 doi: 10.1016/j.nonrwa.2016.05.010
    [4] L. Zhang, C. Zhang, M. Zhao, Dynamic complexities in a discrete predator–prey system with lower critical point for the prey, Math. Comput. Simul., 105 (2014), 119–131. https://doi.org/10.1016/j.matcom.2014.04.010 doi: 10.1016/j.matcom.2014.04.010
    [5] C. Dai, M. Zhao, L. Chen, Dynamic complexity of an Ivlev-type prey-predator system with impulsive state feedback control, J. Appl. Math., 2012 (2012). https://doi.org/10.1155/2012/534276
    [6] C. Dai, M. Zhao, H. Yu, Dynamics induced by delay in a nutrient–phytoplankton model with diffusion, Ecol. Complexity, 26 (2016), 29–36. https://doi.org/10.1016/j.ecocom.2016.03.001 doi: 10.1016/j.ecocom.2016.03.001
    [7] C. Dai, M. Zhao, Mathematical and dynamic analysis of a prey-predator model in the presence of alternative prey with impulsive state feedback control, Discrete Dyn. Nat. Soc., 2012 (2012), 724014. https://doi.org/10.1155/2012/724014 doi: 10.1155/2012/724014
    [8] C. Dai, M. Zhao, Bifurcation and patterns induced by flow in a prey-predator system with Beddington-DeAngelis functional response, Phys. Rev. E, 102 (2020), 012209. https://doi.org/10.1103/PhysRevE.102.012209 doi: 10.1103/PhysRevE.102.012209
    [9] L. J. Wang, C. J. Dai, M. Zhao, Hopf bifurcation in an age-structured prey-predator model with Holling response function, Math. Biosci. Eng., 18 (2021), 3144–3159. https://doi.org/10.3934/mbe.2021156 doi: 10.3934/mbe.2021156
    [10] X. Y. Meng, Y. Q. Wu, Bifurcation analysis in a singular Beddington-DeAngelis predator-prey model with two delays and nonlinear predator harvesting, Math. Biosci. Eng., 16 (2019), 2668–2696. https://doi.org/10.3934/mbe.2019133 doi: 10.3934/mbe.2019133
    [11] P. Feng, On a diffusive predator-prey model with nonlinear harvesting, Math. Biosci. Eng., 11 (2014), 807. https://doi.org/10.3934/mbe.2014.11.807 doi: 10.3934/mbe.2014.11.807
    [12] P. A. Abrams, L. R. Ginzburg, The nature of predation: prey dependent, ratio dependent or neither, Trends Ecol. Evol., 15 (2000), 337–341. https://doi.org/10.1016/S0169-5347(00)01908-X doi: 10.1016/S0169-5347(00)01908-X
    [13] C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the European Pine Sawfly1, Can. Entomol., 91 (1959), 293–320. https://doi.org/10.4039/Ent91293-5 doi: 10.4039/Ent91293-5
    [14] G. Seo, D. L. DeAngelis, A predator-prey model with a Holling type I functional response including a predator mutual interference, J. Nonlinear Sci., 21 (2011), 811–833. https://doi.org/10.1007/s00332-011-9101-6 doi: 10.1007/s00332-011-9101-6
    [15] W. Ko, K. Ryu, Qualitative analysis of a predator-prey model with Holling type II functional response incorporating a prey refuge, J. Differ. Equations, 231 (2006), 534–550. https://doi.org/10.1016/j.jde.2006.08.001 doi: 10.1016/j.jde.2006.08.001
    [16] E. Gonzˊalez-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, Bull. Math. Biol., 73 (2011), 1378–1397. https://doi.org/10.1007/s11538-010-9577-5 doi: 10.1007/s11538-010-9577-5
    [17] X. C. Zhang, G. Q. Sun, Z. Jin, Spatial dynamics in a predator-prey model with Beddington-DeAngelis functional response, Phys. Rev. E, 85 (2012), 021924. https://doi.org/10.1103/PhysRevE.85.021924 doi: 10.1103/PhysRevE.85.021924
    [18] D. Xiao, W. Li, M., Han, Dynamics in a ratio-dependent predator–prey model with predator harvesting, J. Math. Anal. Appl., 324 (2006), 14–29. https://doi.org/10.1016/j.jmaa.2005.11.048 doi: 10.1016/j.jmaa.2005.11.048
    [19] B. Liu, Y. Zhang, L. Chen, The dynamical behaviors of a Lotka–Volterra predator–prey model concerning integrated pest management, Nonlinear Anal.: Real World Appl., 6 (2005), 227–243. https://doi.org/10.1016/j.nonrwa.2004.08.001 doi: 10.1016/j.nonrwa.2004.08.001
    [20] M. Cavani, M. Farkas, Bifurcations in a predator-prey model with memory and diffusion. I: Andronov-Hopf bifurcation, Acta Math. Hung., 63 (1994), 213–229. https://doi.org/10.1007/BF01874129 doi: 10.1007/BF01874129
    [21] R. Arditi, L. R. Ginzburg, Coupling in predator-prey dynamics: ratio-dependence, J. Theor. Biol., 139 (1989), 311–326. https://doi.org/10.1016/S0022-5193(89)80211-5 doi: 10.1016/S0022-5193(89)80211-5
    [22] S. B. Hsu, S. P. Hubbell, P. Waltman, Competing predators, SIAM J. Appl. Math., 35 (1978), 617–625. https://doi.org/10.1137/0135051 doi: 10.1137/0135051
    [23] C. Dai, H. Liu, Z. Jin, Q. Guo, Y. Wang, H. Yu, et al., Dynamic analysis of a heterogeneous diffusive prey-predator system in time-periodic environment, Complexity, 2020 (2020), 1–13. https://doi.org/10.1155/2020/7134869 doi: 10.1155/2020/7134869
    [24] W. S. Yang, Dynamics of a diffusive predator-prey model with general nonlinear functional response, Sci. World J., 2014 (2014), 1–10. https://doi.org/10.1155/2014/721403 doi: 10.1155/2014/721403
    [25] C. Duque, M. Lizana, Partial characterization of the global dynamic of a predator-prey model with non constant mortality rate, Differ. Equations Dyn. Syst., 17 (2009), 63–75. https://doi.org/10.1007/s12591-009-0005-y doi: 10.1007/s12591-009-0005-y
    [26] C. Duque, M. Lizana, On the dynamics of a predator-prey model with nonconstant death rate and diffusion, Nonlinear Anal.: Real World Appl., 12 (2011), 2198–2210. https://doi.org/10.1016/j.nonrwa.2011.01.002 doi: 10.1016/j.nonrwa.2011.01.002
    [27] R. Yang, Hopf bifurcation analysis of a delayed diffusive predator-prey system with nonconstant death rate, Chaos, Solitons Fractals, 81 (2015), 224–232. https://doi.org/10.1016/j.chaos.2015.09.021 doi: 10.1016/j.chaos.2015.09.021
    [28] X. Fauvergue, J. C. Malausa, L. Giuge, F. Courchamp, Invading parasitoids suffer no Allee effect: a manipulative field experiment, Ecology, 88 (2007), 2392–2403. https://doi.org/10.1890/06-1238.1 doi: 10.1890/06-1238.1
    [29] A. M. Kramer, O. Sarnelle, R. A. Knapp, Allee effect limits colonization success of sexually reproducing zooplankton, Ecology, 89 (2008), 2760–2769. https://doi.org/10.1890/07-1505.1 doi: 10.1890/07-1505.1
    [30] Y. Huang, Z. Zhu, Z. Li, Modeling the Allee effect and fear effect in predator-prey system incorporating a prey refuge, Adv. Differ. Equations, 2020 (2020), 1–13. https://doi.org/10.1186/s13662-020-02727-5 doi: 10.1186/s13662-020-02727-5
    [31] D. Sen, S. Ghorai, S. Sharma, M. Banerjee, Allee effect in prey's growth reduces the dynamical complexity in prey-predator model with generalist predator, Appl. Math. Modell., 91 (2021), 768–790. https://doi.org/10.1016/j.apm.2020.09.046 doi: 10.1016/j.apm.2020.09.046
    [32] A. Kumar, B. Dubey, Dynamics of prey-predator model with strong and weak Allee effect in the prey with gestation delay, Nonlinear Anal.: Modell. Control, 25 (2020), 417–442. https://doi.org/10.15388/namc.2020.25.16663 doi: 10.15388/namc.2020.25.16663
    [33] Y. Kang, O. Udiani, Dynamics of a single species evolutionary model with Allee effects, J. Math. Anal. Appl., 418 (2014), 492–515. https://doi.org/10.1016/j.jmaa.2014.03.083 doi: 10.1016/j.jmaa.2014.03.083
    [34] X. Yu, S. Yuan, T. Zhang, Persistence and ergodicity of a stochastic single species model with Allee effect under regime switching, Commun. Nonlinear Sci. Numer. Simul., 59 (2018), 359–374. https://doi.org/10.1016/j.cnsns.2017.11.028 doi: 10.1016/j.cnsns.2017.11.028
    [35] E. Gonzˊalez-Olivares, J. Mena-Lorca, A. Rojas-Palma, J. D. Flores, Dynamical complexities in the Leslie–Gower predator–prey model as consequences of the Allee effect on prey, Appl. Math. Modell., 35 (2011), 366–381. https://doi.org/10.1016/j.apm.2010.07.001 doi: 10.1016/j.apm.2010.07.001
    [36] D. Mukherjee, Study of refuge use on a predator-prey system with a competitor for the prey, Int. J. Biomath., 10 (2017), 1750023. https://doi.org/10.1142/S1793524517500231 doi: 10.1142/S1793524517500231
    [37] R. K. Naji, S. J. Majeed, The dynamical analysis of a prey-predator model with a refuge-stage structure prey population, Int. J. Differ. Equations, 2016 (2016), 1–10. https://doi.org/10.1155/2016/2010464 doi: 10.1155/2016/2010464
    [38] Y. Zhang, X. Rong, J. Zhang, A diffusive predator-prey system with prey refuge and predator cannibalism, Math. Biosci. Eng., 16 (2019), 1445–1470. https://doi.org/10.3934/mbe.2019070 doi: 10.3934/mbe.2019070
    [39] F. Chen, L. Chen, X. Xie, On a Leslie–Gower predator–prey model incorporating a prey refuge, Nonlinear Anal.: Real World Appl., 10 (2009), 2905–2908. https://doi.org/10.1016/j.nonrwa.2008.09.009 doi: 10.1016/j.nonrwa.2008.09.009
    [40] L. Chen, F. Chen, L. Chen, Qualitative analysis of a predator–prey model with Holling type II functional response incorporating a constant prey refuge, Nonlinear Anal.: Real World Appl., 11 (2010), 246–252. https://doi.org/10.1016/j.nonrwa.2008.10.056 doi: 10.1016/j.nonrwa.2008.10.056
    [41] Z. F. Zhang, T. R. Ding, W. Z. Huang, Z. X. Dong, Qualitative theory of differential equation, in Translations of Mathematical Monographs, 1992. https://doi.org/10.1090/mmono/101
    [42] D. Y. Wu, H. Y. Zhao, Y. Yuan, Complex dynamics of a diffusive predator-prey model with strong Allee effect and threshold harvesting, J. Math. Anal. Appl., 469 (2019), 982–1014. https://doi.org/10.1016/j.jmaa.2018.09.047 doi: 10.1016/j.jmaa.2018.09.047
    [43] D. Y. Wu, H. Y. Zhao, Spatiotemporal dynamics of a diffusive predator-prey system with Allee effect and threshold hunting, J. Nonlinear Sci., 30 (2020), 1015–1054. https://doi.org/10.1007/s00332-019-09600-0 doi: 10.1007/s00332-019-09600-0
    [44] L. Perko, Differential Equations and Dynamical Systems, Springer, 2001. https://doi.org/10.1007/978-1-4613-0003-8
    [45] J. Sotomayor, Generic bifurcations of dynamical systems, Dyn. Syst., 1973. https://doi.org/10.1016/B978-0-12-550350-1.50047-3
    [46] X. C. Zhang, G. Q. Sun, Z. Jin, Spatial dynamics in a predator-prey model with Beddington-DeAngelis functional response, Phys. Rev. E, 85 (2012), 021924. https://doi.org/10.1103/PhysRevE.85.021924 doi: 10.1103/PhysRevE.85.021924
    [47] X. B. Zhang, H. Y. Zhao, Stability and bifurcation of a reaction-diffusion predator-prey model with non-local delay and Michaelis-Menten-type prey-harvesting, Int. J. Comput. Math., 93 (2016), 1447–1469. https://doi.org/10.1080/00207160.2015.1056169 doi: 10.1080/00207160.2015.1056169
  • This article has been cited by:

    1. Yuying Liu, Qi Cao, Wensheng Yang, Influence of Allee effect and delay on dynamical behaviors of a predator–prey system, 2022, 41, 2238-3603, 10.1007/s40314-022-02118-4
    2. Ali Al Khabyah, Rizwan Ahmed, Muhammad Saeed Akram, Shehraz Akhtar, Stability, bifurcation, and chaos control in a discrete predator-prey model with strong Allee effect, 2023, 8, 2473-6988, 8060, 10.3934/math.2023408
    3. Kunlun Huang, Xintian Jia, Cuiping Li, Analysis of modified Holling-Tanner model with strong Allee effect, 2023, 20, 1551-0018, 15524, 10.3934/mbe.2023693
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3591) PDF downloads(329) Cited by(3)

Figures and Tables

Figures(9)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog