1,004
Views
1
CrossRef citations to date
0
Altmetric
Research Article

Complex dynamics of a predator–prey model with opportunistic predator and weak Allee effect in prey

, , & ORCID Icon
Article: 2225545 | Received 22 Aug 2022, Accepted 10 Jun 2023, Published online: 20 Jun 2023

ABSTRACT

In this work, we first modify a Lotka–Volterra predator–prey system to incorporate an opportunistic predator and weak Allee effect in prey. The prey will be extinct if the combined effect of hunting and other food resources of predator is large. Otherwise, the dynamic behaviour of the system is extremely rich. A series of bifurcations such as saddle-node bifurcation, Hopf bifurcation, and Bogdanov–Takens bifurcation can happen. The validity of the theoretical results are supported with numerical simulations.

MATHEMATICS SUBJECT CLASSIFICATIONS:

1. Introduction

Predation is an important interaction among species in nature, where each species in a food chain must capture or be captured by other species. Mathematical models described by differential equations have provided insights on understanding the evolution and fitness of species. For example, results on persistence [Citation1,Citation9,Citation19], stability of equilibria [Citation22], and bifurcation [Citation32] have corresponding ecological significance.

Predators can be divided into two types: specialist predator and generalist predator [Citation14–16]. A specialist predator completely depends on the prey while a generalist predator has alternative food resources other than specialized prey. In the former, predators die out in the absence of the prey whereas, in the latter, predators can survive in the absence of the specified prey. What's interesting is that there exists opportunistic predators [Citation6] among specialist predator. Different from traditional specialist predator, this type predator such as Jackal and Martial eagle will snatch the prey that other predators have already captured or steal the leftovers of other predators. Therefore, a simple model on predation with an opportunistic predator is described by the following Lotka-Volterra predator–prey system, (1) x˙=rx(1xK)bxy,y˙=cxy+mydy2,(1) where the term my represents the intrinsic growth due to other food resources in the absence of prey. The dynamical behaviour of system (Equation1) was analysed completely in Chen [Citation8]. It admits at most a positive equilibrium, which is globally asymptotically stable when exists.

Nevertheless, a predator–prey system is affected by many factors. One currently studied is the Allee effect, which can be regarded as a negative density dependence when the population growth rate is decreased at low population size [Citation13]. Allee effect might be caused by difficulties in finding mates, social dysfunction at small population sizes, inbreeding depression [Citation11]. Allee effect is observed not only in animals but also in microorganisms and plants. As the density of the species increases, the influence of Allee effect will be weakened and even vanishes. Allee effect can be divided into two types, strong Allee effect [Citation5,Citation30] and weak Allee effect [Citation7,Citation29]. The strong Allee effect implies that there is a threshold for the population size. If the population size is lower than this size, the per capita growth rate is negative. But the weak Allee effect means that the per capita growth rate decreases but still remains positive.

In recent years, many scholars have extensively studied predator–prey models with both generalist predators and Allee effect [Citation2–4,Citation23–25,Citation27,Citation28]. In [Citation24], Sen et al. studied the impact of Allee effect in a predator–prey model with a generalist predator. The model exhibits simpler dynamics than the corresponding one without Allee effect. It undergoes the same local and global bifurcations. However, the tristability [Citation17] for the one without Allee effect will not occur. Meanwhile, Arancibia-Ibarra and Flores [Citation3] studied a modified Leslie-Gower predator–prey model with Holling type II functional response, strong and weak Allee effects in the prey, and a generalist predator. It is shown that the model with strong Allee effect has at most two positive equilibria, one always being a saddle and the other exhibiting multi-stability phenomenon since the equilibrium can be stable or unstable. The model with weak Allee effect has at most three positive equilibria, where one is always a saddle and the other two can be stable or unstable nodes.

However, predator–prey models with opportunistic predators and Allee effect in prey have been rarely studied. Therefore we modify the following Lotka-Volterra type predator–prey system involving Allee effect studied by Merdan [Citation20], (2) x˙=rx(1x)xβ+xaxy,y˙=ay(xy).(2) Taking into account other food resources for predators, we modify (Equation2) as (3) x˙=rx(1xK)xA+xbxy,y˙=cxy+mydy2,(3) where x and y stand for the densities of the prey and predators at time t, respectively. Here r is the intrinsic fertility rate of the prey in the absence of predators; K is the carrying capacity of prey; the term xA+x describes the Allee effect in prey with A reflecting the intensity of Allee effect; b is the maximal predator per capita capture rate; c is the conversion coefficient; m stands for the intrinsic fertility rate of predators in the absence of prey; d is the intraspecific competition coefficient of predators. All r, K, A, b, c, m, and d are positive constants.

For simplicity of forthcoming calculations and to reduce the number of parameters, we introduce the dimensionless changes of variables x¯=Kx,y¯=rdy,t¯=rtand denote A¯=AK,b¯=bdr,c¯=cKr,m¯=md.After dropping the bars, system (Equation3) becomes (4) x˙=x(1x)xA+xbxy,y˙=cxy+myy2.(4) From the biological point of view, the initial conditions of (Equation4) are (x(0),y(0))R+2. For each such initial condition, it is easy to see that (Equation4) has a unique global solution, which stays in R+2. Moreover, it follows from the first equation of (Equation4) that x˙x(1x),which implies that lim supxx(t)1. Then, for any ϵ>0, there exists T0 such that x(t)1+ϵ for tT. This, together with the second equation of (Equation4), implies that y˙(t)(c(1+ϵ)+m)yy2fortT,which yields lim supty(t)c(1+ϵ)+m. Since ε is arbitrary, we immediately get lim supty(t)c+m by letting ϵ0. Thus we have shown that every solution is bounded and, furthermore, the above differential inequalities also produce the attracting and positively invariant set {(x,y)R+2:0x1, 0yc+m}.The goal of this paper is to analyse the dynamics of (Equation4). We first discuss the existence and stability of equilibria in Section 2. The results indicate the occurrence of various bifurcations. Thus, in Section 3, we provide details on saddle-node bifurcation, Hopf bifurcation, and Bogdanov–Takens bifurcation. The paper concludes with some discussion.

2. Existence and stability of equilibria

2.1. Existence of equilibria

We start with the existence of equilibria of (Equation4).

System (Equation4) always has the three boundary equilibria E0(0,0), E1(1,0), and E2(0,m). Other equilibria are positive ones, which are determined by (1x)xA+xby=0,y=cx+m.Thus x is a positive zero of g(x) in (0,1), where (5) g(x)=(bc+1)x2+(Abc+bm1)x+Abm.(5) Note that g is quadratic in x with g(0)=Abm>0. Therefore, it is necessary that 1−bm>0 and A<A0=1bmbc if g(x)=0 has positive roots. The discriminant of g(x)=0 is Δ=b2c2A22(b2cm+bc+2bm)A+(bm1)2.Regarded as a quadratic function of A, Δ has two zeros A±=cmb+c+2m±2m(c+m)(bc+1)bc2.It is easy to check that A<A0<A+. Thus- when bm<1,

  1. if A>A then g(x)>0 for x>0;

  2. if 0<A<A then g(x) has two distinct positive roots, (6) x3=1AbcbmΔ2(bc+1)andx4=1Abcbm+Δ2(bc+1);(6)

  3. if A=A then g(x)=0 has a unique positive real root (7) x=1Abcbm2(bc+1).(7)

Note that x3<x<x4<1.

Summarizing the above discussion, we have obtained the following result on equilibria of (Equation4).

Theorem 2.1

The following statements on equilibria of (Equation4) hold.

  1. System (Equation4) always has the boundary equilibria E0(0,0), E1(1,0), and E2(0,m).

  2. System (Equation4) can have positive equilibria only if bm<1. In this case,

    1. if A<A then there are two positive equilibria E3(x3,y3) and E4(x4,y4), where x3 and x4 are given by (Equation6), y3,4=cx3,4+m;

    2. if A=A then there is a unique positive equilibrium E(x,y), where x is given by (Equation7) and y=cx+m;

    3. if A>A then there is no positive equilibrium.

In the following, we analyse the stability of the equilibria stated in Theorem 2.1.

2.2. Stability of the boundary equilibria E0, E1, and E2

We first consider the boundary equilibria.

Theorem 2.2

The three boundary equilibria, E0, E1, and E2 of (Equation4) are repelling saddle-node, saddle, and stable node, respectively.

Proof.

First, the Jacobian matrix of (Equation4) at E0(0,0) is JE0=(000m). As 0 is an eigenvalue of JE0, we make a Taylor expansion at (0,0) and the time scale dt=mdτ to change (Equation4) into x˙=1Amx2bmxy+O(|x,y|3),y˙=y+cmxy1my2,(here we still use dot to denote the derivative with respect to τ). Since the coefficient of x2 is 1Am>0, by Zhang et al. [Citation31, Theorem 7.1 in Chapter 2], E0 is a saddle-node and the repelling parabolic is located at the right half plane.

Next, the Jacobian matrices at E1(1,0) and E2(0,m) are respectively JE1=(1A+1b0c+m)andJE2=(bm0cmm).It follows immediately that E1 is a saddle and E2 is a stable node.

Note that the x-axis and the y-axis are positively invariant. By Theorem 2.1, if bm1 or (bm<1 and A>A) then (Equation4) has no positive equilibrium and hence there is no closed orbit in the first quadrant. At the meantime, both E0 and E1 are unstable. This gives the following result.

Corollary 2.3

If bm1 or (bm<1 and A>A), then the equilibrium E2 is globally asymptotically stable (see Figure ).

Figure 1. When A=0.7, b=2, c=0.25, and m=0.05, the equilibrium E2 of (Equation4) is globally asymptotically stable.

Figure 1. When A=0.7, b=2, c=0.25, and m=0.05, the equilibrium E2 of (Equation4(4) x˙=x(1−x)xA+x−bxy,y˙=cxy+my−y2.(4) ) is globally asymptotically stable.

2.3. Types of the positive equilibria E3 and E4

Next, we discuss the types of the positive equilibria E3 and E4. For this purpose, denote m=x4(cA2+2cx4A+cx42+2x4A+x42A)(A+x4)2>0.

Theorem 2.4

Suppose bm<1 and 0<A<A. Then the equilibrium E3 of (Equation4) is always a saddle and E4 is a{sourceif m<m,sinkif m>m,centre or fine focusif m=m.

Proof.

The Jacobian matrix at any positive equilibrium E(x,y) of (Equation4) is given by JE=(x(2Ax+x2A)(A+x)2bxc(cx+m)cxm).Thus det(JE)=x(cx+m)[(bc+1)x2+2A(bc+1)x+bcA2A](A+x)2and tr(JE)=x(12x)A+xx2(1x)(A+x)2cxm.For g defined by (Equation5), we have g(x)=2(bc+1)x+Abc+bm1.As g(x)=0, we have c=x2+(bm1)x+bmAbx(A+x).Then det(JE) and g(x) can be simplified as det(JE)=(bmA2+2bmAx+(bmA1)x2)(x1)x(A+x)3b,g(x)=bmA2+2bmAx+(bmA1)x2x(A+x),which give (8) det(JE)=(1x)x2(A+x)2bg(x).(8) Recall 0<x<1. Therefore, the positive equilibrium E(x,y) is an elementary equilibrium if g(x)0, a hyperbolic saddle if g(x)<0, a degenerate equilibrium if g(x)=0. As g(x3)<0 and g(x4)>0, we know that E3 is a saddle and E4 is an anti-saddle [Citation18]. Furthermore, tr(JE4)=x4(12x4)A+x4x42(1x4)(A+x4)2cx4m.It is easy to see that tr(JE4)>0, tr(JE4)=0, and tr(JE4)<0 if m<m, m=m, and m>m respectively. Then the type of E4 follows and this completes the proof.

Remark 2.1

Suppose bm<1 and A<A. If m>m then, by Theorems 2.2 and 2.4, E2 is a stable node, E3 is a saddle, and E4 is a sink. Thus (Equation4) admits bistable phenomenon and the stable manifold of the saddle E3 is the separatrix (see Figure ).

Figure 2. When A=0.5, b=2, c=0.25, and m=0.05, system (Equation4) admits bistable phenomenon.

Figure 2. When A=0.5, b=2, c=0.25, and m=0.05, system (Equation4(4) x˙=x(1−x)xA+x−bxy,y˙=cxy+my−y2.(4) ) admits bistable phenomenon.

2.4. Stability of the positive equilibrium E

Finally, we consider the positive equilibrium E.

Theorem 2.5

Suppose that bm<1 and A=A. Then the unique positive equilibrium E of (Equation4) is {an attracting saddle-nodeif bcxcxm>0;a repelling saddle-nodeif bcxcxm<0;a cusp of codimension twoif c=c,m=m,b>1, and x<b12b1.Here c=(12b)x+b1b3x and m=(1b)((2b1)x+1b)b3.

Proof.

The Jacobian matrix of system (Equation4) evaluated at E is given by JE=(x(2Ax+x2A)(A+x)2bxc(cx+m)cxm).Note that (Equation8) still holds for E, that is, det(JE)=(1x)x2(A+x)2bg(x). As g(x)=0, it follows that x(2Ax+x2A)(A+x)2=bcx. We perform the transformation (x,y)=(X+x,Y+y)to transform the equilibrium E to the origin and (Equation4) into (9) X˙=a10X+a01Y+a20X2+a11XY+a02Y2+a30X3+a21X2Y+a12XY2+a03Y3+P3(X,Y),Y˙=b10X+b01Y+b20X2+b11XY+b02Y2+b30X3+b21X2Y+b12XY2+b03Y3+Q3(X,Y),(9) where a10=bcx,a01=bx,a20=x3+3Ax2+3A2xA2A+x,a11=b,a30=A2(A+1)(A+x)4,b10=c(cx+m),b01=(cx+m),b11=c,b02=1,b20=a02=a21=a12=a03=0,and P3(X,Y) and Q3(X,Y) are C functions of at least order four in terms of (X,Y).

First, we assume that a10+b01=bcx(cx+m)0. Then the eigenvalues of JE are λ1=0 and λ2=a10+b010. Employing the transformation (X,Y)=(u+v,cu+b01a01v)anddτ=(a10+b01)dt,we can rewrite (Equation9) as u˙=c¯20u2+c¯11uv+c¯02v2+P¯4(u,v),v˙=v+d¯20u2+d¯11uv+d¯02v2+Q¯4(u,v),where c¯20=(cx+m)(Abcbm1)A2(A+x)2((bcc)xm)2,c¯11=1xc(A+x)3((bcc)xm)[(4bAc3bc2m+6Ac2+2c2+cm)x4+(6bA2c3+bAc2m+6A2c2+3Acmm2)x3+(2bA3c3+3bA2c2m2A2c2+3A2cm3Am2)x2+(bA3c2mA3cm2A2cm3A2m2)xA3m2],c¯02=cx+mbxc2(A+x)3((bcc)xm)[(2bAc2+3Ac+c+m)x3+(3bA2c2+2bAcm+3A2c+3Am)x2+(2bA3c2+bA2c2+3bA2cm+A3c+3A2m)x+bA3cm+A3m],d¯20=cbx(A+x)3((bcc)xm)[(bc+1)x3+(3bAc+3A)x2+(3bA2c+3A2)x+bA3cA2],d¯11=1(A+x)3((bcc)xm)[(b2c2+bc2+bc+c)x4+(3b2Ac2+3bAc2+3bAcbm+3Ac+m)x3+(3b2A2c2+3bA2c2+3bA2c3bmA+3A2c+3Am)x2+(b2A3c2+bA3c2bA3c2bA2c3bA2m+A3c+3A2m)xbA3m+A3m],d¯02=1bxc2(A+x)3((bcc)xm)[(b2c2m+2bAc3+3Ac2+c2+2cm)x4+(3b2Ac2m+3bA2c3+2bAc2m+3A2c2+6Acm+m2)x3+(b2A3c3+b2A2c3+3b2A2c2m+bA3c3+3bA2c2m+A3c2+6A2cm+3Am2)x2+(b2A3c2m+bA3c2m+2A3cm+3A2m2)x+A3m2],and P4¯(u,v), Q4¯(u,v) are C functions of at least order three in terms of (u,v). According to the centre manifold method [Citation21] and noticing that c¯20>0, we can approximate the equation near the centre manifold by dudt=(cx+m)(Abcbm1)A2(A+x)2((bcc)xm)2u2+O(|u|3).Thus the degenerate equilibrium E is a saddle-node by Theorem 7.1 in Zhang et al. [Citation31].

Next, assume that (a10+b01)=0. Recall that x=1Abcbm2(bc+1)andA=cmb+c+2m2m(c+m)(bc+1)bc2.Solving c, m, and A from these relations gives us c=c,m=m,A=A=(b1)x2+x(12b)x+b1.As all c, m, and A are positive, we have (10) b>1and0<x<b12b1.(10) Since a010, we use the linear transformation X=u,Y=a10a01u+1a01vto transform (Equation9) into (11) u˙=v+c20u2c11uv+O(|u,v|3),v˙=d20u2+d11uvd02v2,(11) where c20=(b1)(bxx+1)(2bxbx+1)b3x(x1),c11=x,d20=(b1)(bxx+1)(2bxbx+1)2x(x1)b5,d11=(b1)((2b1)x+1b)b3x,d02=1bx.By Remark 1 of Section 2.13 in Perko [Citation21], we obtain an equivalent system of system (Equation11) in a small neighbourhood of (0,0), u˙=v+O(|u,v|3),v˙=e20u2+e11uv+O(|u,v|3),where e20=d20=(b1)(bxx+1)(2bxbx+1)2x(x1)b5,e11=2c20+d11=(b1)(2bxx+1)(2bxbx+1)b3x(1x).Obviously, e20<0 and e11<0. Therefore, the equilibrium E is a cusp of codimension 2.

3. Bifurcation analysis

This section is devoted to the analysis of possible bifurcations of (Equation4), which include saddle-node bifurcation, Hopf bifurcation, and Bogdanov–Takens bifurcation.

3.1. Saddle-node bifurcation

Theorem 2.1 tells us that when bm<1, the number of equilibria of (Equation4) depends on the Allee effect parameter A. Due to Theorem 2.5, the degenerate equilibrium E is a saddle-node if bcxcxm0. Therefore, it is possible for (Equation4) to undergo a saddle-node bifurcation at A=ASN=A, which is confirmed in the following theorem.

Theorem 3.1

System (Equation4) undergoes a saddle-node bifurcation at E when A=ASN and bcx(cx+m)0.

Proof.

We apply the Sotomayor's theorem [Citation21] to finish the proof. We have seen before that JE has zero as a simple eigenvalue if A=ASN and bcx(cx+m)0. Choose V=(V1V2)=(1c)andW=(W1W2)=(1bxcx+m),which are - eigenvectors of JE and JET corresponding to the eigenvalue 0, respectively. Then FA(E,ASN)=((ASNc+c+m)b(ASNbc+bm1)2(ASN+1bm)2(bc+1)0)and D2F(E,ASN)(V,V)=(2F1x2V12+22F1xyV1V2+2F1y2V222F2x2V12+22F2xyV1V2+2F2y2V22)(E;ASN)=(ASN(ASNbcbm1)(ASN+x)20).Obviously, V and W satisfy the transversality conditions on the occurrence of saddle-node bifurcation at E when A=ASN, that is, WTFA(E,ASN)=(ASNc+c+m)b(ASNbc+bm1)2(ASN+1bm)2(bc+1)0,WT[D2F(E,ASN)(V,V)]=ASN(ASNbcbm1)(ASN+x)20.

When the parameter A varies to cross the surface from one side to the other side, the number of positive equilibria of (Equation4) changes from zero to two and hence the saddle-node bifurcation yields two equilibria. The phase of the saddle-node bifurcation, obviously, corresponds with the ecological system having critical Allee value ASN. When A>ASN, the prey will face the risk of extinction; When A<ASN, the dynamic behaviour of (Equation4) becomes very complex since the saddle and anti-saddle come out in the phase, which means that the density of prey must be large enough for its survival.

3.2. Hopf-bifurcation

Next, we consider Hopf bifurcation. By Theorem 2.4, E3 is always a saddle and E4 can be stable or unstable. If tr(JE4)=0, then the eigenvalues of E4 are a pair of purely imaginary complex numbers since det(JE4)>0. Thus E4 becomes a non-hyperbolic equilibrium and a Hopf-bifurcation may occur around E4(x4,y4). This is what we are going to do in the following.

For simplicity, inspired by Dai et al. [Citation12], we first do the state variable scaling, x~=xx4,y~=yy4,τ=y4t,and denote A~=Ax4,p~=1y4,q~=x4y4,b~=b,c~=cx4y4m~=my4.After dropping the tildes, system (Equation4) becomes (we still denote τ as t) (12) x˙=x(pqx)xA+xbxy,y˙=cxy+myy2,(12) where p, q, A, b, c, and m are all positive constants and p>q.

Since system (Equation12) has E4~(1,1) as an equilibrium, we have b=pqA+1>0 and m=1−c. Substituting them into (Equation12) and taking dt=1A+xdτ (still denote τ as t), we get (13) x˙=x2(pqx)pqA+1x(A+x)y,y˙=[cxy+(1c)yy2](A+x),(13) where p>q and c<1. Since (Equation4) and (Equation13) have the same topological structures, we can investigate the Hopf bifurcation around E4~(1,1) in (Equation13) to get the information on Hopf bifurcation around E4(x4,y4) in (Equation4).

The Jacobian matrix of (Equation13) at E4~ is JE4~=(pA2qAqA+1p+qc(A+1)A1).Then det(JE4~)=(A+1)(pq)cpA+2qA+qand tr(JE4~)=A2pA+2qA+2A+q+1A+1.It follows that det(JE4~)>0 for c>max{A(pq)q(A+1)A(pq)+(pq),0}c and tr(JE4~){=0if p=p,>0if p>p,<0if p<p,where p=A2+2qA+2A+q+1A.

Theorem 3.2

Suppose that p>q and c<c<1. Then the equilibrium E4~(1,1) of (Equation13) is {a stable hyperbolic node or focus if p<p;an unstable hyperbolic node or focus if p>p;a fine focus or centreif p=p.

Next, we study Hopf bifurcation around E4~ in (Equation13). Obviously, dtr(JE4~)dp|p=p=AA+10,namely, the transversality condition on the occurrence of Hopf bifurcation is satisfied. Through calculating the first Lyapunov coefficient, we can determine the stability of the limit cycles around the equilibrium E4~.

We first translate E4~(1,1) to (0,0) by (xˆ,yˆ)=(x1,y1) and then use the linear transformation uˆ=xˆ,vˆ=1D(pA2qAqA+1xˆ+(qp)yˆ)and the time scaling dtD=dτ to transform (Equation13) into uˆ˙=vˆ+f(uˆ,vˆ),vˆ˙=uˆ+g(uˆ,vˆ),where D=(A+1)2(cA+cqA+c)A>0,f(uˆ,vˆ)=cˆ20uˆ2+cˆ11uˆvˆ+cˆ02vˆ2+cˆ30uˆ3+aˆ21uˆ2vˆ+aˆ12uˆvˆ2+aˆ03vˆ3+O(|uˆ,vˆ|4),g(uˆ,vˆ)=dˆ20uˆ2+dˆ11uˆvˆ+dˆ02vˆ2+dˆ30uˆ3+dˆ21uˆ2vˆ+dˆ12uˆvˆ2+dˆ03vˆ3+O(|uˆ,vˆ|4),and cˆ20=q+1D,cˆ11=A+2A+1,cˆ30=q+1D,cˆ21=1A+1,dˆ20=(c1)A3+(cq+3c+q1)A2+(3cq+q2+3c+q)A+cq2+2cq+c(A+1)(A+q+1)(Ac+cqA+c),dˆ11=AD(cA+cqA+c+q+1)(A+1)(A+q+1)(cA+cqA+c),dˆ02=AA+q+1,dˆ30=A(cA+qA+cq+q2+c+2q+1)(A+1)(A+q+1)(cA+cqA+c),dˆ21=AD(cA+cqA+c+q+1)(A+1)2(A+q+1)(cA+cqA+c),dˆ12=A(A+1)(A+q+1),dˆ03=cˆ02=cˆ12=cˆ03=0.Making use of the formal series method in [Citation31], we get the first-order Lyapunov number, l1=18AD(q+1)((A2+2qAq1)cA22qA)(A+1)3(cA+cqA+c)2.If A2+2qAq10 then l1>0 and if A2+2qAq1>0 then l1>0 because AA+q+1<c<1. This shows that l1 is always positive.

Theorem 3.3

System (Equation13) experiences subcritical Hopf bifurcation at E4~ with p=p and E4~ is surrounded by an unstable limit cycle.

Now, we present some numerical simulations to demonstrate the feasibility of the above results on Hopf bifurcation. We fix A=1, c=23, and q=1.

  1. With p=7+0.1, E4~ is a hyperbolic unstable focus (see Figure (a)).

  2. With p=7, E4~ is an unstable weak focus (see Figure (b)).

  3. With p=7−0.1, E4~ is a hyperbolic stable focus surrounded by an unstable limit cycle (see Figure (c)).

  4. With p=7−0.26, system (Equation13) admits an unstable homoclinic loop which consists of an unstable homoclinic orbit and a saddle (see Figure (d)).

  5. With p=7−0.4, E4~ is a stable focus (see Figure (e)).

Figure 3. Phase portraits of (Equation13) when p differs with A=1, c=23, and q=1. (a) E4~ is a hyperbolic unstable focus when p=7+0.1. (b) E4~ is an unstable weak focus when p=7. (c) E4~ is a hyperbolic stable focus surrounded by an unstable limit cycle when p=7−0.1. (d) System (Equation13) admits an unstable homoclinic loop consisting of an unstable homoclinic orbit and a saddle when p=7−0.26. (e) E4~ is a stable focus when p=7−0.4.

Figure 3. Phase portraits of (Equation13(13) x˙=x2(p−qx)−p−qA+1x(A+x)y,y˙=[cxy+(1−c)y−y2](A+x),(13) ) when p differs with A=1, c=23, and q=1. (a) E4~ is a hyperbolic unstable focus when p=7+0.1. (b) E4~ is an unstable weak focus when p=7. (c) E4~ is a hyperbolic stable focus surrounded by an unstable limit cycle when p=7−0.1. (d) System (Equation13(13) x˙=x2(p−qx)−p−qA+1x(A+x)y,y˙=[cxy+(1−c)y−y2](A+x),(13) ) admits an unstable homoclinic loop consisting of an unstable homoclinic orbit and a saddle when p=7−0.26. (e) E4~ is a stable focus when p=7−0.4.

As the bifurcation parameter p changes, the stability of E4~ switches from unstable to stable, accompanied by an unstable limit cycle. Therefore, system (Equation13) undergoes subcritical Hopf bifurcation at p=7.

3.3. Bogdanov–Takens bifurcation

When bm<1 and A=A, by Theorem 2.5, the degenerate equilibrium E of (Equation4) is a cusp of codimension two if c=c, m=m, b>1, and x<b12b1. This indicates the possible occurrence of Bogdanov–Takens bifurcation around E.

Theorem 3.4

With A and m as the bifurcation parameters, system (Equation4) undergoes a Bogdanov–Takens bifurcation of codimension two around the positive equilibrium E when A=A, c=c, m=m, and condition (Equation10) holds.

Proof.

We perturb m and A with m+ϵ2 and A+ϵ1 respectively to get the unfolding system of (Equation4), (14) x˙=x(1x)xA+ϵ1+xbxy,y˙=cxy+(m+ϵ2)yy2,(14) where ϵ=(ϵ1,ϵ2) is a parameter vector in a small neighbourhood of the origin. Obviously, when ϵ1=ϵ2=0, system (Equation14) has a cusp of codimension two.

First, the transformation (x,y)=(u1+x,u2+y) translates the equilibrium (x,y) to (0,0) and transforms (Equation14) into (15) u˙1=a~10(ϵ)u1+a~01(ϵ)u2+a~00(ϵ)+a~20(ϵ)u12+a~11(ϵ)u1u2+O(|u1,u2|3),u˙2=b~10(ϵ)u1+b~01(ϵ)u2+b~00(ϵ)+b~20(ϵ)u12+b~11(ϵ)u1u2+b~02(ϵ)u22+O(|u1,u2|3),(15) where a~10(ϵ)=3x2+1(A+x+ϵ1)2+x2(1x)A+x+ϵ1by,a~01(ϵ)=bx,b~10(ϵ)=cy,b~01(ϵ)=y+ϵ2,a~00(ϵ)=x2(1x)A+ϵ1+xbxy,a~20(ϵ)=3x+1A+x+ϵ1+x(3Ax+2x2+3xϵ12Ax2ϵ1)(A+x+ϵ1)3,a~11(ϵ)=b,b~00(ϵ)=ϵ2y,b~20(ϵ)=0,b~11(ϵ)=c,b~02(ϵ)=1,Next, with the C change of coordinates in a small neighbourhood of the origin, (u3,u4)=(u1,a~10(ϵ)u1+a~01(ϵ)u2+a~00(ϵ)+a~20(ϵ)u12+a~11(ϵ)u1u2+O(|u1,u2|3)),system (Equation15) becomes (16) u˙3=u4,u˙4=c~00(ϵ)+c~10(ϵ)u3+c~01(ϵ)u4+c~20(ϵ)u32+c~11(ϵ)u3u4+c~02(ϵ)u42+O(|u3,u4|3),(16) where c~00(ϵ)=b~01(ϵ)a~00(ϵ),c~10(ϵ)=a~01(ϵ)b~10(ϵ)a~10(ϵ)b~01(ϵ),c~01(ϵ)=a~10(ϵ)+b~01(ϵ)a~00(ϵ)a~11(ϵ)a~01(ϵ),c~20(ϵ)=b~01(ϵ)a~20(ϵ)+b~10(ϵ)a~11(ϵ),c~11(ϵ)=2a~20(ϵ)+a~00(ϵ)a~112(ϵ)a~10(ϵ)a~11(ϵ)a~012(ϵ),c~02(ϵ)=a~11(ϵ)a~01(ϵ).Now, introducing a new time variable τ with dt=(1c~02(ϵ)u3)dτ and still denoting τ as t for simplicity, we rewrite system (Equation16) as (17) u˙3=u4(1c~02(ϵ)u3),u˙4=(1c~02(ϵ)u3)[c~00(ϵ)+c~10(ϵ)u3+c~01(ϵ)u4+c~20(ϵ)u32+c~11(ϵ)u3u4+c~02(ϵ)u42+O(|u3,u4|3)].(17) Then (u5,u6)=(u3,u4(1c~02(ϵ)u3) transforms (Equation17) into (18) u˙5=u6,u˙6=d~00(ϵ)+d~10(ϵ)u5+d~01(ϵ)u6+d~20(ϵ)u52+d~11(ϵ)u5u6+O(|u5,u6|3),(18) where d~00(ϵ)=c~00(ϵ),d~10(ϵ)=c~10(ϵ)2c~00(ϵ)c~02(ϵ),d~01(ϵ)=c~01(ϵ),d~20(ϵ)=c~20(ϵ)2c~02(ϵ)c~10(ϵ)+c~00(ϵ)c~02(ϵ)2,d~11(ϵ)=c~01(ϵ)c~02(ϵ)+c~11(ϵ).As d~20(0)=(b1)(bxx+1)(2bxbx+1)2b5x(x1)<0, we have d~20(ϵ)<0 when ϵ1 and ϵ2 are small enough. With the following change of coordinates and time scaling, (u7,u8)=(u5,u6d~20(ϵ))andτ=d~20(ϵ)t,we can transform (Equation18) into (19) u˙7=u8,u˙8=e~00(ϵ)+e~10(ϵ)u7+e~01(ϵ)u8u72+e~11(ϵ)u7u8+O(|u7,u8|3),(19) where e~00(ϵ)=d~00(ϵ)d~20(ϵ),e~10(ϵ)=d~10(ϵ)d~20(ϵ),e~01(ϵ)=d~01(ϵ)d~20(ϵ),e~11(ϵ)=d~11(ϵ)d~20(ϵ).Furthermore, the transformation (u9,u10)=(u7e~10(ϵ)2,u8)transforms (Equation19) into u˙9=u10,u˙10=f~00(ϵ)+f~01(ϵ)u10u92+f~11(ϵ)u9u10+O(|u9,u10|3),where f~00(ϵ)=e~00(ϵ)+14e~10(ϵ)2,f~01(ϵ)=e~01(ϵ)+12e~10(ϵ)e~11(ϵ),f~11(ϵ)=e~11(ϵ).It follows from f~11(0)=(2bxbx+1)(b1)(2bxx+1)b3x(x1)<0 that f~11(ϵ)0 when ϵ1 and ϵ2 are sufficiently small. Employing the change of variables and time scaling, (u11,u12)=(f~11(ϵ)2u9,f~11(ϵ)3u10)andτ=1f~11(ϵ)t,we obtain the versal unfolding of system (Equation14), u˙11=u12,u˙12=g~00(ϵ)+g~01(ϵ)u12+u112+u11u12+O(|u11,u12|3),where g~00=f~00(ϵ)f~11(ϵ)4,g~01=f~01(ϵ)f~11(ϵ).Notice that |(g~00,g~01)(ϵ1,ϵ2)|ϵ1=ϵ2=0=12(b1)(2bxx+1)5b2(bxx+1)5(x1)2x0.Therefore, the Bogdanov–Takens bifurcation theory [Citation10,Citation21] implies that (Equation14) undergoes a subcritical Bogdanov–Takens bifurcation of codimension two when the parameters ϵ1 and ϵ2 vary in a small neighbourhood of (0,0), which includes a sequence of bifurcations with codimension one saddle-node bifurcation, subcritical Hopf bifurcation, and Homoclinic bifurcation [Citation26].

According to the versal unfolding of system (Equation14) and results in [Citation10], the Bogdanov–Takens bifurcation diagram consists of saddle-node, Hopf, and homoclinic bifurcation curves whose local expressions in the small neighbourhood of (0,0) are as follows.

  1. The saddle-node bifurcation curve SN={(ϵ1,ϵ2):g~00(ϵ1,ϵ2)=0, g~01(ϵ1,ϵ2)0};

  2. the Hopf-bifurcation curve H={(ϵ1,ϵ2):g~00(ϵ1,ϵ2)<0, g~01(ϵ1,ϵ2)=g~00(ϵ1,ϵ2)};

  3. the homoclinic curve HL={(ϵ1,ϵ2):g~00(ϵ1,ϵ2)<0, g~01(ϵ1,ϵ2)=57g~00(ϵ1,ϵ2)}.

We give some numerical simulations on Bogdanov–Takens bifurcation to end this section. Choosing (A,b,c,m)=(0.6,2,0.25,0.05), we can draw the bifurcation diagram of (Equation14) including three bifurcation curves H, HL, and SN, which divide the small neighbourhood of (0,0) in the ϵ1ϵ2-plane into four regions I, II, III, and IV (see Figure (a)). Further, the values of ϵ1 and ϵ2 will change phase portrait of system (Equation14) and we have the following results:

  1. when (ϵ1,ϵ2)=(0.1,0.001) lies in region I, system (Equation14) has no positive equilibrium, which means that predator and prey will not coexist over time (see Figure (b));

  2. when (ϵ1,ϵ2)=(0.1,0.01135) lies in region II, system (Equation14) has an unstable focus E4 and a saddle E3 as shown in Figure (c);

  3. when (ϵ1,ϵ2)=(0.1,0.0116) lies in region III, Figure (d) indicates that system (Equation14) has an unstable limit cycle with a stable focus and a saddle, which infer that populations of predator and prey present periodical changes;

  4. when (ϵ1,ϵ2)=(0.1,0.01171) lies on the curve HL, an unstable homoclinic loop appears in the phase portrait of system (Equation14) (see Figure (e));

  5. finally, let (ϵ1,ϵ2)=(0.1,0.0118) lie in region IV. Then Figure (f) tells us that a stable focus and a saddle come out in the phase portrait of system (Equation14). From an ecological point of view, whether predator and prey coexist or not depends on their initial population values.

4. Conclusion

In this paper, we studied the dynamic behaviour of a modified Lotka-Volterra predator–prey model where predators have other food resources and the prey is influenced by weak Allee effect. As the Allee parameter, A, changes, the model can have at most two positive equilibria. When there is a unique positive equilibrium, E, the equilibrium is degenerate. It can be a saddle-node of codimension 1 or a cusp of codimension 2. When there are two positive equilibria, one is always a saddle (E3) and the other is an anti-saddle (E4).

Figure 4. Phase portraits of system (Equation14). (a) Bifurcation diagram of system (Equation14). (b) System (Equation14) admits no positive equilibrium when (ϵ1,ϵ2)=(0.1,0.001). (c) System (Equation14) admits an unstable focus and a saddle when (ϵ1,ϵ2)=(0.1,0.01135). (d) System (Equation14) admits an unstable limit cycle with a stable focus and a saddle when (ϵ1,ϵ2)=(0.1,0.0116). (e) System (Equation14) admits an unstable homoclinic orbit and a saddle when (ϵ1,ϵ2)=(0.1,0.01171). (f) System (Equation14) admits a stable focus and a saddle when (ϵ1,ϵ2)=(0.1,0.0118).

Figure 4. Phase portraits of system (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ). (a) Bifurcation diagram of system (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ). (b) System (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ) admits no positive equilibrium when (ϵ1,ϵ2)=(0.1,−0.001). (c) System (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ) admits an unstable focus and a saddle when (ϵ1,ϵ2)=(0.1,−0.01135). (d) System (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ) admits an unstable limit cycle with a stable focus and a saddle when (ϵ1,ϵ2)=(0.1,0.0116). (e) System (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ) admits an unstable homoclinic orbit and a saddle when (ϵ1,ϵ2)=(0.1,−0.01171). (f) System (Equation14(14) x˙=x(1−x)xA−∗+ϵ1+x−bxy,y˙=c∗xy+(m∗+ϵ2)y−y2,(14) ) admits a stable focus and a saddle when (ϵ1,ϵ2)=(0.1,−0.0118).

The system is not structurally stable when parameters change values. As a result, bifurcation phenomenon is very important from the point of ecosystems. Choosing A as the bifurcation parameter, the system can have saddle-node bifurcation at E where saddle-node appears in the phase diagram. The modification of the bifurcation parameter gives the variation in the number of equilibria. Therefore, by changing the bifurcation parameter, we should get that the nullclines intersect at two points, then these two points are the same, and finally, the nullclines do not intersect. Theorem 2.4 tells us that the stability of the anti-saddle E4 will switch from unstable to stable as the intrinsic fertility rate of the predator, m, increases and crosses m. Especially, the anti-saddle E4 is a weak focus when m=m and there can be a subcritical Hopf bifurcation, where unstable limit cycles will appear. Moreover, when both A and m are regarded as bifurcation parameters, non-degenerate Bogdanvo-Takens bifurcation of codimension 2 occurs at the cusp E. Thus, as A and m vary, the system produces rich dynamic behaviour, which includes saddle-node bifurcation, Hopf bifurcation, and homoclinic bifurcation.

From the findings, we observe that if the Allee effect is strong enough then the prey will become extinct and the system has no positive equilibrium. All orbits in the first quadrant tend to the prey-free equilibrium, E2(0,m), which is globally asymptotically stable (see Figure ). This is because predators have other food resources. Moreover, E2 is always a stable node (locally asymptotically stable). For appropriate values of A and m, the system has two positive equilibria. When one of them is stable, there is a bistable phenomenon (see Figure ).

Compared with the work of Merdan [Citation20], the modified system has extra food resources and hence it exhibits more complicated dynamical behaviour such as the change of the number of equilibria, limit cycles bifurcating via Hopf bifurcation, and appearance of homoclinic loops.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported partially by the Natural Science Foundation of Fujian Province [2020J01499] and NSERC of Canada [RGPIN-2019-05892].

References

  • A.S. Ackleh, M.I. Hossain, A. Veprauskas, and A. Zhang, Persistence and stability analysis of discrete-time predator–prey models: A study of population and evolutionary dynamics, J. Difference Equ. Appl. 25 (2019), pp. 1568–1603.
  • C. Arancibia-Ibarra and J. Flores, Modelling and analysis of a modified May–Holling–Tanner predator–prey model with Allee effect in the prey and an alternative food source for the predator, Math. Biosci. Eng. 17 (2020), pp. 8052–8073.
  • C. Arancibia-Ibarra and J. Flores, Dynamics of a Leslie–Gower predator–prey model with Holling type II functional response, Allee effect and a generalist predator, Math. Comput. Simulation 188 (2021), pp. 1–22.
  • C. Arancibia-Ibarra, M. Bode, J. Flores, G. Pettet, and P. Van Heijster, Turing patterns in a diffusive Holling–Tanner predator–prey model with an alternative food source for the predator, Commun. Nonlinear Sci. Numer. Simul. 99 (2021), Article ID 105802.
  • D. Bai, Y. Kang, S. Ruan, and L. Wang, Dynamics of an intraguild predation food web model with strong Allee effect in the basal prey, Nonlinear Anal. Real World Appl. 58 (2021), Article ID 103206, 36 pp.
  • A.A. Berryman and P. Kindlmann, Population Systems. A General Introduction, Springer, 2008.
  • S. Biswas, S.K. Sasmal, S. Samanta, Md. Saifuddin, N. Pal, and J. Chattopadhyay, Optimal harvesting and complex dynamics in a delayed eco-epidemiological model with weak Allee effects, Nonlinear Dynam. 87 (2017), pp. 1553–1573.
  • L.S. Chen, Mathematical Ecology Model and Research Method, Science Press, Beijing, 2017.
  • Y.-S. Chen, T. Giletti, and J.-S. Guo, Persistence of preys in a diffusive three species predator–prey system with a pair of strong-weak competing preys, J. Differ. Equ. 281 (2021), pp. 341–378.
  • S.-N. Chow, C.Z. Li, and D. Wang, Normal Forms and Bifurcation of Planar Vector Fields, Cambridge University Press, Cambridge, 1994.
  • F. Courchamp, T. Clutton-Brock, and B. Grenfell, Inverse dependence and the Allee effect, Trends Ecol. Evol. 14 (1999), pp. 405–410.
  • Y. Dai, Y. Zhao, and B. Sang, Four limit cycles in a predator–prey system of Leslie type with generalized Holling type III functional response, Nonlinear Anal. Real World Appl. 50 (2019), pp. 218–239.
  • B. Dennis, Allee effects: Population growth, critical density, and the chance of extinction, Nat. Resour. Model. 3 (1989), pp. 481–538.
  • E. González-Olivares and A. Rojas-Palma, Global stability in a modified Leslie-Gower type predation model assuming mutual interference among generalist predators, Math. Biosci. Eng. 17 (2020), pp. 7708–7731.
  • E. González-Olivares, C. Arancibia-Ibarra, A. Rojas-Palma, and B. González-Yañez, Bifurcations and multistability on the May-Holling-Tanner predation model considering alternative food for the predators, Math. Biosci. Eng. 16 (2019), pp. 4274–4298.
  • E. González-Olivares, C. Arancibia-Ibarra, A. Rojas-Palma, and B. González-Yañez, Dynamics of a modified Leslie-Gower predation model considering a generalist predator and the hyperbolic functional response, Math. Biosci. Eng. 16 (2019), pp. 7995–8024.
  • J. Jiao, W. Zhang, and P. Yu, Tristable phenomenon in a predator–prey system arising from multiple limit cycles bifurcation, Internat. J. Bifur. Chaos Appl. Sci. Eng. 30 (2020), Article ID 2050129, 23 pp.
  • Yu.A. Kuznetsov, Elements of Applied Bifurcation Theory, 3rd ed., Applied Mathematical Sciences Vol. 112, Springer-Verlag, New York, USA, 2004.
  • L. Meng, C. Du, and M. Deng, Persistence and extinction of a modified Leslie–Gower Holling-type II stochastic predator–prey model with impulsive toxicant input in polluted environments, Nonlinear Anal. Hybrid Syst. 27 (2018), pp. 177–190.
  • H. Merdan, Stability analysis of a Lotka–Volterra type predator–prey system involving Allee effects, ANZIAM J. 52 (2010), pp. 139–145.
  • L. Perko, Differential Equations and Dynamical Systems, 3rd ed., Springer-Verlag, New York, 2001.
  • G. Ren and B. Liu, Global existence and convergence to steady states for a predator–prey model with both predator–and prey–taxis, Discrete Contin. Dyn. Syst. 42 (2022), pp. 759–779.
  • D. Sen, S. Petrovskii, S. Ghorai, and M. Banerjee, Rich bifurcation structure of prey-predator model induced by the Allee effect in the growth of generalist predator, Internat. J. Bifur. Chaos Appl. Sci. Eng. 30 (2020), Article ID 2050084, 22 pp.
  • D. Sen, S. Ghorai, S. Sharma, and M. Banerjee, Allee effect in prey's growth reduces the dynamical complexity in prey-predator model with generalist predator, Appl. Math. Model. 91 (2021), pp. 768–790.
  • G. Seo and G.S.K. Wolkowicz, Pest control by generalist parasitoids: A bifurcation theory approach, Discrete Contin. Dyn. Syst. Ser. S 13 (2020), pp. 3157–3187.
  • Y. Tian and M. Han, Hopf and homoclinic bifurcations for near-Hamiltonian systems, J. Differ. Equ.262 (2017), pp. 3214–3234.
  • J.P. Tripathi, P.S. Mandal, and A. Poonia, A widespread interaction between generalist and specialist enemies: The role of intraguild predation and Allee effect, Appl. Math. Model. 89 (2021), pp. 105–135.
  • K. Vishwakarma and M. Sen, Role of Allee effect in prey and hunting cooperation in a generalist predator, Math. Comput. Simulation 190 (2021), pp. 622–640.
  • W. Wang, Y.-n. Zhu, Y. Cai, and W. Wang, Dynamical complexity induced by Allee effect in a predator–prey model, Nonlinear Anal. Real World Appl. 16 (2014), pp. 103–119.
  • C. Zhang, Pattern formation with jump discontinuity in a macroalgae-herbivore model with strong Allee effect in macroalgae, J. Math. Anal. Appl. 504 (2021), Article ID 125371, 26 pp.
  • Z.F. Zhang, T.R. Ding, W.Z. Huang, and Z.X. Dong, Qualitative Theory of Differential Equations, American Mathematical Society, Providence, 1992.
  • Z. Zhu, Y. Chen, Z. Li, and F. Chen, Stability and bifurcation in a Leslie-Gower predator–prey model with Allee effect, Internat. J. Bifur. Chaos Appl. Sci. Eng. 32 (2022), Article ID 2250040, 25 pp.