
Citation: Xuehui Ji, Sanling Yuan, Tonghua Zhang, Huaiping Zhu. Stochastic modeling of algal bloom dynamics with delayed nutrient recycling[J]. Mathematical Biosciences and Engineering, 2019, 16(1): 1-24. doi: 10.3934/mbe.2019001
[1] | Da Song, Meng Fan, Ming Chen, Hao Wang . Dynamics of a periodic stoichiometric model with application in predicting and controlling algal bloom in Bohai Sea off China. Mathematical Biosciences and Engineering, 2019, 16(1): 119-138. doi: 10.3934/mbe.2019006 |
[2] | Wenjie Yang, Qianqian Zheng, Jianwei Shen, Linan Guan . Bifurcation and pattern dynamics in the nutrient-plankton network. Mathematical Biosciences and Engineering, 2023, 20(12): 21337-21358. doi: 10.3934/mbe.2023944 |
[3] | Qi Zhou, Huaimin Yuan, Qimin Zhang . Dynamics and approximation of positive solution of the stochastic SIS model affected by air pollutants. Mathematical Biosciences and Engineering, 2022, 19(5): 4481-4505. doi: 10.3934/mbe.2022207 |
[4] | Ying He, Yuting Wei, Junlong Tao, Bo Bi . Stationary distribution and probability density function analysis of a stochastic Microcystins degradation model with distributed delay. Mathematical Biosciences and Engineering, 2024, 21(1): 602-626. doi: 10.3934/mbe.2024026 |
[5] | Frédéric Mazenc, Gonzalo Robledo, Daniel Sepúlveda . A stability analysis of a time-varying chemostat with pointwise delay. Mathematical Biosciences and Engineering, 2024, 21(2): 2691-2728. doi: 10.3934/mbe.2024119 |
[6] | 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 |
[7] | Tingting Yu, Sanling Yuan . Dynamics of a stochastic turbidostat model with sampled and delayed measurements. Mathematical Biosciences and Engineering, 2023, 20(4): 6215-6236. doi: 10.3934/mbe.2023268 |
[8] | Yan Xie, Zhijun Liu, Ke Qi, Dongchen Shangguan, Qinglong Wang . A stochastic mussel-algae model under regime switching. Mathematical Biosciences and Engineering, 2022, 19(5): 4794-4811. doi: 10.3934/mbe.2022224 |
[9] | Ying He, Junlong Tao, Bo Bi . Stationary distribution for a three-dimensional stochastic viral infection model with general distributed delay. Mathematical Biosciences and Engineering, 2023, 20(10): 18018-18029. doi: 10.3934/mbe.2023800 |
[10] | Jean-Jacques Kengwoung-Keumo . Competition between a nonallelopathic phytoplankton and an allelopathic phytoplankton species under predation. Mathematical Biosciences and Engineering, 2016, 13(4): 787-812. doi: 10.3934/mbe.2016018 |
It is known that many shallow lakes in China have been experiencing a problem of algal bloom resulting in shortage of drinking water supply and degradation of lake ecosystem [1]. For example, it is reported that in the spring of 2007, the accumulation of algal blooms in Tai Lake, the third largest freshwater lake in China, left approximately 2 million people of Wuxi City without drinking water for at least a week, and led to massive and rapid fish deaths [2,3,4]. The bloom of algal is a complex process [5] and is due to the increased input of nutrients, such as nitrogen and phosphorus, which directly affect the growth of algal [6] and nutrient-rich sediment (dead algae) [7], which sinks to the bottom of the lake forming detritus, and then is converted into nutrients due to the decomposition by bacteria [8,9].
It should be noted from [8,9]. that the conversion of detritus into nutrients is not instantaneous and requires time for the completion of bacterial decomposition resulting a delay. To describe the conversion of detritus into nutrients, authors of [8,9] proposed a nutrient-phytoplankton model
{˙S=D(S0−S)−mU(S)x+μD1∫∞0f(s)x(t−s)ds,˙x=x[−(D+D1)+γmU(S)], | (1.1) |
where S(t) and x(t) denote the concentrations of a limiting nutrient (nitrogen or phosphorus) and algae density at time t, respectively; S0 is the input concentration of the nutrient, D is the washout rate, m is the maximum specific uptake rate of nutrient, D1 is the death rate of the algae, γ∈(0,1), is the fraction of nutrient conversion, the delay kernel f(s) is a non-negative bounded function defined on [0,∞) describing the contribution of the biomass dead in the past to the nutrient recycling at time t, μ∈(0,1) is the fraction of the nutrient recycled by bacterial decomposition of the dead algae, U(S) is the specific growth function. It is found in [9]that the delay should be relatively small to have global attractivity of positive equilibrium. Recently, Misra and Chandra [11] studied the effect of delayed nutrient recycling on a nutrient-algae-dissolved oxygen model. They found that if the delay is lager than some threshold then the concentration of algae fluctuates and may increase drastically; meanwhile, the concentration of dissolved oxygen may reduce drastically, resulting massive death of fish population in the lake. To avoid the massive death of fish population and other losses, the detritus from the lake should be regularly removed before the threshold value. To provide sound scientific advices, various studies using mathematical models have been conducted for algal bloom caused by nutrients in the lakes, where both nutrient and algae population dynamics are described by a system of coupled deterministic continuous processes [12,13,14,15].
It is known that nutrient inputs from runoff vary not only in quantity (influenced by rainfall and other environmental factors), but also in composition (based on the form of fertilizer in use) [16]. These variations are random in nature and generally affect the nutrient balances, influence mass transfer efficiencies along food chains [17], and therefore play a key role in bloom initiation and maintenance [18]. To better understand algal bloom phenomenon, stochastic models are needed, describing the variation of nutrients and environmental forcing. Several studies have already addressed the stochastic modelling of the algal bloom. For example, according to the experimental data from Bohai Bay of China, Huang et al. [19] constructed a nonlinear stochastic model on densities of two kinds of typical HAB (harmful algae blooms) algae: diatom and dianoflagellate, and analyzed their characters of stability and Hopf bifurcation; Das et al. [20] proposed a stochastic model with the main purpose of considering the severity and duration of algal blooms in the ecological arena; Mandal et al. [21] established stochastic models of allelopathic interactions between two competing phytoplankton species as a continuous time Markov chain model and as an Itô stochastic differential equation model, in which approximate extinction probabilities for both species were obtained analytically for the continuous time Markov chain model.
Our goals of this paper are to extend the delayed model (1.1) into a stochastic one for algal bloom and to explore how the random fluctuations and the delay affect the dynamics of algae population. Recall that system (1.1) has a washout equilibrium E0=(S0,0) when
U(S0)<D+D1γm | (1.2) |
and a positive equilibrium E∗(S∗,x∗) when
D+D1<γm and S0>S∗, | (1.3) |
with
S∗=U−1(D+D1γm), x∗=D(S0−S∗)mU(S∗)−μD1 |
and here U−1 is the inverse function of U [8,9]. Under random fluctuations, one naturally concerns whether or not there exists a steady state for the stochastic system, and how the dynamics of algae population can be affected.
To derive a reasonable stochastic analogue of the deterministic model (1.1), it generally needs the help of Markov chain [22]. But this approach is not justified for (1.1) because the delayed stochastic system is non-Markovian. So, in this paper, we assume that f(s) takes the following special family of memory functions:
fnα(s)=αn(n−1)!sn−1e−αs, | (1.4) |
where α>0 is a constant, n is an nonnegative integer. In particular, we call (1.4) the weak kernel when n=1 and strong kernel when n=2. Notice also that in the limit case when n→∞, the delay kernel fnα(s) converges to a Dirac delta function f(s)=δ(s−τf). Then (1.1) becomes a model with discrete time delay τf. To properly propose our model, we first convert (1.1) into a system of ordinary differential equations by using chain-trick, and then derive a stochastic analogue by means of Markov chain. Based on the derived stochastic model we will investigate the effects of delay and environmental random fluctuations on the dynamics of the model.
The paper is organized as follows. In Section 2, a detailed derivation of the stochastic model is performed by using discrete Markov chain. In Section 3, we show the uniqueness and global existence of a positive solution of the stochastic model. In Section 4, we carry out the long term behavior analysis of the model: we first investigate the sufficient conditions for the stochastic stability of the washout equilibrium; then the spectral densities of the nutrient and the algae population are estimated by using Fourier transform method. Our study shows that the algae population can be extinct if experiencing a sufficiently large noise. Finally, numerical simulations to illustrate the results obtained and a brief discussion are presented in Section 5.
By taking into account of random effects, we consider a stochastic analogue of the deterministic model (1.1)
{dSdt=D(S0−S)−mU(S)x+μD1∫∞0f(s)x(t−s)ds+σ1Sξ1(t),dxdt=x[−(D+D1)+γmU(S)]+σ2xξ2(t), | (2.1) |
with initial value
S(θ,ω)=φ1(θ)≥0, x(θ,ω)=φ2(θ)≥0, θ∈(−∞,0] and φi(0)>0 (i=1,2), |
where φ1(θ), φ2(θ)∈C((−∞,0],R+), the families of continuous functions from (−∞,0] to R+. (ξ1(t),ξ2(t)) is a two dimensional Gaussian white noise process satisfying
E[ξi(t)=0], i=1,2; E[ξi(t)ξj(t′)]=δijδ(t−t′), i,j=1,2 |
where δij is the Kronecker symbol and δ is the Dirac delta function. Since ξi(t) is delta-correlated, so ξi(t)dt can be written as dBi(t), where B(t)=(B1(t),B2(t) is an independent Brownian motion or Wiener process. Thus (2.1) can be rewritten as the form
{dS=[D(S0−S)−mU(S)x+μD1∫∞0f(s)x(t−s)ds]dt+σ1SdB1(t),dx=x[−(D+D1)+γmU(S)]dt+σ2xdB2(t). | (2.2) |
We now show in detail that model (2.2) or (2.1) is a reasonable stochastic analogue of the deterministic model (1.1). To this end, introduce
yi=∫∞0fiα(s)x(t−s)ds, i=1,2,…,n, |
then (1.1) becomes the following system of coupled ordinary differential equations
˙S=D(S0−S)−mU(S)x+μD1yn,˙x=x[−(D+D1)+γmU(S)],˙y1=−α(y1−S),˙yi=−α(yi−yi−1), i=2,3,…,n. | (2.3) |
Follow the idea and techniques in [22], next we derive a reasonable stochastic analogue of (2.3) using a discrete time Markov chain. For a fixed time increment Δt>0 and t=0,Δt,2Δt,⋯ define a process
XΔt(t)=(SΔt(t),xΔt(t),yΔt1,yΔt2,…,yΔtn)T. |
Here SΔt(t) denotes the nutrient concentration, xΔt(t) denotes the concentration of algae population, and yΔti(t), i=1,2,…,n, are auxiliaries. Let the initial value be
XΔt(0)=(S(0),x(0),y1(0),y2(0),…,yn(0))T |
and {RΔtj(k)}∞k=0, j=1,2,…,n+2 denote the n+2 sequences of random variables. Suppose that these variables are jointly independent and that within each sequence the variables are identically distributed such that
ERΔtj(k)=0, E[RΔtj(k)]2=σ2jΔt, j=1,2, k=0,1,..., | (2.4) |
where σj≥0 are constants reflecting the size of the stochastic effects, and
RΔtj(k)=0, j=3,4,…,n+2, k=0,1,... | (2.5) |
Variable RΔt2(k) is supposed to capture the effect of random influences (due to the external factors such as variation of nutrients, environmental forcing, or inherent factors, e.g. mutations) on the concentration of the algae during the period [kΔt,(k+1)Δt). We assume that xΔt grows within the time period according to the deterministic system (2.3) and by the random amount RΔt2(k)xΔt(kΔt). Random effects on SΔt can be similarly modelled as RΔt1(k)SΔt(kΔt). Specifically, for k=0,1,..., we set
SΔt((k+1)Δt)=SΔt(kΔt)+Δt{D(S0−SΔt(kΔt))−mU(SΔt(kΔt))xΔt(kΔt)+μD1yΔtn(kΔt)}+RΔt1(k)SΔt(kΔt), |
xΔt((k+1)Δt)=xΔt(kΔt)+Δt{xΔt(kΔt)[−(D+D1)+γmU(SΔt(kΔt))]}+RΔt2(k)xΔt(kΔt), |
yΔt1((k+1)Δt)=yΔt1(kΔt)−Δtα(yΔt1(kΔt)−SΔt(kΔt)) |
and
yΔti((k+1)Δt)=yΔti(kΔt)−Δtα(yΔti(kΔt)−yΔti−1(kΔt)), i=2,3,…,n. |
We next show that XΔt converges to a diffusion process as Δt→0. To this end, we first determine the drift coefficients of the diffusion. Let PΔt(u,dv) denote the transition probabilities of the homogeneous Markov chain {XΔt(kΔt)}∞k=0, that is
PΔt(u,A)=P{XΔt((k+1)Δt)∈A | XΔt(kΔt)=u} |
for all u=(u1,…,un+2)∈Rn+2 and all Borel sets A⊂Rn+2. Let u1=S, u2=x, u3=y1,…,un+2=yn. Then, by (2.4) and (2.5), we have
1Δt∫(v1−u1)PΔt(u,dv)=D(S0−S)−mU(S)x+μD1yn+SΔtERΔt1(0)=D(S0−S)−mU(S)x+μD1yn, | (2.6) |
1Δt∫(v2−u2)PΔt(u,dv)=x[−(D+D1)+γmU(S)]+xΔtERΔt2(0)=x[−(D+D1)+γmU(S)], | (2.7) |
1Δt∫(v3−u3)PΔt(u,dv)=−α(y1−S) | (2.8) |
and
1Δt∫(vi+2−ui+2)PΔt(u,dv)=−α(yi−yi−1), i=2,…,n. | (2.9) |
To determine the diffusion coefficients, we consider the moments
gΔtij(u)=1Δt∫(vi−ui)(vj−uj)PΔt(u,dv), i,j=1,2,…,n+2. |
It follows from (2.4) and (2.5) that
|gΔt22(u)−σ22x2|=|1ΔtE[Δt[−(D+D1)+γmU(S)]x+RΔt2(0)x]2−σ22x2|=|Δt[−(D+D1)+γmU(S)]2x2+2[−(D+D1)+γmU(S)]x2ERΔt2(0)+1Δtx2E[RΔt2(0)]2−σ22x2|=Δt[−(D+D1)+γmU(S)]2x2. |
Thus,
limΔt→0+sup‖u‖≤K|gΔt22(u)−σ22x2|=0 | (2.10) |
for all K∈(0,∞). Similarly, for all K∈(0,∞), we can obtain
limΔt→0+sup‖u‖≤K|gΔt11(u)−σ21S2|=0, | (2.11) |
limΔt→0+sup‖u‖≤K|gΔti+2,i+2(u)|=0,i=1,2,…,n | (2.12) |
and
limΔt→0+sup‖u‖≤K|gΔtij(u)|=0, for i,j=1,…,n+2 and i≠j. | (2.13) |
Assuming that E[RΔti(k)]4=o(Δt) for i=1,2, one may verify that for all K∈(0,∞),
limΔt→0+sup‖u‖≤K1Δt∫‖u−v‖3PΔt(u,dv)=0. | (2.14) |
Extend the definition of XΔt(t) to all t≥0 by setting XΔt(t)=XΔt(kΔt) for t∈[kΔt,(k+1)Δt) and let X(t) be the solution of system
dS=[D(S0−S)−mU(S)x+μD1yn]dt+σ1SdB1(t),dx=x[−(D+D1)+γmU(S)]dt+σ2xdB2(t),dy1=−α(y1−S)dt,dyi=−α(yi−yi−1)dt, i=2,3,…,n. | (2.15) |
with initial condition X(0)=(S(0),x(0),y1(0),…,yn(0))T. Then according to Theorem 7.1 of [23], we can from (2.6)-(2.15) obtain the following lemma.
Lemma 2.1. Given initial condition X(0)=(S(0),x(0),y1(0),…,yn(0))T, assume X(t) is the unique solution of (2.15). Then XΔt(t) converges weakly to X(t) as Δt→0.
Notice that if we take n→∞ for the kernel fnα(s), then it converges to a Dirac delta function. Thus (2.1) becomes a stochastic model with discrete time delay. Then, conditional probability density can be used to determine the "conditional average drift" and "conditional average diffusion" of the process (see Chapter 6 in [24]).
We should point out that (2.1) and (2.2) are two equivalent stochastic counterparts of deterministic model (1.1). In the sequel, we will select the specific form of the random model according to the specific needs. In the following sections, we will first show the uniqueness of positive solution of system (2.1), and then study the dynamics of the system on a complete probability space (Ω,F,{Ft}t≥0,P) with a filtration {Ft}t≥0 satisfying the usual conditions, i.e., it is right continuous and F0 contains all P-null sets. Before that, we introduce a differential operator L associated with a general n-dimensional stochastic functional differential equation [25]:
dx=f(t,xt)dt+g(t,xt)dB(t) | (2.16) |
with initial condition x0=φ∈H, where H is the space of F0-adapted random variables φ, with φ(s)∈Rn for s≤0, and
‖φ‖=sups≤0|φ(s)|, ‖φ‖21=supE(|φ(s)|2). |
B(t) denotes m-dimensional standard Brownian motions. The differential operator L is defined as
L=∂∂t+f(xt,t)T⋅∂∂xt+12Tr[gT(xt,t)⋅∂2∂x2t⋅g(xt,t)]. | (2.17) |
The existence and uniqueness theorem of solutions for stochastic functional differential equations had been studied by Mao [25], see also the related results in [26,27]. However, these results are all in the case that the delay is finite. Recently, Wei and Wang [28] generalized the theorem to the case of infinite delay under the linear growth condition and local Lipschitz condition. For our model (2.1), obviously, the coefficients are locally Lipschitz continuous, but they do not satisfy the linear growth condition. So, the solution of system (2.1) may explode at a finite time. In this section, we shall show that the solution of system (2.1) with any positive initial value remains positive for all t≥0.
Theorem 3.1. Given any initial value (φ1(θ),φ2(θ))∈C((−∞,0],R2+), model (2.1) has a unique solution (S(t),x(t)) for all t≥0; furthermore, the solution remains positive for all t>0 with probability 1, namely S(t)>0 and x(t)>0 for all t≥0 almost surely, if the specific growth function U(S) satisfies
U(0)=0, U′(S)>0, U″(S)<0 for S≥0 and limS→∞U(S)=1. |
Proof. It is easy to verify that for any given initial value (φ1(θ),φ2(θ))∈C((−∞,0],R2+), system (2.1) has a unique local solution (S(t),x(t)) on t∈[0,τe), where τe is the explosion time. To prove this theorem, we only need to show this solution is also global, that is τe=∞ a.s.
Let k0>0 be sufficiently large so that 1k0≤φ1(0),φ2(0)≤k0, and for each integer k≥k0 define
τk=inf{t∈[0,τe):S(t)∉(1k,k) or x(t)∉(1k,k)}, |
which is known as the stopping time and increasing as k→∞. Set τ∞=limk→∞τk, whence τ∞≤τe a.s.. We next show τ∞=∞ a.s. by contradiction, which implies τe=∞ a.s..
If τ∞≠∞, then there is a pair of constants T>0 and δ∈(0,1) such that
P(τ∞≤T)>δ. |
Hence there is an integer k1≥k0 such that for all k≥k1
P(τk≤T)≥δ. | (3.1) |
Consider a Lyapunov function
V(S,x)=γ(S−C1−C1lnSC1)+(x−1−lnx)+γμD1∫∞0f(s)∫tt−sx(τ)dτds, | (3.2) |
where C1=D+D1−γμD1γmU′(0). By Itô's formula, one has
dV(S,x)=γdS−γC1SdS+γC12S2(dS)2+dx−1xdx+12x2(dx)2+γμD1(x(t)−∫∞0f(s)x(t−s)ds)dt=LV(S,x)dt+γσ1(S−C1)dB1+σ2(x−1)dB2, | (3.3) |
where
LV(S,x)=γDS0+γC1D+D+D1+12γC1σ21+12σ22+γμD1∫∞0f(s)x(t−s)ds−γDS+γC1mU(S)Sx−γC1DS0S−γC1μD1∫∞0f(s)x(t−s)dsS−(D+D1)x−γmU(S)+γμD1(x(t)−∫∞0f(s)x(t−s)ds). | (3.4) |
Notice for S>0
U′(S)<U(S)S<U′(0). |
Then from (3.4), we deduce that
LV≤γDS0+γC1D+D+D1+12γC1σ21+12σ22−(D+D1−γμD1−γC1mU′(0))x≤γDS0+γC1D1+D+D1+12γC1σ21+12σ22≜K2. |
Therefore,
dV(S,x)≤K2dt+γσ1(S−C1)dB1+σ2(x−1)dB2. |
Integrating both sides of the above inequality from 0 to τk∧T, and taking expectation, yields
E(V(S(τk∧T),x(τk∧T)))≤V(φ1(0),φ2(0))+K2E(τk∧T). |
One then has
E(V(S(τk∧T),x(τk∧T)))≤V(φ1(0),φ2(0))+K2T. | (3.5) |
Set Ωk={τk≤T}(k≥k1) and by (3.1), P(Ωk)≥δ. Note that for every ω∈Ωk, there is at least one of S(τk,ω), x(τk,ω) which equals either k or 1k. Then
V(S(τk,ω),x(τk,ω))≥γ(k−C1−C1lnkC1)∧γ(1k−C1−C1ln1kC1)∧(k−1−lnk)∧(1k−1−ln1k). |
It follows from (3.5) that
V(φ1(0),φ2(0))+K2T≥E(1Ωk(ω)V(S(τk∧T),x(τk∧T)))≥δ[γ(k−C1−C1lnkC1)∧γ(1k−C1−C1ln1kC1)∧(k−1−lnk)∧(1k−1−ln1k)], |
where 1Ωk is the indicator function of Ωk. Let k→∞, then
∞>V(φ1(0),φ2(0))+K2T=∞, |
which leads to a contradiction. So τ∞=∞. This completes the proof.
In this section, we will study the long term dynamical behavior of model (2.1), particularly near the equilibria of the corresponding deterministic model (1.1). Obviously, when σ1=0 and σ2≠0, E0 is still an equilibrium of stochastic model (2.1), but E∗ is not; when σ1 and σ2 are positive, neither E0 or E∗ is the equilibrium of model (2.1). Therefore, we will investigate
(a) the stochastic stability of E0 when σ1 is zero; and
(b) the spectral densities of the nutrient and the algae population when σi>0,i=1,2.
When σ1=0, substituting u1=S−S0, u2=x into model (2.2), one obtains
{du1=[−Du1−mU(u1+S0)u2+μD1∫∞0f(s)u2(t−s)ds]dt,du2=u2[−(D+D1)+γmU(u1+S0)]dt+σ2u2dB2, | (4.1) |
which has the linearized system
{du1=[−Du1−mU(S0)u2+μD1∫∞0f(s)u2(t−s)ds]dt,du2=u2[−(D+D1)+γmU(S0)]dt+σ2u2dB2. | (4.2) |
For the trivial solution of system (4.2), one has the following stability result. The method of constructing Liyapunov functions in its proof will help us to construct suitable Liyapunov functions in investigating the stability of the trivial solution of nonlinear system (4.1).
Proposition 4.1. Assume C2=μD1[D+D1−γmU(S0)] and σ1=0. Then the trivial solution of system (4.2) is asymptotically mean square stable if
mU(S0)+μD1+C2Tf<2D, and | (4.3) |
2γmU(S0)+σ22<2(D+D1), | (4.4) |
where Tf is the average delay with value Tf=nα.
Proof. Consider the function
V1(u1,u2)=u21. | (4.5) |
It follows from (4.2) and Itô's formula that
dV1(u1,u2)=2u1du1+(du1)2=[−2Du21−2mU(S0)u1u2+2μD1u1∫∞0f(s)u2(t−s)ds]dt=[−2Du21−2mU(S0)u1u2+2μD1u1u2−2μD1u1∫t0f(s)∫tt−sdu2(τ)ds+h(t)]dt, |
where
h(t)=−2μD1u1∫∞tf(s)(u2(t)−u2(t−s))ds. | (4.6) |
Then we have
LV1(u1,u2)=−2Du21−2mU(S0)u1u2+2μD1u1u2+h(t)−2μD1u1∫∞0f(s)∫tt−s[−(D+D1)+γmU(S0)]u2(τ)dτds≤−2Du21−2mU(S0)u1u2+2μD1u1u2+h(t)+μD1[D+D1−γmU(S0)]∫∞0f(s)∫tt−s(u21(t)+u22(τ))dτds=−2Du21−2mU(S0)u1u2+2μD1u1u2+h(t)+μD1[D+D1−γmU(S0)](Tfu21(t)+∫∞0f(s)∫tt−su22(τ)dτds). | (4.7) |
Define function
V2(u1,u2)=C2∫∞0f(s)∫tt−s∫tru22(τ)dτdrds | (4.8) |
which is well defined since ∫∞0s2f(s)ds=n(n+1)α2. Then by Itô's formula, we have
LV2(u1,u2)=C2Tfu22(t)−C2∫∞0f(s)∫tt−su22(τ)dτds. | (4.9) |
It then follows from (4.7) and (4.9) that
L(V1+V2)≤−2Du21−2mU(S0)u1u2+2μD1u1u2+C2Tfu21+C2Tfu22+h(t). | (4.10) |
We now consider a function
V3(u1,u2)=C3u22, |
where C3=2D2(D+D1)−2γmU(S0)−σ22. By Itô's formula, we have
LV3(u1,u2)=2C3[−(D+D1)+γmU(S0)]u22+C3σ22u22. | (4.11) |
For function
V(u1,u2)=V1(u1,u2)+V2(u1,u2)+V3(u1,u2), |
we have from (4.10) and (4.11) that
LV(u1,u2)≤−(2D−mU(S0)−μD1−C2Tf−2μD1∫∞tf(s)ds)u21+[mU(S0)+μD1+C2Tf−2C3(D+D1)+2C3γmU(S0)+C3σ22+μD1∫∞tf(s)ds]u22+μD1‖φ2‖2∫∞tf(s)ds≤−(2D−mU(S0)−μD1−C2Tf−2μD1∫∞tf(s)ds)(u21+u22)+μD1‖φ2‖2∫∞tf(s)ds. |
By (4.3), we choose ε>0 such that
mU(S0)+μD1+C2Tf+2μD1ε<2D. |
Let T=T(ε)>0 such that ∫∞tf(s)ds<ε for all t≥T. Then for all t≥T, one has
LV(u1,u2)≤−(2D−mU(S0)−μD1−C2Tf−2μD1ε)(u21+u22)+μD1‖φ2‖2∫∞tf(s)ds. |
Integrating both sides of the above from T to t≥T yields
E(V(t))+(2D−mU(S0)−μD1−C2Tf−2μD1ε)∫tTE(u21(s)+u22(s))ds≤V(T)+μD1‖φ2‖2∫tT∫∞sf(u)duds≤V(T)+μD1‖φ2‖2∫∞0sf(s)ds=V(T)+μD1‖φ2‖2Tf<∞. |
Using the similar discussion as that in [9] and the Barbǎlat lemma, we conclude that E(u21(t)+u22(t))→0 as t→∞. Applying the definition of mean square stability of the solution [25], we obtain the conclusion.
Next, we give the result about the stability of the trivial solution of non-linear system (4.1), that is, the stability of E0(S0,0) of system (2.1).
Theorem 4.1. Let σ1=0 and assume conditions (4.3) and (4.4) hold. Then the trivial solution (0,0) of system (4.1) or the equilibrium E0(S0,0) of system (2.1) is stochastically stable.
Proof. Let (u1,u2) be any solution of system (4.1). Define the stopping time
Tε1=inf{t≥0:u21+u22≥ε21}. |
For the Lyapunov function V1(u1,u2) defined in (4.5), we obtain
LV1(u1,u2)=−2Du21−2mU(u1+S0)u1u2+2μD1u1u2+h(t)−2μD1u1∫∞0f(s)∫tt−s[−(D+D1)+γmU(u1+S0)]u2(τ)dτds≤−2Du21−2mU(u1+S0)u1u2+2μD1u1u2+h(t)+μD1[D+D1−γmU(S0)]∫∞0f(s)∫tt−s(u21(t)+u22(τ))dτds=−2Du21−2mU(u1+S0)u1u2+2μD1u1u2+h(t)+μD1[D+D1−γmU(S0)]×(Tfu21(t)+∫∞0f(s)∫tt−su22(τ)dτds), | (4.12) |
where h(t) is defined as in (4.6). For V2(u1,u2) in (4.8), one then has
L(V1+V2)≤−2Du21−2mU(u1+S0)u1u2+2μD1u1u2+C2Tfu21+C2Tfu22+h(t). | (4.13) |
Now define
V3(u1,u2)=C4u22, |
where C4 is a constant to be determined later. We have that
LV3(u1,u2)=2C4[−(D+D1)+γmU(u1+S0)]u22+C4σ22u22. | (4.14) |
Therefore, for the function
V(u1,u2)=V1(u1,u2)+V2(u1,u2)+V3(u1,u2), |
it follows from (4.13) and (4.14) that
LV(u1,u2)≤−2Du21−2mU(u1+S0)u1u2+2μD1u1u2+C2Tfu21+C2Tfu22+2C4[−(D+D1)+γmU(u1+S0)]u22+h(t)+C4σ22u22. | (4.15) |
By (4.4), one can find a constant δ>0 such that
2γmU(δ+S0)+σ22<2(D+D1). | (4.16) |
Choose
C4=2D2(D+D1)−2γmU(δ+S0)−σ22. |
Then, it follows from (4.15) and (4.16) that
LV(u1,u2)≤−(2D−mU(S0)−μD1−C2Tf−2μD1∫∞tf(s)ds)u21+[mU(S0)+μD1+C2Tf−2C3(D+D1)+2C3γmU(S0)+C3σ22+μD1∫∞tf(s)ds]u22+μD1‖φ2‖2∫∞tf(s)ds≤−(2D−mU(S0)−μD1−C2Tf−2μD1∫∞tf(s)ds)(u21+u22)+μD1‖φ2‖2∫∞tf(s)ds. | (4.17) |
Integrating both sides of the above formula from 0 to t∧Tε1, yields
E(V(t∧Tε1))≤V(0)+μD1‖φ2‖2∫t∧Tε10∫∞sf(τ)dτds≤‖φ1‖2+[12μD1(D+D1−γmU(S0))∫∞0s2f(s)ds+C4]‖φ2‖2+μD1‖φ2‖2∫∞0sf(s)ds≤(1∨p)(‖φ1‖2+‖φ2‖2), |
where p=12μD1(D+D1−γmU(S0))∫∞0s2f(s)ds+C4+μD1Tf. Now for ε1, ε2∈(0,1), let
δ=min{(1∧C41∨pε2)12ε1, ε12}. |
Then, if ‖φ1‖2+‖φ2‖2<δ2, it follows that
E(V(t∧Tε1))≤(1∨p)δ2≤(1∧C4)ε21ε2. |
On the other hand, we have
E(V(t∧Tε1))≥E[1{Tε1≤t}V(t∧Tε1)]=E[1{Tε1≤t}V(Tε1)]=P{Tε1≤t}V(Tε1)≥(1∧C4)ε21P{Tε1≤t}. |
Hence, we have P{Tε1≤t}≤ε2. Letting t→∞ gives
P{Tε1<∞}≤ε2, |
which is equivalent to
P{u21+u22<ε21}≥1−ε2. |
Then the definition of the stochastic stability of the solution implies the conclusion.
In this subsection, we study the dynamics of system (2.1) near E∗, the positive equilibrium of (1.1). To this end, we assume condition (1.3) holds. Thus E∗ is a stable positive stable positive equilibrium for system (1.1) but not the equilibrium for system (2.1) when σi≠0. We aim to investigate the spectral densities, denoting the intensities of fluctuations, of the nutrient and the algae population by Fourier transform method.
To perform the spectral density analysis on model (2.1) with a general delay kernel function is very difficult, so we turn to consider the limit case when n→∞ in (1.4), i.e., f(s)=δ(s−τf). Introducing S(t)=exp(u1(t)) and x(t)=exp(u2(t)), one converts (2.1) into
{du1dt=DS0e−u1−D−mU(eu1)eu2−u1+μD1eu2(t−τf)−u1+σ1ξ1,du2dt=−(D+D1)+γmU(eu1)+σ2ξ2. | (4.18) |
Substituting
u1=v1+u∗1, u2=v2+u∗2 |
into (4.18), yields
{dv1dt=DS0e−(v1+u∗1)−D−mU(ev1+u∗1)ev2+u∗2−v1−u∗1+μD1ev2(t−τf)+u∗2−v1−u∗1+σ1ξ1,dv2dt=−(D+D1)+γmU(ev1+u∗1)+σ2ξ2, | (4.19) |
where (u∗1,u∗2)=(lnS∗,lnx∗). The linearized system of (4.19) is
{dv1dt=a11v1+a12v2+a13v2(t−τf)+σ1ξ1,dv2dt=a21v1+σ2ξ2, | (4.20) |
where
a11=−DS0e−u∗1−mU′(eu∗1)eu∗2+mU(eu∗1)eu∗2−u∗1−μD1eu∗2−u∗1,a12=−mU(eu∗1)eu∗2−u∗1, a13=μD1eu∗2−u∗1, a21=γmU′(eu∗1)eu∗1. |
Given continuous function v(t) over the interval −T2≤t≤T2 (T>0), define function
˜v(ω)=∫T2−T2v(t)e−iωtdt. | (4.21) |
Since ˜v(ω) is the Fourier transform of v(t), we know
v(t)=12π∫∞−∞˜v(ω)eiωtdω, | (4.22) |
which implies that 12π˜v(ω) is the amplitude density of the components of v(t) in the angular frequency interval ω to ω+dω. Thus 12π˜v(ω)dω can be considered as an estimate of the amplitude of the component of v(t) with angular frequency ω. Taking Fourier transform of system (4.20), we obtain
{σ1˜ξ1(ω)=(−a11+iω)˜v1(ω)−(a12+a13e−iωτf)˜v2(ω),σ2˜ξ2(ω)=−a21˜v1(ω)+iω˜v2(ω), | (4.23) |
where
∫∞−∞ξj(t−τf)e−iωtdt=e−iωτf˜ξj(ω) j=1, 2. |
Write (4.23) in the matrix form
˜η(ω)=A(ω)˜v(ω), | (4.24) |
where
A(ω)=(−a11+iω−(a12+a13e−iωτf)−a21iω) |
˜η(ω)=(~η1(ω),~η2(ω))T=(σ1˜ξ1(ω),σ2˜ξ2(ω))T, ˜v(ω)=(˜v1(ω),˜v2(ω))T. Let |A(ω)|=det(A(ω)) and we assume |A(ω)|≠0. Then from (4.24) we have
˜v(ω)=A−1(ω)˜η(ω), | (4.25) |
where
A−1(ω)=1|A(ω)|(iωa12+a13e−iωτfa21−a11+iω)≜(a′11a′12a′21a′22). | (4.26) |
Then
˜vi=2∑j=1a′ij˜ηj, i=1,2. | (4.27) |
If the function v(t) has zero mean value then the fluctuation intensity (variance) of the components in the frequency band ω and ω+dω is Sv(ω)dω, where the spectral density Sv(ω) is formally defined by
Sv(ω)dω=limT→∞¯|v(ω)|2T. |
Hence,
Sη(ω)dω=limT→∞1T∫T2−T2∫T2−T2¯η(t)η(t′)exp(iω(t′−t))dtdt′. | (4.28) |
From (4.27) and (4.28), we have
Svi(ω)=2∑j=1|a′ij|2σjSξj(ω), i=1,2, | (4.29) |
since ¯ξi(t)=0 and ¯ξi(t)ξj(t′)=δijδ(t−t′). Therefore, the fluctuation intensity (variance) in vi is
σ2vi=12π∫∞−∞Svi(ω)dω=12π2∑j=1∫∞−∞|a′ij|2σjSξj(ω)dω=12π2∑j=1∫∞−∞|a′ij|2σjdω, i=1,2. | (4.30) |
It follows from (4.26) that
σ2vi=12π∫∞−∞Pi(ω)M(ω)dω, i=1,2, | (4.31) |
where
P1(ω)=σ1ω2+σ2(a12+a13cosωτf)2+σ2a213sin2ωτf,P2(ω)=σ1a221+σ2(a211+ω2),M(ω)=(−ω2+a12a21+a13a21cosωτf)2+(a11ω−a13a21sinωτf)2. | (4.32) |
When τf=0, we have
P1(ω)=σ1ω2+σ2(a12+a13)2,P2(ω)=σ1a221+σ2(a211+ω2),M(ω)=(−ω2+a12a21+a13a21)2+(a11ω)2=(−ω2+a12a21+a13a21+a11iω)×(−ω2+a12a21+a13a21−a11iω). | (4.33) |
Following Gradshteyn and Ryzhik [29], the general integral encountered in calculations of fluctuation is of the type
In=∫∞−∞gn(ω)dωhn(ω)hn(−ω), | (4.34) |
where
gn(ω)=b0ω2n−2+b1ω2n−4+⋯+bn−1,hn(ω)=a0ωn+a1ωn−1+⋯+an. | (4.35) |
When n=2, the integral is given by
I2=πi(a0b1−a2b0)a0a1a2. |
Then from the relation between g2(ω) and Pi(ω)(i=1,2) and h2(ω)h2(−ω) with M(ω) we obtain
a0=−1, a1=a11, a2=a12a21+a13a21, b0(1)=σ1, b0(2)=σ2, |
b1(1)=σ2(a12+a13)2, b1(2)=σ1a221+σ2a211. |
Hence,
σ2v1=σ2(a12+a13)2+σ1(a12a21+a13a21)2a11(a12a21+a13a21),σ2v2=σ1a221+σ2a211+σ2(a12a21+a13a21)2a11(a12a21+a13a21). | (4.36) |
Now, the variances have been obtained in (4.31) for τf>0 and in (4.36) for τf=0. But we should point out that the computed variance σ2vi is for the linear system (4.20), while for the nonlinear system (4.19), it is difficult to obtain. So, σ2vi in (4.31) is an estimate of the variance for system (4.19). Notice that v1=lnS−lnS∗ and v2=lnx−lnx∗, so the fluctuation intensities of the nutrient and the algae population can be estimated as (S∗)2σ2v1 and (x∗)2σ2v2, respectively. Thus, the results obtained so far in this subsection can be summarized as follow.
Theorem 4.2. Assume (1.3) holds. Then the fluctuation intensities of the nutrient and the algae population can be estimated as (S∗)2σ2v1 and (x∗)2σ2v2, respectively, where σ2vi is given in (4.31) for τf>0 and in (4.36) for τf=0.
The following theorem shows that a sufficiently large noise can make the algae population extinct exponentially with probability one.
Theorem 4.3. For any given initial value (S(0),x(0))∈R2+, the solution (S(t),x(t)) of system (2.1) has the following property:
lim supt→∞lnx(t)t≤−(D+D1)+γm−12σ22. |
In particular, if σ22≥2γm−2(D+D1), then
lim supt→∞lnx(t)t≤0. |
Proof. Define function V(x)=lnx. Then by Itô's formula, we have
dV(x)=1xdx−12x2(dx)2=[−(D+D1)+γmU(S)−12σ2]dt+σ2dB2. |
It follows that
dV(x)≤[−(D+D1)+γm−12σ22]dt+σ2dB2, |
integrating both sides of which from 0 to t yields
lnx(t)−lnφ2(θ)≤[−(D+D1)+γm−12σ22]t+M(t), | (4.37) |
where M(t)=σ2B2(t). Obviously, M(t) is a locally continuous martingale and
⟨M(t),M(t)⟩t=σ22t, |
which implies that
lim supt→∞⟨M(t),M(t)⟩tt<∞. |
By Strong Law of Large Numbers, we obtain
limt→∞M(t)t=0 a.s. |
Dividing t on the both sides of (4.37) and letting t→∞, we have
lim supt→∞lnx(t)t≤−(D+D1)+γm−12σ22. |
This completes the proof of the Theorem.
Theorem 4.3 reveals that a large noise may induce the extinction of algae population even though its corresponding deterministic model (1.1) has a stable positive equilibrium E∗.
In this paper, a delayed stochastic model (2.1) for algal bloom with nutrient recycling has been proposed and investigated. Such a stochastic model is useful in investigating and understanding various ecologically realistic features. We have focused our attention on two aspects:
(ⅰ) the random influences incorporated through perturbation on the nutrient and the algae population, and
(ⅱ) the conversion delay of detritus into nutrients.
For model (2.1), we first showed its reasonability by means of an approximate Markovian system of it under the assumption that the delay kernel f(s) takes the family of generic delay kernel (1.4). Then we carried out the analysis of the uniqueness and the global existence of its positive solution with the help of the result in [28], since the incorporated delay in the system is infinite. Next, we analyzed its long time behaviors around the various equilibria of its corresponding deterministic model (1.1). Our findings in Theorem 4.1 reveal that E0 is stochastically stable provided the intensity of the noise σ2 and the average delay Tf=nα are small. Though E∗ is not an equilibrium of model (2.1) when σi≠0, it is shown in Theorem 4.4 that the fluctuations intensities of the nutrient and the algae population can be estimated in a neighbourhood of E∗. In addition, our result in Theorem 4.4 reveals that sufficiently large noise can make the algae population extinct exponentially with probability one, even if its corresponding deterministic model (1.1) has a stable positive equilibrium.
To illustrate the theoretic results obtained, numerical simulations are carried out by using Milstein scheme [30]. Here we assume that the specific growth function U(S) is of Michaelis-Menten type
U(S)=Sa1+S, |
where a1 is the half-saturation constant.
The first given example below concerns the effects of the random influence and the delay on the long time behavior around the washout equilibrium. Here we take n=1 in (1.4), that is, the weak delay kernel f(s)=αe−αs. Then the discretization of model (2.1) for t=0,Δt,2Δt,…,nΔt takes the form
{Si+1=Si+[D(S0−Si)−mU(Si)xi+μD1y1,i]Δt++σ1Si√Δtξ1i,xi+1=xi+xi[−(D+D1)+γmU(Si)]Δt+σ2xi√Δtξ2i,y1,i+1=y1,i−α(y1,i−Si)Δt, | (5.1) |
where time increment Δt>0 and ξi are N(0,1)−distributed independent random variables which can be generated numerically by pseudorandom number generators.
Example 5.1. Consider model (2.1) with D=0.3,D1=0.1,S0=3,m=0.54,a1=0.4,μ=0.3,γ=0.8, σ1=0, σ2=0.05 and α=10.
It is easy to see that Tf=0.1. Simple computations show that
μd1+mU(S0)+C2Tf=0.5065<2D |
and
2γmU(S0)+σ2m2U2(S0)=0.7649<2(D+D1). |
Then by Theorem 4.1, the equilibrium E0 of model (2.1) is stochastically stable. Our simulation supports this result as shown in Figure 1. To examine the effect of the delay in nutrient recycling, increasing the value of Tf from 0.1 to 10 (i.e., decreasing the value of α from 10 to 0.1), we find that the equilibrium E0 continues to be stochastically stable, but the levels of the nutrient and the algae population will be lower in the beginning of time (see Figure 1). It is because that a large delay in nutrient recycling can make the algae grow slowly.
To study effects of the random influence and delay on E∗, we give the following example with different values of noise intensities and delay. For the kernel function f(s)=δ(s−τf), the discretization of model (2.1) for t=0,Δt,2Δt,…,nΔt takes the form
{Si+1=Si+[D(S0−Si)−mU(Si)xi+μD1xi(i−τf/Δt)]Δt++σ1Si√Δtξ1i,xi+1=xi+xi[−(D+D1)+γmU(Si)]Δt+σ2xi√Δtξ2i, | (5.2) |
where time increment Δt>0 and ξi are N(0,1)−distributed independent random variables which can be generated numerically by pseudorandom number generators.
Example 5.2. Consider model (2.1) with D=0.3, D1=0.1, S0=5, m=0.7, a1=0.4, μ=0.3 and γ=0.8.
It is easy to see example 5.2 satisfies condition (1.3). By Theorem 4.2, the fluctuation intensities of the nutrient and the algae population can be estimated as σ2S and σ2x, respectively, where σ2S=(S∗)2σ2v1 and σ2x=(x∗)2σ2v2. The variations of σ2S and σ2x with respect to the delay τf and noise intensities σi are drawn in Figure 2 for different τf and σi, and the corresponding trajectories of the nutrient and the algae around E∗ are drawn in Figure 3.
Figures 2(a) and (b) show the effect of the delay on σ2S and σ2x, which are drawn by taking σ1=0.01 and σ2=0.01. From Figure 2(a), σ2S has a little decrease with the increase of the delay in the beginning, then it increases as the delay increases and finally it remains unchanged when the delay increases to 30. From Figure 2(b), σ2x first decreases as the delay increases, then it remains unchanged when the delay increases to 30. These two sub-figures reveal that the delay has a different effect on σ2S and σ2x when it is small, while it does not affect the fluctuation intensities when it is large. The corresponding trajectories of the nutrient and the algae around E∗ are drawn by taking τf=0.1 and τf=10, respectively, please see Figure 3(a) and (b). These two sub-figures reveal that with a different value of τf, trajectories of the nutrient and the algae oscillate ultimately around E∗ with a certain amplitude, but with a larger value of τf, the levels of the nutrient and the algae will be lower in the beginning.
Figures 2(c) and (d) show the effect of noise intensities on σ2S and σ2x, which are drawn by taking τf=σ2=0 in Figure 2(c) and τf=σ1=0 in Figure 2(d). From these two sub-figures, the fluctuation intensities of both the nutrient and the algae increase as σ1 or σ2 increases, and σ2S increases faster in the absence of σ2 and σ2x grows faster in the absence of σ1. The corresponding trajectories of the nutrient and the algae around E∗ are drawn by taking two different values of σ1 (Figure 3(c)-(d)) and of σ2 (Figure 3(e)-(f)).
From Figures 2 and 3, one can find that the fluctuation intensities σ2S and σ2x are more sensitive to the noise intensities σi than the delay τf, since the fluctuation intensities change within a narrow range with respect to τf (Figures 2 and 3(a)-(b)), and they change within a wide range with respect to σ2S and σ2x (Figures 2(c)-(d) and 3(c)-(f)). Furthermore, the fluctuation intensity of the nutrient is more sensitive to σ1 than σ2, since in the absence of σ2, σ2S has a faster increase with the increase of σ1. Rather, the fluctuation intensity of the algae population is more sensitive to σ2.
To further study the effects of the random influence on the extinction of system (2.1), we give the following example with a sufficiently large noise. Here the discretization of model (2.1) takes the form as in (5.1).
Example 5.3. Consider model (2.1) with α=10 (i.e., Tf=0.1), σ1=0 and σ2=0.5 and all other parameters have the same values as in Example 5.2.
Simple computation shows that
2γm−2(D+D1)=0.064≤σ22=0.25. |
By Theorem 4.3, the algae population will go to extinction (see Figure 4).
Examples 5.2 and 5.3 reveal that the algae population is more sensitive to σ2. Under certain parametric conditions, there is fundamentally different behavior for the algae with different value of σ2. If σ2=0.01, the algae population will survive. Whereas, when σ2 increase to 0.5, the extinction of the algae will occur.
To sum up, this paper presents an investigation on the effect of the environmental noise and the delay occurred in the nutrient recycling on a nutrient-algae system. Our findings are useful for a better understanding of the dynamics of algal blooms. We should point out there are other interesting topics meriting further investigation, for example, the stationary distribution of the system. It is also very interesting to study the long time behavior of the multi-nutrient multi-algae system with noise. We leave these for future considerations.
Research is supported by the National Natural Science Foundation of China (11671260) and Shanghai Leading Academic Discipline Project (No. XTKX2012).
The authors declare no conflict of interest in this paper.
[1] | B. Qin, Z. Liu and K. Havens, Eutrophication of shallow lakes with special reference to Lake Taihu, China, Hydrobiologia, 2007. |
[2] | B. Qin, Lake Taihu, China, Springer, New York, 2008. |
[3] | B. Qin, G. Zhu and G. Gao, A drinking water crisis in Lake Taihu, China: linkage to climatic variability and lake management. Environ. Manage., 45 (2010), 105–112. |
[4] | L. Xu, J. Shen and D. Marinova, Changes of public environmental awareness in response to the Taihu blue- green algae bloom incident in China, Environ. Dev. Sustain., 15 (2013), 1281–1302. |
[5] | C. Qiuwen and E. Mynett, Modelling algal blooms in the Dutch coastal waters by integrated numerical and fuzzy cellular automata approaches, Ecol. Model., 199 (2006), 73–81. |
[6] | B. Wang, Cultural eutrophication in the Changjiang (Yangtze River) plume: history and perspective, Estuar. Coast. Shelf. S., 69 (2006), 474–477. |
[7] | W. He, J. Shang, X. Lu and C. Fan, Effects of sludge dredging on the prevention and control of algae-caused black bloom in Taihu Lake, China, J. Environ. Sci., 25 (2013), 430–440. |
[8] | E. Beretta, G. Bischi and F. Solimano, Stability in chemostat equations with delayed nutrient recycling, J. Math. Biol., 28 (1990), 99–111. |
[9] | X. He, S. Ruan and H. Xia, Global stability in chemostat-type plankton models with delayed nutrient recycling, J. Math. Biol., 37 (1998), 253–271. |
[10] | J. Caperon, Time lag in population growth response of isochrysis galbana to a variable nitrate environment, Ecology, 50 (1969), 188–192. |
[11] | A. Misra, P. Chandra and V. Raghavendra, Modeling the depletion of dissolved oxygen in a lake due to algal bloom: effect of time delay, Adv. Water Resour., 34 (2011), 1232–1238. |
[12] | A. Huppert, B. Blasius and R. Olinky, A model for seasonal phytoplankton blooms, J. Theor. Biol., 236 (2005), 276–290. |
[13] | S. Chen, X. Chen and Y. Peng, A mathematical model of the effect of nitrogen and phosphorus on the growth of blue-green algae population, Appl. Math. Model., 33 (2009), 1097–1106. |
[14] | S. Kunikane and M. Kaneko, Growth and nutrient uptake of green alga, Scenedesmus dimorphus, under a wide range of nitrogen/phosphorus ratio-II, Kinetic model, Water Res., 18 (1984), 1313–1326. |
[15] | D. Trolle, J. Elliott and W. Mooij, Advancing projections of phytoplankton responses to climate change through ensemble modelling, Environ. Modell. Softw., 61 (2014), 371–379. |
[16] | D. Anderson, J. Burkholder andW. Cochlan, Harmful algal blooms and eutrophication: examining of linkages from selected coastal regions of the United States, Harmful Algae, 8 (2008), 39–53. |
[17] | R. Sarkar, A stochastic model for autotroph-herbivore system with nutrient reclycing, Ecol. Model., 178 (2004), 429–440. |
[18] | M. Zhu, H. Paerl and G. Zhu, The role of tropical cyclones in stimulating cyanobacterial (Microcystis spp.) blooms in hypertrophic Lake Taihu, China, Harmful Algae, 39 (2014), 310–321. |
[19] | D Huang, H Wang and J. Feng, Modelling algal densities in harmful algal blooms (HAB) with stochastic dynamics, Appl. Math. Model., 32 (2008), 1318–1326. |
[20] | K. Das and N. Gazi, Random excitations in modelling of algal blooms in estuarine systems, Ecol. Model., 222 (2011), 2495–2501. |
[21] | P. Mandal, L. Allen and M. Banerjee, Stochastic modeling of phytoplankton allelopathy, Appl. Math. Model., 38 (2014), 1583–1596. |
[22] | L. Imhof and S. Walcher, Exclusion and persistence in deterministic and stochastic chemostat models, J. Differ. Equations, 217 (2005), 26–53. |
[23] | R. Durrett, Stochastic calculus: a practical introduction, CRC press, Boston, 1996. |
[24] | M. Frtihcan, Complex time-delay systems: theory and applications, Springer, 2010. |
[25] | X. Mao, Stochastic differential equations and applications, Elsevier, 2007. |
[26] | L. Arnold, Stochastic differential equations: theory and applications, New York, 1974. |
[27] | S. Mohammed, Stochastic functional differential equations, Boston Pitman, 1984. |
[28] | F. Wei and K. Wang, The existence and uniqueness of the solution for stochastic functional differential equations with infinite delay, J. Math. Anal. Appl., 331 (2007), 516–531. |
[29] | I. Gradshteyn and I. Ryzhik, Table of integrals, series and products, Academic press, New York, 1980. |
[30] | D. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. |
1. | Da Song, Meng Fan, Shihan Yan, Meng Liu, Dynamics of a nutrient-phytoplankton model with random phytoplankton mortality, 2020, 488, 00225193, 110119, 10.1016/j.jtbi.2019.110119 | |
2. | Yahong Peng, Yujing Li, Tonghua Zhang, Global bifurcation in a toxin producing phytoplankton–zooplankton system with prey-taxis, 2021, 61, 14681218, 103326, 10.1016/j.nonrwa.2021.103326 | |
3. | Qing Guo, Chuanjun Dai, Hengguo Yu, He Liu, Xiuxiu Sun, Jianbing Li, Min Zhao, Stability and bifurcation analysis of a nutrient‐phytoplankton model with time delay, 2020, 43, 0170-4214, 3018, 10.1002/mma.6098 | |
4. | Tingting Yu, Sanling Yuan, Dynamics of a stochastic turbidostat model with sampled and delayed measurements, 2023, 20, 1551-0018, 6215, 10.3934/mbe.2023268 | |
5. | 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, 2023, 20, 1551-0018, 15496, 10.3934/mbe.2023692 | |
6. | Bruce E. Kurtz, James E. Landmeyer, James K. Culter, Detection of periodic peaks in Karenia brevis concentration consistent with the time-delay logistic equation, 2024, 946, 00489697, 174061, 10.1016/j.scitotenv.2024.174061 | |
7. | Jyoti Maurya, A. K. Misra, Santo Banerjee, Bifurcation analysis of fish-algae-nutrient interactions in aquatic ecosystems, 2024, 0924-090X, 10.1007/s11071-024-10312-8 | |
8. | Da Song, Wentao Fu, Meng Fan, Bifurcation Analysis of a Stochastic Phytoplankton Growth Model Under Photoinhibition, 2025, 85, 0036-1399, 875, 10.1137/24M1684311 |