
This paper examines a diffusive toxic-producing plankton system with delay. We first show the global attractivity of the positive equilibrium of the system without time-delay. We further consider the effect of delay on asymptotic behavior of the positive equilibrium: when the system undergoes Hopf bifurcation at some points of delay by the normal form and center manifold theory for partial functional differential equations. Global existence of periodic solutions is established by applying the global Hopf bifurcation theory.
Citation: Hong Yang. Global dynamics of a diffusive phytoplankton-zooplankton model with toxic substances effect and delay[J]. Mathematical Biosciences and Engineering, 2022, 19(7): 6712-6730. doi: 10.3934/mbe.2022316
[1] | Xinyou Meng, Jie Li . Stability and Hopf bifurcation analysis of a delayed phytoplankton-zooplankton model with Allee effect and linear harvesting. Mathematical Biosciences and Engineering, 2020, 17(3): 1973-2002. doi: 10.3934/mbe.2020105 |
[2] | Zhichao Jiang, Xiaohua Bi, Tongqian Zhang, B.G. Sampath Aruna Pradeep . Global Hopf bifurcation of a delayed phytoplankton-zooplankton system considering toxin producing effect and delay dependent coefficient. Mathematical Biosciences and Engineering, 2019, 16(5): 3807-3829. doi: 10.3934/mbe.2019188 |
[3] | Ruiqing Shi, Jianing Ren, Cuihong Wang . Stability analysis and Hopf bifurcation of a fractional order mathematical model with time delay for nutrient-phytoplankton-zooplankton. Mathematical Biosciences and Engineering, 2020, 17(4): 3836-3868. doi: 10.3934/mbe.2020214 |
[4] | Elvira Barbera, Giancarlo Consolo, Giovanna Valenti . A two or three compartments hyperbolic reaction-diffusion model for the aquatic food chain. Mathematical Biosciences and Engineering, 2015, 12(3): 451-472. doi: 10.3934/mbe.2015.12.451 |
[5] | Zhenyao Sun, Da Song, Meng Fan . Dynamics of a stoichiometric phytoplankton-zooplankton model with season-driven light intensity. Mathematical Biosciences and Engineering, 2024, 21(8): 6870-6897. doi: 10.3934/mbe.2024301 |
[6] | Juan Li, Yongzhong Song, Hui Wan, Huaiping Zhu . Dynamical analysis of a toxin-producing phytoplankton-zooplankton model with refuge. Mathematical Biosciences and Engineering, 2017, 14(2): 529-557. doi: 10.3934/mbe.2017032 |
[7] | Yuanpei Xia, Weisong Zhou, Zhichun Yang . Global analysis and optimal harvesting for a hybrid stochastic phytoplankton-zooplankton-fish model with distributed delays. Mathematical Biosciences and Engineering, 2020, 17(5): 6149-6180. doi: 10.3934/mbe.2020326 |
[8] | Saswati Biswas, Pankaj Kumar Tiwari, Yun Kang, Samares Pal . Effects of zooplankton selectivity on phytoplankton in an ecosystem affected by free-viruses and environmental toxins. Mathematical Biosciences and Engineering, 2020, 17(2): 1272-1317. doi: 10.3934/mbe.2020065 |
[9] | He Liu, Chuanjun Dai, Hengguo Yu, Qing Guo, Jianbing Li, Aimin Hao, Jun Kikuchi, Min Zhao . Dynamics induced by environmental stochasticity in a phytoplankton-zooplankton system with toxic phytoplankton. Mathematical Biosciences and Engineering, 2021, 18(4): 4101-4126. doi: 10.3934/mbe.2021206 |
[10] | Xin Zhao, Lijun Wang, Pankaj Kumar Tiwari, He Liu, Yi Wang, Jianbing Li, Min Zhao, Chuanjun Dai, Qing Guo . Investigation of a nutrient-plankton model with stochastic fluctuation and impulsive control. Mathematical Biosciences and Engineering, 2023, 20(8): 15496-15523. doi: 10.3934/mbe.2023692 |
This paper examines a diffusive toxic-producing plankton system with delay. We first show the global attractivity of the positive equilibrium of the system without time-delay. We further consider the effect of delay on asymptotic behavior of the positive equilibrium: when the system undergoes Hopf bifurcation at some points of delay by the normal form and center manifold theory for partial functional differential equations. Global existence of periodic solutions is established by applying the global Hopf bifurcation theory.
In aquatic systems such as rivers, lakes and oceans, harmful algal blooms (HABs) have attracted considerable scientific attention in recent years. The effect of toxin-producing phytoplankton (TPP) on zooplankton is one reason for such attention. Researchers have made great progress in mathematical modeling in this field [1,2,3,4,5,6]. To fully understand the mechanism of planktonic blooms and to formulate reasonable control measures, Chattopadhyay et al.[3] build a ODE phytoplankton-zooplankton model with toxin effect, using data gathered from the coastal region of West Bengal and part of Orissa, India. These researchers investigated plankton models with different predational response functions and toxin liberation processes to obtain rich dynamics. They also considered the diffusivity of plankton influenced by ocean currents and tides, and the time delay effect of the phytoplankton toxin on zooplankton. Wang et al. [7] consider a ODE plankton system with Holling II response function and linear toxin processes. These authors concluded that the system undergoes Hopf bifurcation at positive equilibrium and Hopf-transcritical bifurcation when the parameters satisfy a particular condition. In spatial plankton models, for example, Chaudhuri et al. [4] consider that the system includes both non-toxic and toxic phytoplankton and the system shows toxic phytoplankton-induced spatiotemporal patterns when one phytoplankton releases toxin. Further, to describe the reduction of zooplankton due to toxin-producing phytoplankton, plankton systems with discrete delay are presented in References [2,8] to show the effect of delay. There is o large body of literatures describing the dynamics of aquatic models [6,9,10,11].
In fact, the problem we analyze is based on Chen et al.[1], which considers a model as follows:
{du(t)dt=(1−u(t))u(t)−αv(t)u(t)1+cu(t),t>0,dv(t)dt=βu(t)v(t)1+cu(t)−θu(t)v(t)−rv(t),t>0. | (1.1) |
They conclude that the plankton model (1.1) occurs as a bistable phenomenon.
In real-world conditions, plankton affected by tides and turbulence may move and diffuse across lakes and seas. Thus, such diffusion should be considered in studying the dynamics of plankton models. Meanwhile, the effect of phytoplankton toxin on zooplankton is in the form of a time-delay. We assume that no plankton species enter or leave at the boundaries of their environments. Based on these considerations, we present the plankton model as follows:
{∂u(x,t)∂t=Δu(x,t)+(1−u(x,t))u(x,t)−v(x,t)u(x,t)1+cu(x,t),∂v(x,t)∂t=Δv(x,t)+βv(x,t)u(x,t)1+cu(x,t)−θu(x,t−τ)v(x,t)−rv(x,t), | (1.2) |
for t>0,x∈Ω=[0,lπ]. u represents the TPP population, and v represents the zooplankton population. Here, parameters β,r,θ and c are positive. β denotes the conversion ratio, r denotes the mortality rate of zooplankton due to natural death and higher predation, θ denotes the rate of toxin liberation by TPP population, and c denotes the half-saturation constant, here, we consider the case c>1. d1 and d2 represent the diffusion coefficients of phytoplankton and zooplankton, respectively.
Here, the initial conditions of system (1.2) are considered as
u0(x)=φ1(x,ϑ)≥0,v0(x)=φ2(x,ϑ)≥0,x∈Ω,ϑ∈[−τ,0] | (1.3) |
where φ=(φ1,φ2) is uniformly continuous, and the homogeneous Neumann boundary conditions are imposed as
∂u∂ν=∂v∂ν=0,t>0,x∈∂Ω, | (1.4) |
where ∂∂ν denotes the outward normal derivative on ∂Ω.
This paper proceeds as follows. Section 2 gives the uniform boundedness of solutions. Section 3, by constructing the upper and lower solutions, establishes the global attractivity of positive equilibrium of the system without time-delay. Further, taking time delay as a branch parameter, we give the sufficient condition for which the system undergoes Hopf bifurcation at the interior equilibrium. In Section 4, under the assumption condition, we analyze the existence of Hopf bifurcation. Section 5 considers the global existence of these bifurcating periodic solutions. Finally, in Section 6 presents several numerical simulations.
This section shows the boundedness of the solutions of system (1.2).
Lemma 2.1. Under the initial boundary conditions (1.3)–(1.4), all nontrivial solutions of system (1.2) are positive and uniformly bounded.
Proof. For v(x,t) note that, on 0≤t≤τ, v(x,t−τ)=φ2(x,t−τ) and u(x,t) is bounded on x∈ˉΩ. For further proof, we consider the following auxiliary system
{mt=Δm+βum1+cu−θM1m−rm,x∈Ω,t>0,mν=0,x∈∂Ω,t>0,m(x,0)=φ2(x,0),x∈Ω. |
where M1=maxx∈ˉΩ,t∈[−τ,0]u(x,t). By comparison principle, we know v(x,t)≥m(x,t)≥0 for x∈Ω,t>0. To proceed in this way, for x∈Ω,t∈[τ,2τ], we have v(x,t)≥m(x,t)≥0. Then, by mathematical induction, we obtain the positivity of v(x,t) in x∈Ω,t>0. Similarly, we can prove the positivity of u(x,t).
Let
u(x1,t1)=maxx∈ˉΩ,t>0u(x,t),βu(x2,t2)+v(x2,t2)=maxx∈ˉΩ,t>0u(x,t). |
Then, in view of the Hopf boundary lemma and the boundary condition, it follows that
u(x1,t1)(1−u(x1,t1))−u(x1,t1)v(x1,t1)1+cu(x1,t1)≥0forx1∈Ω. |
This implies
u(x1,t1)<1. |
By adding the two equations in system (1.2) with the form βu+v, it follows that
(∂t−Δ)(βu+v)=βu(1−u)−rv−θutv=βu(1−u)+rβu−r(βu+v)−θutv≤βu(1−u)+rβu−r(βu+v). |
By the maximum principle in Reference [12], this implies that
βu+v<β+4rβ4r. |
This completes the proof.
Clearly, system (1.2) has two boundary equilibria, E0=(0,0) and E1=(1,0).
Let us now analyze the stability of plankton system (1.2) at the interior equilibrium. By referring to [1,Lemma 2.3], we know system (1.2) has a unique positive equilibrium E∗(u−,v−) when the parameters satisfy the following condition
(H1)c>1,√rcθ>1,andβ>(c+1)(r+θ). |
When τ=0, system (1.2) becomes
{∂u(x,t)∂t=Δu(x,t)+(1−u(x,t))u(x,t)−v(x,t)u(x,t)1+cu(x,t),∂v(x,t)∂t=Δv(x,t)+βv(x,t)u(x,t)1+cu(x,t)−θu(x,t)v(x,t)−rv(x,t). | (3.1) |
Applying the upper and lower solutions method, we show that the interior equilibrium E∗(u−,v−) for system (3.1) is globally attractive under assumption (H1). We denote E∗(u−,v−)=(δ,vδ), where
δ=β−(cr+θ)−√(cr+θ−β)2−4crθ2cθ,vδ=(1−δ)(1+cδ)(0<δ<1). |
In this section, note that (u1,v1)>(u2,v2) means u1>u2 and v1>v2.
Theorem 3.1. Under assumption (H1), the interior equilibrium E∗ for the plankton system (3.1) is globallyattractive.
Proof. Firstly, we can know from system (3.1)
∂u∂t−Δu=(1−u)u−vu1+cu≤u(1−u), |
by the convergence of the logistic equation, the comparison principle of parabolic equation and Lemma 2.1, then for a sufficiently small positive number ε, there must be a t1>0 such that, for t≥t1,
u<1+ε2 and v<β(1+4r)4r−ε2. |
Moreover, for any initial value (u0,v0)>(0,0), there exists a t2>t1 such that, for t≥t2, (u,v)>(ε2,ε2)>(0,0). Then the solutions (u,v) of system (3.1) satisfy
(ε2,ε2)<(u,v)<(1+ε2,β(1+4r)4r−ε2). |
Denote (˜c1,˜c2)=(1+ε2,β(1+4r)4r−ε2) and (ˆc1,ˆc2)=(ε2,ε2), then (0,0)<(ˆc1,ˆc2)<(˜c1,˜c2) and
˜c1(1−˜c1)−˜c1ˆc21+c˜c1≤0,βˆc1˜c21+cˆc1−r˜c2−θˆc1˜c2≤0,ˆc1(1−ˆc1)−ˆc1˜c21+cˆc1≥0,β˜c1ˆc21+c˜c1−rˆc2−θ˜c1ˆc2≥0, |
and it is evident that there exists a K>0 such that
|u1(1−u1)−u1v11+cu1−(u2(1−u2)−u2v21+cu2)|≤K(|u1−u2|+|v1−v2|),|βu1v11+cu1−rv1−θu1v1−(βu2v21+cu2−rv2−θu2v2)|≤K(|u1−u2|+|v1−v2|). |
In applying the upper and lower method in References [13,14], we define the two sequences (˜cm1,˜cm2) and (ˆcm1,ˆcm2) as follows:
˜cm1=˜cm−11+1K((1−˜cm−11)˜cm−11−˜cm−11ˆcm−121+c˜cm−11),˜cm2=˜cm−12+1K(βˆcm−11˜cm−121+cˆcm−11−r˜cm−12−θˆcm−11˜cm−12),ˆcm1=ˆcm−11+1K(ˆcm−11(1−ˆcm−11)−ˆcm−11˜cm−121+cˆcm−11),ˆcm2=ˆcm−12+1K(β˜cm−11ˆcm−121+c˜cm−11−rˆcm−12−θ˜cm−11ˆcm−12), |
where (˜c01,˜c02)=(˜c1,˜c2),(ˆc01,ˆc02)=(ˆc1,ˆc2).
Therefore, it follows that
(ˆc1,ˆc2)≤(ˆcm1,ˆcm2)≤(ˆcm+11,ˆcm+12)≤(˜cm+11,˜cm+12)≤(˜cm1,˜cm2)≤(˜c1,˜c2) |
then there exist (ˉc1,ˉc2)>(0,0) and (ˇc1,ˇc2)>(0,0) satisfying
limm→∞˜cm1=ˉc1,limm→∞˜cm2=ˉc2,limm→∞ˆcm1=ˇc1,limm→∞ˆcm2=ˇc2, |
and
ˉc1(1−ˉc1)−ˉc1ˇc21+cˉc1=0,βˇc1ˉc21+cˇc1−rˉc2−θˇc1ˉc2=0,ˇc1(1−ˇc1)−ˇc1ˉc21+cˇc1=0,βˉc1ˇc21+cˉc1−rˇc2−θˉc1ˇc2=0. |
Since (δ,vδ) is the unique equilibrium of system (3.1), we observe (ˉc1,ˉc2)=(ˇc1,ˇc2). Thus, by employing in Reference [14,Theorem 2.2], the solutions (u,v) of system (3.1) converge uniformly to E∗ as t→∞.
Further, for the dynamics of system (1.2), we fix the parameters d1,d2,r,θ and c, and pick delay τ as the bifurcation parameter.
Firstly, in the phase space C=C([−τ,0],X), we linearize system (1.2) to analyze the stability of the positive equilibrium (δ,vδ), and have
˙Z(t)=DΔZ(t)+L(Zt), | (3.2) |
where D=(d100d2) and define L:C→X as
L(φt)=L1(φ1(0)φ2(0))+L2(φ1(−τ)φ2(−τ)), |
and the corresponding L1 and L2 are
L1=(1−2δ−B1−A1βB10)andL2=(00−θvδ0), |
where A1(δ)=δ1+cδ and B1(δ)=1−δ1+cδ. Then by a simple derivation, the characteristic equation of the system (1.2) at (u−,v−) is
λ2−λTn+Dn−e−λτθδ(1−δ)=0,n=0,1,2,⋯, | (3.3) |
where
{Tn=1−2δ−B1−(d1+d2)n2l2,Dn=d1d2n4l4−d2n2l2(1−2δ−B1)+βA1B1. | (3.4) |
When τ=0, Eq (3.3) becomes
λ2−λTn+Dn−θδ(1−δ)=0,n=0,1,2,⋯, | (3.5) |
and then Eq (3.5) has two roots given by
λ±n=Tn±√T2n−4(Dn−θδ(1−δ))2,n=0,1,2,⋯. |
Here, we establish the definition:
(H2)1−2δ−B1<0. |
Therefore, if (H1) holds, then βA1B1−θδ(1−δ)>0, and adding the condition (H2), and we obtain that Dn−θδ(1−δ)>0. Thus, we obtain the following results.
Lemma 3.2. Under the assumptions (H1) and (H2), the positive equilibrium E∗ of system (1.2) is locally asymptotically stable when τ=0.
Remark 3.3. When the parameters of system (1.2) satisfy the condition (H1) and (H2), λ=0 is not the characteristic root of Eq (3.5).
In the following, with the help from Ruan and Wei [15], we analyze the existence of purely imaginary eigenvalues λ=±iω(ω>0) to study the stability of the positive equilibrium E∗. Plugging λ=iω into Eq (3.3), we obtain
−ω2−Tniω+Dn−e−iωτθδ(1−δ)=0,n=0,1,2,⋯, |
then it follows from separating the real and imaginary parts that
ω4+(T2n−2Dn)ω2+D2n−θ2δ2(1−δ)2=0,n=0,1,2,⋯, | (3.6) |
we rewrite the above Eq (3.6) by z=ω2, and get
z2+(T2n−2Dn)z+D2n−θ2δ2(1−δ)2=0,n=0,1,2,⋯, | (3.7) |
and Eq (3.7) has a pair of roots given by
z±n=2Dn−T2n±√T4n−4T2nDn+4θ2δ2(1−δ)22. |
When assumptions (H1) and (H2) hold, we find that
D2n−θ2δ2(1−δ)2>0,n=0,1,2,⋯. |
Define
N:={n0∈N0|Eq(3.7)has two positive roots withn=n0}. |
By the above calculation, we have
τ±n,k=1ω±n(−arccos(ω±n)2−Dnθδ(δ−1)+2(k+1)π),n∈N,k∈N0. | (3.8) |
Lemma 3.4. When the parameters of system (1.2) satisfy conditions (H1) and (H2), the following conclusions hold.
(i) If 2Dn−T2n<0 for all n∈N0, then all the roots of Eq (3.3) have negative real parts.
(ii) If 2Dn−T2n>0 and T4n−4T2nDn+4θ2δ2(1−δ)2≥0, thenthe (n+1)th of Eq (3.3) has a pair of simple pure imaginary roots ±iω+n(±iω−n) at τ=τ+n,k(τ=τ−n,k),n∈N,k∈N0.
Lemma 3.5. When the parameters of system (1.2) satisfy conditions (H1) and (H2), if
T2n−2Dn≤0andT4n−4T2nDn+4θ2δ2(1−δ)2>0, |
then
Reλ′(τ+n,k)>0,Reλ′(τ−n,k)<0forn∈N,k∈N0. |
Proof. Taking the differential of both sides of Eq (3.3) with respect to τ, we obtain
(dλdτ)−1=Tn−2λ−τθδ(1−δ)e−λτθδ(1−δ)λe−λτ, |
then we simply calculate to obtain
Re(dλdτ)−1|τ=τ±n,k=Re((Tn−2iω±n)(cosω±nτ+isinω±nτ)−τθδ(1−δ)θδ(1−δ)iω±n)=T2n−2(Dn−(ω±n)2)θ2δ2(1−δ)2=±√T4n−4T2nDn+4θ2δ2(1−δ)2θ2δ2(1−δ)2. |
Completing the proof.
Thus, it is straightforward that
τ+n,0≤τ−n,0for alln∈N, |
and we know that ˜τ=τ+n,0 is the smallest point changing the stability of the linearized system (1.2) when the other parameters are fixed.
Theorem 3.6. Assume that (H1) and (H2) hold.
(i) If 2Dn−T2n<0 for all n∈N0, then the positive equilibrium E∗ of system (1.2) is locally asymptotically stable.
(ii) If 2Dn−T2n>0 and T4n−4T2nDn+4θ2δ2(1−δ)2<0, thenthe system (1.2) undergoes a Hopf bifurcation at E∗ for τ=τ+n,k(τ=τ−n,k),n∈N,k∈N0.Further,
Reλ′(τ+n,k)>0,Reλ′(τ−n,k)<0forn∈N,k∈N0. |
In this section, by employing the bifurcation theory in References [16,17], we analyze the property of local Hopf bifurcation near the interior equilibrium E∗. For fixed k∈N0 and n∈N, we denote ˆτ=τ±n,k. Setting τ=ˆτ+μ, then μ=0 is the Hopf bifurcation value of system (1.2). We rescale the time by t→tˆτ to normalize the delay. Meanwhile, let
z1(x,t)=u(x,τt)−δ,z2(x,t)=v(x,τt)−vδ. |
System (1.2) becomes
{∂z1∂t=ˆτ(Δz1+(1−2δ−B1)z1−A1z2+f1((z1)t,(z2)t)),∂z2∂t=ˆτ(Δz2+βB1z1−θvδz2+f2((z1)t,(z2)t)), | (4.1) |
furthermore, for simply research, system (4.1) is again transformed abstractly into
dZ(t)dt=ˆτΔZ(t)+ˆτL(Zt)+G(Zt,μ), | (4.2) |
where
L(φ)=((1−2δ−B1)φ1(0)−A1φ2(0)βB1φ1(0)−θvδφ1(−1)),G(φ,μ)=μΔZ(t)+μˆτL(Zt)+(ˆτ+μ)F0(φ),F0(φ)=(f1(φ1,φ2)f2(φ1,φ2))=((φ1(0)+δ)−(φ1(0)+δ)2−(φ1(0)+δ)(φ2(0)+vδ)1+c(φ1(0)+δ)−(1−2δ−B1)φ1(0)+A1φ2(0)β(φ1(0)+δ)(φ2(0)+vδ)1+c(φ1(0)+δ)−rvδ−θ(φ1(−1)+δ)(φ2(0)+vδ)−βB1φ1(0)+θvδφ1(−1)) | (4.3) |
for φ=(φ1,φ2)T∈C([−1,0],X).
We know that ±iω±n0ˆτ are two pairs of simple, purely imaginary eigenvalues corresponding to the following linear system at (0, 0)
dZ(t)dt=ˆτDΔZ(t)+ˆτL(Zt), | (4.4) |
and the linear functional differential equation
dZ(t)dt=ˆτL(Zt). | (4.5) |
By Riesz representation theorem, there exists a 2×2 matrix η(ϑ,ˆτ)(ϑ∈[−1,0]), we know that the elements of η(ϑ,ˆτ) are of bounded variation functions such that
(ˆτ+μ)L(φ)=∫0−1dη(ϑ,μ)φ(ϑ)forφ(ϑ)∈C([−1,0],R2), | (4.6) |
and we have
dη(ϑ,μ)=(ˆτ+μ)L1ϕ(ϑ)+(ˆτ+μ)L2ϕ(ϑ+1), |
where L1 and L2 are defined in the previous section.
Here, for the sake of derivation, we give the following notation:
an=cos(nx/l)‖cos(nx/l)‖L2,ξn={an(1,0)T,an(0,1)T} |
and
φn=⟨φ,ξn⟩=(⟨φ1,an⟩,⟨φ2,an⟩)T |
for φ=(φ1,φ1)T∈C.
Define a bilinear form
(ψ,φ)=∞∑i,j=0(ψi,φj)∫Ωaiajdx, |
where
ψ=∞∑n=0ψnan∈C∗=C([0,1],X∗),φ=∞∑n=0φnan∈C, |
and
ψn=C([0,1],(R2)∗),φn=C([−1,0],R2). |
Since
∫Ωaiajdx=0fori≠j, |
then
(ψ,φ)=∞∑n=0(ψn,φn)|an|2, |
with the bilinear form
(ψn,φn)=ˉψTn(0)φn(0)−∫0−1∫ϑξ=0ˉψTn(ξ−ϑ)dηn(0,ϑ)φn(ξ)dξ. |
For φ(ϑ)∈C1([−1,0],R2), denote
A(0)(φ(ϑ))={dφ(ϑ)dϑ,ϑ∈[−1,0),∫0−1dη(ϑ,0)φ(ϑ),ϑ=0, |
for ψ, define A∗ as
A∗ψ(s)={−dψ(s)ds,s∈(0,1],∞∑n=0∫0−1dηTn(0,t)ψn(−t)an,s=0. |
Here,
q∗(s)=1¯M(q∗1,1)eiω+n0ˆτs(s∈[0,1])andq(ϑ)=(1,q1)Teiω+n0ˆτϑ(ϑ∈[−1,0]) |
are the eigenvectors of A∗ and A(0) corresponding to the eigenvalue −iω+n0ˆτ and iω+n0ˆτ, respectively, where
q1=β(iω+n0−1+2δ+vδ(1+cδ)2+d1n2l2)r+θδ,q∗1=β(iω+n0+d2n2l2)r+θδ,¯M=ˉq∗1+q1−θvδˆτq1e−iω+n0ˆτ. |
In view of {±iω+n0ˆτ}, the center subspace of linear system (4.4) is given as
P={zqbn0+¯zqbn0|z∈C}, |
and C=P⊕Q, where Q is the stable subspace.
In the following, when we let μ=0 in Eq (4.2), there exists a center manifold
W(t,ϑ)=W(z,ˉz,ϑ)=W20(ϑ)z22+W11zˉz+W02(ϑ)ˉz22+⋯, | (4.7) |
from the theory of [16,17], system (4.4) can be rewritten as
Zt=zq(⋅)an0+ˉzˉq(⋅)an0+W(t,⋅), |
then it follows that
˙z=iωn0z+ˉq∗T(0)⟨G(0,Zt),ξn0⟩, | (4.8) |
with ⟨G,ξn⟩=(⟨G1,an⟩,⟨G2,an⟩)T. We can also write Eq (4.9) as
˙z=iωn0z+g(z,ˉz), | (4.9) |
with
g(z,ˉz)=g20z22+g11zˉz+g02ˉz22+g21z2ˉz2+⋯. |
According to the Taylor formula,
G(φ,0)=ˆτ(G1G2)=ˆτ(a20φ21(0)+a11φ1(0)φ2(0)b20φ21(0)+b11φ1(0)φ2(0)+b′11φ1(−1)φ2(0)+b30φ31(0)+b21φ21(0)φ2(0)+O(4)) |
where
a20=−(1+vδ(1+cδ)2),a11=−1(1+cδ)2,b20=−2cvδ(1+cδ)3,b11=β(1+cδ)2,b′11=−θ,b30=−2c2vδ(1+cδ)4,b21=−2c(1+cδ)3. |
By simple calculation, we have
g20=2¯M[ˉq∗1(a20+a11q1)+(b11q1+b′11e−2iωn0ˆτq1+b20)],g02=2¯M[ˉq∗1(a20+a11ˉq1)+(b11ˉq1+b′11e2iωn0ˆτˉq1+b20)],g11=1¯M[ˉq∗1(2a20+a11(q1+ˉq1))+(b11(ˉq1+q1)+b′11(e−iωn0ˆτˉq1+eiωn0ˆτq1)+2b20)],g21=1¯M[ˉq∗1(a201lπ∫lπ0(2W111+W102)dx+a111lπ∫lπ0(W2202+W1202+W111q1+W211)dx)+b111lπ∫lπ0(W2202+W1202+W111q1+W211)dx+b′111lπ∫lπ0(e−iωn0ˆτW211+eiωn0ˆτW2202+W120(−τ)2ˉq1+W111(−τ)q1)dx+b201lπ∫lπ0(W120+2W211)dx+b301lπ∫lπ0(3+3W211)dx+b21(2q1+ˉq1)]. |
In the following, we further compute W11(ϑ) and W20(ϑ) to obtain g21. Here,
˙W=˙Zt−˙zqan0−˙ˉzˉqan0=AW+H(z,ˉz,ϑ), | (4.10) |
where
H(z,ˉz,ϑ)=H20(ϑ)z22+H11(ϑ)zˉz+H02(ϑ)ˉz22+⋯. |
Therefore, we have
(2iωn0ˆτI−A0)W20=H20,−A0W11=H11,(−2iωn0ˆτ−A0)W02=H02. | (4.11) |
and from Eq (4.11), obtain
W20(ϑ)=−g20iωn0ˆτ(1q1)eiωn0ˆτϑan0−ˉg023iωn0ˆτ(1ˉq1)e−iωn0ˆτϑan0+E1e2iωn0ˆτϑ, | (4.12) |
and
W11(ϑ)=g11iωn0ˆτ(1q1)eiωn0ˆτϑan0−ˉg11iωn0ˆτ(1ˉq1)e−iωn0ˆτϑan0+E2. | (4.13) |
For Eq (4.13), when ϑ=0, we have
(2iωn0ˆτI−A0)E1=−Fzz,−A0E2=−Fzˉz, | (4.14) |
where E1=∑∞n=0En1bn and E1=∑∞n=0En1bn. Further, we obtain
(2iωn0ˆτI−A0)En1bn=−⟨Fzz,ξn⟩an,−A0En2bn=−⟨Fzˉz,ξn⟩an,n=0,1,⋯, |
where
⟨Fzz,ξn⟩={1√lπF20,n0≠0,n=0,1√2lπF20,n0≠0,n=2n0,1√lπF20,n0=0,n=0,0,other,⟨Fzˉz,ξn⟩={1√lπF11,n0≠0,n=0,1√2lπF11,n0≠0,n=2n0,1√lπF11,n0=0,n=0,0,other. |
We then compute En1 and En2 given by
En1=En11⋅En12,En2=En21⋅En22, |
where
En11=(2iωn0ˆτ−(1−2δ−B1)A1−βB1+θvδe−2iωn0ˆτ2iωn0ˆτ)−1,En12=(a20+a11q1b11q1+b′11e−2iωn0ˆτq1+b20), |
and
En21=(−(1−2δ−B1)A1−βB1+θvδ0)−1,En22=(2a20+a11(q1+ˉq1)b11(ˉq1+q1)+b′11(e−iωn0ˆτˉq1+eiωn0ˆτq1)+2b20). |
Thus, by a series of derivations and calculations, we can determine the value of g21.
Then, we analyze the bifurcation property according to the following expression:
c1(0)=i2ωn0ˆτ(g20g11−2|g11|2−13|g02|2)+12g21,μ2=Re(c1(0))Re(λ′(ˆτ)),β2=2Re(c1(0)),T2=−Im(c1(0))+μ2Im(λ′(ˆτ))ωn0ˆτ. | (4.15) |
Theorem 4.1. For system (1.2), the following statements hold.
(i) If μ2>0(<0), then the direction of the Hopf bifurcation is forward(backward), that is, there exist bifurcating periodic solutions for τ>ˆτ(τ<ˆτ);
(ii) If β2<0(>0), then the bifurcating periodic solutions on the center manifolds are orbitally stable(unstable);
(iii) If T2>0(<0), then the bifurcating periodic solutions are period increases(decreases).
Further, by Theorem 3.6, we obtain the following result.
Corollary 4.2. If Re(c1(0))<0(>0), then μ2>0(<0) and β2<0(>0).
In the above section, the sufficient condition for the occurence of Hopf bifurcation at E∗ is given. In this section, we analyze the global dynamics of system (1.2) near the equilibrium E∗. First, we cite the global Hopf bifurcation result in Reference[17].
Lemma 5.1. Let S denote the closure of the set {(z,α,β)∈E×R×(0,∞)};u(t)=z(βt) is a nontrivial 2π/β periodic solutions. Then for each connected component C, at least one of the following holds:
(1) C is unbounded, i.e. sup{maxt∈R|z(t)|+|α|+β+β−1;(z,α,β)∈C}=∞;
(2) C∩(M∗×(0,∞)) is finite and for all k≥1, one has the equality Σ(x0,α0,β0)∈C∩(M∗×(0,∞))μk(x0,α0,β0)=0.
For convenience, let zt=(ut,vt), the system (1.2) is rewriren as follows
˙z(t)=F(zt,τ,p), | (5.1) |
where zt(ϑ)=z(t+ϑ)∈C([−τ,0],X), X is the Banach space. With the help of the bifurcation theory in Reference [17], we define
Y=C([−τ,0],X),Σ=Cl{(z,τ,p)∈Y×R×R+:zis ap-periodic solution of (1.2)},N={(ˉz,ˉτ,ˉp):F(ˉz,ˉτ,ˉp)=0}, |
and l(z∗,τ+n0,k,2πω+n0) is the connected component of (z∗,τ+n0,k,2πω+n0) in Σ, where τ+n0,k and ω+n0 are defined in (3.8).
Lemma 5.2. Assume condition (H1) holds. Then system (1.2) has no any nontrivial τ-periodic solution.
Proof. By contradiction, assume that system (1.2) has τ-periodic solution, in other words, system (3.1) then has a periodic solution. We know that system (3.1) also has the same equilibria as system (1.2), that is,
E0=(0,0),E1=(1,0), |
and an interior equilibrium E∗. For system (3.1), it is straightforward that the u-axis and v-axis are the invariable manifold, and the orbits of the system do not intersect each other, thus, the solutions can not cross the coordinate axes. Hence, it follows that it must be the interior equilibrium E∗ if there exists any periodic solution within the first quadrant. From the above discussion in Lemma 2.1, the positive equilibrium E∗ is globally attractive, then system (3.1) has no nontrivial positive periodic solution, this means that system (1.2) has no nontrivial τ-periodic solution. This leads to contradiction. Completing the proof.
Theorem 5.3. Assume condition (H1) holds and 2Dn−T2n>0, T2n−4Dn+4θ2δ2(1−δ)2>0. When λ=iω+n0, then for each τ>τ+n0,k(k=0,1,…) system (1.2) has at least k+1 periodic solutions.
Proof. It is straightforward to calculate that the characteristic matrix of system (1.2) at a constant steady state ˜E=(˜u,˜v) is shown by
Υ(˜E,τ,p)(λ)=(λ+n2l2−(1−2˜u−B1)A1θ˜v−βB1λ+n2l2−(β˜u1+c˜u−r−θ˜u)). | (5.2) |
Then we find that system (1.2) has no the center of the form (E0,τ,p) and (E1,τ,p).
For discussion of E∗, we first know that (E∗,τ+n0,k,2πω+n0) is a isolated center (see [32,Definition 3.1]) by Lemma 3.4. When λ(τ+n0,k)=iω+n0, we show the result that Reλ′(τ+n0,k)>0 in Lemma 3.5. We define the smooth curve of the solution of the characteristic equation corresponding to the characteristic matrix (5.2) λ:(τ+n0,k−ζ,τ+n0,k+ζ)→C satisfying as follows: there exist ε>0,ζ>0 such that, for all τ∈[τ+n0,k−ζ,τ+n0,k+ζ]
det(Υ(λ(τ)))=0,|λ(τ)−iω+n0|<ε, |
Furthermore, we define
˜Ωε,2πω+n0={(ς,p):0<ς<ε,|p−2πω+n0|<ε}. |
and
H±(E∗,τ+n0,k,2πω+n0)(ς,p)=det(Υ(E∗,τ+n0,k±ζ,p)(ς+i2πp)). |
Then we obtain that the crossing number of the centers (E∗,τ+n0,j,2πω+n0) is
γ(E∗,τ+n0,k,2πω+n0)=deg(H−(E∗,τ+n0,k,2πω+n0),˜Ωε,2πω+n0)−deg(H+(E∗,τ+n0,k,2πω+n0),˜Ωε,2πω+n0)=−1. |
Consequently, by [32,Theorem 3.3], the connected component l(z∗,τ+n0,k,2πω+n0) of (z∗,τ+n0,k,2πω+n0) in Σ is unbounded. Then, from (3.3), we have
τ+n,k=1ω+n0(−arccos(ω+n0)2−Dnθδ(1−δ)+2(k+1)π),n0∈N,k∈N0, |
and combining with Lemmas 2.1 and 5.2, it follows that for each τ>τ+n0,k(k=0,1,…) system (1.2) has at least k+1 periodic solutions.
In the following, let Ω=(0,π), and we choose the set of parameters as
c=1.6,β=0.8,r=0.1,θ=0.05 |
to simulate the dynamics of system (1.2). For the above values and n=0, we can verify that conditions (H1) and (H2) hold, 2D0−T20≈0.0014>0 and T40−4T20D0+4θ2δ2(1−δ)2≈0.002>0, by calculation, Eq (3.8) is obtained as
τ+0,k=12.15+22.60k,τ−0,k=24.13+24.98k,k∈N0, |
then the simulations are shown as (see Figures 1 and 2)
This paper's main contribution is that it provides analytic results for the reaction-diffusion TPP-zooplankton model with Holling II response function. Here, for system (3.1), when time delay is considered, it is found that time delay is a factor that causes the dynamic behavior of the system to destabilize, that is, the positive equilibrium of system (1.2) changes from globally asymptotically stable into unstable. Furthermore, from a biological point of view, plankton populations fluctuates periodically over time. Finally, numerical simulation shows that the plankton system (1.2) with time delay discussed in this paper better describes real-world problems.
The author was supported by the China Postdoctoral Science Foundation (No. 2020M681521), Postdoctoral Science Foundation of Jiangsu Province (No. 2021K456C), Postdoctoral Science Foundation of Lianyungang City (No. LYG20210012) and Open Project of Jiangsu Key Laboratory of Marine Biological Resources and Environment (No. SH20201209).
The author declare there is no conflict of interest.
Data sharing not applicable to this article as no datasets were generated or analysed during the current study.
[1] |
S. Chen, H. Yang, J. Wei, Global dynamics of two phytoplankton-zooplankton models with toxic substances effect, J. Appl. Anal. Comput., 9 (2019), 1–14. https://doi.org/10.11948/2156-907X.20180187 doi: 10.11948/2156-907X.20180187
![]() |
[2] |
J. Chattopadhyay, R. Sarkar, A delay differential equation model on harmful algal blooms in the presence of toxic substances, IMA J. Math. Appl. Med. Biol., 19 (2002), 137–161. https://doi.org/10.1093/imammb/19.2.137 doi: 10.1093/imammb/19.2.137
![]() |
[3] |
J. Chattopadhayay, R. Sarkar, S. Mandal, Toxin-producing plankton may act as a biological control for planktonic blooms—field study and mathematical modelling, J. Theor. Biol., 215 (2002), 333–344. https://doi.org/10.1006/jtbi.2001.2510 doi: 10.1006/jtbi.2001.2510
![]() |
[4] |
S. Chaudhuri, J. Chattopadhyay, E. Venturino, Toxic phytoplankton-induced spatiotemporal patterns, J. Biol. Phys., 38 (2012), 331–348. https://doi.org/10.1007/s10867-011-9251-7 doi: 10.1007/s10867-011-9251-7
![]() |
[5] |
Y. Lv, Y. Pei, S. Gao, C. Li, Harvesting of a phytoplankton-zooplankton model, Nonlinear Anal. Real World Appl., 11 (2010), 3608–3619. https://doi.org/10.1016/j.nonrwa.2010.01.007 doi: 10.1016/j.nonrwa.2010.01.007
![]() |
[6] |
B. Mukhopadhyay, R. Bhattacharyya, Modelling phytoplankton allelopathy in a nutrient-plankton model with spatial heterogeneity, Ecol.l Modell., 198 (2006), 163–173. https://doi.org/10.1016/j.ecolmodel.2006.04.005 doi: 10.1016/j.ecolmodel.2006.04.005
![]() |
[7] |
Y. Wang, H. Wang, W. Jiang, Hopf-transcritical bifurcation in toxic phytoplankton-zooplankton model with delay, J. Math. Anal. Appl., 415 (2014), 574–594. https://doi.org/10.1016/j.jmaa.2014.01.081 doi: 10.1016/j.jmaa.2014.01.081
![]() |
[8] |
J. Zhao, J. Wei, Dynamics in a diffusive plankton system with delay and toxic substances effect, Nonlinear Anal. Real World Appl., 22 (2015), 66–83. https://doi.org/10.1016/j.nonrwa.2014.07.010 doi: 10.1016/j.nonrwa.2014.07.010
![]() |
[9] |
T. Saha, M. Bandyopadhyay, Dynamical analysis of toxin producing phytoplankton-zooplankton interactions, Nonlinear Anal. Real World Appl., 10 (2009), 314–332. https://doi.org/10.1016/j.nonrwa.2007.09.001 doi: 10.1016/j.nonrwa.2007.09.001
![]() |
[10] |
J. Zhao, J. P. Tian, J. Wei, Minimal model of plankton systems revisited with spatial diffusion and maturation delay, Bull. Math. Biol., 78 (2016), 381–412. https://doi.org/10.1007/s11538-016-0147-3 doi: 10.1007/s11538-016-0147-3
![]() |
[11] |
Y. Zhao, Z. Feng, Y. Zheng, X. Cen, Existence of limit cycles and homoclinic bifurcation in a plant-herbivore model with toxin-determined functional response, J. Differe. Equations, 258 (2015), 2847–2872. https://doi.org/10.1016/j.jde.2014.12.029 doi: 10.1016/j.jde.2014.12.029
![]() |
[12] | R. Peng, J. Shi, M. Wang, On stationary patterns of a reaction-diffusion model with autocatalysis and saturation law, Nonlinearity, 21 (2008), 1471–1488. |
[13] |
C. Pao, Dynamics of nonlinear parabolic systems with time delays, J. Math. Anal. Appl., 198 (1996), 751–779. https://doi.org/10.1006/jmaa.1996.0111 doi: 10.1006/jmaa.1996.0111
![]() |
[14] | C. Pao, Convergence of solutions of reaction-diffusion systems with time delays, Nonlinear Anal. Theory Methods Appl., 48 (2002), 349–362. |
[15] | S. Ruan, J. Wei, On the zeros of transcendental functions with applications to stability of delay differential equations with two delays, Dyn. Contin. Discrete Impulsive Syst. Series A Math. Anal., 10 (2003), 863–874. |
[16] | B. Hassard, N. Kazarinoff, Y. Wan, Theory and Applications of Hopf Bifurcation, Cambridge University Press, Cambridge, 1981. |
[17] | J. Wu, Theory and applications of partial functional-differential equations, in Applied Mathematical Sciences, Springer, New York. 1996. |
[18] | S. Chen, J. Wei, J. Yu, Stationary patterns of a diffusive predator-prey model with Crowley-Martin functional response, Nonlinear Anal. Real World Appl., 39 (2018), 33-57. |
[19] | D. Gilbarg, N. Trudinger, Elliptic Partial Differential Equation of Second Order, Springer-Verlag, Berlin. 2001. |
[20] |
T. Kar, K. Chaudhuri, On non-selective harvesting of two competing fish species in the presence of toxicity, Ecol. Modell., 161 (2003), 125–137. https://doi.org/10.1016/S0304-3800(02)00323-X doi: 10.1016/S0304-3800(02)00323-X
![]() |
[21] |
C. Lin, W. M. Ni, I. Takagi, Large amplitude stationary solutions to a chemotaxis system, J. Differ. Equations, 72 (1988), 1–27. https://doi.org/10.1016/0022-0396(88)90147-7 doi: 10.1016/0022-0396(88)90147-7
![]() |
[22] | Y. Lou, W. M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Differ. Equations, 131 (1996), 79–131. |
[23] |
Y. Lv, J. Cao, J. Song, R. Yuan, Y. Pei, Global stability and Hopf-bifurcation in a zooplankton-phytoplankton model, Nonlinear Dyn., 76 (2014), 345–366. https://doi.org/10.1007/s11071-013-1130-2 doi: 10.1007/s11071-013-1130-2
![]() |
[24] |
A. B. Medvinsky, S. V. Petrovskii, I. A. Tikhonova, H. Malchow, B. L. Li, Spatiotemporal complexity of plankton and fish dynamics, SIAM Rev., 44 (2002), 311–370. https://doi.org/10.1137/S0036144502404442 doi: 10.1137/S0036144502404442
![]() |
[25] |
W. Ni, M. Wang, Dynamics and patterns of a diffusive Leslie-Gower prey-predator model with strong Allee effect in prey, J. Differ. Equations, 261 (2016), 4244–4274. https://doi.org/10.1016/j.jde.2016.06.022 doi: 10.1016/j.jde.2016.06.022
![]() |
[26] |
R. Peng, J. Shi, Non-existence of non-constant positive steady states of two Holling type-II predator-prey systems: Strong interaction case, J. Differ. Equations, 247 (2009), 866–886. https://doi.org/10.1016/j.jde.2009.03.008 doi: 10.1016/j.jde.2009.03.008
![]() |
[27] |
R. Peng, F. Yi, X. Q. Zhao, Spatiotemporal patterns in a reaction-diffusion model with the Degn-Harrison reaction scheme, J. Differ. Equations, 254 (2013), 2465–2498. https://doi.org/10.1016/j.jde.2012.12.009 doi: 10.1016/j.jde.2012.12.009
![]() |
[28] |
J. P. Shi, X. F. Wang, On the global bifurcation for quasilinear elliptic systems on bounded domains, J. Differ. Equations, 246 (2009), 2788–2812. https://doi.org/10.1016/j.jde.2008.09.009 doi: 10.1016/j.jde.2008.09.009
![]() |
[29] | J. Smoller, A. Wasserman, Global bifurcation of steady-state solutions, J. Differ. Equations, 39 (1981), 269–290. |
[30] |
J. Wang, J. Shi, J. Wei, Predator-prey system with strong Allee effect in prey, J. Math. Biol., 62 (2011), 291–331. https://doi.org/10.1007/s00285-010-0332-1 doi: 10.1007/s00285-010-0332-1
![]() |
[31] |
J. Wang, J. Shi, J. Wei, Dynamics and pattern formation in a diffusive predator-prey system with strong Allee effect in prey, J. Differ. Equations, 251 (2011), 1276–1304. https://doi.org/10.1016/j.jde.2011.03.004 doi: 10.1016/j.jde.2011.03.004
![]() |
[32] |
J. Wu, Symmetric functional differential equations and neural networks with memory, Trans. Amer. Math. Soc., 350 (1998), 4799–4838. https://doi.org/10.1090/S0002-9947-98-02083-2 doi: 10.1090/S0002-9947-98-02083-2
![]() |
[33] |
X. Xu, J. Wei, Bifurcation analysis of a spruce budworm model with diffusion and physiological structures, J. Differ. Equations, 262 (2017), 5206–5230. https://doi.org/10.1016/j.jde.2017.01.023 doi: 10.1016/j.jde.2017.01.023
![]() |
[34] | H. Yang, Analysis of stationary patterns and bifurcations of a diffusive phytoplankton-zooplankton model with toxic substances effect, J. Math. Anal. Appl., Forthecoming, 2023. |
[35] |
F. Yi, J. Liu, J. Wei, Spatiotemporal pattern formation and multiple bifurcations in a diffusive bimolecular model, Nonlinear Anal. Real World Appl., 11 (2010), 3770–3781. https://doi.org/10.1016/j.nonrwa.2010.02.007 doi: 10.1016/j.nonrwa.2010.02.007
![]() |
[36] | F. Yi, J. Wei, J. Shi, Bifurcation and spatio-temporal patterns in a diffusive homogenous predator-prey system, J. Differ. Equations, 246 (2009), 1944–1977. |
1. | Lidan Liu, Meng Fan, Yun Kang, Effect of nutrient supply on cell size evolution of marine phytoplankton, 2022, 20, 1551-0018, 4714, 10.3934/mbe.2023218 | |
2. | Sobirjon Shoyimardonov, NEIMARK-SACKER BIFURCATION AND STABILITY ANALYSIS IN A DISCRETE PHYTOPLANKTON-ZOOPLANKTON SYSTEM WITH HOLLING TYPE Ⅱ FUNCTIONAL RESPONSE, 2023, 13, 2156-907X, 2048, 10.11948/20220345 | |
3. | Haixia Li, Gaihui Guo, Lijuan Wang, Aili Wang, The effects of toxin and mutual interference among zooplankton on a diffusive plankton–fish model with Crowley–Martin functional response, 2024, 1747-6933, 1, 10.1080/17476933.2024.2409882 | |
4. | Yulong Li, Long Zhou, Fengjie Geng, Dynamics on semi-discrete Mackey-Glass model, 2025, 10, 2473-6988, 2771, 10.3934/math.2025130 |