486
Views
0
CrossRef citations to date
0
Altmetric
Review Article

Hunting cooperation in a prey–predator model with maturation delay

, &
Article: 2332279 | Received 25 Apr 2023, Accepted 12 Mar 2024, Published online: 22 Mar 2024

ABSTRACT

We investigate the dynamics of a prey–predator model with cooperative hunting among specialist predators and maturation delay in predator growth. First, we consider a model without delay and explore the effect of hunting time on the coexistence of predator and their prey. When the hunting time is long enough and the cooperation rate among predators is weak, prey and predator species tend to coexist. Furthermore, we observe the occurrences of a series of bifurcations that depend on the cooperation rate and the hunting time. Second, we introduce a maturation delay for predator growth and analyse its impact on the system's dynamics. We find that as the delay becomes larger, predator species become more likely to go extinct, as the long maturation delay hinders the growth of the predator population. Our numerical exploration reveals that the delay causes shifts in both the bifurcation curves and bifurcation thresholds of the non-delayed system.

AMS SUBJECT CLASSIFICATIONS:

1. Introduction

Classical Gause-type prey–predator models with prey-dependent functional response and constant death rate of specialist predator can exhibit only two types of coexistence, stable steady-state coexistence and oscillatory coexistence [Citation1–4]. The consideration of prey–predator-dependent functional response [Citation5–9], predator's density-dependent death rate [Citation10,Citation11], Allee effect in prey growth [Citation12,Citation13] induces various types of interesting bifurcation structures which some time indicates the possible scenario of partial or total extinction [Citation14,Citation15]. The total extinction scenario is quite prominent for the models with strong Allee effect and the predator is a specialist predator [Citation16,Citation17]. The extinction scenario arises through mainly global bifurcations. Extinction scenario is not prominent in case of weak Allee effect and when the predator is a generalist predator [Citation18,Citation19].

Prey–predator models are building blocks of several long food chains and food webs. Classical prey-dependent functional responses are replaced by prey–predator-dependent functional responses to understand the dynamics of the prey–predator interaction where various types of social mechanism among the predators influence the rate of capture of the focal prey. Mutual interference among the predators lead to Beddington–DeAngelis-type functional response [Citation20–22]. Similarly, the cooperation among the predators is proved to enhance the rate of successful catch for certain specialist predators. This leads to the investigation of the dynamics of prey–predator models with hunting cooperation among the predators [Citation23,Citation24]. Group hunting mechanism among certain species of the predators increases the success rate of catching the targeted victim and an increase in predation rate is partly proportional to the predator density [Citation25]. Alves and Hilker [Citation26] described the modifications of four different functional responses by incorporating the cooperative hunting among the predators. Cooperative hunting mechanism is beneficial among the predators when the prey abundance is quite high, however, accelerated killing rate in turn induces Allee affect due to over predation.

Dynamics and growth of prey and predator population are governed by various positive and negative feedbacks resulting from a variety of biological and demographic actions. It is evident that the response of these actions may not affect the net population growth instantaneously but may be mediated by some time lag [Citation27–30]. The effect of the time lagged mechanisms on the dynamics is studied with the help of delay differential equation models [Citation31–34]. Most of the research with delayed models aim to study the destabilizing effect of time delay [Citation35–38] but at several situation we find the delay may not alter the dynamics significantly. In a recent work, the authors have shown that the appropriately formulated prey–predator model with maturation delay can capture the stabilizing effect of time delay [Citation16]. Long maturation time for the specialist predator reduces the slower recruitment of adult compartment which reduces the grazing pressure and hence stabilizes the dynamics. Motivated by this fact, here we are interested to explore the dynamics of a delayed prey–predator model with cooperative hunting among the specialist predators and maturation time delay in predator growth.

The organization of this paper is as follows. In Section 2, we investigate the dynamics of a prey–predator model without delays incorporating hunting cooperation. The number of equilibria is classified and stability of a trivial equilibrium, a predator-free equilibrium and a coexistence equilibrium is studied. Bifurcation scenario with respect to the hunting time is also discussed. The bifurcation diagrams displaying the occurrence of Hopf bifurcation, transcritical bifurcation and saddle-node bifurcation is presented. In Section 3, we investigate the dynamics of the model with delays. The number of equilibria is classified and stability of a trivial equilibrium and a predator-free equilibrium is studied. Bifurcation scenario with respect to the hunting time and delay is also discussed. The bifurcation diagrams displaying the occurrence of Hopf bifurcation, transcritical bifurcation, saddle-node bifurcation, Bogdanov–Takens bifurcation and cusp bifurcation is presented. In Section 4, analytical results are validated using extensive numerical simulations, and we explain how cooperative hunting and maturation delay affect the system's dynamics. In Section 5, we offer a brief discussion.

2. Model without delay

We consider a prey–predator model with hunting cooperation among the predators. The basic model, without delay term, is proposed by Alves and Hilker [Citation26]; it can be represented by the following nonlinear coupled ordinary differential equations: (1a) dXdT=rX(1XK)(λ+aY)1+H1(λ+aY)XXY,(1a) (1b) dYdT=e(λ+aY)1+H1(λ+aY)XXYmY,(1b) where X(T) and Y(T) represent the prey density and predator density at time T; r and K represent the intrinsic prey growth rate and carrying capacity of prey respectively; λ, H1 and e represent encounter rate of predators; handling time per prey per predator and conversion efficiency of predators respectively; aY is the cooperation term and m denotes the predator death rate. Using the dimensionless variables n=mX; p=λmY, t = mT and dimensionless parameters σ=rm; κ=eλKm; α=amλ2 and h1=H1me, we obtain the following system in terms of dimensionless variables and dimensionless parameters (2a) dndt=n(σ(1nκ)(1+αp)p1+h1(1+αp)n)f(n,p),(2a) (2b) dpdt=((1+αp)n1+h1(1+αp)n1)pg(n,p),(2b) subject to non-negative initial conditions. The prey and predator population densities at any instant of time t are denoted by n and p respectively. Here we interpret the dimensionless parameters in the context of ecology. σ and κ parametrize the logistic growth of prey species. α is the measure of cooperation among the predators for hunting prey and h1 is the dimensionless hunting time which is the sum of the times required for searching, killing and consuming the victim [Citation1,Citation26,Citation39].

2.1. Equilibria and stability

For the convenience of mathematical notations, we can write f(n,p)=nf1(n,p) and g(n,p)=pg1(n,p), where (3) f1(n,p)=σ(1nκ)(1+αp)p1+h1(1+αp)n,g1(n,p)=(1+αp)n1+h1(1+αp)n1.(3) The well-posedness of the model (2) is proved in Appendix 1. The number of coexistence equilibrium points depends upon the number of point of intersections between the curves f1(n,p)=0 and g1(n,p)=0 within the first quadrant. One can immediately see that the model (2) always has a trivial equilibrium E0(0,0) and a predator-free equilibrium E1(κ,0). Let us hereafter investigate the existence of a non-trivial equilibrium. The non-trivial prey nullcline is given by the equation f1(n,p)=0. This curve passes through two points (κ,0) and (0,p¯) where p¯ is the unique positive root of the equation αp2+pσ=0.The continuous curve joining the two points (κ,0) and (0,p¯) is the feasible prey-nullcline. With tedious algebraic calculations, we can prove that the non-trivial prey nullcline is either monotone decreasing in (0,κ) or is monotone increasing for (0,n1) and then decreasing for (n1,κ) where 0<n1<κ. The non-trivial predator nullcline is given by the equation g1(n,p)=0. This curve intersects n-axis at (11h1,0) and to be relevant in the context of ecology, we should have h1<1. The dimensionless parameter h1<1 implies that m<eH1 in terms of original parameters. This indicates that for the feasible coexistence of prey and their predators, the intrinsic death rate of the generalist predator should not be high. This finding aligns with the conditions required for the feasible coexistence of two species as outlined in the classical Rosenzweig–MacArthur model. For h11, the entire non-trivial predator nullcline will be in the second quadrant. Solving the equation g1(n,p)=0 for p, we can write p=1n+h1nα(1h1)n.If we differentiate this equation with respect to n, we find dpdn=1α(1h1)n2<0,for n>0 and 0h1<1. Clearly the non-trivial predator nullcline passes through (11h1,0), has vertical asymptote n = 0 and is monotone decreasing in the first quadrant.

Based upon the nature of two nullclines f1(n,p) and g1(n,p), we find unique point of intersection between two non-trivial nullclines when 11h1<κ. For 11h1>κ, we find either two coexistence equilibrium or no feasible equilibrium. The situation changes from two coexistence equilibrium to no coexistence equilibrium through a saddle-node bifurcation (for details, see Section 2.2). The cases of 0, 1 and 2 coexistence equilibrium point are demonstrated in Figures  and . As is indicated in Figures  and  for two cases, one can observe the following two scenarios in terms of the changing patterns of the number of nontrivial equilibrium points by increasing magnitude of h1:

Figure 1. Plots of two non-trivial nullclines f1(n,p) and g1(n,p) for σ=0.5,κ=2,α=1 and h1 varying from 0.1 to 0.6. As h1 increases, the number of equilibrium point changes from 1 to 0.

Figure 1. Plots of two non-trivial nullclines f1(n,p) and g1(n,p) for σ=0.5,κ=2,α=1 and h1 varying from 0.1 to 0.6. As h1 increases, the number of equilibrium point changes from 1 to 0.

Figure 2. Plots of two non-trivial nullclines f1(n,p) and g1(n,p) for σ=10,κ=1.2,α=1 and h1 varying from 0.1 to 0.75. As h1 increases, the number of equilibrium point changes from 1 to 0 ‘through 2’.

Figure 2. Plots of two non-trivial nullclines f1(n,p) and g1(n,p) for σ=10,κ=1.2,α=1 and h1 varying from 0.1 to 0.75. As h1 increases, the number of equilibrium point changes from 1 to 0 ‘through 2’.

Scenario 1 (σ=0.5, Figure )

  • When h1=0.25 (Blue curves), there exists a unique coexistence equilibrium Theorem 2.1(i).

  • When h1=0.5 (Red curves), there exists a unique coexistence equilibrium at transcritical bifurcation threshold (11κ(1h1)=0).

  • When h1=0.6 (Green curves), there are no coexistence equilibrium Theorem 2.1(ii) or (v).

Scenario 2 (σ=10, Figure )

  • When h1=0.1 (Blue curves), there exists a unique coexistence equilibrium Theorem 2.1(i).

  • When h1=0.166666667 (Red curves), there exists a unique coexistence equilibrium at transcritical bifurcation threshold (11κ(1h1)=0).

  • When h1=0.4 (Green curves), there exist two coexistence equilibria Theorem 2.1(iii).

  • When h1=0.66186 (Magenta curves), there exists a unique instantaneous coexistence equilibrium at saddle-node bifurcation threshold Theorem 2.1(iv).

  • When h1=0.75 (Black curves), there are no coexistence equilibrium points Theorem 2.1(v).

For further increase in magnitude of h1 leads to situation with no coexistence equilibrium. Now, for the positive equilibrium point E+(n,p), we can get the component p by solving the second equation of (2) which is given as p=h1nn+1αn(1h1) where n is the positive root of the cubic equation given by ασ(1h1)n3ακσ(1h1)n2κ(1h1)n+κ=0.In particular, for the existence of the coexistence equilibrium E+(n,p), the following theorem holds.

Theorem 2.1

The following statement holds true.

  1. If κ(1h1)>1, the model (2) has a unique coexistence equilibrium.

  2. If α1h1σ and κ(1h1)1, the model (2) has no coexistence equilibrium.

  3. If α>1h1σ and 1l(ασ1h1)<κ(1h1)1, the model (2) has two coexistence equilibria.

  4. If α>1h1σ and κ(1h1)=1l(ασ1h1)1, the model (2) has a unique coexistence equilibrium.

  5. If α>1h1σ and κ(1h1)<1l(ασ1h1)1, the model (2) has no coexistence equilibrium,

where the function l is positive and strictly monotone increasing on [1,] defined as l(u)=13+233u+13+227u+293uu+13l(1)=1(u1).

Proof.

When h1=0, n and p satisfy the following equations: (4) σ(1nκ)(1+αp)p1+h1(1+αp)n=0,(1+αp)n1+h1(1+αp)n1=0.(4) We note that (Equation4) yields σ(11κ(1h1)(1+αp))(1h1)(1+αp)p=0, that is, σ(11κ(1h1))+(ασ(1h1))p=2α(1h1)(p)2+α2(1h1)(p)3. Let us consider the case κ(1h1)>1. Since 11κ(1h1)>0, it is straightforward to prove that (i) holds true. Let us consider the case κ(1h1)1, that is, 11κ(1h1)0. If α1h1σ, then σ(11κ(1h1))+(ασ(1h1))p0 and 2α(1h1)p2+α2(1h1)p3>0 for all p>0. This implies that (ii) holds true. If α>1h1σ, let us define w(p)=σ(11κ(1h1))+(ασ(1h1))p2α(1h1)p2α2(1h1)p3.We note that w(0)=σ(11κ(1h1))0. By w(p)=(1h1)(ασ1h114αp3α2p2), we have w(p)=13pw(p)23α(1h1)p2+23(ασ(1h1))p+σ(11κ(1h1))=(13p+29α)w(p)+(23ασ+29(1h1))p29α(ασ(1h1))+σ(11κ(1h1)).As w(p)=0 if p=23α+ασ1h1+133α2=:p+>0, the function w has a local maximum: w(p+)=(23ασ+29(1h1))(23α+ασ1h1+133α2)29α(ασ(1h1))+σ(11κ(1h1))=σ(13+233ασ1h1+13+227ασ(1h1)+293ασ(1h1)ασ1h1+131κ(1h1))=σ(l(ασ1h1)1κ(1h1))at p=p+. Since the function l(u) is positive and strictly monotone increasing in u1 (due to the fact that (1+13u)u+13+133u=1u((u+13)32+133) is strictly monotone increasing in u1) with l(u)>l(1)=1 for all u>1, we obtain l(ασ1h1)>1. This implies that w(p+)>0 if l(ασ1h1)>1κ(1h1)>1 and w(p+)<0 if l(ασ1h1)<1κ(1h1). Therefore, w(p)=0 has two positive roots if l(ασ1h1)>1κ(1h1)>1 and no real roots if l(ασ1h1)<1κ(1h1). In addition, w(p)=0 has a double real root if l(ασ1h1)=1κ(1h1). Hence, the proof of (iii)–(v) is complete.

Remark 2.1

Theorem 2.1(ii) indicates that the predator cannot have a chance to coexist with the prey when the degree of cooperation among the predators is low.

Let us investigate the stability of equilibria. It is straightforward to prove that a trivial equilibrium E0 is always unstable, because the Jacobian matrix for the model (2) evaluated at E0 has at least one positive eigenvalue. The Jacobian matrix for the model (2) evaluated at E1 is given by (5) J1=[σκ1+h1κ0κ1+h1κ1].(5) E1 is stable for κ1+h1κ<1 and is unstable otherwise. The predator-free equilibrium loses stability through a transcritical bifurcation at h1TC=11κ (satisfying κ1+h1TCκ1=0). In terms of h1TC, E1 is stable for h1>h1TC and is unstable for h1<h1TC.

2.2. Saddle-node bifurcation

Now we prove that the system experiences a transition of having two coexistence equilibria from no coexistence equilibrium point or vice versa through saddle-node bifurcation by proving the required transversality conditions [Citation40]. If n denotes the first component of a feasible coexistence equilibrium point, then n is a positive root of the cubic equation (6) G(n)σα(1h1)n3σακ(1h1)n2κ(1h1)n+κ=0.(6) Let us assume that the equation G(n)=0 has a double root at the parametric threshold h1SN and the double root is denoted by n02. Then the following conditions should be satisfied: (7) G(n02)=0,G(n02)=0,G′′(n02)0.(7) In terms of the non-trivial nullclines, we can say that the two curves f1(n,p)=0 and g1(n,p)=0 touch each other at ESN(n02,p02) when h1=h1SN.

Evaluating the Jacobian matrix for the model (2) at (n02,p02), we find (8) JSN=[nf1nnf1ppg1npg1p](n02,p02).(8) Note that f1(n02,p02)=0 and g1(n02,p02)=0. Two curves f1(n,p)=0 and g1(n,p)=0 are tangential at (n02,p02) implies (9) f1nf1p|(n02,p02)=g1ng1p|(n02,p02).(9) Clearly, the Jacobian matrix JSN has a zero eigenvalue. If V and W denote the eigenvectors of JSN and JSNT respectively corresponding to zero eigenvalue then (10) V=[11+αp02αn02],W=[αp02(1+αp02)2n02+αp02+11].(10) Let F=[f,g]t. Now we can calculate WtFh1(ESN; h1=h1SN)=Wt[fh1gh1]ESN,h1SN=(1+αp02)3n023p02(1+h1SN(1+αp02)n02)2((1+αp02)2n02+αp02)0,and WtD2F(ESN; h1=h1SN)(V,V)=Wt[2fn2v12+22fn∂pv1v2+2fp2v222gn2v12+22gn∂pv1v2+2gp2v22 ]ESN,h1SN=αp02(1+αp02)2n02+αp02+1(2fn2v12+22fn∂pv1v2+2fp2v22)+(2gn2v12+22gn∂pv1v2+2gp2v22)=2κ(1+αp02)n02+2κ(1αp02)(1h1SNn02)2α2σn023p02ακn023((1+αp02)2n02+αp02)0.Hence all the transversality conditions for saddle-node bifurcation are satisfied. Figure  is the bifurcation diagram with respect to the parameter h1 with σ=10,κ=1.2,α=1.

Figure 3. The bifurcation diagram with respect to the parameter h1: (a) σ=10,κ=1.2,α=1 and (b) σ=10,κ=1.2,α=2.217.

Figure 3. The bifurcation diagram with respect to the parameter h1: (a) σ=10,κ=1.2,α=1 and (b) σ=10,κ=1.2,α=2.217.

2.3. Transcritical bifurcation

Proposition 2.2

A coexistence equilibrium point E+ bifurcates from the prey only equilibrium point E1 through transcritical bifurcation for the bifurcation parameter h1 and at the bifurcation threshold h1TC=11κ, provided κ>1 holds.

Proof.

At E1 and for h1=h1TC, Jacobian matrix of the model (2) is J(E1:h1=h1TC)=[σ100].It is evident that the Jacobian matrix has a simple zero eigenvalue. Now, corresponding to this zero eigenvalue, eigenvectors of the Jacobian matrix J(E1:h1TC=11κ) and its transpose are evaluated as (11) v=[1σ1],w=[01],(11) respectively. Now, according to Sotomayor's conditions [Citation40] the transversality conditions for transcritical bifurcation are wtFh1(E1; h1=h1TC)=0,wtDFh1(E1; h1=h1TC)v=1(1+h1TC)20,wtD2F(E1; h1=h1TC)(v,v)=(2gn21σ2+22gn∂p1σ+2gp2)|E1,h1TC=2(1+h1TC)2(1σ+α)0.All of the transversality conditions hold here. As a result, we deduce from Sotomayor's theorem that a transcritical bifurcation occurs for the model (2) at h1TC=11κ.

2.4. Hopf bifurcation

The Jacobian matrix for the model (2) evaluated at E+ is given by J+=[σ(12nκ)1+αp(1+h1(1+αp)n)2p(1+αp)n1+h1(1+αp)nαn(1+h1(1+αp)n)2p1+αp(1+h1(1+αp)n)2p(1+αp)n1+h1(1+αp)n1+αn(1+h1(1+αp)n)2p]=[σ(12nκ)p(1+αp)(1h1)21αnp(1h1)2p(1+αp)(1h1)2αnp(1h1)2]=[σ(12nκ)p(1+αp)(1h1)21αp(1h1)1+αpp(1+αp)(1h1)2αp(1h1)1+αp].Moreover, σ(12nκ)=(1+αp)p1+h1(1+αp)nσnκ=p(1h1)(1+αp)σ1κ(1h1)(1+αp).This yields (12) trJ+=σ(12nκ)(1+αp)p(1h1)2+αp(1h1)1+αp=(1+αp)p(1h1)h1σ1κ(1h1)(1+αp)+αp(1h1)1+αp(12) and (13) detJ+=σ(12nκ)αp(1h1)1+αp+p(1+αp)(1h1)2=(p(1h1)(1+αp)σ1κ(1h1)(1+αp))αp(1h1)1+αp+p(1+αp)(1h1)2.(13) The necessary conditions so that the model (2) undergoes a Hopf bifurcation at a positive equilibrium point E+ is that trJ+=0 and detJ+>0, where J+ is the Jacobian matrix evaluated at E+. In fact, one can see that detJ+>0 is always valid when h1=0 (see Appendix 2). Now from the condition trJ+=0, we have obtained the value of h1 as h1=12κp4(αp+1)3(κ2n)σ+α2κp+2p(α2p2+2αpα2+1)κ(αp+1)2kp.At h1=h1, the trace of the Jacobian matrix evaluated at E1 is 0. Assume that the corresponding determinant of the Jacobian matrix evaluated at E1 for the parametric value of h1=h1 is positive. Then the characteristic equation of the corresponding Jacobian matrix will have two purely imaginary eigenvalues ±ih10, where h10=detJ+|(E+, h1=h1).The transversality condition for Hopf bifurcation at E+ is given by d(trJ+)dh1|h1=h1=2(αp+1)p(1h1)αpαp+10.Therefore, the model (2) undergoes a Hopf bifurcation at the positive equilibrium point E+ for the parametric value h1=h1.

2.4.1. Nature of Hopf bifurcating periodic solution

We need to derive the normal form of Hopf bifurcation to know about the direction and stability of the Hopf bifurcation. To derive the normal form, we first need to transform the positive equilibrium point to origin. Let us take the transformations N=nn,P=pp. Using this transformations, the model (2) is transformed into dNdt=(N+n)(σ(1N+nκ)(1+α(P+p))(P+p)1+h(1+α(P+p))(N+n))F(N,P),dPdt=(N+n)(1+α(P+p))(P+p)1+h(1+α(P+p))(N+n)PpG(N,P).Taylor series expansion of F(N,P) and G(N,P) at (N,P)=(0,0) up to order 3 gives dNdt=a10N+a01P+a11NP+a20N2+a02P2+a30N3+a21N2P+a12NP2+a03P3+O(|N|4),dPdt=b10N+b01P+b11NP+b20N2+b02P2+b30N3+b21N2P+b12NP2+b03P3+O(|P|4).The expressions of aij and bij are given in Appendix 3. Neglecting the higher order terms of degree 4 and above, the system of equations transform into (14) Z˙=JEZ+A(Z),(14) where Z=[NP] and [A1A2]=[a11NP+a20N2+a02P2+a30N3+a21N2P+a12NP2+a03P3b11NP+b20N2+b02P2+b30N3+b21N2P+b12NP2+b03P3].The eigenvector w of the Jacobian matrix JE+ with respect to the eigenvalue ih10 when h1=h1 is w=[a01ih10a10].Define Q=[(w), (w)]=[a010a10h10].Take Z = QX or X=Q1Z, where X=[X1X2]. These transformations transform the system into the following equation: X˙=(Q1JEQ)X+Q1A(QX),that is, [X1˙X2˙]=[0h10h100][X1X2]+[N1(X1,X2;h1=h1)N2(X1,X2;h1=h1)].N1 and N2 are non-linear functions in the variable X1 and X2 and the corresponding forms are N1(X1,X2;h1=h1)=A1a01,N2(X1,X2;h1=h1)=a10A1+a01A2h10a01.The expression of A1 and A2 are given in Appendix 3.

To confer about the direction and stability of the bifurcated periodic solution, it is necessary to compute the first Lyapunov coefficient (l1) which is computed as below l1=116[NX1X1X11+NX1X2X21+NX1X1X21+NX2X2X21](0,0;h10)+116h10[NX1X21(NX1X11+NX2X21)NX1X22(NX1X22+NX2X22)NX1X11NX1X12+NX2X21NX2X22](0,0;h10).The Hopf bifurcation is supercritical if l1<0 and subcritical if l1>0. For l1=0, the system experiences a generalized Hopf bifurcation.

3. Delayed model

In this section, we consider the delayed model corresponding to the model (2) by introducing maturation time delay in predator growth. The maturation time delay is the time required for the juvenile predator individuals to become adult and capable to contribute to the growth of their own population [Citation16,Citation41–44]. With the maturation delay parameter τ, we can define the delayed model (15a) dn(t)dt=n(t)(σ(1n(t)κ)(1+αp(t))p(t)1+h1(1+αp(t))n(t)),(15a) (15b) dp(t)dt=exp(βτ)(1+αp(tτ))n(tτ)1+h1(1+αp(tτ))n(tτ)p(tτ)p(t),(15b) where β is the dimensionless mortality rate of the juvenile predators (see [Citation16,Citation45,Citation46] for derivation of the delayed model). Both the parameters τ and β are positive and dimensionless. Juvenile predators survive at a rate exp(βτ) and the proportion of juvenile predators that were able to survive over the time period [tτ,t] is added to the adult predator class at time t. It should be observed that juvenile predator density pj is not appearing in the growth equation for prey and growth equation for adult predator. Therefore, growth equation for juvenile predator can be decoupled. The growth equation for juvenile predators can be rewritten (for detail see [Citation46,Citation47]) as follows: pj(t)=τ0(1+αp(t+u))n(t+u)p(t+u)1+h1(1+αp(t+u))n(t+u)du,and pj(0) is given by pj(0)=τ0(1+αp(u))n(u)p(u)1+h1(1+αp(u))n(u)du.As n(t), p(t) completely determines the juvenile class pj(t); therefore, we only consider the model (15) for our further study (for detail derivation, see [Citation16,Citation45]). The delayed model is subjected to the non-negative continuous history functions n(s),p(s)0 for τs0. Here n(s) and p(s) are two continuous non-negative functions of s for s[τ,0]. Along with the similar discussion in Proposition A.1 in Appendix 1, one can obtain the well-posedness of the model (15).

3.1. Equilibria and stability

Similar to Theorem 2.1, for the existence of the coexistence equilibrium point E+(n,p) of the model (15), the following theorem holds.

Theorem 3.1

The following statement holds true.

  1. If κ(exp(βτ)h1)>1, the model (15) has a unique coexistence equilibrium.

  2. If α1h1exp(βτ)σ and κ(exp(βτ)h1)1, the model (15) has no coexistence equilibrium.

  3. If α>1h1exp(βτ)σ and 1l(ασ1h1exp(βτ))<κ(exp(βτ)h1)1, the model (15) has two coexistence equilibria.

  4. If α>1h1exp(βτ)σ and κ(exp(βτ)h1)=1l(ασ1h1exp(βτ))1, the model (15) has a unique coexistence equilibrium.

  5. If α>1h1exp(βτ)σ and κ(exp(βτ)h1)<1l(ασ1h1exp(βτ))1, the model (15) has no coexistence equilibrium,

where l(u) is defined in Theorem 2.1.

Proof.

One can see that n and p satisfy the following equations: (16) σ(1nκ)(1+αp)p1+h1(1+αp)n=0,exp(βτ)(1+αp)n1+h1(1+αp)n1=0.(16) We note that (Equation16) yields σ(11κ(exp(βτ)h1)(1+αp))(1h1exp(βτ))(1+αp)p=0, that is, σ(11κ(exp(βτ)h1))+(ασ(1h1exp(βτ)))p=2α(1h1exp(βτ))(p)2+α2(1h1exp(βτ))(p)3.Let us consider the case κ(exp(βτ)h1)>1. Since 11κ(exp(βτ)h1)>0, it is straightforward to prove that (i) holds true. Let us consider the case κ(exp(βτ)h1)1, that is, 11κ(exp(βτ)h1)0. If α1h1exp(βτ)σ, then σ(11κ(exp(βτ)h1))+(ασ(1h1exp(βτ)))p0 and 2αp2+α2p3>0 for all p>0. This implies that (ii) holds true. If α>1h1exp(βτ)σ, let us define wτ(p)=σ(11κ(exp(βτ)h1))+(ασ(1h1exp(βτ)))p2α(1h1exp(βτ))p2α2(1h1exp(βτ))p3.We note that wτ(0)=σ(11κ(exp(βτ)h1))0. By wτ(p)=ασ(1h1exp(βτ))4α(1h1exp(βτ))p3α2(1h1exp(βτ))p2 and the similar discussion in the proof of Theorem 2.1, one can see that wτ(p)=0 has two positive roots if 1l(ασ1h1exp(βτ))<κ(exp(βτ)h1) and no real roots if 1l(ασ1h1exp(βτ))>κ(exp(βτ)h1). In addition, wτ(p)=0 has a double real root if κ(exp(βτ)h1)=1l(ασ1h1exp(βτ)). Hence, the proof of (iii)–(v) is complete.

Let us investigate the stability of equilibria. It is straightforward to prove that a trivial equilibrium E0 is always unstable, because the Jacobian matrix for the model (15) evaluated at E0 has at least one positive eigenvalue. The Jacobian matrix for the model (15) evaluated at E1 is given by det(J1τλI)=0,where J1τ=[σκ1+h1κ0exp((β+λ)τ)κ1+h1κ1].Therefore E1 is stable for exp(βτ)κ1+h1κ<1 and is unstable otherwise.

3.2. Delay-induced Hopf bifurcation

The linearization of the model (15) around E+ gives us the following equivalent system in the vicinity of the positive equilibrium point: dudt=ϕ1u(t)+ϕ2v(t),dvdt=ψ1u(tτ)+ψ2v(tτ)+ψ3v(t),where ϕ1=n(1+αp)(σn2+κp)κpκn2(1+αp),ϕ2=n(1+αp)2αpn(1+αp)2,ψ1=pn2(1+αp),ψ2=n+(2n+1)αp+α2np2n(1+αp)2,ψ3=1.The associated characteristic equation is (17) Υ(λ,τ)=λ2(ϕ1+ψ3)λ+ϕ1ψ3+eλτ(ϕ1ψ2ϕ2ψ1ψ2λ)=0.(17) For simplicity, let Υ(λ,τ)=Υ1(λ,τ)+eλτΥ2(λ,τ), where Υ1(λ,τ)=λ2(ϕ1+ψ3)λ+ϕ1ψ3 and Υ2(λ,τ)=(ϕ1ψ2ϕ2ψ1ψ2λ).

To examine the occurrence of delay-induced Hopf bifurcation, specifically to determine the conditions under which a pair of purely imaginary eigenvalues of the characteristic equation (Equation17) exist, we can substitute λ= in (Equation17) which gives Υ1(,τ)+eiωτΥ2(,τ)=0,where ωR+ and Υ(,τ)=0. Sorting out the real and imaginary portions, we have Υ1R(,τ)+Υ2R(,τ)cosωτ+Υ2I(,τ)sinωτ=0,Υ1I(,τ)+Υ2I(,τ)cosωτΥ2R(,τ)sinωτ=0,where R and I indicate the real and imaginary parts of the corresponding components, respectively. As a result, we get (18) sin(ωτ)=Υ1I(,τ)Υ2R(,τ)Υ1R(,τ)Υ2I(,τ)Υ2R2(,τ)+Υ2I2(,τ),(18) (19) cos(ωτ)=Υ1R(,τ)Υ2R(,τ)+Υ1I(,τ)Υ2I(,τ)Υ2R2(,τ)+Υ2I2(,τ).(19) Assume that there is a common solution θ(τ) for both Equations (Equation18) and (Equation19), and that 0θ(τ)<2π. With this, we may deduce that τ0=θ(τ)ω is the smallest Hopf bifurcation threshold. Solving the equation τ=τn(τ) yields the consecutive Hopf bifurcation thresholds τc when switching of stability occurs where τn(τ) is the general solution of Equations (Equation18) and (Equation19) given by τn(τ)=θ(τ)+2ω,n=0,1,2,.The least value of τn(τ) gives us the desired Hopf bifurcation threshold.

3.3. Numerical saddle-node bifurcation results

We need to check the change in shape and position of the non-trivial predator nullcline with the variation of τ. The non-trivial predator nullcline is (20) g1(n,p,τ):=exp(βτ)(1+αp)n1+h1(1+αp)n1=0.(20) This curve intersects n-axis at (1exp(βτ)h1,0) and to be relevant in the context of ecology, we should have h1<exp(βτ). Solving above equation for p, we can write p=1exp(βτ)n+h1nα(exp(βτ)h1)nsince it follows from (Equation20) that 1+αp=1(exp(βτ)h1)n. If we differentiate this equation with respect to n, we find dpdn=1α(exp(βτ)h1)n2<0,for n>0 and 0<h1<exp(βτ). Clearly the non-trivial predator nullcline passes through (1exp(βτ)h1,0), has vertical asymptote n = 0 and is monotone decreasing in the first quadrant.

Based upon the nature of two nullclines, we find unique point of intersection between two non-trivial nullclines when 1exp(βτ)h1<κ. For 1exp(βτ)h1>κ we find either two coexistence equilibria or no feasible equilibrium. The situation changes from two coexistence equilibria to no feasible equilibrium through a saddle-node bifurcation.

4. Numerical results

Here, using numerical examples, we numerically illustrate the bifurcation results associated with the considered model dynamics. The positions of nullclines are significantly affected by the parameters h1 and α, which also have an impact on the stability of the equilibrium point. As a result, the parameters h1 and α are an obvious choice for bifurcation parameters. It should be noted that the inclusion of the delay parameter in the expression of the non-trivial predator nullcline results in a new dependence between the parameter τ and the number and location of interior equilibrium points.

Figure  is a one parameter bifurcation diagram where we choose h1 as the bifurcation parameter and for two fixed but different values of parameter α. In both cases, the non-delayed system (2) has exactly one positive equilibrium point before the transcritical threshold value h1=0.1667. The extinction state E0 is always unstable for any permissible value of h1, implying that the system can never settle down in the extinction state. The prey only equilibrium point E1 remains unstable before the transcritical bifurcation threshold value h1=0.1667. These results show that the coexistence equilibrium point becomes the only attractor, implying that the system is globally stable at the positive equilibrium point up to the transcritical threshold value h1=0.1667. The system experiences coexistence in two positive equilibrium points for 0.1667<h1<0.661 when α=1 and 0.1667<h1<0.818 when α=2.217 after which the coexistence lost through saddle-node bifurcation. For α=1, h1 induces a bi-stability situation between one coexistence equilibrium point and the prey only equilibrium point. The basin of attractions of these two attractors is separated by the stable manifold of a coexistence equilibrium which is a saddle point.

For α=2.217, we observe two types of bi-stability. The first type is node-cycle bi-stability which occurs between a locally stable positive equilibrium point and a locally stable limit cycle when h1 crosses the transcritical bifurcation threshold 0.1667. The second type is node–node bi-stability which occurs between the prey only equilibrium point and a coexistence equilibrium point when h1 crosses the Hopf bifurcation threshold 0.752 and continues up to saddle node bifurcation threshold 0.818. After the saddle-node bifurcation, the prey only equilibrium point E1 becomes the only stable attractor of the system implying E0 becomes globally stable. These imply that, in the case of cooperative hunting among predators, less hunting time is advantageous for the coexistence of prey and their predators. On the other hand, the system will eventually enter a predator extinction state if the predators require more time to hunt their prey. Note that increasing the value of α induces Hopf bifurcations in the model dynamics through which the prey and their predators start to coexist in oscillatory mode for some intermediate value of hunting cooperation parameter h1.

The model analysis for the delayed system (15) reveals that the number of positive equilibrium points has a dependence on the magnitude of the delay parameter τ. We can check the plots of the non-trivial nullcline for a representative set of parameter values σ=10, κ=1.2, α=1, h1=0.1, β=0.1 and different values of τ. Plots of the non-trivial predator nullcline g1(n,p,τ) for three different values of τ are in Figure . Now we are interested to understand the shifting of bifurcation thresholds with the variation of the delay parameter τ.

Figure 4. Plots of non-trivial prey nullcline f1(n,p) and predator nullcline g1(n,p,τ) for σ=10, κ=1.2, α=1, h1=0.1, β=0.1 and three different values of τ: τ=0, τ=1.5, τ=3.

Figure 4. Plots of non-trivial prey nullcline f1(n,p) and predator nullcline g1(n,p,τ) for σ=10, κ=1.2, α=1, h1=0.1, β=0.1 and three different values of τ: τ=0, τ=1.5, τ=3.

Two bifurcation diagrams are presented in Figure  with respect to the parameter h1 and some different but fixed values of the delay parameter τ as mentioned at the caption of the figure. If we compare the bifurcation thresholds between the Figures (a) and , we find that the transcritical bifurcation threshold shifts from 0.1667(τ=0) to 0.02737(τ=0.3,β=0.5) and the saddle-node bifurcation threshold shifts from 0.661(τ=0) to 0.485(τ=0.3,β=0.5) to 0.326(τ=0.6,β=0.5). Note that less hunting time is required to induce the dynamical shift from single coexistence to double coexistence and from double coexistence to no coexistence if juvenile predators require longer maturation times. Similar comparisons can be made between the thresholds in Figures (b) and (a) for transcritical, saddle-node and Hopf bifurcations. For α=2.217, we find that the first Hopf bifurcation threshold shifts from 0.132(τ=0) to 0.044(τ=0.05,β=0.1). The transcritical bifurcation threshold shifts from 0.1667(τ=0) to 0.1617(τ=0.05,β=0.1). Second Hopf bifurcation threshold shifts from 0.752(τ=0) to 0.748(τ=0.05,β=0.1). The saddle node bifurcation threshold shifts from 0.818(τ=0) to 0.812(τ=0.05,β=0.1). It is worth noting that increasing the magnitude of the delay causes all of the above bifurcations to occur for a lower value of the parameter h1 (see Figures b,  and ). It is noted that for an increased maturation time of juvenile predators, reduced hunting time induces the system to exhibit oscillatory coexistence and further ruling out the oscillations to have steady-state coexistence of prey and their predators.

Figure 5. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=1, β=0.5: (a) τ=0.3 and (b) τ=0.6.

Figure 5. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=1, β=0.5: (a) τ=0.3 and (b) τ=0.6.

Figure 6. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, β=0.1: (a) τ=0.05 and (b) τ=0.06.

Figure 6. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, β=0.1: (a) τ=0.05 and (b) τ=0.06.

Figure 7. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, β=0.1: (a) τ=0.0701 and (b) τ=0.11.

Figure 7. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, β=0.1: (a) τ=0.0701 and (b) τ=0.11.

Figure 8. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, τ=0.2, β=0.5.

Figure 8. The bifurcation diagram with respect to the parameter h1 with σ=10, κ=1.2, α=2.217, τ=0.2, β=0.5.

We can also check the bifurcation diagram with the time delay τ as bifurcation parameter (Figure ). This bifurcation diagram shows that the coexistence of prey and their predator of the considered model system is only possible within a certain range of delay parameter magnitude. For a unique coexistence equilibrium point, a short maturation delay is required. For other fixed parameter values as in the caption of Figure , when the maturation time is substantially low then due to increment in the magnitude of the delay parameter the system starts to exhibit oscillatory coexistence after crossing the first Hopf bifurcation threshold τH1=0.03. The originated limit cycle through Hopf bifurcation is the only stable attractor of the system for τH1<τ<τTC, where τTC=0.138 is the transcritical bifurcation threshold. This is due to initial negative feedback on the growth of predator population. There is a node-cycle bi-stability between the prey only equilibrium point E1 and the locally stable limit cycle for τTC<τ<τH2, where τH2=0.98 is the second Hopf bifurcation threshold through which the associated unstable positive equilibrium point becomes locally stable and remain stable until becomes infeasible through saddle-node bifurcation at τSN=1.557. We also attempt to comprehend the shift of bifurcation curves and corresponding co-dimension two bifurcations in the h1α parametric plane when maturation delay is included in the corresponding predator–prey model (Figure ). It is observed when a permissible amount of maturation delay is allowed in the model system, then both the saddle-node bifurcation curve and the Hopf bifurcation curve shift to a left-upper position in h1α parametric domain. The transcritical bifurcation curve moves to left in the parametric domain if delay is incorporated. When delay is taken into account, the cusp bifurcation point, which is the meeting point of the saddle-node bifurcation curve and transcritical bifurcation, shifts from (0.17,2.487)(τ=0) to (0.11,2.499)(τ=0.6), i.e. moves to the left-upper position in the parametric plane. The Bogdanov–Takens bifurcation point, where the saddle-node bifurcation curve intersects the Hopf bifurcation curve, shifts from (0.394,9.241)(τ=0) to (0.5,17.77)(τ=0.6), i.e. moves to the right-upper of the h1α parametric domain.

Figure 9. The bifurcation diagram with respect to the parameter τ with σ=10, κ=1.2, α=2, h1=0.1, β=0.5.

Figure 9. The bifurcation diagram with respect to the parameter τ with σ=10, κ=1.2, α=2, h1=0.1, β=0.5.

Figure 10. (Shift of bifurcation curves and bifurcation thresholds) σ=0.4, κ=1.2, β=0.1. Black and magenta colour solid curves represent saddle node bifurcation curves (SNC) for τ=0 and τ=0.6 respectively. Red and blue colour broken curves represent Hopf bifurcation curves (HC) for τ=0 and τ=0.6 respectively. Black and magenta colour vertical dash–dot curves represent transcritical bifurcation curves for (TC) for τ=0 and τ=0.6 respectively. BT and CP stand for Bogdanov–Takens bifurcation threshold and cusp bifurcation thresholds respectively.

Figure 10. (Shift of bifurcation curves and bifurcation thresholds) σ=0.4, κ=1.2, β=0.1. Black and magenta colour solid curves represent saddle node bifurcation curves (SNC) for τ=0 and τ=0.6 respectively. Red and blue colour broken curves represent Hopf bifurcation curves (HC) for τ=0 and τ=0.6 respectively. Black and magenta colour vertical dash–dot curves represent transcritical bifurcation curves for (TC) for τ=0 and τ=0.6 respectively. BT and CP stand for Bogdanov–Takens bifurcation threshold and cusp bifurcation thresholds respectively.

5. Discussion

In this article, we have studied the dynamics of prey–predator models incorporating hunting cooperation among predator species, for which a hunting time is considered in a predation term. In Section 2, for the model (2), when a hunting time h1 is large (such that κ1+h1κ1 holds), the number of coexistence equilibria is varied by the product of a cooperation rate and an intrinsic rate of growth of prey species ασ. Therefore, the introduction of hunting cooperation among predators emerges as a pivotal factor in shaping both the ecological equilibrium states (Figures  and ) and thus the dynamical stability of the underlying system. By varying h1, one can also observe a characteristic bifurcation dynamics including not only transcritical bifurcation at a predator-free equilibrium but also saddle-node bifurcation at a coexistence equilibrium for small α and saddle-node, Hopf bifurcation at a coexistence equilibrium for large α (see Figure ). This suggests that, when cooperation among predator species is weak, prey species tend to evade extinction and instead coexist with predators, especially when the predators have sufficiently long hunting times. In Section 3, we have introduced the model (15) in which the maturation delay for predator species τ is incorporated. Even if κ1+h1κ exceeds 1, a predator-free equilibrium is still asymptotically stable as long as it does not exceeds eβτ. This implies that predator species tend to go extinct when τ is large because the long maturation delay suppresses the increment of the predation term for predator species. Figure  represents that a brief hunting time (h1) suffices to trigger shifts in ecological coexistence when there is a substantial maturation delay (τ). Even with short hunting time, a large τ prompts transitions: from single stable state to double stable states and from double stable states to no coexistence stable state. As maturation delay (τ) increases, oscillations in population size initially intensify (large amplitude) at intermediate τ, showcasing a balance between predator-free and coexistence states. Subsequently, a state of bi-stability emerges between these points. However, with further τ values, coexistence collapses, leaving the predator-free equilibrium as the sole stable state. This implies that extended maturation delays can disrupt ecological balance, favouring a less diverse and stable state dominated by the predator-free equilibrium (see Figures  and ). It is worth noting from Figures  and  that reduced cooperation among predators results in intricate bifurcation dynamics when hunting time serves as the bifurcation parameter. This suggests that when predators lack cooperation in hunting, the population dynamics of predators and prey become highly sensitive to changes in hunting time, leading to unpredictable variations. From a biological point of view, such unforeseeable dynamics will arise from variation in behaviour of the predators for hunting due to the less cooperation effect. Further theoretical analysis on the bifurcation dynamics of the model (15) is our future work.

Figure 11. (Shift of bifurcation curves and bifurcation thresholds) σ=0.4, κ=1.2, β=0.1. Black, magenta and green colour solid curves represent saddle node bifurcation curves (SNC) for τ=0, τ=0.6 and τ=3 respectively. Red, blue and brown colour broken curves represent Hopf bifurcation curves (HC) for τ=0, τ=0.6 and τ=3 respectively. Black and magenta colour vertical dash–dot curves represent transcritical bifurcation curves for (TC) for τ=0 and τ=0.6 respectively. BT and CP stand for Bogdanov–Takens bifurcation threshold and cusp bifurcation thresholds respectively.

Figure 11. (Shift of bifurcation curves and bifurcation thresholds) σ=0.4, κ=1.2, β=0.1. Black, magenta and green colour solid curves represent saddle node bifurcation curves (SNC) for τ=0, τ=0.6 and τ=3 respectively. Red, blue and brown colour broken curves represent Hopf bifurcation curves (HC) for τ=0, τ=0.6 and τ=3 respectively. Black and magenta colour vertical dash–dot curves represent transcritical bifurcation curves for (TC) for τ=0 and τ=0.6 respectively. BT and CP stand for Bogdanov–Takens bifurcation threshold and cusp bifurcation thresholds respectively.

Acknowledgements

The authors wish to express their gratitude to the editors and anonymous referees for helpful comments and valuable suggestion which improved the quality of this paper.

Disclosure statement

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

Additional information

Funding

This research was partially supported by JSPS KAKENHI Grant Number 19K14598 (YE).

References

  • Kot M. Elements of mathematical ecology. Cambridge: Cambridge University Press; 2001.
  • Kuang Y, Freedman HI. Uniqueness of limit cycles in gause-type models of predator-prey systems. Math Biosci. 1988;88:67–84. doi: 10.1016/0025-5564(88)90049-1
  • Přibylová L, Berec L. Predator interference and stability of predator–prey dynamics. J Math Biol. 2015;71:301–323. doi: 10.1007/s00285-014-0820-9
  • Tyutyunov YV, Titova LI. From Lotka–Volterra to Arditi–Ginzburg: 90 years of evolving trophic functions. Biol Bull Rev. 2020;10:167–185. doi: 10.1134/S207908642003007X
  • Arditi R, Ginzburg LR. Coupling in predator–prey dynamics: ratio-dependence. J Theor Biol. 1989;139:311–326. doi: 10.1016/S0022-5193(89)80211-5
  • Cosner C, DeAngelis DL, Ault JS, et al. Effects of spatial grouping on the functional response of predators. Theor Popul Biol. 1999;56:65–75. doi: 10.1006/tpbi.1999.1414
  • Holling CS. Some characteristics of simple types of predation and parasitism. Can Entomol. 1959;91:385–398. doi: 10.4039/Ent91385-7
  • Kuang Y, Beretta E. Global qualitative analysis of a ratio-dependent predator–prey system. J Math Biol. 1998;36:389–406. doi: 10.1007/s002850050105
  • Xiao D, Ruan S. Global dynamics of a ratio-dependent predator–prey system. J Math Biol. 2001;43:268–290. doi: 10.1007/s002850100097
  • Bazykin AD. Nonlinear dynamics of interacting populations. Singapore: World Scientific; 1998.
  • Lu M, Huang J. Global analysis in Bazykin's model with Holling II functional response and predator competition. J Differ Equ. 2021;280:99–138. doi: 10.1016/j.jde.2021.01.025
  • Courchamp F, Berec L, Gascoigne J. Allee effects in ecology and conservation. Oxford: OUP Oxford; 2008.
  • Stephens PA, Sutherland WJ, Freckleton RP. What is the Allee effect?. Oikos. 1999;87:185–190. doi: 10.2307/3547011
  • Hilker FM. Population collapse to extinction: the catastrophic combination of parasitism and Allee effect. J Biol Dyn. 2010;4:86–101. doi: 10.1080/17513750903026429
  • Morozov A, Petrovskii S, Li BL. Spatiotemporal complexity of patchy invasion in a predator–prey system with the Allee effect. J Theor Biol. 2006;238:18–35. doi: 10.1016/j.jtbi.2005.05.021
  • Banerjee M, Takeuchi Y. Maturation delay for the predators can enhance stable coexistence for a class of prey–predator models. J Theor Biol. 2017;412:154–171. doi: 10.1016/j.jtbi.2016.10.016
  • Sen M, Banerjee M, Morozov A. Bifurcation analysis of a ratio-dependent prey–predator model with the Allee effect. Ecol Complex. 2012;11:12–27. doi: 10.1016/j.ecocom.2012.01.002
  • Erbach A, Lutscher F, Seo G. Bistability and limit cycles in generalist predator-prey dynamics. Ecol Complex. 2013;14:48–55. doi: 10.1016/j.ecocom.2013.02.005
  • Sen D, Petrovskii S, Ghorai S, et al. Rich bifurcation structure of prey–predator model induced by the Allee effect in the growth of generalist predator. Int J Bifurcat Chaos. 2020;30:Article ID 2050084. doi: 10.1142/S0218127420500844
  • Beddington JR. Mutual interference between parasites or predators and its effect on searching efficiency. J Anim Ecol. 1975;44:331–340. doi: 10.2307/3866
  • Cantrell RS, Cosner C. On the dynamics of predator–prey models with the Beddington–DeAngelis functional response. J Math Anal Appl. 2001;257:206–222. doi: 10.1006/jmaa.2000.7343
  • DeAngelis DL, Goldstein RA, O'Neill RV. A model for tropic interaction. Ecology. 1975; 56:881–892. doi: 10.2307/1936298
  • Nilsson PA, Lundberg P, Brönmark C, et al. Behavioral interference and facilitation in the foraging cycle shape the functional response. Behav Ecol. 2007;18:354–357. doi: 10.1093/beheco/arl094
  • Zhang Y, Richardson JS. Unidirectional prey–predator facilitation: apparent prey enhance predators' foraging success on cryptic prey. Biol Lett. 2007;3:348–351. doi: 10.1098/rsbl.2007.0087
  • Berec L. Impacts of foraging facilitation among predators on predator–prey dynamics. Bull Math Biol. 2010;72:94–121. doi: 10.1007/s11538-009-9439-1
  • Alves MT, Hilker FM. Hunting cooperation and Allee effects in predators. J Theor Biol. 2017;419:13–22. doi: 10.1016/j.jtbi.2017.02.002
  • Beretta E, Kuang Y. Convergence results in a well-known delayed predator–prey system. J Math Anal Appl. 1996;204:840–853. doi: 10.1006/jmaa.1996.0471
  • Kuang Y. Delay differential equations: with applications in population dynamics. Boston: Academic Press; 1993.
  • MacDonald N. Time lags in biological models. Vol. 27. Springer Science & Business Media; 2013.
  • Xiao Y, Chen L. Modeling and analysis of a predator–prey model with disease in the prey. Math Biosci. 2001;171:59–82. doi: 10.1016/S0025-5564(01)00049-9
  • Cushing JM. Integrodifferential equations and delay models in population dynamics. Vol. 20. Springer Science & Business Media; 2013.
  • Hadeler KP, Mackey MC, Stevens A. Topics in mathematical biology. Berlin: Springer; 2017.
  • Ruan S. On nonlinear dynamics of predator-prey models with discrete delay. Math Model Nat Phenom. 2009;4:140–188. doi: 10.1051/mmnp/20094207
  • Volterra V. Variations and fluctuations of the number of individuals in animal species living together. Anim Ecol. 1926;409–448.
  • Mackey MC, Glass L. Oscillation and chaos in physiological control systems. Science. 1977;197:287–289. doi: 10.1126/science.267326
  • Nakaoka S, Saito Y, Takeuchi Y. Stability, delay, and chaotic behavior in a Lotka–Volterra predator–prey system. Math Biosci Eng. 2006;3:173–187. doi: 10.3934/mbe.2006.3.173
  • Tang S, Chen L. Global qualitative analysis for a ratio-dependent predator–prey model with delay. J Math Anal Appl. 2002;266:401–419. doi: 10.1006/jmaa.2001.7751
  • Xiao D, Ruan S. Multiple bifurcations in a delayed predator–prey system with nonmonotonic functional response. J Differ Equ. 2001;176:494–510. doi: 10.1006/jdeq.2000.3982
  • Turchin P. Complex population dynamics. Princeton University Press; 2013.
  • Perko L. Differential equations and dynamical systems. New York: Springer-Verlag; 2000.
  • Martin A, Ruan S. Predator–prey models with delay and prey harvesting. J Math Biol. 2001;43:247–267. doi: 10.1007/s002850100095
  • Roy J, Banerjee M. Global stability of a predator-prey model with generalist predator. Appl Math Lett. 2023;142:Article ID 108659. doi: 10.1016/j.aml.2023.108659
  • Sen M, Banerjee M, Morozov A. Stage-structured ratio-dependent predator-prey models revisited: when should the maturation lag result in systems' destabilization?. Ecol Complex. 2014;19:23–34. doi: 10.1016/j.ecocom.2014.02.001
  • Wangersky PJ, Cunningham WJ. Time lag in prey–predator population models. Ecology. 1957;38:136–139. doi: 10.2307/1932137
  • Gourley SA, Kuang Y. A stage structured predator–prey model and its dependence on maturation delay and death rate. J Math Biol. 2004;49:188–200. doi: 10.1007/s00285-004-0278-2
  • Liu S, Beretta E. A stage-structured predator–prey model of Beddington–DeAngelis type. SIAM J Appl Math. 2006;66:1101–1129. doi: 10.1137/050630003
  • Liu S, Chen L, Liu Z. Extinction and permanence in nonautonomous competitive system with stage structure. J Math Anal Appl. 2002;274:667–684. doi: 10.1016/S0022-247X(02)00329-3

Appendices

Appendix 1.

Well-posedness

In this appendix, we prove the well-posedness of the model (2).

Proposition A.1

The model (2) is well-posed.

Proof.

The right-hand side of (2) is completely continuous and locally Lipschitzian on R+2. It follows that the solution of (2) exists and is unique on [0,T) for some T>0. One can obtain that n(t)>0,p(t)>0 for all t[0,T). Furthermore, for t[0,T), we obtain ddt(n(t)+p(t))=σn(t)(1n(t)κ)p(t)σ(κn(t))p(t)σκmin{σ,1}(n(t)+p(t)),which implies that (n(t),p(t)) is uniformly bounded on [0,T) by the positivity of n(t) and p(t) on [0,T). Thus T=, that is, (n(t),p(t)) exists and is unique and n(t)>0,p(t)>0 for all t>0. One can easily prove that the solution of (2) is continuous with respect to the initial values. This completes the proof.

Appendix 2.

Hopf bifurcation for h1=0 for (2)

Assume that κ>1 holds. Substituting h1=0 into (Equation12) and (Equation13), one can see that trJ+=σ1κ(1+αp)+αp1+αp=αpσκ1+αp,detJ+=(p(1+αp)σ1κ(1+αp))αp1+αp+p(1+αp)=11+αp((p(1+αp)2σκ)αp1+αp+p(1+αp)2)=σ1+αp((12κ+αp)αp1+αp+11κ+αp).It is noting that detJ+>0 always holds true because (12κ+αp)αp1+αp+11κ+αp>(1+2αp)(11κ)>0. We have (trJ+)24detJ+=(αpσκ1+αp)24(p(1+αp)σ1κ(1+αp))αp1+αp4p(1+αp)=1(1+αp)2((αp+σκ)24σ(11κ+αp)(1+2αp))=1(1+αp)2((18σ)(αp)2+σ(10κ12)αp+σκ2(σ4κ(κ1))).Since σ(11κ)+(ασ1)p2α(p)2α2(p)3=0,i.e. σ(11κ)+σαp=p(1+αp)2,αp(α) is strictly monotone increasing on α. Due to the characteristic equation det(J+λI)=λ2(trJ+)λ+detJ+=0, if σ>max{18,4κ24κ}, then there exists a unique α¯>0 such that (18σ)(α¯p(α¯))2+σ(10κ12)α¯p(α¯)+σκ2(σ4κ(κ1))=0and the following statement holds true:

  1. If 0αp<σκ, it holds that trJ+<0.

  2. If αp>σκ, it holds that trJ+>0.

Therefore, if σ>max{18,4κ24κ}, then (2) undergoes Hopf bifurcation at E=E+ when α crosses a critical value αH. Here αH is a unique solution of αHp(αH)=σκ.

Appendix 3.

Notations in Section 2.4.1

The notations used in Section 2.4.1 are as follows: a10=n(σk+(αp+1)2ph1(npαh1+h1n+1)(1+h1(αp+1)n))+σ(1nk)(αp+1)p1+h1(αp+1)n,a01=nnpαh1+h1n+1(2αp+1(αp+1)ph1αnnpαh1+h1n+1),a20=n(p3α2h12p2αh1ph1)h1(αp+1)(npαh1+h1n+1)3σk+(αp+1)2ph1(npαh1+h1n+1)2,a11=(αp1)h1n2αp1(1+h1(αp+1)n)3,a02=n1+h1(αp+1)n(α(np2α2h1+2npαh1+nh1+2αp+1)h1αn(npαh1+nh1+1)2),a30=n(p4α3h123p3α2h123p2αh12ph12)h1(αp+1)(npαh1+nh1+1)3(1+h1(αp+1)n)+(p3α2h12p2αh1ph1)h1(αp+1)(npαh1+nh1+1)2(1+h1(αp+1)n),a21=(αp+1)h1(h(αp+1)p+3αp+1)(1+h1(αp+1)p)4,a12=(1+h12(αp+1)n2+2npαh1)α(1+h1(αp+1)n)4, a03=(nαh1α)h1αn2(npαh1+nh1+1)3(1+h1(αp+1)n),b10=11+h1(αp+1)n((αp+1)pn(αp+1)2ph1npαh1+nh1+1),b01=11+h1(αp+1)n(n(αp+1)+npαn2(αp+1)ph1αnpαh1+nh1+1)1,b20=(p2α+p)h1(αp+1)(npαh1+nh1+1)2(1+h1(αp+1)n),b11=npαh1+nh1+2αp+1(npαh1+nh1+1)3,b02=nα(nh1+1)(npαh1+nh1+1)3,b30=(p3α2h12p2αh1ph1)h1(αp+1)(npαh1+nh1+1)3(1+h1(αp+1)n),b21=(αp+1)h1(h1(αp+1)n+3αp+1)(1+h1(αp+1)n)4,b12=α(n2pαh12+n2h12+2npαh11)(npαh1+nh1+1)4,b03=(n2αh1+nα)h1αn(npαh1+nh1+1)3(1+h1(αp+1)n).In addition, A1 and A2 are defined as follows: A1=(a20a012a11a01a10+a02a102)X12+h10(2a02a10a11a01)X1X2+h102a02X22+(a12a01a102a03a103+a30a013a21a012a10)X13+h10(2a12a10a01a21a0123a03a102)X12X2+h102(a12a013a03a10)X1X22h10a03X23,A2=(b20a012b11a01a10+b02a102)X12+h10(2b02a10b11a01)X1X2+h102b02X22+(b30a013+b12a01a102b21a012a10b03a103)X13+h10(2b12a10a01b21a0123b03a102)X12X2+h102(b12a013b03a10)X1X22h10b03X23.