1,055
Views
14
CrossRef citations to date
0
Altmetric
Research Article

Dynamics of the interaction of plankton and planktivorous fish with delay

& | (Reviewing Editor)
Article: 1074337 | Received 29 Oct 2014, Accepted 12 Jul 2015, Published online: 12 Aug 2015

Abstract

This paper is devoted to the study of a plankton–fish ecosystem model. The model represents the interaction between phytoplankton, zooplankton, and fish with Holling II functional response consisting of carrying capacity and constant intrinsic growth rate of phytoplankton. It is observed that if the carrying capacity of phytoplankton population crosses a certain critical value, the system enters into Hopf bifurcation. We have introduced discrete time delay due to gestation in the functional response term involved with the growth equation of planktivorous fish. We have studied the effect of time delay on the stability behavior. In addition, we have obtained an estimate for the length of time delay to preserve the stability of the model system. Existence of Hopf bifurcating small amplitude periodic solutions is derived by considering time delay as a bifurcation parameter. It is observed that constant intrinsic growth rate of phytoplankton and mortality rate of planktivorous fish play an important role in changing one steady state to another steady state and oscillatory behavior of the system. Computer simulations illustrate the results.

AMS Subject Classifications:

Public Interest Statement

Economically important fish species have long been regarded in isolation from each other and their habitat. In order to comprehensively assess the impacts of fisheries, the entire habitat must be considered. Only then will a sustainable and economic fishery system be possible. Predation by fish determines the abundance of herbivorous zooplankton which in turn regulates the level of phytoplankton. Also, changes in the abundance of planktivorous fish do affect both the phytoplankton and zooplankton. The relationship between phytoplankton, zooplankton, and fish culture is of paramount importance in determining the water quality, on one hand, and the natural productivity and the fish production, on the other. It is observed that constant intrinsic growth rate of phytoplankton and mortality rate of planktivorous fish play an important role in changing one steady state to another steady state and oscillatory behavior of the system.

1. Introduction

It is well known that plankton plays an integral role in marine ecosystem. Phytoplankton and zooplankton are the main two types of plankton. Many models have already been built to simulate zooplankton–phytoplankton interactions. In marine ecosystem, much attention has been paid to the effect of fish and plankton biomass. Top-down and bottom-up effects on plankton–fish dynamics have been discussed by Scheffer (Citation1991) and Pal and Chatterjee (Citation2012). The authors have discussed that at high fish densities, zooplankton is controlled by fish predation and algal biomass is light or nutrient limited, whereas at low fish densities, zooplankton is food limited and phytoplankton density is controlled by zooplankton grazing. Scheffer, Rinaldi, and Kuznetsov (Citation2000) have discussed about isocline analysis and simulations to show how an increase in fish predation may affect plankton dynamics in the model and to explain which type of bifurcations may arise. Role of gestation delay in a plankton–fish model under stochastic fluctuations has been discussed by Mukhopadhyay and Bhattacharyya (Citation2008). Realistic patterns of patchiness in plankton–fish dynamics have been studied by Upadhyay, Kumari, and Rai (Citation2009). The authors also studied both one-dimensional and two-dimensional reaction–diffusion model of nutrient–phytoplankton–zooplankton–fish interaction. As studied by Biktashev, Brindley, and Horwood (Citation2003), the food source for fish larvae depends on the zooplankton dynamics, which in turn is coupled to larval populations through the zooplankton mortality term. A conceptual mathematical model of fish and zooplankton populations inhabiting the lakes Naroch and Myastro has been developed and studied by Medvinsky et al. (Citation2009).

In general, delay differential equations exhibit much more complicated dynamics than ordinary differential equations since a time delay can cause destabilization of equilibria and can induce various oscillations and periodic solutions. As observed by Bischi (Citation1992), the effects of the time delay involved nutrient recycling on resilience, that is, the rate at which a system returns to a stable steady state following a perturbation. The researchers have suggested that characterizing a system by oscillatory behavior, an increase in the distributed time delay, can have a stabilizing effect. But it has been found that the introduction of time delays is a destabilizing process, in the sense that increasing the time delay could cause a stable equilibrium to become unstable and/or cause the populations to fluctuate (Cushing, Citation1977; Freedman & Rao, Citation1983). Chemostat-type models incorporating discrete delays have been investigated by Freedman, So, and Waltman (Citation1989). Also, delayed models in population biology have been discussed in Gopalsamy (Citation1984), Kuang (Citation1993), Khare, Misra, Singh, and Dhar (Citation2011) and Liu, Beretta, and Breda (Citation2010). A discrete time delay term to the model of Beretta, Bischi, and Solimano (Citation1991) was introduced in Ruan (Citation1995). This discrete time delay term may be considered as delay due to gestation. The author also allowed the washout rates for nutrient and plankton to be different. Recently, effect of delay on nutrient cycling in phytoplankton–zooplankton interaction model has been studied by Das and Ray (Citation2008). Rehim and Imran (Citation2012) have induced a discrete time delay to both the consume response function and distribution of toxic substance terms in phytoplankton–zooplankton model. They have found out a range of gestation delays which initially impart stability, then induce instability and ultimately lead to periodic behavior. Due to gestation of prey, a discrete time delay toxin producing phytoplankton–zooplankton system shows rich dynamic behavior including chaos and limit cycles (Singh & Gakkhar, Citation2012). In particular, Gazi and Bandyopadhyay (Citation2006) have introduced discrete time delay due to recycling of dead organic matters and gestation of nutrients to the growth equations of various trophic levels. They have studied the effect of time delay on the stability behavior and obtained an estimate for the length of time delay to preserve the stability of the model system. Effects of time delay on a harvested predator–prey model have discussed in Gazi and Bandyopadhyay (Citation2008), Maity, Patra, and Samanta (Citation2007) and Yongzhen, Min, and Changguo (Citation2011).

Here, an open system is considered with three interacting components consisting of phytoplankton (P), zooplankton (Z), and planktivorous fish (F). In this paper, a plankton–fish interaction model is described. The stability of equilibrium point is described. We have derived the conditions for instability of the system around the interior equilibrium and Hopf bifurcation in presence of delay and non-delay models. Direction of Hopf bifurcation is discussed. Numerical simulations under a set of parameter values have been performed to support our analytical results.

2. The mathematical model

Let P(t) be the concentration of the phytoplankton at time t with carrying capacity K and constant intrinsic growth rate r. Let Z(t) be the concentration zooplankton biomass and F(t) be the concentration of planktivorous fish biomass present at any instant of time t.

Let α1 be the maximal zooplankton ingestion rate and α2 maximal zooplankton conversion rate for the growth of zooplankton, respectively (α2α1). Let γ1 be the maximal planktivorous fish ingestion rate and γ2(γ2γ1) be the maximal planktivorous fish conversion rate due to grazing of herbivorous zooplankton. Further, μ1, μ2 , μ3 denote the mortality rates of the phytoplankton, zooplankton, and planktivorous fish biomass, respectively. We choose Holling type II functional form to describe the grazing phenomena with K1 and K2 as half saturation constant.

The basic equations with all of the parameters are:(1) dPdt=rP(1-PK)-α1PZK1+P-μ1PJ1(P,Z,F)dZdt=α2PZK1+P-γ1ZFK2+Z-μ2ZJ2(P,Z,F)dFdt=γ2ZFK2+Z-μ3FJ3(P,Z,F).(1)

Set X=(P,Z,F)TR+3 and J(X)=J1(X),J2(X),J3(X)T, with J:R+3R3. The system (Equation 1) can be written in compact form as X˙=J(X). Its Jacobian is(2) V=r-2rPK-α1K1Z(K1+P)2-μ1-α1PK1+P0α2K1Z(K1+P)2α2PK1+P-γ1K2F(K2+Z)2-μ2-γ1ZK2+Z0γ2K2F(K2+Z)2γ2ZK2+Z-μ3.(2)

3. Some preliminary results

3.1. Positive invariance

The system (Equation 1) is not homogeneous. However, it is easy to check that whenever choosing X(0)R+3 with Xi=0, for i=1,2,3, then Ji(X)Xi=00. This ensures that the solution remains within the positive orthant, ensuring the biological well posedness of the system.

3.2. Boundedness of the system

Theorem 1

All the solutions of (Equation 1) are ultimately bounded.

Proof

We define a functionw=P+Z+F.

We have dPdtrP(1-rK) which gives Pc1Kc1+e-rtK as t, where c1 is constant.

For the large value of t, we get dwdtrP-D0w, where D0=min{μ1,μ2,μ3}.

Let x(t) be the solution of dxdt+xD0=rK, satisfying x(0)=w(0).

Then, x(t)=rKD0+(w(0)-rKD0)e-D0trKD0 as t. By comparison, it follows that lim supt[P(t)+Z(t)+F(t)]rKD0, proving the theorem.

3.3. Equilibria

The system (Equation 1) possesses the following four equilibria: the plankton free equilibrium E0=(0,0,0), the zooplankton free equilibrium E01=K(r-μ1)r,0,0, the planktivorous fish free equilibrium E1=(P1,Z1,0), and possibly the coexistence of the three populations E=(P,Z,F).

3.3.1. Plankton free equilibrium

E0(0,0,0) is always feasible; the Jacobian (Equation 2) evaluated at this equilibrium has the eigenvalues -μ2<0, -μ3<0 and r-μ1>0. Clearly, E0 is always unstable.

3.3.2. Zooplankton free equilibrium

At E01, the Jacobian (Equation 2) factorizes to give three explicit eigenvalues -(r-μ1)<0, -μ3<0 and μ2(R0-1), where R0 is defined below. Thus, E01 is locally asymptotically stable if and only if(3) R0=α2K(r-μ1)μ2[r(K1+K)-Kμ1]<1,(3)

3.3.3. Planktivorous fish free equilibrium

At E1, the population levels areP1=μ2K1α2-μ2,Z1=α2K1[K(α2-μ2)(r-μ1)-rμ2K1]α1K(α2-μ2)2.

This equilibrium is feasible if it is satisfying(4) α2>μ2,K>rμ2K1(α2-μ2)(r-μ1).(4)

At E1, the Jacobian (Equation 2) factorizes to give one explicit eigenvalue γ2Z1(K2+Z1)-1-μ3 and the quadratic λ2+rP1K-P1α1Z1(K1+P1)2λ+α1α2K1P1Z1(K1+P1)3=0. The Routh–Hurwitz conditions for the latter are easily seen to hold, but only when (Equation 4) is satisfied. Thus, stability of E1 is then ensured by(5) R1=α2K1[K(α2-μ2)(r-μ1)-rK1μ2](γ2-μ3)μ3K2(α2-μ2)2<1.(5)

3.3.4. The coexistence equilibrium

The coexistence equilibrium E=(P,Z,F) cannot be found explicitly since Z=μ3K2γ2-μ3, F=[(α2-μ2)P-μ2K1]K2γ2γ1(γ2-μ3)(K1+P) and P=-B+B2+4AC2A, where A=r(γ2-μ3), B=-[(K-K1)r-Kμ1](γ2-μ3), C=-K[(r-μ1)K1-α1μ3K2].

Thus, the condition for the existence of the interior equilibrium point E(P,Z,F) is given by, P>0,Z>0,F>0. For feasibility, we certainly needγ2>μ3and0<r<minKμ1K-K1,α1μ3K2+μ1K1K1.Stability analysis of the positive interior equilibrium of the system (Equation 1)

The variational matrix of system (Equation 1) around the positive equilibrium E=(P,Z,F) isV=n11n120n21n22n230n320,

where n11=α1PZ(K1+P)2-rPK,n12=-α1PK1+P<0, n21=K1α2Z(K1+P)2>0, n22=γ1ZF(K2+Z)2>0, n23=-γ1ZK2+Z<0, n32=K2γ2F(K2+Z)2>0.

The characteristic equation is Q3+A1Q2+A2Q+A3=0,

where A1=-(n11+n22), A2=n11n22-n12n21-n32n23, A3=n11n32n23.

  • Case1: If n11>0, it implies that A3<0. Thus, E is unstable.

  • Case2: If n11<0, it implies that A3>0.

Here, A1>0, if rPK>α1PZ(K1+P)2+γ1ZF(K2+Z)2.

Also, A2>0 if -n12n21-n32n23>-n11n22 where n23n32<0, n12n21<0 and n11n22 is negative. So, A1A2-A3>0 if A1A2>A3. Therefore, according the Routh–Hurwitz criteria, all roots of the above cubic equation have negative real parts satisfying A1>0,A3>0, and A1A2-A3>0. Then, the system becomes locally asymptotically stable around E. Thus, depending upon system parameters, the system may exhibit stable or unstable behavior in this case.

Table 1. The table representing thresholds and stability of steady states

The analytical results are summarized in Table .

3.4. Hopf bifurcation at coexistence

Theorem 2

When the carrying capacity K crosses a critical value, say K, the system (Equation 1) enters into a Hopf bifurcation around the positive equilibrium, which induces oscillations of the populations.

Proof

The necessary and sufficient conditions for the existence of the Hopf bifurcation for K=K, if it exists, are (i) Ai(K)>0, i=1,2,3 (ii) A1(K)A2(K)-A3(K)=0 and (iii) A1(K)A2(K)+A1(K)A2(K)-A3(K)0. The transversality condition, the third (iii), is obtained observing that the eigenvalues of the characteristic equation, of the form χi=ui+ivi, must satisfy the transversality condition duidKK=K0.

We will now verify the Hopf bifurcation condition (iii); putting χ=u+iv in the characteristic equation, we get(6) (u+iv)3+A1(u+iv)2+A2(u+iv)+A3=0,(6)

On separating the real and imaginary parts and eliminating v, we get(7) 8u3+8A1u2+2u(A12+A2)+A1A2-A3=0.(7)

It is clear from the above that u(K)=0 if A1(K)A2(K)-A3(K)=0. Further, at K=K, u(K) is the only root since the discriminate 8u2+8A1u+2(A12+A2)=0 if  64A12-64(A12+A2)<0.

Further, differentiating (Equation 7) with respect to K, we have 24u2dudK+16A1ududK+2(A12+A2)dudK+2u[2A1dA1dK+dA2dK]+dSdK=0, where S=A1A2-A3.

At K=K, we have u(K)=0, so that above equation becomes dudKK=K=-dSdK2(A12+A2)0, providing the third condition (iii).

This ensures that the above system has a Hopf bifurcation around the positive interior equilibrium E.

Next, we perform detailed analysis of the bifurcation solutions to study the nature of Hopf bifurcation. We start by rewriting system (Equation 1) in the form X˙=MX+F(X)+OX4 where X=pzf : p=P-P, z=Z-Z, f=F-F and M=V.

Here, F=F1F2F3,

where F1=-α1Z(K1+P)3p3+α1K1(K1+P)3p2z+α1K1Z(K1+P)3-rKp2-α1K1(K1+P)2pz, F2=α2Z(K1+P)3p3-γ1F(K2+Z)3z3-α2K1(K1+P)3p2z+γ1K2(K2+Z)3z2f-α2K1Z(K1+P)3p2+γ1K2F(K2+Z)3z2+α2K1(K1+P)2pz-γ1K2(K2+Z)2zf, F3=γ2F(K2+Z)3z3-γ2K2(K2+Z)3z2f-γ2K2F(K2+Z)3z2+γ2K2(K2+Z)2zf.

Now, let the normalized eigenvectors of M and MT corresponding to the eigenvalues iω and -iω be q and p, respectively, where q=1|q|-n12n11-iω1n32iω, p=1|p|-n21n11+iω1-n23iω. Let us define the following multi-linear functions for the three vectors x=(x1,x2,x3)R3, y=(y1,y2,y3)R3, and w=(w1,w2,w3)R3.Bi(x,y)=j,k=13[δ2Fiδξjδξk]ξ=0xjykCi(x,y,w)=j,k,l=13[δ3Fiδξjδξkδξl]ξ=0xjykwl.

where B(x,y)=2α1K1Z(K1+P)3-rKx1y1-α1K1(K1+P)2x1y2-2α2K1Z(K1+P)3x1y1+2γ1K2F(K2+Z)3x2y2+α2K1(K1+P)2x1y2-γ1K2(K2+Z)2x2y3-2γ2K2F(K2+Z)3x2y2+γ2K2(K2+Z)2x2y3,C(x,y,w)=-6α1Z1(K1+P)3x1y1w1+2α1K1(K1+P)3x1y1w26α2Z(K1+P)3x1y1w1-6γ1F(K2+Z)3x2y2w2-2α2K1(K1+P)3x1y1w2+2γ1K2(K2+Z)3x2y2w36γ2F(K2+Z)3x2y2w2-2γ2K2(K2+Z)3x2y2w3.

With these, F(X) will be of the form F(X)=12B(X,X)+12C(X,X,X).

Solving the corresponding linear system, we getM-1B(q,q¯)=s1s2s3,

wheres1=-1n11n32n23-n23n322α1K1Z(K1+P)3-rKn122n112+ω2+α1K1(K1+P)2n12(n11+iω)n112+ω2+n12n23γ2K2(K2+Z)2in32ω-2γ2K2F(K2+Z)3,s2=-1n11n32n23-n11n23γ2K2(K2+Z)2in32ω-2γ2K2F(K2+Z)3,s3=-1n11n32n23[n21n32[2α1K1Z(K1+P)3-rKn122n112+ω2+α1K1(K1+P)2n12(n11+iω)n112+ω2]-n11n32[-2γ1K1Z(K1P)3n122n112+ω2+2γ1K2F(K2+Z)3-α2K1(K1+P)2n12(n11+iω)n112+ω2-γ1K2(K2+Z)2in32ω]+(n11n22-n12n21)γ2K2(K2+Z)2in32ω-2γ2K2F(K2+Z)3].

Again,(2iωI3-M)-1=1|2iωI3-M|-4ω2-2n22iω-n23n322iωn12n12n232iωn21-4ω2-2n11iω2n23iω-n11n23n21n32n32(2iω-n11)E,

where E=(2iω-n11)(2iω-n12)-n12n21.

Therefore,(2iωI3-M)-1b11q12+b12q1q2b21q12+b22q22+b23q1q2+b24q2q3b31q2q3+b32q22=1|2iωI3-M|(-4ω2-2n22iω-n23n32)τ1+2iωn12τ2+n12n23τ32iωn21τ1-(4ω2+2n11iω)τ2+(2n23iω-n11n23)τ3n21n32τ1+n32(2iω-n11)τ2+Eτ3,

where τ1=b11q12+b12q1q2; τ2=b21q12+b22q22+b23q1q2+b24q2q3τ3=b31q2q3+b32q22.

Using the invariant expression for the first lyapunov coefficient l1(0), we get the first lyapunov coefficients as Mukhopadhyay and Bhattacharyya (Citation2011). l1(0)=12ωRe[<p,c(q,q,q¯)-2<p,B(q,M-1,B(q,q¯)))>+<p,B(q¯,(2iωI-M)-1B(q,q))>].

Now, if l1(0)<0, the Hopf bifurcation will be of supercritical nature and there will be emergence of stable limit cycles.

4. The mathematical model with time delay

We have already discussed the usefulness of time delay in realistic modeling of ecosystems. Here, we consider the following modification of the model (Equation 1) incorporating discrete time delay in it. In this paper, we have introduced a discrete time delay term to the above model; this term may be considered as delay due to gestation. Here, τ is the discrete time delay.(8) dPdt=rP(1-PK)-α1PZK1+P-μ1PdZdt=α2PZK1+P-γ1ZFK2+Z-μ2ZdFdt=γ2Z(t-τ)F(t-τ)K2+Z(t-τ)-μ3F.(8)

The system (Equation 8) has to be analyzed with the following initial conditions,(9) P(0)>0,Z(0)>0,F(0)>0,t[-τ,0].(9)

4.1. Qualitative analysis of the model with time delay

 V¯=r-2rP¯K-α1K1Z¯(K1+P¯)2-μ1-α1P¯K1+P¯0α2K1Z¯(K1+P¯)2α2P¯K1+P¯-γ1K2F¯(K2+Z¯)2-μ2-γ1Z¯K2+Z¯0γ2K2F¯(K2+Z¯)2e-λτγ2Z¯K2+Ze-λτ-μ3.

Remark 1

The characteristic equation for the variational matrix V0 about the plankton free steady states E0 and E01 remains the same as obtained for the non-delayed system (Equation 8). Thus, in our model system, the delay has no effect on the stability nature of the system about E0 and E01.

To find conditions for the other local asymptotic stability of system (Equation 8), we use the following theorem from Gopalsamy (Citation1992).

Theorem 3

If DE(λ,τ)=0 denotes the characteristic equation, then a set of necessary and sufficient conditions for the equilibrium to be asymptotically stable for all τ0 are

(i)

The real parts of all the roots of DE(λ,0)=0 are negative.

(ii)

For all real ω and any τ0, DE(iω,τ)0, where i=-1.

The characteristic equation for the variational matrix V1 about E1 takes the following form(10) DE1(λ,τ)=λ3+lλ2+mλ+n-e-λτ(fλ2+gλ+h)=0,(10)

where l=rP1K-α1P1Z1(K1+P1)2+μ3, m=α1α2K1P1Z1(K1+P1)3+μ3(rP1K-α1P1Z1(K1+P1)2), n=α1α2μ3K1P1Z1(K1+P1)3, f=γ2Z1K2+Z1, g=γ2Z1K2+Z1(rP1K-α1P1Z1(K1+P1)2), h=α1α2γ2K1P1Z12(K1+P1)3(K2+Z1).

Here, DE1(λ,0)=0 has roots with negative real parts, provided system (Equation 8) is locally asymptotically stable about equilibrium E1 (for the condition, see Section 3.3.3).

For ω=0, DE1(0,τ)=n-h0.

Now, for ω0,

DE1(iω,τ)=-iω3-lω2+miω+n-e-iωτ(-fω2+giω+h).

Let DE1(iω,τ)=0 and separating the real and imaginary parts, we get -lω2+n=-fω2cosωτ+hcosωτ+gωcosωτ,

-ω3+mω=fω2sinωτ-hsinωτ+gωcosωτ.

Squaring and adding the above two equations, we get

ω6++A11ω4+A22ω2+A33=0, where A11=l2-f2-2m, A22=m2-2ln+2fh-g2, A33=n2-h2.

Sufficient conditions for the non-existence of a real number ω satisfying DE1(iω,τ)=0 can be written as ω6+A11ω4+A22ω2+A33>0,

which can be transformed to ω6+A11ω2+A222A112+A33-A2224A11>0.

Therefore, a sufficient condition for E1 to be stable is (i) A11>0 and (ii) A33>A2224A11.

4.2. Estimation of the length of delay to preserve stability at E1

In this section, we assume that in the absence of delays, E1 is locally asymptotically stable. Let u1(t),v1(t),andw1(t) be the respective linearized variables of this model. The system (Equation 8) becomes(11) du1dt=A1^u1(t)+A2^v1(t)dv1dt=B1^u1(t)+B2^w1(t)dw1dt=C1^w1(t)+C2^w1(t-τ),(11)

where A1^=α1P1Z1(K1+P1)2-rP1K, A2^=-α1P1K1+P1, B1^=α2K1Z1(K1+P1)2, B2^=-γ1Z1K2+Z1, C1^=-μ3, C2^=γ2Z1K2+Z1.

Let u^1(L), v^1(L), and w^1(L) be the Laplace transform of u1(t), v1(t), and w1(t), respectively. Taking the Laplace transformation of system (Equation 11), we have(L-A1^)u^1(L)=A2^v^1(L)+u1(0)Lv^1(L)=B1^u^1(L)+B2^w^1(L)+v1(0)(L-C1^)w^1(L)=C2^M1(L)e-Lτ+C2^e-Lτw^1(L),

where M1(L)=-τ0e-Ltw1(t)dt.

The inverse Laplace transform of u^1(L), v^1(L), and w^1(L) will have terms which exponentially increase with time, if u^1(L), v^1(L), and w^1(L) have pole with positive real parts. Since E1 needs to be locally asymptotically stable, it is necessary and sufficient that all poles of u^1(L), v^1(L), and w^1(L) have negative real parts. We shall employ the Nyquist criterion which states that if L is the arc length of a curve encircling the right half-plane, the curve of u^1(L), v^1(L), and w^1(L) will encircle the origin a number of times equal to the difference between the number of poles and the number of zeroes of u^1(L), v^1(L), and w^1(L) in the right half-plane.

Let G(L)=L3+lL2+mL+n-e-Lτ(fL2+gL+h) from (Equation 10). Then, E1 is locally asymptotically stable if(12) ReG(iv2)=0,(12) (13) ImG(iv2)>0,(13)

where v2 is the smallest positive root of Equation 12 (Freedman, Erbe, & Rao, Citation1986).

Now, (Equation 12) and (Equation 13) become-lv22+n=-fv22cosv2τ+hcosv2τ+gv2sinv2τ,-v23+mv2>fv22sinv2τ-hsinv2τ+gv2cosv2τ.

To get an estimation on the length of delay, we utilize the following conditions,(14) -lv2+n=-fv2cosvτ+hcosvτ+gvsinvτ,(14) (15) -v3+mv>fv2sinvτ-hsinvτ+gvcosvτ.(15)

Therefore, E1 will be stable if the above inequality holds at v=v2, where v2 is the first positive root of Equation 14. We shall now estimate an upper bound v++ of v2, which would be independent of τ. Then, we estimate τ so that (Equation 15) holds for all 0vv++, and hence in particular at v=v2.

Maximizing -fv2cosvτ+hcosvτ+gvsinvτ, subject to |sinvτ|1, |cosvτ|1, we obtain(16) -lv2+nfv2+h+gv.(16)

Thus, the unique positive solution of  (f+l)v2+gv-(n-h)=0 is denoted by v++. Hence, if v++=-g+g2+4(f+l)(n-h)2(f+l), then from (Equation 16) we have v2v++ which ensures that v++ is independent of τ. Then, we estimate τ so that (Equation 15) holds for all values 0vv++; now, rearranging (Equation 15) by |sinvτ|τv and |1-cosvτ|12τ2v2, we have(17) g2τ2v2+(h-fv2)τ+(m-g-v2)>0.(17)

Thus, (Equation 15) will be satisfied if Aθ1τ2+Bθ1τ+Cθ1>0, where Aθ1=g2v++2, Bθ1=h-fv++2, Cθ1=m-g-v++2.

Then, the Nyquist criterion holds for 0ττ1, where τ1=12Aθ1(-Bθ1+Bθ12+4Aθ1Cθ1) and τ1 gives an estimate for the length of delay for which stability is preserved.

Theorem 4

If there exist τ in 0ττ1 such that Aθτ2+Bθτ+Cθ>0, then τ1 is the maximum value (length of delay) of τ for which E1 is asymptotically stable.

4.3. Qualitative analysis of the model at E with gestation time delay (τ0)

The characteristic equation for the variational matrix V about E takes the following form:(18) DE(λ1,τ)=λ13+l¯λ12+m¯λ1+n¯-e-λ1τ(f¯λ12+g¯λ1+h¯)=0,wherel¯=rPK-α1PZ(K1+P)2-γ1ZF(K2+Z)2+μ3,m¯=α1PZ(K2+Z)2-rPKγ1ZF(K2+Z)2-μ3+α1α2K1PZ(K1+P)3-γ1μ3ZF(K2+Z)3,n¯=α1PZ(K1+P)2-rPKγ1μ3ZF(K2+Z)3+α1α2K1μ3PZ(K1+P)3,f¯=γ2ZK2+Z,g¯=(rPK-α1PZ(K1+P)2)γ2ZK2+Z-γ1γ2Z2F(K2+Z)3-γ1γ2K2ZF(K2+Z)3,h¯=α1PZ(K1+P)2-rPKγ1γ2Z2F(K2+Z)3+γ1γ2K2ZF(K2+Z)3-α1α2γ2K1PZ2(K1+P)3(K2+Z).(18)

Now, DE(iω¯,τ)=-iω¯3-l¯ω¯2+m¯iω¯+n¯-e-ω¯λτ(-f¯ω¯2+g¯iω¯+h¯)(Using Theorem3).

Let DE(iω¯,τ)=0 and separating the real and imaginary parts, we get-l¯ω¯2+n¯=-f¯ω¯2cosω¯τ+h¯cosω¯τ+g¯ω¯sinω¯τ,-ω¯3+m¯ω¯=f¯ω¯2sinω¯τ-h¯sinω¯τ+g¯ω¯cosω¯τ.

Squaring and adding the above two equations, we getω¯6+Q1¯ω¯4+Q2¯ω¯2+Q3¯=0,

whereQ1¯=l¯2-f¯2-2m¯,Q2¯=m¯2-2l¯n¯+2f¯h¯-g¯2,Q3¯=n¯2-h¯2.

The sufficient conditions for the non-existence of a real number ω satisfying DE=0 can be written as ω¯6+Q1¯ω¯4+Q2¯ω¯2+Q3¯>0, which can be transformed toω¯6+Q1¯ω¯2+Q2¯2Q1¯2+Q3¯-Q2¯24Q1¯>0.

Therefore, a sufficient condition for E to be stable is (a) Q1¯>0  and (b) Q3¯>Q2¯24Q1¯.

4.4. Estimation of the length of delay to preserve stability at E

In this section, we assume that in the absence of delays, E is locally asymptotically stable. Let u(t),v(t),andw(t) be the respective linearized variables of this model. The system (Equation 8) becomes(19) dudt=A1¯u(t)+A2¯v(t),dvdt=B1¯u(t)+B2¯v(t)+B3¯w(t),dwdt=C1¯v(t-τ)+C2¯w(t)+C3¯w(t-τ),(19)

where A1¯=α1PZ(K1+P)2-rPK, A2¯=-α1PK1+P, B1¯=α2K1Z(K1+P)2, B2¯=γ1ZFK2+Z, B3¯=-γ1ZK2+Z, C1¯=γ2K2F(K2+Z)2, C2¯=-μ3, C3¯=γ2ZK2+Z.

Let u¯(L), v¯(L), and w¯(L) be the Laplace transform of u(t), v(t), and w(t), respectively. Taking the Laplace transformation of system (Equation 19), we have(L-A1¯)u¯(L)=A2¯v¯(L)+u(0),(L-B2¯)v¯(L)=B1¯u¯(L)+B3¯w¯(L)+v(0),(L-C2¯)w¯(L)=C1¯e-LτK1+C1¯e-Lτv¯(L)+C3¯e-LτK2+C3¯e-Lτw¯(L)+w(0),

where K1(L)=-τ0e-Ltv(t)dt, K2(L)=-τ0e-Ltw(t)dt .

Let F¯(L)=L3+l¯L2+m¯L+n¯-e-Lτ(f¯L2+g¯L+h¯).

Then, E is locally asymptotically stable if ReF¯(iv¯0)=0 and ImF¯(iv¯0)>0 i.e.-l¯v¯02+n¯=-f¯v¯02cosv¯0τ+h¯cosv¯0τ+g¯v¯0sinv¯0τ,-v¯03+m¯v¯0>f¯v¯02sinv¯0τ-h¯sinv¯0τ+g¯v¯0cosv¯0τ,

where v¯0 be the smallest positive root of Re(iv¯0)=0.

To get an estimation on the length of delay, we use the following conditions,(20) -l¯v2+n¯=-f¯v2cosvτ+h¯cosvτ+g¯vsinvτ,(20) (21) -v3+m¯v>f¯v2sinvτ-h¯sinvτ+g¯vcosvτ.(21)

Therefore, E will be stable if above inequality holds at v=v¯0, where v¯0 is the first positive root of Equation 20. We shall now estimate an upper bound v+ of v0, which would be independent of τ. Then, we estimate τ so that (Equation 21) holds for all 0vv+, and hence in particular at v¯=v0.

Maximizing -f¯v2cosvτ+h¯cosvτ+g¯vsinvτ, subject to |sinvτ|1, |cosvτ|1, we obtain(22) -l¯v2+n¯f¯v2+h¯+g¯v.(22)

Thus, the unique positive solution of  (f¯+l¯)v2+g¯v-(n¯-h¯)=0 is denoted by v+. Hence, if v+=-g¯+g¯2+4(f¯+l¯)(n¯-h¯)2(f¯+l¯), then from above equation we have v0v+ which indicates that v+ is independent of τ. Then, we estimate τ so that (Equation 21) holds for all values 0vv+; now, rearranging (Equation 21) by |sinvτ|τv and |1-cosvτ|12τ2v2, we have(23) g¯2τ2v2+(h¯-f¯v2)τ+(m¯-g¯-v¯2)>0.(23)

Thus, Equation 21 will be satisfied ifAθτ2+Bθτ+Cθ>0,whereAθ=g¯2v+2,Bθ=h¯-f¯v+2,Cθ=m¯-g¯-v+2.

Then, the Nyquist criterion holds for 0ττ+, where τ+=12Aθ(-Bθ+Bθ2+4AθCθ) and τ+ gives an estimate for the length of delay for which stability is preserved. Thus, we are now in a position to state the following theorem.

Theorem 5

If there exist τ in 0ττ+ such that Aθτ2+Bθτ+Cθ>0, then τ+ is the maximum value (length of delay) of τ for which E is asymptotically stable.

5. Bifurcation analysis in the presence of discrete delay

Let us consider τ0  and assume λ1=μ+iν in Equation 18. Then, separating the real and imaginary parts, we get a system of transcendental equations as(24) μ3-3μν2-l¯(μ2-ν2)+m¯μ+n¯=f¯(μ2-ν2)cosντe-μτ+2fνμ¯sinντe-μτ+g¯μcosντe-μτ+h¯cosντe-μτ+g¯νsinντe-μτ,(24) (25) -ν3+3μ2ν+2l¯μν+m¯ν=-f¯(μ2-ν2)sinντe-μτ+2f¯νμcosντe-μτ-g¯μsinντ-h¯sinντe-μτ+g¯νcosντe-μτ.(25)

Let us consider λ1, and hence μ and ν are functions of τ. We want to know the change of stability of E which will occur at the values of τ=τ^ for μ = 0 and ν0.

Then, the above two equations become(26) -l¯ν2+n¯=-f¯ν2cosντ^+h¯cosντ^+g¯νsinντ^-ν3+m¯ν=f¯ν2sinντ^-h¯sinντ^+g¯νcosντ^,(26)

Now, eliminating τ^, we have(27) ν6+(l¯2-f¯2-2m¯)ν4+(m¯2-2l¯n¯+2f¯h¯-g¯2)ν2+n¯2-h¯2=0.(27)

From (Equation 26), we get τ^n=1ν^tan-1{ν^(f¯ν4^-f¯m¯ν2^-h¯ν2^-l¯g¯ν2^-h¯m¯-n¯g¯)g¯ν4^-g¯m¯ν2^+f¯n¯ν2^-h¯l¯ν2^+f¯l¯ν4^-h¯n¯}+nπν^. The smallest τn^ is given by n=0 and we take it as τ^n=1ν^tan-1{ν^(f¯ν4^-f¯m¯ν2^-h¯ν2^-l¯g¯ν2^-h¯m¯-n¯g¯)g¯ν4^-g¯m¯ν2^+f¯n¯ν2^-h¯l¯ν2^+f¯l¯ν4^-h¯n¯}. In order to establish the Hopf bifurcation at τ=τ^, we need to show that dμdτ0 at τ=τ^. We differentiate Equations 24 and 25 with respect to τ and setting τ=τ^, μ=0 and ν=ν^, we get(28) L1dμdτ(τ^)+M1dνdτ(τ^)=X1-M1dμdτ(τ^)+L1dνdτ(τ^)=Y1,(28)

whereL1=-3ν^2+m¯-g¯cosν^τ^-2f¯ν^sinν^τ^+τ^(h¯cosν^τ^+g¯ν^sinν^τ^-f¯ν^2cosν^τ^),M1=-2l¯ν^+2f¯ν^cosν^τ^-g¯sinν^τ^+τ^(h¯sinν^τ^-g¯ν^cosν^τ^-f¯ν^2sinν^τ^),X1=f¯ν^3sinν^τ^-h¯ν^sinν^τ^+g¯ν^2cosν^τ^,Y1=f¯ν^3cosν^τ^-h¯ν^cosν^τ^-g¯ν^2sinν^τ^.

Solving (Equation 28),

we get dμdτ(τ^)=L1X1-M1Y1L12+M12,

where dμdτ(τ^) has the same sign as that of L1X1-M1Y1.

Substituting the values of L1,M1,X1,andY1 and using (Equation 26), we getL1X1-M1Y1=ν^2[3ν^4+2(l¯2-f¯2-2m¯)ν^2+(m¯2-2l¯n¯+2f¯h¯-g¯2)].

Let H(z)=z3+B1z2+B2Zz+B3, where B1=l¯2-f¯2-2m¯,  B2=m¯2-2l¯n¯+2f¯h¯-g¯2, B3=(n¯2-h¯2) which is left side of (Equation 27) with ν^2=z. Then, H(ν^2)=0 and we note that(29) dμdτ(τ^)=ν^2L12+M12dHdz(ν^2).(29)

Hence, we can describe the criteria for the preservation of stability (instability) geometrically as follows:

(1)

If the polynomial H(z) has no positive roots, there is no change of stability.

(2)

If H(z)is decreasing (increasing) at all its positives roots, stability (instability) is preserved.

We note the following facts:
(i)

For the existence of ν^, H(z) must have at least one positive real root.

(ii)

Since H(z) is cubic in z , limzH(z)=.

(iii)

If H(z) has a unique positive root, then it must increase at that point to satisfy (ii).

(iv)

If H(z) has two or three distinct positive real roots, then it must decrease at one root and increase at other; hence, (ii) is not satisfied.

(v)

If B3<0, then H(z) has only one root.

(vi)

If B1>0, B3>0, and B2<0, then (i) will be satisfied.

Now, if B1>0, B3>0, and B2<0, then the minimum of H(z) will exist at zmin=-B1+B12-3B23

and (i) will be satisfied if H(zmin)>0, i.e.(30) 2B13-9B1B2+27B3>2(B12-3B2)3/2,(30)

or 2B1(B12-3B2)+27B3-3B1B2>2(B12-3B2)3/2.

Since 27B3-3B1B2>27B3 (since B1>0, B3>0, and B2<0), and B12-3B2>B12, hence 2B1(B12-3B2)+27B3-3B1B2>27B3+2B13.

Thus, for Equation 30 to hold, it is sufficient that 27B3+2B13>2(B12-3B2)3/2B2>13B12-27B3+2B1322/3.

Table 2. A set of parametric values

Now, we can state the following theorem.

Theorem 6

If B1>0, B3>0, and relation (Equation 30) hold, then the stable positive equilibrium E remains stable for all τ>0.

6. Numerical simulations

In this section, we focus our attention on the occurrence and termination of the fluctuating plankton population. We begin with a parameter set (see Table , Chatterjee & Pal, Citation2011) for which the existence condition of the coexistence equilibrium point E is satisfied and the coexistence equilibrium point E=(0.40,0.10,0.54) is locally asymptotically stable in the form of a stable focus with eigenvalues -0.8003,-0.0129±i0.2064 (cf. Figure ).

6.1. Effect of r

We have observed that the system shows oscillatory behavior around the positive interior equilibrium E for increasing the value of r=1.6 with eigenvalues -1.2305,0.0013±i0.0.2057 (cf. Figure ). Decreasing the value of r from 1.2 to 0.3 with the same set of parametric values in Table , the system shifts to planktivorous fish free equilibrium E1=(0.0250,0.0766,0) in the form of a stable focus with eigenvalues -0.0173,-0.0026±i0.1188 (cf. Figure ). Figure (a–c) depicts the different steady-state behaviors of phytoplankton, zooplankton, and planktivorous fish in the system (Equation 1) for the parameter r. Here, we see a Hopf bifurcation point at rc0=1.583292 (denoted by a red star (H)) with first Lyapunov coefficient being –0.2999282. Clearly, the first Lyapunov coefficient is negative. This means that a stable limit cycle bifurcates from the equilibrium, when it loses stability.

Figure 1. The equilibrium point E is stable for the parametric values as given in Table .

Figure 1. The equilibrium point E∗ is stable for the parametric values as given in Table 2.

Figure 2. The figure depicts oscillatory behavior around the positive interior equilibrium point E of system (Equation 1) for increasing r, from 1.2 to 1.6 with other parametric values as given in Table .

Figure 2. The figure depicts oscillatory behavior around the positive interior equilibrium point E∗ of system (Equation 1) for increasing r, from 1.2 to 1.6 with other parametric values as given in Table 2.

6.2. Effects of K

For K = 0.8, leaving all other parameters unaltered, the system exhibits oscillation around the positive interior equilibrium E with eigenvalues -0.8599,0.0069±i0.2370 (cf. Figure ). Figure (a–c) depicts the different steady-state behaviors of phytoplankton, zooplankton, and planktivorous fish in the system (Equation 1) for the parameter K. Here, we see a Hopf bifurcation point at K=0.711977 (denoted by a red star (H)) with first Lyapunov coefficient being –0.3651997. Clearly, the first Lyapunov coefficient is negative. This means that a stable limit cycle bifurcates from the equilibrium, when it loses stability.

Figure 3. The figure depicts stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for decreasing r, from 1.2 to 0.3 with other parametric values as given in Table .

Figure 3. The figure depicts stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for decreasing r, from 1.2 to 0.3 with other parametric values as given in Table 2.

Figure 4. (a) The figure depicts different steady-state behaviors of phytoplankton for the effect of r. (b) The figure depicts different steady-state behaviors of zooplankton for the effect of r with other parametric values as given in Table . (c) The figure depicts different steady-state behaviors of planktivorous fish for the effect of r with other parametric values as given in Table .

Figure 4. (a) The figure depicts different steady-state behaviors of phytoplankton for the effect of r. (b) The figure depicts different steady-state behaviors of zooplankton for the effect of r with other parametric values as given in Table 2. (c) The figure depicts different steady-state behaviors of planktivorous fish for the effect of r with other parametric values as given in Table 2.

Figure 5. The figure depicts oscillatory behavior around the positive interior equilibrium point E of system (Equation 1) for increasing K, from 0.5 to 0.8 with other parametric values as given in Table .

Figure 5. The figure depicts oscillatory behavior around the positive interior equilibrium point E∗ of system (Equation 1) for increasing K, from 0.5 to 0.8 with other parametric values as given in Table 2.

Figure 6. (a) The figure depicts different steady-state behaviors of phytoplankton for the effect of K. (b) The figure depicts different steady-state behaviors of zooplankton for the effect of K with other parametric values as given in Table . (c) The figure depicts different steady-state behaviors of planktivorous fish for the effect of K with other parametric values as given in Table .

Figure 6. (a) The figure depicts different steady-state behaviors of phytoplankton for the effect of K. (b) The figure depicts different steady-state behaviors of zooplankton for the effect of K with other parametric values as given in Table 2. (c) The figure depicts different steady-state behaviors of planktivorous fish for the effect of K with other parametric values as given in Table 2.

Figure 7. The figure depicts oscillatory behavior around the positive interior equilibrium point E of system (Equation 1) for decreasing μ3, from 0.08 to 0.02 with same set of parametric values as use in Table .

Figure 7. The figure depicts oscillatory behavior around the positive interior equilibrium point E∗ of system (Equation 1) for decreasing μ3, from 0.08 to 0.02 with same set of parametric values as use in Table 2.

Figure 8. The figure depicts stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for increasing μ3, from 0.08 to 0.3 with same set of parametric values as use in Table .

Figure 8. The figure depicts stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for increasing μ3, from 0.08 to 0.3 with same set of parametric values as use in Table 2.

Figure 9. The figure depicts asymptotically stable behavior at positive interior equilibrium point E of system (Equation 1) for r = 1.6 and μ3=0.12 with other parametric values as given in Table ( Blue solid line). The trajectory (Black dotted line) shows stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for r = 1.6 and μ3=.3 with other parametric values as given in Table .

Figure 9. The figure depicts asymptotically stable behavior at positive interior equilibrium point E∗ of system (Equation 1) for r = 1.6 and μ3=0.12 with other parametric values as given in Table 2 ( Blue solid line). The trajectory (Black dotted line) shows stable behavior around the planktivorous fish free equilibrium point E1 of system (Equation 1) for r = 1.6 and μ3=.3 with other parametric values as given in Table 2.

Figure 10. The figure depicts stable behavior around the positive interior equilibrium point E of system (Equation 1) for K=0.8 and μ3=0.12 with other parametric values as given in Table .

Figure 10. The figure depicts stable behavior around the positive interior equilibrium point E∗ of system (Equation 1) for K=0.8 and μ3=0.12 with other parametric values as given in Table 2.

Figure 11. The figures depict oscillatory behavior of phytoplankton, Zooplankton, and planktivorous fish population for τ=1 with same set of other parametric values as given in Table .

Figure 11. The figures depict oscillatory behavior of phytoplankton, Zooplankton, and planktivorous fish population for τ=1 with same set of other parametric values as given in Table 2.

Figure 12. The coexistence equilibrium is locally asymptotically stable for τ=1 and r=1.

Figure 12. The coexistence equilibrium is locally asymptotically stable for τ=1 and r=1.

Figure 13. The coexistence equilibrium is locally asymptotically stable for τ=1 and K=0.35.

Figure 13. The coexistence equilibrium is locally asymptotically stable for τ=1 and K=0.35.

Figure 14. The coexistence equilibrium is locally asymptotically stable for τ=1 and μ3=0.11.

Figure 14. The coexistence equilibrium is locally asymptotically stable for τ=1 and μ3=0.11.

Figure 15. The bifurcation diagram of all the populations with K as the bifurcation parameter.

Figure 15. The bifurcation diagram of all the populations with K as the bifurcation parameter.

Figure 16. The bifurcation diagram of all the populations with r as the bifurcation parameter.

Figure 16. The bifurcation diagram of all the populations with r as the bifurcation parameter.

Figure 17. The Kr two parameters bifurcation diagram.

Figure 17. The K–r two parameters bifurcation diagram.

Figure 18. The K-μ3 two parameters bifurcation diagram.

Figure 18. The K-μ3 two parameters bifurcation diagram.

Figure 19. The r-μ3 two parameters bifurcation diagram.

Figure 19. The r-μ3 two parameters bifurcation diagram.

Figure 20. The bifurcation diagram for τ with all parametric values as given in Table .

Figure 20. The bifurcation diagram for τ with all parametric values as given in Table 2.

6.3. Effects of μ3

Keeping the other parameters fixed and decreasing the value of μ3 from 0.08 to 0.02, the system exhibits oscillatory behaviors around the positive interior equilibrium E with eigenvalues -1.0833,0.0002±i0.1084 (cf. Figure ). But the system shifts to planktivorous fish free equilibrium E1=(0.0250,0.3437,0) with eigenvalues -0.0289,-0.0080±i0.2516, when μ3 increase from 0.08 to 0.3 with the same set of parametric values as in Table (cf. Figure ).

6.4. Combined effect of r and μ3

It has already been shown that the system depicts oscillatory behavior around the positive interior equilibrium E for r=1.6. If the mortality rate of planktivorous fish μ3 in increased from 0.08 to 0.12, the system becomes asymptotically stable at positive interior equilibrium E=(0.3868,0.1588,0.5591) in the form of a stable focus with eigenvalues -1.0114,-0.0107±i0.2431 (cf. Figure ). But if the mortality rate of planktivorous fish μ3 is increased from 0.12 to 0.3, the system shifts to planktivorous fish free equilibrium E1=(0.025,0.4625,0) with eigenvalues -0.0284,-0.0104±i0.2918 (cf. Figure ).

6.5. Combined effect of K and μ3

It has already been shown that the system depicts oscillatory behavior around the positive interior equilibrium E for K=0.8 (cf. Figure ). But the system shifts to asymptotically stable behavior at E=(0.5960,0.1590,0.7282) in the form of a stable focus with eigenvalues -0.6461,-0.0065±i0.2876 for increasing the value of μ3 from 0.08 to 0.12 (cf. Figure ).

6.6. Effects of τ

Now, we are to introduce gestation delay, τ, in system (Equation 1). Keeping the other parameters fixed and increasing the value of the delay parameter τ from 0 to 1, we observe that the solution of system (Equation 8) changes from stable behavior to oscillatory behavior around the positive interior equilibrium E (cf. Figure ) for same set of parametric values as in Figure . We see that decreasing the value of r from 1.2 to 1, the system shifts to stable behavior around the positive interior equilibrium (cf. Figure ). Similar cases happen when decreasing the carrying capacity and increasing mortality rate of planktivorous fish from 0.5 to 0.35 and 0.08 to 0.11, respectively, (cf. Figure and cf. Figure ) for same set of parametric values as in Figure .

6.7. Hopf bifurcation

For a clear understanding of the dynamical changes due to change in constant nutrient input, K, a bifurcation diagram is plotted with this parameter as the bifurcation parameter with other three species (cf. Figure ). We have also plotted another single parameter bifurcation diagram for r (cf. Figure ).

Now, we have plotted a two parameters bifurcation diagrams, Kr, K-μ3 and r-μ3 (cf. Figures ). Here, we see a generalized Hopf (GH) point, where the first Lyapunov coefficient (l1) vanishes but the second lyapounov coefficient is non-zero. The bifurcation point separates branches of sub and supercritical Andronov–Hopf bifurcations in the parameter plain. The Bogdanov–Takens (BT) point is the common point for the limit point curve and the curve corresponding to equilibria. Actually, at BT point, the system has an equilibrium with a double zero eigenvalues. For nearby parameter values, the system has two equilibria (a saddle and a non-saddle) which collide and disappear via a saddle-node bifurcation. The non-saddle equilibrium undergoes an Andronov–Hopf bifurcation, generating a limit cycle. Finally, a bifurcation diagram is plotted with τ as the bifurcation parameter with other three species (cf. Figure ).

7. Discussion

A phytoplankton–zooplankton–fish interaction model is considered with nutrient uptake functions. Firstly, the model is studied analytically and the threshold conditions for the existence and stability of various steady states are worked out (see Table ). It is observed that for constant intrinsic growth rate r, there is a chance for the population to fluctuate. But further increase in the value of constant intrinsic growth rate r may lead to extinction of planktivorous fish population.

It is observed that the system shifts to oscillatory behavior in presence of high value of carrying capacity. We have also observed that there is a chance of extension of planktivorous fish population for high value of mortality rate of planktivorous fish. Our study indicates that low value of mortality rate of planktivorous fish may lead to oscillatory behavior around the positive interior equilibrium.

Next, we have established that the system remains locally asymptotically stable and all solutions approach toward E in presence of discrete delay whenever its magnitude lies below some threshold condition. Further, we have also observed the maximum value (length of delay) of τ (i.e. τ+) for which a locally asymptotically stable interior equilibrium E will remain asymptotically stable, where τ+=12Aθ(-Bθ+Bθ2+4AθCθ) (see Theorem 5). Moreover, we analyzed conditions for bifurcation of the positive interior equilibrium. We observed that a positive interior equilibrium remains stable if B1>0 and B3>0 for all τ>0 (see Theorem 6). Further, numerical simulations demonstrate the following conclusions:

(a)

The system exhibits dynamic instability due to higher gestation delay.

(b)

Low levels of constant intrinsic growth rate r induce stability around the positive interior equilibrium in presence of gestation delay.

(c)

Low levels of carrying capacity of phytoplankton K help the system (Equation 8) to approach stability at E in presence of gestation delay.

(d)

High value of mortality rate of planktivorous fish prevents instability behavior in presence of gestation delay.

Throughout the article of this model (analytically and numerically), an attempt has been made to search for a suitable way to control the system and maintain a stable coexistence between all the species. It has been observed that to control the fracturing population and to maintain stability around the coexistence equilibrium, we have to balance the rate of carrying capacity of phytoplankton, mortality rate of planktivorous fish population, and constant intrinsic growth rate.

Acknowledgements

The authors would like to thank the referees very much for their valuable comments and suggestions.

Additional information

Funding

The research is partially supported by University Grants Commission, New Delhi [grant number MRP-MAJ-MATH-2013-609].

Notes on contributors

Samares Pal

Our research group is headed by Samares Pal and includes Anal Chatterjee as a post-doctoral researcher. Research in the group addresses a wide range of questions broadly concerning the dynamics of plankton ecosystems under environmental conditions. The main emphasis of our work includes mathematical modeling of marine ecosystems affected by fish induced, overfishing of zooplankton. The research reported in this paper will help in studying the interspecies between phytoplankton and zooplankton in presence of fish and the subsequent changes the stability behavior.

Notes

Corresponding author: Samares Pal, Department of Mathematics, University of Kalyani, Kalyani 741235, India. E-mail: [email protected].

References

  • Beretta, E., Bischi, G. I., & Solimano, F. (1991). Stability in chemostat equations with delayed nutrient recycling. Journal of Mathematical Biology, 28, 99–111. doi:10.1007/BF00171521
  • Biktashev, V. N., Brindley, J., & Horwood, J. W. (2003). Phytoplankton blooms and fish recruitment rate. Journal of plankton Research, 25, 21–33. doi:10.1093/plankt/25.1.21
  • Bischi, G. I. (1992). Effects of time lags on transient characteristics of a nutrient cycling model. Mathematical Biosciences, 109, 151–175. doi:10.1016/0025-5564(92)90043-V
  • Chatterjee, A., & Pal, S. (2011). Effect of dilution rate on the predictability of a realistic ecosystem model with instantaneous nutrient recycling. Journal of Biological Systems, 19, 629–650. doi:10.1142/S021833901100410X
  • Cushing, J. M. (1977). lntegrodifferential equations and delay models in population dynamics (Lecture notes in biomathematics, Vol. 20). Berlin: Springer-Verlag. doi:10.1007/978-3-642-93073-7
  • Das, K., & Ray, S. (2008). Effect of delay on nutrient cycling in phytoplankton--zooplankton interactions in estuarine system. Ecological Modelling, 215, 69–76. doi:10.1016/j.ecolmodel.2008.02.019
  • Freedman, H. I., Erbe, L. H., & Rao, V. S. H. (1986). Three species food chain models with mutual interference and time delays. Mathematical Biosciences, 80, 57–80. doi:10.1016/0025-5564(86)90067-2
  • Freedman, H. I., & Rao, V. S. H. (1983). The trade-off between mutual interference and time lags in predator-prey systems. Bulletin of Mathematical Biology, 45, 991–1004. doi:10.1007/BF02458826
  • Freedman, H. I., So, J., & Waltman, P. (1989). Coexistence in a model of competition in the chemostat incorporating discrete delay. SIAM Journal on Mathematical Analysis, 49, 859–870. doi:10.1137/0149050
  • Gazi, N. H., & Bandyopadyay, M. (2006). Effect of time delay on a detritus-based ecosystem. International Journal of Mathematics and Mathematical Sciences, Article ID: 25619, 1–28. doi:10.1155/IJMMS/2006/25619
  • Gazi, N. H., & Bandyopadhyay, M. (2008). Effect of time delay on a harvested predator-prey model. Journal of Applied Mathematics and Computing, 26, 263–280. doi:10.1007/s12190-007-0015-2
  • Gopalsamy, K. (1984). Delayed responses and stability in two-species systems. The Journal of the Australian Mathematical Society Series B Applied Mathematics, 25, 473–500. doi:10.1017/S0334270000004227
  • Gopalsamy, K. (1992). Stability and oscillations in delay differential equations of population dynamics (Mathematics and its applications, Vol. 74). Dordrecht: Kluwer. doi:10.1007/978-94-015-7920-9
  • Khare, S., Misra, O. P., Singh, C., & Dhar, J. (2011). Role of delay on planktonic ecosystem in the presence of a toxic producing phytoplankton. International Journal of Differential Equations, Article ID: 603183, doi:10.1155/2011/603183
  • Kuang, Y. (1993). Delay differential equations with applications in popular dynamics. New York, NY: Academic Press.
  • Liu, S., Beretta, E., & Breda, D. (2010). Predator-prey model of Beddington--DeAngelis type with maturation and gestation delays. Nonlinear Analysis: Real World Applications, 11, 4072–4091. doi:10.1016/j.nonrwa.2010.03.013
  • Maity, A., Patra, B., & Samanta, G. P. (2007). Persistence and stability in a ratio-dependent predator-prey system with delay and harvesting. Natural Resource Modeling, 20, 575–600. doi:10.1111/j.1939-7445.2007.tb00221.x
  • Medvinsky, A. B., Rusakov, A. V., Bobyrev, A. E., Burmensky, V. A., Kriksunov, A. E., Nurieva, N. I., & Gonik, M. M. (2009). A conceptual mathematical model of aquatic communities in lakes Naroch and Myastro (Belarus). Biophysics, 54, 90–93. doi:10.1134/S0006350909010151
  • Mukhopadhyay, B., & Bhattacharyya, R. (2008). Role of gestation delay in a plankton--fish model under stochastic fluctuations. Mathematical Biosciences, 215, 26–34. doi:10.1016/j.mbs.2008.05.007
  • Mukhopadhyay, B., & Bhattacharyya, R. (2011). A stage-structured food chain model with stage dependent predation: Existence of codimension one and codimension two bifurcations. Nonlinear Analysis: Real World Applications, 12, 3056–3072. doi:10.1016/j.nonrwa.2011.05.007
  • Pal, S., & Chatterjee, A. (2012). Role of constant nutrient input and mortality rate of planktivorous fish in plankton community ecosystem with instantaneous nutrient recycling. Canadian Applied Mathematics Quarterly, 20, 179–207.
  • Rehim, M., & Imran, M. (2012). Dynamical analysis of a delay model of phytoplankton--zooplankton interaction. Applied Mathematical Modelling, 36, 638–647. doi:10.1016/j.apm.2011.07.018
  • Ruan, S. (1995). The effect of delays on stability and persistence in plankton models. Nonlinear Analysis, Theory, Methods and Applications, 24, 575–585. doi:10.1016/0362-546X(95)93092-I
  • Scheffer, M. (1991). Fish and nutrients interplay determines algel biomass: A minimal model. Oikos, 62, 271–282. Retrieved from http://www.jstor.org/stable/3545491
  • Scheffer, M., Rinaldi, S., & Kuznetsov, A. Y. (2000). Effects of fish on plankton dynamics: A theoretical analysis. Canadian Journal of Fisheries and Aquatic Sciences, 57, 1208–1219.
  • Singh, A., & Gakkhar, S. (2012). Analysis of delayed toxin producing phytoplankton--zooplankton system. International Journal of Modeling and Optimization, 2, 677–680. doi:10.7763/IJMO.2012.V2.208
  • Upadhyay, R. K., Kumari, N., & Rai, V. (2009). Wave of chaos in a diffusive system: Generating realistic patterns of patchiness in planktonfish dynamics. Chaos, Solitons and Fractals, 40, 262–276. doi:10.1016/j.chaos.2007.07.078
  • Yongzhen, P., Min, G., & Changguo, L. (2011). A delay digestion process with application in a three-species ecosystem. Communications in Nonlinear Science and Numerical Simulation, 16, 4365–4378. doi:10.1016/j.cnsns.2011.03.018