College of Life and Environmental Science, Wenzhou University, Wenzhou 325035, China
2.
School of Intelligent Manufacturing and Vehicle Engineering, Sichuan Institute of Industrial Technology, Deyang 618500, China
3.
Department of Basic Science and Humanities, Indian Institute of Information Technology, Bhagalpur 813210, India
4.
College of Mathematics and Physics, Wenzhou University, Wenzhou 325035, China
5.
Environmental Engineering Program, University of Northern British Columbia, Prince George V2N 4Z9, Canada
6.
WZU-UNBC Joint Research Institute of Ecology and Environment, Wenzhou University, Wenzhou 325035, China
7.
National & Local Joint Engineering Research Center for Ecological Treatment Technology of Urban Water Pollution, Wenzhou University, Wenzhou 325035, China
Academic Editor: Yang Kuang
Received:
06 May 2023
Revised:
03 July 2023
Accepted:
04 July 2023
Published:
26 July 2023
In this paper, we investigate a stochastic nutrient-plankton model with impulsive control of the nutrient concentration and zooplankton population. Analytically, we find that the population size is nonnegative for a sufficiently long time. We derive some sufficient conditions for the existence of stable periodic oscillations, which indicate that the plankton populations will behave periodically. The numerical results show that the plankton system experiences a transition from extinction to the coexistence of species due to the emergence of impulsive control. Additionally, we observe that the nutrient pulse has a stronger relationship with phytoplankton growth than the zooplankton pulse. Although the frequency of impulsive control and appropriate environmental fluctuations can promote the coexistence of plankton populations, an excessive intensity of noise can result in the collapse of the entire ecosystem. Our findings may provide some insights into the relationships among nutrients, phytoplankton and zooplankton in a stochastic environment.
Citation: 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[J]. Mathematical Biosciences and Engineering, 2023, 20(8): 15496-15523. doi: 10.3934/mbe.2023692
Related Papers:
[1]
Zhiwei Huang, Gang Huang .
Mathematical analysis on deterministic and stochastic lake ecosystem models. Mathematical Biosciences and Engineering, 2019, 16(5): 4723-4740.
doi: 10.3934/mbe.2019237
[2]
Zhihao Ke, Chaoqun Xu .
Structure analysis of the attracting sets for plankton models driven by bounded noises. Mathematical Biosciences and Engineering, 2023, 20(4): 6400-6421.
doi: 10.3934/mbe.2023277
[3]
Xuehui Ji, Sanling Yuan, Tonghua Zhang, Huaiping Zhu .
Stochastic modeling of algal bloom dynamics with delayed nutrient recycling. Mathematical Biosciences and Engineering, 2019, 16(1): 1-24.
doi: 10.3934/mbe.2019001
[4]
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
[5]
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
[6]
Yue Song, Yi Zhang, Song Yang, Na Li .
Investigation on stability and controller design for singular bio-economic systems with stochastic fluctuations. Mathematical Biosciences and Engineering, 2021, 18(3): 2991-3005.
doi: 10.3934/mbe.2021150
[7]
Fathalla A. Rihan, Hebatallah J. Alsakaji .
Analysis of a stochastic HBV infection model with delayed immune response. Mathematical Biosciences and Engineering, 2021, 18(5): 5194-5220.
doi: 10.3934/mbe.2021264
[8]
Yan Zhang, Shujing Gao, Shihua Chen .
Modelling and analysis of a stochastic nonautonomous predator-prey model with impulsive effects and nonlinear functional response. Mathematical Biosciences and Engineering, 2021, 18(2): 1485-1512.
doi: 10.3934/mbe.2021077
[9]
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
[10]
Yan Wang, Tingting Zhao, Jun Liu .
Viral dynamics of an HIV stochastic model with cell-to-cell infection, CTL immune response and distributed delays. Mathematical Biosciences and Engineering, 2019, 16(6): 7126-7154.
doi: 10.3934/mbe.2019358
Abstract
In this paper, we investigate a stochastic nutrient-plankton model with impulsive control of the nutrient concentration and zooplankton population. Analytically, we find that the population size is nonnegative for a sufficiently long time. We derive some sufficient conditions for the existence of stable periodic oscillations, which indicate that the plankton populations will behave periodically. The numerical results show that the plankton system experiences a transition from extinction to the coexistence of species due to the emergence of impulsive control. Additionally, we observe that the nutrient pulse has a stronger relationship with phytoplankton growth than the zooplankton pulse. Although the frequency of impulsive control and appropriate environmental fluctuations can promote the coexistence of plankton populations, an excessive intensity of noise can result in the collapse of the entire ecosystem. Our findings may provide some insights into the relationships among nutrients, phytoplankton and zooplankton in a stochastic environment.
1.
Introduction
Over the past several decades, researchers have shown an increased interest in plankton dynamics due to the numerous negative effects caused by algal blooms [1,2,3], which can threaten entire aquatic systems and result in economic losses, altering long-term species coexistence and influencing ecological stability. On the other hand, under conditions of nutrient shortage, the plankton community may suffer from nutrient deficiency and the aquatic system may evolve into species extinction. As important components of aquatic food webs, nutrients and the zooplankton population are widely recognized as the most important factors for phytoplankton growth [4]. Previous studies have established that a high nutrient load can be a major trigger for the increased frequency, intensity and duration of the plankton blooms [5,6]. Moreover, the experiment by Vanni [7] has demonstrated that even small changes in the zooplankton size can enormously affect the phytoplankton community. Although numerous experimental and theoretical studies have been devoted to the study of phytoplankton growth response to the changes in nutrients and the zooplankton population [8,9], the relationships between phytoplankton growth, nutrient concentration and the grazer zooplankton are still not well understood due to the fact that many events are pulsed and short-lived, such as nutrient delivery [10].
The aquatic ecosystems are inevitably subjected to instantaneous perturbations, leading to abrupt changes in nutrient concentration and zooplankton biomass. For example, heavy rainfall triggers increased nutrient input through river discharge, resulting in algal blooms in coastal areas [11]. These phenomena are commonly described as pulsed instantaneous behavior, also known as impulsive control, and they are characterized by the introduction or removal of some members from a population at a certain moment of time. Such pulsed instantaneous behavior cannot be considered continuously [12]. Therefore, mathematical models of ecological population dynamics have been developed in the study of plankton dynamics as they not only capture the ubiquitous stoichiometric constraints for the growth and interactions of species [13], but they also provide quantitative insights into population growth dynamics [14]. Actually, the theory of dynamical systems is generally recognized as two basic types: continuous and discontinuous [15]. In the natural world, many systems are characterized by the fact that they are subjected to pulse perturbations, causing discontinuous behaviors. Thus, impulsive differential equations have drawn increasing attention due to the advantage that the theory of impulsive differential equations can accurately capture the instantaneous state of a system. Based on the previous study of impulsive differential equations [16,17,18], in recent years, impulsive differential equations have been widely used to describe such instantaneous perturbations in the study of plankton dynamics [19,20,21]. Moreover, modelling approaches may have important implications for how plankton population growth responds to pulse perturbations of the nutrients and zooplankton population.
Impulsive control refers to the application of sudden and short-duration interventions or perturbations to a system. Impulsive control studies can help researchers to gain insight into how pulse perturbations influence the plankton dynamics. Additionally, studying impulsive control in plankton dynamics allows researchers to explore potential control strategies to achieve desired outcomes in terms of regulating plankton populations or managing ecological processes. In recent years, plankton dynamics induced by nutrient and zooplankton pulses have been reported in many aquatic food chain models [9,22,23]. The experimental study conducted by Spatharis et al. [24] emphasized that a sudden pulse of high nutrient concentration may promote the dominance of a single species and yield a negative effect on the aquatic ecological diversity. However, Cottingham et al. [25] demonstrated that the phytoplankton show low sensitivity to nutrient pulses in systems with a high density of large zooplankton [26]. Furthermore, Zhao et al. [27] reported that both phytoplankton and zooplankton can coexist if the amount of the released zooplankton is within certain ranges. Their results illustrated that both a nutrient pulse and zooplankton pulse may alter the dynamical behavior of an ecosystem, and that some proper impulsive control strategies can maintain the balance of the ecosystem [12,22,28]. Despite all of our knowledge on the relationships between phytoplankton, nutrient and zooplankton, there is limited research on whether phytoplankton growth is more sensitive to a nutrient pulse or zooplankton pulse.
To the best of our knowledge, lots of studies on plankton dynamics are conducted in a stable environment [29,30,31]. However, numerous studies support that almost all natural aquatic ecosystems are often subject to environmental fluctuations, such as variation in temperature, light and water depth. Environmental fluctuations have the potential to disturb the steady state through direct or indirect effects on density or parameter values [32]. Specifically, Mao et al. [33] developed a simple stochastic population growth model, and reported that the stochastic noise has the capability to suppress the explosion of population. The findings of Deng et al. [34] demonstrated that noise can suppress/enhance the exponential growth of populations. These results revealed that environmental fluctuations can significantly affect the dynamical behavior of biological populations. Furthermore, algal blooms may exhibit stochastic scales and timing due to the fact that the growth responses of plankton are inevitably affected by random fluctuations [35,36]. In recent years, several attempts have been made to investigate how environmental noise affects ecosystems [37,38]. For instance, Yu et al. [39] demonstrated that the Markov chain is beneficial for the survival of plankton. Biologically, it is more realistic to consider environmental noise in the species dynamical model.
The interaction between nutrient pulses and phytoplankton growth under stochastic disturbance has been investigated in [40]. However, far too little attention has been paid to the synergistic effects of a nutrient pulse and zooplankton pulse on the plankton dynamics under environmental fluctuation. Understanding how a phytoplankton population responds to nutrient pulses and zooplankton pulses under stochastic fluctuations may provide insights into the complex interactions among phytoplankton, nutrients and zooplankton. In addition, nutrient pulse control and phytoplankton pulse control are common measures to control phytoplankton blooms. Thus, investigating the interacting dynamics of nutrient pulses, zooplankton pulses and phytoplankton population may allow us to develop better strategies to mitigate eutrophication and its associated bloom events. In order to provide a better understanding of how phytoplankton respond to a nutrient pulse and zooplankton pulse in a random environment, in the present paper, an aquatic food chain model comprised of nutrient, phytoplankton and zooplankton species is investigated under the conditions of a random environment. In the model, nutrient pulses and zooplankton pulses are adopted to respectively describe the instantaneous changes in nutrient concentration and zooplankton biomass. Here, we aim to explore whether phytoplankton growth is more affected by pulsed nutrients or zooplankton under environmental fluctuations.
The rest of this article is organized in the following way. A mathematical model for the nutrient-plankton model of the aquatic habitat with impulsive control and stochastic disturbance is introduced in next section. In Section 3, the stochastic-impulsive coupled model is analyzed analytically. Some numerical results are presented in Section 4 to deeply explore the effects of impulsive control and stochasticity on the interplay of nutrients and planktons in the aquatic habitat. Section 5 covers a discussion of this study and biologically implications of the obtained results. At the end, we close the paper with the conclusion in Section 6.
2.
The mathematical model
Numerous mathematical and ecological studies have been conducted to explore the growth dynamics of plankton populations in the aquatic ecosystems [41,42,43,44]. By considering the nutrients, phytoplankton and zooplankton as dynamical variables, we develop here a mathematical model to comprehend the responses of plankton populations to the pulsed nutrients and zooplankton in a fluctuating environment. In a real aquatic system, the predation of phytoplankton by zooplankton is very complex, e.g., copepods [45]. Thus, we adopt the Holling type Ⅱ functional response to model the hunting behavior of zooplankton for phytoplankton. Moreover, some species of phytoplankton can produce toxins, which contribute to the death of the zooplankton population. Both the theoretical and experimental studies have proven that the toxin-producing phytoplankton can greatly influence the occurrence as well as termination of the plankton blooms [46]. Indeed, the liberation of toxins by phytoplankton can reduce the growth of zooplankton species and substantially contribute to their mortality. Actually, toxin-producing phytoplankton are not easily accessible by the zooplankton. Hence, the Holling type Ⅱ functional form is a more common and intuitively obvious choice to describe the grazing phenomenon of phytoplankton by zooplankton [46]. Moreover, we consider the intraspecific competition among the phytoplankton communities and the zooplankton communities due to the limitation of the resources in the aquatic biome. At any time t>0, let N(t), P(t) and Z(t) respectively represent the concentration of nutrients, phytoplankton population and zooplankton population in a considered aquatic system. Based on the following ecological assumptions, we will formulate our mathematical model for the nutrient-plankton system.
(1) The nutrient concentration in the aquatic system is determined by its washout and uptake by phytoplankton population at the rates b and n, respectively.
(2) The growth of phytoplankton depends on the nutrient availability and the predation by zooplankton.
(3) We represent the biomass conversion rate and the death rate of the phytoplankton population by β and m, respectively. The zooplankton predates the phytoplankton at a rate followed by the Holling type Ⅱ functional form [47], which is given by cPZh+P, where c denotes the phytoplankton capture rate for zooplankton and h is the half-saturation constant for such an uptake.
(4) We denote the biomass conversion rate and the natural mortality of zooplankton by d and k, respectively. The effect of toxin-producing phytoplankton on the zooplankton population is modeled by the term ρPZh+P, where ρ is the rate by which the phytoplankton population releases the toxic chemicals.
(5) Evidently, intraspecific competition exists in a real natural aquatic ecosystem due to the limitation of resources [48]. Here, we also model the intraspecific competition among the phytoplankton communities and the zooplankton communities and denote the strengths of competition by r and v, respectively.
(6) The natural ecosystems are affected by environmental fluctuations [49], so we use white noise to represent their impact on the nutrient-plankton model.
By considering the aforementioned assumptions, we come up with the following mathematical model that involves stochastic fluctuation:
In model (2.1), Bi(t) (i=1,2,3) denotes the independent standard Brownian motions and σ2i denotes the intensities of the white noises. We assume that all of the parameters involved in model (2.1) are positive, bounded and continuous θ-periodic functions in [0,+∞). In Table 1, we provide the biological meanings of variables and parameters describing model (2.1) in a constant environment.
Table 1.
Biological explanations of variables/parameters in model (2.1) and the numerical values of parameters used for simulation results.
The aquatic ecosystems are inevitably affected by instantaneous perturbations, which leads to abrupt changes in the nutrient concentration and zooplankton biomass, so we adopt nutrient and zooplankton pulses in model (2.1) and obtain the following implusive-stochastic differential equations:
Note that model (2.2) shares common biological meanings with model (2.1). Further, there exists a sequence of real numbers with 0<t1<t2<⋯<tk<⋯, limt→∞tk=+∞. For biological reasons, we impose the following restriction on αik(i=1,2):
1+αik>0,i=1,2,k=1,2,⋯.
For αik>0(i=1,2), the impulsive effects represent the process of species introductions whereas αik<0(i=1,2) denotes harvesting. Here, we only consider αik>0(i=1,2). Following the approach of [19,50,51], we assume that there exists a positive integer p such that tk+p=tk+T and αi(k+p)=αik,i=1,2,k∈Z+. We also consider that [0,T)∩{tk,k∈Z+}={t1,t2,⋯,tp}.
3.
The main results
We assume that (Ω,F,{Ft}t≥0,P) is a complete probability space with a filtration Ft(t≥0) satisfying the usual normal conditions i.e., it is right continuous and F0 contains all of the P-null sets.
3.1. Preliminaries
Assume that a product equals unity if the number of factors is zero. For a bounded and continuous function f:[0,+∞)→R+, we define
fu=supt≥0f(t),fl=inft≥0f(t).
Moreover, if the function f is integrable on [0,+∞)→R+, then
⟨f⟩=1t∫t0f(s)ds,⟨f⟩T=1T∫T0f(s)ds.
Now, we present the following definitions and lemmas in order to facilitate our further discussions. Consider the n-dimensional stochastic differential equation
dx(t)=f(x(t),t)dt+g(x(t),t)dB(t),t≥t0
(3.1)
with the initial value x(t0)=x0∈Rn, where B(t) is an n-dimensional standard Brownian motion. The differential operator L associated with Eq (3.1) is defined as
where X(0)∈Rn. A stochastic process X(t)=(X1(t),X2(t),⋯,Xn(t))T,t∈[0,∞) is said to be a solution of model (3.3) if the following conditions hold:
(i) X(t) is Ft adapted and continuous on the intervals (0,t1) and (tk,tk+1)∈R+,k∈N; F(t,X(t))∈L1(R+;Rn) and G(t,X(t))∈L2(R+;Rn), where Lk(R+;Rn) is the set of all Rn valued measurable Ft-adapted processes f(t) satisfying ∫T0∣f(t)∣kdt<∞ a.s. for every T.
(ii) For any tk,k∈N, X(t+k)=limt→t+kX(t) exists and X(t−k)=limt→t−kX(t) with a probability of one.
(iii) For almost every t∈(0,t1), X(t) satisfies the following integral equation:
X(t)=X(0)+∫t0F(s,X(s))ds+∫t0G(s,X(s))dB(s).
Further, for all t=[tk,tk+1],k∈N,X(t) satisfies the following integral equation:
X(t)=X(t+k)+∫ttkF(s,X(s))ds+∫ttkG(s,X(s))dB(s).
Lemma 3.1.[53]. For any u>0, the following inequality holds:
u≤2(u+1−logu)−(4−2log2).
Definition 3.2[54,55]. A stochastic process ξ(t)=ξ(t,ω),(−∞<t<+∞) is said to be periodic with period T if for every finite sequence of numbers t1,t2,⋯,tn, the joint distribution of random variables ξ(t1+h),ξ(t2+h),⋯,ξ(tn+h) is independent of h, where h=kT,k=±1,±2,⋯.
Remark 1.[56] A Markov process ξ(t) is T-periodic if and only if its transition probability function is T-periodic and the function P0(t,A)=P{ξ(t)∈A} satisfies the following equation:
P0(s,A)=∫RlP0(s,dx)P(s,x,s+T,A)≡p0(s+T,A)
for every A∈B.
For the following Itˆo differential equation:
dx(t)=b(t,x(t))dt+σ(t,x(t))dB(t),
(3.4)
let all of the coefficients be T-periodic in t and satisfy the linear growth condition and the Lipschitz condition in every cylinder Ul×R+ for l>0, where Ul={x:‖x‖≤l}. If there exists a function v=v(t,x) which is twice continuously differentiable with respect to x and once continuously differentiable with respect to t in Rn×R+, and if T is periodic in t and satisfies the following conditions:
inf‖x‖>Rv(t,x)→+∞asl→+∞,
(3.5)
Lv(t,x)≤−1outsidesomecompactset,
(3.6)
then there exists a solution of model (3.4) which is a T-periodic Markovian process [56].
3.2. Existence and uniqueness of the global positive solution
From the biological perspective, the population size should be nonnegative. Following the approach of [40], we will show that model (2.2) has a unique global positive solution. First, we consider the model without impulses as
According to [54], A1(t) and A2(t) are periodic functions with periodic T. Thus, we obtain the following lemma.
Lemma 3.2.Let N(t)=A1(t)y1(t),P(t)=y2(t) and Z(t)=A2(t)y3(t).
(1) If (N(t),P(t),Z(t)) is the solution of model (2.2), then (y1(t),y2(t),y3(t)) is the solution of model (3.7).
(2) If (y1(t),y2(t),y3(t)) is the solution of model (3.7), then (N(t),P(t),Z(t)) is the solution of model (2.2).
The proof easily follows from [52], hence, it is omitted. Now, we state the following theorem.
Theorem 3.1.For any initial value (N(0),P(0),Z(0))∈R3+, model (2.2) exhibits a unique positive solution (N(t),P(t),Z(t)), which will remain in R3+ with a probability of one.
Proof. Obviously, for any fixed t≥0, there exists a nonnegative integer n∈{0,1,2,⋯} such that nT≤t≤(n+1)T. In view of the equalities tk+p=tk+T and αk+p=αk, one can get
tk+np=tt+(n−1)p+T=⋯=tk+nT,αk+np=αk+(n−1)p=⋯=αk.
Owing to [0,∞)∩{tk,k∈N}={t1,t2,⋯,tp}, there exists a positive integer r∈{1,2,⋯,p} such that
Now, we will prove that model (3.7) has a unique globally positive solution (y1(t),y2(t),y3(t)). Obviously, the coefficients of model (3.7) satisfy the local Lipschitz condition. Thus, there exists a unique local solution on [0,τe), where τe is the explosion time. This solution will be global if τe=∞ almost surely. We choose γ0>1 sufficiently large so that y1(0),y2(0) and y3(0) all lie in the interval [1γ0,γ0]. For any positive integer γ(γ≥γ0), we define the stopping time as follows:
We set inf∅=∞. Clearly, τγ increases as γ→∞. Let τ∞=limγ→+∞τγ. Obviously, if τ∞=∞, then τ∞≤τe almost surely, τe=∞ and (y1(t),y2(2),y3(t))∈R3+ almost surely.
where IΩγ is the indicator function of Ωγ. Let γ→∞; then, limγ→∞P{τγ≤T}=0 implying that P{τγ≤T}=0. T>0 is arbitrary, so P{τ∞<∞}=0. This implies that P{τ∞=∞}=1, as required. Thus, one can conclude that model (3.7) has a unique positive solution (y1(t),y2(t),y3(t)) for all t≥0 almost surely.
Let N(t)=A1(t)y1(t),P(t)=y2(t) and Z(t)=A2(t)y3(t). From Lemma 3.2, it follows that (N(t),P(t),Z(t)) is the solution of model (2.2). Apparently, N(t),P(t) and Z(t) are continuous on the intervals (0,t1) and (tk,tk+1), k∈N. For t≠tk, we have
with the initial value (u(0),v(0),w(0))=(logN0,logP0,logZ0). Obviously, the coefficients of model (3.9) satisfy the local Lipschitz condition. Thus, by the theory of stochastic differential equation, model (3.9) has a unique solution (u(0),v(0),w(0)). In view of the Itˆo formula, (N(t),P(t),Z(t))=(eu(t),ev(t),ew(t)) is the unique solution of model (3.8).
Now, for t∈(t1,t2], model (2.2) can be rewritten as follows:
Using the same argument as above, one can show that model (3.10) admits a unique solution for t∈(t1,t2]. By the analogy, we can say that model (2.2) has a unique solution with the initial value (N(0),P(0),Z(0)) for each interval (tk,tk+1]. □
One can conclude from the above theorem that for model (2.2), the population size will remain nonnegative for sufficiently large values of t.
3.3. Existence of positive T-periodic solution
Here, we derive conditions for the existence and uniqueness of a positive periodic solution of model (2.2).
Lemma 3.3.[39]. For any initial value (y1(0),y2(0),y3(0))∈R3+, the solution (y1(t),y2(t),y3(t)) of model (3.7) has the following properties:
Theorem 3.2.If ϕ1, ϕ2, ϕ3>0, then model (2.2) has a positive T-periodic solution.
Proof. In view of Theorem 3.1, it suffices to show that a periodic solution exists for the equivalent model (3.7) without impulses. Thus, we only need to verify conditions (3.5) and (3.6).
From Eqs (3.12) and (3.13), we can take ϵ sufficiently small such that
LV≤−1,for(y1,y2,y3)∈R3+∖U,
where U=[ϵ,1ϵ]×[ϵ,1ϵ]×[ϵ,1ϵ]. Therefore, for the initial values y1(0)>0, y2(0)>0 and y3(0)>0, the solution y(t)=(y1(t),y2(t),y3(t)) of model (3.7) is a positive T-periodic Markov process. Based on the previous analysis, one can get that
where A1(t) and A2(t) are T-periodic functions. Thus, one can say that (N(t),P(t),Z(t)) is a positive T-periodic solution of model (2.2). □
From a biological point of view, the above theorem tells that model (2.2) exhibits a positive T-periodic solution for an appropriate noise intensity, and that the phytoplankton population will oscillate periodically.
4.
Numerical simulations
In order to have better insight into how impulsive control and environmental fluctuation affects the distributions of plankton populations in aquatic ecosystems, we present here some numerical results. To solve the stochastic system, we have used the Euler-Maruyama method whereas the Runge-Kutta method is employed to solve the deterministic system. If not specified in the text, we adopt the values of the parameters as in Table 1. To keep the model tractable, we set a common period of 40 d−1 for all parameters in model (2.2). Further, to capture the impact of periodic environment, we set the following forms of a model parameters:
Here, we mainly focus on the impulsive control and the noise intensity as the key factors in controlling the plankton blooms.
4.1. Effects of impulsive control
In real aquatic ecosystems, the oscillatory behavior of population density is observed due to the existence of ubiquitous pulse perturbations and stochastic fluctuation. For instance, industrial wastewater discharge can lead to rapid increment in the nutrient concentrations in water bodies, which produces transient effects on the state of aquatic ecosystems and may trigger the rapid growth of algae. In recent years, the importance of impulsive control has been widely recognized by researchers worldwide. Here, we present the effects of impulsive control on the plankton populations. The paths of nutrients, phytoplankton and zooplankton are depicted in Figure 1 for a fixed noise intensity, pulse intensity and number of pulses per period (σ1=σ2=σ3=0.02+0.01sin(πt/20), α1=exp(2),α2=exp(2), p=3). One can see in the figure that without impulsive fluctuation, both nutrient and plankton populations tend to extinct. However, when the impulsive control is applied on the stochastic model, nutrient and plankton populations can coexist. Obviously, the results displayed by Figure 1 are induced by impulsive control, indicating that impulsive control has the potential to alter population dynamics and may promote the coexistence of plankton populations.
Figure 1.
Time evolutions of (a) nutrients, (b) phytoplankton and (c) zooplankton in models (2.1) and (2.2).
In order to investigate how plankton populations respond to the combined effects of nutrient and zooplankton pulses, in Figure 2, we present the mean for plankton biomasses in the (α1,α2) plane. Apparently, for relatively small nutrient pulses, the phytoplankton population disappears from the aquatic system even if a high pulse of zooplankton is applied (see Figure 2(c)). One possible explanation for this behavior of phytoplankton might be the deficiency of nutrients in the aquatic system. Figure 2 depicts that an increment in the nutrient pulse may cause a significant increase in the mean of plankton biomasses indicating that as the nutrient pulse increases, so does the oscillatory behavior of plankton populations. The variance and the third order central moment of plankton populations are presented in Figure 3 by taking three different sets of pulse intensity. Apparently, the plankton populations disappear from the aquatic system if there is no nutrient pulse. One can also see that the variance of plankton biomasses show overall an increasing trend as the nutrient pulse increases, but there is a sharp increase in the skewness of the phytoplankton population. These results indicate that the nutrient pulse has the potential to alter the survival probability of plankton populations in the aquatic system. Next, we set the nutrient pulse as α1=3 and show the variance and the third order central moment of plankton populations by differing the zooplankton impulse; see Figure 3(c)–(d). Notably, the variance and the skewness of the phytoplankton population is unlikely to change significantly with a rise in the zooplankton impulse. Thus, we can say that the nutrient concentration in the aquatic reservoir is an important driver in schemes to control phytoplankton growth. Here, our findings concur with the results reported by the authors of [57]. Specifically, in comparison to the impulsive control of the zooplankton population, the phytoplankton growth has a stronger relationship with the nutrient pulse.
Figure 2.
Results for model (2.2) with σ1=σ2=σ3=0.05+0.01sin(πt/20). (a) The mean for phytoplankton biomass in the (α1,α2) plane; (b) the mean for zooplankton biomass in the (α1,α2) plane; (c) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=0.1,α2=8; (d) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=8,α2=0.1; (e) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=5,α2=5.
Figure 3.
Results for σ1=σ2=σ3=0.05+0.01sin(πt/20). (a) The variance and the third-order central moment of phytoplankton population and (b) zooplankton population versus different nutrient pulses with α2=4; (c) the variance and the third order central moment of the phytoplankton population and (d) zooplankton population versus different zooplankton pulses with α1=3.
Further, the frequency of pulses of plankton populations is also expected to rise due to the number of pulses per period, which may play an important role in stabilizing the plankton systems [19]. Figure 4 shows the mean and the variance of plankton biomasses with respect to the frequency of pulses. We observed that plankton populations tend to extinction when the number of pulses per period is small. This might happen because a small number of pulses per period can lessen the concentration of nutrients in the aquatic ecosystem, which may not support the survival of plankton populations. However, an increment in the number of pulses may induce a significant increase in the mean and the variance of plankton biomasses indicating that the densities of plankton populations show stronger oscillatory behavior. These results reveal that an appropriate frequency of pulses can contribute to the survival of plankton populations in the aquatic ecosystem.
Figure 4.
Results for model (2.2) with σ1=σ2=σ3=0.05+0.01sin(πt/20) and α1=α2=exp(1), showing the mean and the variance of phytoplankton biomass and zooplankton biomass with respect to number of pulses per period. The path of plankton populations indicates the changes of plankton biomasses for p=2 and p=5, respectively.
It is well known that all populations show substantial stochasticity in their ecological dynamics [37]. Importantly, the random fluctuation plays a key role in determining the survivorship of plankton populations in aquatic systems [39]. But, comparatively little attention has been given to the effect of environmental fluctuation on an aquatic system with pulse perturbations. As one can see in Figure 5, the means of nutrient concentration and plankton biomass are less affected by stochastic perturbations, but environmental noise can significantly affect the variance of nutrient concentration and phytoplankton biomass. It means that the presence of environmental noise can promote the oscillatory behavior of nutrients and phytoplankton population.
Figure 5.
Results for model (2.2) with σ1=σ2=σ3=0.1+0.01sin(πt/20) and α1=α2=exp(2). (a) The mean and the variance of phytoplankton and zooplankton biomasses; evolution of a single path of (b) nutrient concentration, (c) phytoplankton population, (e) zooplankton population.
Studies have shown that an excessive noise intensity can potentially affect the species richness and completely destroy the plankton system of aquatic reservoirs [8]. For example, a dramatic decrease in the temperature can cause the complete disappearance of the Cylindrospermopsis filaments [58]. Figure 6, depicts that for a smaller range of noise intensity, the amplitude of oscillations in the plankton densities increases as the noise intensity increases. But, when the noise intensity is high, the nutrient concentration becomes zero while the plankton populations disappear from the aquatic system. Similar results were reported by the authors of [44]. These findings indicate that the control of environmental noise can have a dramatic role in the survival of both phytoplankton and zooplankton populations in aquatic ecosystems.
Figure 6.
Time evolutions of nutrients (first column), phytoplankton (second column) and zooplankton (third column) for model (2.2) and the corresponding deterministic model. Here, we have chosen α1=α2=exp(2) and p=3; σ1=σ2=σ3=0.2+0.01sin(πt/20) in (a−c) and σ1=σ2=σ3=0.5+0.01sin(πt/20) in (d−f).
Over the past few decades, the control of algal blooms has received considerable attention among research scientists. There has been an increasing focus on the roles of nutrients and zooplankton in the growth of phytoplankton [59,60,61]. Due to the advantage of impulsive differential equations, modeling approaches have become a widely used tool to enhance our understanding of pulsed instantaneous behaviors that exist in aquatic ecosystems. Here, we have investigated a nutrient-plankton model with impulsive control and stochastic disturbance. Our theoretical results showed the existence of a global positive solution for the considered model. Moreover, we found that the amplitudes of oscillations in the plankton density exhibit a more or less periodic nature under certain conditions.
Numerically, we saw the combined effects of nutrient and zooplankton pulses on the plankton population growth; see Figure 2. We observed that nutrient pulses have a more significant and enduring impact on the phytoplankton biomass compared to the zooplankton pulse, and that plankton populations will tend to extinction when the nutrient pulse and zooplankton pulse are low. Importantly, when the nutrient pulse is low, the phytoplankton population has become extinct in the aquatic system due to a lack of nutrients, as shown in Figure 3(a). This situation may arise in the aquatic systems because an inadequate nutrient pulse leads to insufficient nutrient concentration for the uptake of phytoplankton. However, as the nutrient pulse increases, the variance of phytoplankton and zooplankton populations exhibit an overall increasing trend, indicating stronger oscillatory behavior. One possible explanation is that an increment in the phytoplankton biomass due to a higher nutrient pulse may provide sufficient food for the grazer zooplankton [5]. We also noted that the density of the phytoplankton population does not change much in response to changing the pulse intensity of zooplankton even if a high pulse of zooplankton is applied, as shown in Figure 3(c). These results further emphasize that the nutrient pulse is strongly related to phytoplankton growth [62]. Overall, Figure 3 shows a positive association between the nutrient pulse and the amplitudes of oscillations in the densities of plankton populations. Notably, the zooplankton pulse does not significantly affect the phytoplankton population. These results suggest that the growth of phytoplankton is directly correlated with the concentration of nutrients in the aquatic system.
Apart from the impulsive intensity, the frequency of impulsive control also has a direct role in regulating plankton dynamics [14]. Our simulation results showed a significant positive correlation between the number of pulses per period and the intensities of oscillations in the densities of plankton populations, as shown in Figure 4. Biologically, Figure 4 tells that a low number of pulses per period leads to stagnation in the growth of phytoplankton as the nutrient concentration declines in the aquatic system [63]. However, a high frequency of impulsive control leads to an increased input of nutrients into aquatic ecosystems, which can support the survivorship of the plankton populations. Thus, one can say that an increasing number of pulses per period may contribute to the survival of plankton populations in an aquatic habitat.
Environmental disturbances could not be neglected in the process of phytoplankton growth as environmental noise has the potential to alter population dynamics [37]. For instance, in [58], it was experimentally demonstrated that a dramatic drop in the temperature can promote the disappearance of the Cylindrospermopsis filaments. Therefore, we present Figures 5 and 6 to show the effects of stochastic fluctuations on the impulsive control. It follows from Figure 5 that the noise intensity has a significant effect on the variance of nutrients and phytoplankton, which implies that oscillations in the densities of nutrient concentration and phytoplankton population have a positive relationship with the environmental fluctuations. However, the results in Figure 6 suggest that the entire plankton population may collapse in response to increased noise intensity. This finding is in line with the results reported in [58]. It suggests that the environmental fluctuations are likely to influence the plankton dynamics whereas the noise control could be of practical significance for the mitigation of planktonic growth. From the results obtained here, one can conclude that nutrient pulses may be more likely to affect phytoplankton dynamics, and that the environmental fluctuations could have larger impacts on aquatic ecosystems.
6.
Conclusions
In this paper, a stochastic-impulsive coupled nutrient-plankton model was investigated. Our theoretical analysis showed that model (2.2) has a unique globally positive solution. Some sufficient conditions have been derived for the existence of a positive T-periodic solution. Numerical results revealed that due to the existence of impulsive control, the plankton dynamics experienced a transition from extinction to species coexistence. The results indicated that the nutrient pulse is more effective than the zooplankton pulse in controlling the phytoplankton growth. We observed that an increased impulse frequency and an appropriate noise intensity can increase the amplitudes of oscillations in the plankton density whereas an excessive noise intensity may cause a collapse of the entire aquatic ecosystem. It is worth noting that our results may not predict the specific events within a particular lake/ocean, but it contributes to the studies on how the coupling between stochasticity and impulse influences the growth of phytoplankton species. Overall, the findings of this study may provide some insights into the research on the relationships between the phytoplankton population, nutrients and zooplanktonic predators in a stochastic environment.
Use of AI tools declaration
The authors declare that they have not used artificial intelligence tools in the creation of this article.
Acknowledgments
The authors are grateful to the anonymous referees for their careful reading, valuable comments and helpful suggestions, which have contributed to improve the presentation of this work significantly. This work was supported by the National Key Research and Development Program of China (Grant No. 2018YFE0103700) and the National Natural Science Foundation of China (Grants No. 61901303, 12202319).
Conflict of interest
The author declares that there are no conflicts of interest.
References
[1]
D. M. Anderson, Prevention, control and mitigation of harmful algal blooms: Multiple approaches to HAB management, Harmful Algae Manage. Mitigation, 2004 (2004), 123–130. https://doi.org/10.1007/s10311-022-01457-2 doi: 10.1007/s10311-022-01457-2
[2]
B. Balaji-Prasath, Y. Wang, Y. P. Su, D. P. Hamilton, H. Lin, L. Zheng, et al., Methods to control harmful algal blooms: A review, Environ. Chem. Lett., 20 (2022), 3133–3152. https://doi.org/10.1007/s10311-022-01457-2 doi: 10.1007/s10311-022-01457-2
[3]
J. J. Gallardo-Rodríguez, A. Astuya-Villalón, A. Llanos-Rivera, V. Avello-Fontalba, V. Ulloa-Jofré, A critical review on control methods for harmful algal blooms, Rev. Aquacult., 11 (2019), 661–684. https://doi.org/10.1111/raq.12251 doi: 10.1111/raq.12251
[4]
D. M. Anderson, Approaches to monitoring, control and management of harmful algal blooms (HABs), Ocean Coastal Manage., 52 (2009), 342–347. https://doi.org/10.1016/j.ocecoaman.2009.04.006 doi: 10.1016/j.ocecoaman.2009.04.006
[5]
A. Burson, M. Stomp, E. Greenwell, J. Grosse, J. Huisman, Competition for nutrients and light: testing advances in resource competition with a natural phytoplankton community, Ecology, 99 (2018), 1108–1118. https://doi.org/10.1002/ecy.2187 doi: 10.1002/ecy.2187
[6]
D. W. Schindler, R. E. Hecky, D. L. Findlay, M. P. Stainton, B. R. Parker, M. J. Paterson, et al., Eutrophication of lakes cannot be controlled by reducing nitrogen input: Results of a 37-year whole-ecosystem experiment, Proc. Natl. Acad. Sci., 105 (2008), 11254–11258. https://doi.org/10.1073/pnas.0805108105 doi: 10.1073/pnas.0805108105
[7]
M. J. Vanni, Effects of nutrients and zooplankton size on the structure of a phytoplankton community, Ecology, 68 (1987), 624–635. https://doi.org/10.2307/1938467 doi: 10.2307/1938467
[8]
X. H. Ji, S. L. Yuan, T. H. Zhang, H. P. Zhu, Stochastic modeling of algal bloom dynamics with delayed nutrient recycling, Math. Biosci. Eng., 16 (2019), 1–24. https://doi.org/10.3934/mbe.2019001 doi: 10.3934/mbe.2019001
[9]
G. D. Liu, X. Z. Meng, S. Y. Liu, Dynamics for a tritrophic impulsive periodic plankton–fish system with diffusion in lakes, Math. Methods Appl. Sci., 44 (2021), 3260–3279. https://doi.org/10.1002/mma.6938 doi: 10.1002/mma.6938
[10]
P. M. Glibert, V. Kelly, J. Alexander, L. A. Codispoti, W. C. Boicourt, T. M. Trice, et al., In situ nutrient monitoring: A tool for capturing nutrient variability and the antecedent conditions that support algal blooms, Harmful Algae, 8 (2008), 175–181. https://doi.org/10.1016/j.hal.2008.08.013 doi: 10.1016/j.hal.2008.08.013
[11]
H. W. Han, R. S. Xiao, G. D. Gao, B. S. Yin, S. K. Liang, X. Q. lv, Influence of a heavy rainfall event on nutrients and phytoplankton dynamics in a well-mixed semi-enclosed bay, J. Hydrol., 617 (2023), 128932. https://doi.org/10.1016/j.jhydrol.2022.128932 doi: 10.1016/j.jhydrol.2022.128932
[12]
X. D. Li, X. Y. Yang, T. W. Huang, Persistence of delayed cooperative models: Impulsive control method, Appl. Math. Comput., 342 (2019), 130–146. https://doi.org/10.1016/j.amc.2018.09.003 doi: 10.1016/j.amc.2018.09.003
[13]
M. K. Alijani, H. Wang, J. J. Elser, Modeling the bacterial contribution to planktonic community respiration in the regulation of solar energy and nutrient availability, Ecol. Complexity, 23 (2015), 25–33. https://doi.org/10.1016/j.ecocom.2015.05.002 doi: 10.1016/j.ecocom.2015.05.002
[14]
C. J. Dai, M. Zhao, H. G. Yu, Dynamics induced by delay in a nutrient–phytoplankton model with diffusion, Ecol. Complexity, 26 (2016), 29–36. https://doi.org/10.1016/j.ecocom.2016.03.001 doi: 10.1016/j.ecocom.2016.03.001
[15]
P. Feketa, V. Klinshov, L. Lücken, A survey on the modeling of hybrid behaviors: How to account for impulsive jumps properly, Commun. Nonlinear Sci. Numer. Simul., 103 (2021), 105955. https://doi.org/10.1016/j.cnsns.2021.105955 doi: 10.1016/j.cnsns.2021.105955
[16]
A. M. Samoilenko, N. A. Perestyuk, Impulsive Differential Equations, World Scientific, 1995. https://doi.org/10.1142/2892
[17]
T. Yang, Impulsive Control Theory, Springer Science & Business Media, 2001.
[18]
V. Lakshmikantham, P. S. Simeonov, Theory of Impulsive Differential Equations, World Scientific, 1989.
[19]
H. Liu, C. J. Dai, H. G. Yu, Q. Guo, J. B. Li, A. M. Hao, et al., Dynamics of a stochastic non-autonomous phytoplankton–zooplankton system involving toxin-producing phytoplankton and impulsive perturbations, Math. Comput. Simul., 203 (2023), 368–386. https://doi.org/10.1016/j.matcom.2022.06.012 doi: 10.1016/j.matcom.2022.06.012
[20]
X. W. Yu, S. L. Yuan, T. H. Zhang, Survival and ergodicity of a stochastic phytoplankton–zooplankton model with toxin-producing phytoplankton in an impulsive polluted environment, Appl. Math. Comput., 347 (2019), 249–264. https://doi.org/10.1016/j.amc.2018.11.005 doi: 10.1016/j.amc.2018.11.005
[21]
D. Z. Li, Y. Liu, H. D. Cheng, Dynamic complexity of a phytoplankton-fish model with the impulsive feedback control by means of Poincaré map, Complexity, 2020 (2020), 8974763. https://doi.org/10.1155/2020/8974763 doi: 10.1155/2020/8974763
[22]
J. Yang, M. Zhao, A mathematical model for the dynamics of a fish algae consumption model with impulsive control strategy, J. Appl. Math., 2012 (2012), 452789. https://doi.org/10.1155/2012/452789 doi: 10.1155/2012/452789
[23]
W. Li, T. H. Zhang, Y. F. Wang, H. D. Cheng, Dynamic analysis of a plankton–herbivore state-dependent impulsive model with action threshold depending on the density and its changing rate, Nonlinear Dyn., 107 (2022), 2951–2963. https://doi.org/10.1007/s11071-021-07022-w doi: 10.1007/s11071-021-07022-w
[24]
S. Spatharis, G. Tsirtsis, D. B. Danielidis, T. Do Chi, D. Mouillot, Effects of pulsed nutrient inputs on phytoplankton assemblage structure and blooms in an enclosed coastal area, Estuarine Coastal Shelf Sci., 73 (2007), 807–815. https://doi.org/10.1016/j.ecss.2007.03.016 doi: 10.1016/j.ecss.2007.03.016
[25]
K. L. Cottingham, S. Glaholt, A. C. Brown, Zooplankton community structure affects how phytoplankton respond to nutrient pulses, Ecology, 85 (2004), 158–171. https://doi.org/10.1890/02-0570 doi: 10.1890/02-0570
[26]
J. dos Santos Severiano, V. L. dos Santos Almeida-Melo, E. M. de Melo-Magalhães, M. do Carmo Bittencourt-Oliveira, A. do Nascimento Moura, Effects of zooplankton and nutrients on phytoplankton: An experimental analysis in a eutrophic tropical reservoir, Mar. Freshwater Res., 68 (2016), 1061–1069. https://doi.org/10.1071/MF15393 doi: 10.1071/MF15393
[27]
Z. Zhao, C. G. Luo, L. Y. Pang, Y. Chen, Nonlinear modelling of the interaction between phytoplankton and zooplankton with the impulsive feedback control, Chaos Solitons Fractals, 87 (2016), 255–261. https://doi.org/10.1016/j.chaos.2016.04.011 doi: 10.1016/j.chaos.2016.04.011
[28]
H. J. Guo, L. S. Chen, X. Y. Song, Qualitative analysis of impulsive state feedback control to an algae-fish system with bistable property, Appl. Math. Comput., 271 (2015), 905–922. https://doi.org/10.1016/j.amc.2015.09.046 doi: 10.1016/j.amc.2015.09.046
[29]
C. R. Tian, S. G. Ruan, Pattern formation and synchronism in an allelopathic plankton model with delay in a network, SIAM J. Appl. Dyn. Syst., 18 (2019), 531–557. https://doi.org/10.1137/18M1204966 doi: 10.1137/18M1204966
[30]
N. K. Thakur, A. Ojha, D. Jana, R. K. Upadhyay, Modeling the plankton–fish dynamics with top predator interference and multiple gestation delays, Nonlinear Dyn., 100 (2020), 4003–4029. https://doi.org/10.1007/s11071-020-05688-2 doi: 10.1007/s11071-020-05688-2
[31]
Q. Guo, C. J. Dai, H. G. Yu, H. Liu, X. X. Sun, J. B. Li, et al., Stability and bifurcation analysis of a nutrient-phytoplankton model with time delay, Math. Methods Appl. Sci., 43 (2020), 3018–3039. https://doi.org/10.1002/mma.6098 doi: 10.1002/mma.6098
[32]
S. Q. Zhang, T. H. Zhang, S. L. Yuan, Dynamics of a stochastic predator-prey model with habitat complexity and prey aggregation, Ecol. Complexity, 45 (2021), 100889. https://doi.org/10.1016/j.ecocom.2020.100889 doi: 10.1016/j.ecocom.2020.100889
[33]
X. R. Mao, M. Glenn, R. Eric, Environmental Brownian noise suppresses explosion in population dynamics, Stochastic Process Their Appl., 97 (2002), 95–110. https://doi.org/10.1016/s0304-4149(01)00126-0 doi: 10.1016/s0304-4149(01)00126-0
[34]
F. Q. Deng, Q. Luo, X. R. Mao, S. L. Pang, Noise suppresses or expresses exponential growth, Syst. Control Lett., 57 (2008), 262–270. https://doi.org/10.1016/j.sysconle.2007.09.002 doi: 10.1016/j.sysconle.2007.09.002
[35]
Q. Guo, Y. Wang, C. J. Dai, L. J. Wang, H. Liu, J. B. Li, et al., Dynamics of a stochastic nutrient–plankton model with regime switching, Ecol. Modell., 477 (2023), 110249. https://doi.org/10.1016/j.ecolmodel.2022.110249 doi: 10.1016/j.ecolmodel.2022.110249
[36]
J. A. Freund, S. Mieruch, B. Scholze, K. Wiltshire, U. Feudel, Bloom dynamics in a seasonally forced phytoplankton–zooplankton model: Trigger mechanisms and timing effects, Ecol. Complexity, 3 (2006), 129–139. https://doi.org/10.1016/j.ecocom.2005.11.001 doi: 10.1016/j.ecocom.2005.11.001
[37]
J. P. DeLong, C. E. Cressler, Stochasticity directs adaptive evolution toward nonequilibrium evolutionary attractors, Ecology, 104 (2022), e3873. https://doi.org/10.1002/ecy.3873 doi: 10.1002/ecy.3873
[38]
H. Liu, C. J. Dai, H. G. Yu, Q, Guo, J. B. Li, A. M. Hao, et al., Dynamics induced by environmental stochasticity in a phytoplankton-zooplankton system with toxic phytoplankton, Math. Biosci. Eng., 18 (2021), 4101–4126. https://doi.org/10.3934/mbe.2021206 doi: 10.3934/mbe.2021206
[39]
X. W. Yu, S. L. Yuan, T. H. Zhang, Asymptotic properties of stochastic nutrient-plankton food chain models with nutrient recycling, Nonlinear Anal. Hybrid Syst., 34 (2019), 209–225. https://doi.org/10.1016/j.nahs.2019.06.005 doi: 10.1016/j.nahs.2019.06.005
[40]
X. M. Feng, J. X. Sun, L. Wang, F. Q. Zhang, S. L. Sun, Periodic solutions for a stochastic chemostat model with impulsive perturbation on the nutrient, J. Biol. Syst., 29 (2021), 849–870. https://doi.org/10.1142/S0218339021500200 doi: 10.1142/S0218339021500200
[41]
Q. Guo, C. J. Dai, L. J. Wang, H. Liu, Y. Wang, J. B. Li, et al., Stochastic periodic solution of a nutrient–plankton model with seasonal fluctuation, J. Biol. Syst., 30 (2022), 695–720. https://doi.org/10.1142/S0218339022500255 doi: 10.1142/S0218339022500255
[42]
C. J. Dai, H. G. Yu, Q. Guo, H. Liu, Q. Wang, Z. L. Ma, et al., Dynamics induced by delay in a nutrient-phytoplankton model with multiple delays, Complexity, 2019 (2019), 3879626. https://doi.org/10.1155/2019/3879626 doi: 10.1155/2019/3879626
[43]
D. Song, M. Fan, S. H. Yan, M. Liu, Dynamics of a nutrient-phytoplankton model with random phytoplankton mortality, J. Theor. Biol., 488 (2020), 110119. https://doi.org/10.1016/j.jtbi.2019.110119 doi: 10.1016/j.jtbi.2019.110119
[44]
H. Wang, M. Liu, Stationary distribution of a stochastic hybrid phytoplankton–zooplankton model with toxin-producing phytoplankton, Appl. Math. Lett., 101 (2020), 106077. https://doi.org/10.1016/j.aml.2019.106077 doi: 10.1016/j.aml.2019.106077
[45]
A. Mandal, P. K. Tiwari, S. Pal, A nonautonomous model for the effects of refuge and additional food on the dynamics of phytoplankton-zooplankton system, Ecol. Complexity, 46 (2021), 100927. https://doi.org/10.1016/j.ecocom.2021.100927 doi: 10.1016/j.ecocom.2021.100927
[46]
J. Chattopadhayay, R. 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
[47]
C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the European pine sawfly, Can. Entomol., 91 (1959), 293–320. https://doi.org/10.4039/Ent91293-5 doi: 10.4039/Ent91293-5
[48]
A. M. Siepielski, A. Nemirov, M. Cattivera, A. Nickerson, Experimental evidence for an eco-evolutionary coupling between local adaptation and intraspecific competition, Am. Nat., 187 (2016), 447–456. https://doi.org/10.1086/685295 doi: 10.1086/685295
[49]
R. M. May, Stability and Complexity in Model Ecosystems, Princeton university press, 2019.
[50]
Y. Zhang, S. H. Chen, S. J. Gao, X. Wei, Stochastic periodic solution for a perturbed non-autonomous predator–prey model with generalized nonlinear harvesting and impulses, Phys. A, 486 (2017), 347–366. http://dx.doi.org/10.1016/j.physa.2017.05.058 doi: 10.1016/j.physa.2017.05.058
[51]
W. J. Zuo, D. Q. Jiang, Periodic solutions for a stochastic non-autonomous Holling–Tanner predator–prey system with impulses, Nonlinear Anal. Hybrid Syst., 22 (2016), 191–201. http://dx.doi.org/10.1016/j.nahs.2016.03.004 doi: 10.1016/j.nahs.2016.03.004
[52]
S. W. Zhang, D. J. Tan, Dynamics of a stochastic predator–prey system in a polluted environment with pulse toxicant input and impulsive perturbations, Appl. Math. Modell., 39 (2015), 6319–6331. https://doi.org/10.1016/j.apm.2014.12.020 doi: 10.1016/j.apm.2014.12.020
[53]
N. Dalal, D. Greenhalgh, X. R. Mao, A stochastic model for internal HIV dynamics, J. Math. Anal. Appl., 341 (2008), 1084–1101. https://doi.org/10.1016/j.hal.2020.101845 doi: 10.1016/j.hal.2020.101845
[54]
Y. Zhang, S. H. Chen, S. J. Gao, X. Wei, Stochastic periodic solution for a perturbed non-autonomous predator–prey model with generalized nonlinear harvesting and impulses, Phys. A, 486 (2017), 347–366. https://doi.org/10.1016/j.physa.2017.05.058 doi: 10.1016/j.physa.2017.05.058
[55]
J. X. Zhao, Y. F. Shao, Stochastic periodic solution and permanence of a holling–leslie predator-prey system with impulsive effects, J. Math., 2021 (2021), 6694479. https://doi.org/10.1155/2021/6694479 doi: 10.1155/2021/6694479
J. J. Elser, M. Kyle, L. Steger, K. R. Nydick, J. S. Baron, Nutrient availability and phytoplankton nutrient limitation across a gradient of atmospheric nitrogen deposition, Ecology, 90 (2009), 3062–3073. https://doi.org/10.1890/08-1742.1 doi: 10.1890/08-1742.1
[58]
G. Borics, I. Grigorszky, S. Szabó, J. Padisák, Phytoplankton associations in a small hypertrophic fishpond in East Hungary during a change from bottom-up to top-down control, in Developments in Hydrobiology (eds. C. S. Reynolds, M. Dokulil and J. Padisák), 150 (2000), 79–90. https://doi.org/10.1007/978-94-017-3488-2_7
[59]
Y. Kang, F. Koch, C. J. Gobler, The interactive roles of nutrient loading and zooplankton grazing in facilitating the expansion of harmful algal blooms caused by the pelagophyte, Aureoumbra lagunensis, to the Indian River Lagoon, FL, USA, Harmful Algae, 49 (2015), 162–173. https://doi.org/10.1016/j.hal.2015.09.005 doi: 10.1016/j.hal.2015.09.005
[60]
X. D. Wang, B. Q. Qin, G. Gao, H. W. Paerl, Nutrient enrichment and selective predation by zooplankton promote Microcystis (Cyanobacteria) bloom formation, J. Plankton Res., 32 (2010), 457–470. https://doi.org/10.1093/plankt/fbp143 doi: 10.1093/plankt/fbp143
[61]
A. P. Belfiore, R. P. Buley, E. G. Fernandez-Figueroa, M. F. Gladfelter, A. E. Wilson, Zooplankton as an alternative method for controlling phytoplankton in catfish pond aquaculture, Aquacult. Rep., 21 (2021), 100897. https://doi.org/10.1016/j.aqrep.2021.100897 doi: 10.1016/j.aqrep.2021.100897
[62]
D. M. Anderson, A. D. Cembella, G. M. Hallegraeff, Progress in understanding harmful algal blooms: paradigm shifts and new technologies for research, monitoring, and management, Ann. Rev. Mar. Sci., 4 (2012), 143–176. https://doi.org/10.1146/annurev-marine-120308-081121 doi: 10.1146/annurev-marine-120308-081121
[63]
V. Ittekkot, U. Brockmann, W. Michaelis, E. T. Degens, Dissolved free and combined carbohydrates during a phytoplankton bloom in the northern North Sea, Mar. Ecol. Progress Ser., 4 (1981), 299–305. https://doi.org/10.1016/j.amc.2015.09.046 doi: 10.1016/j.amc.2015.09.046
[64]
M. Rehim, Z. Z. Zhang, A. Muhammadhaji, Mathematical analysis of a nutrient–plankton system with delay, SpringerPlus, 5 (2016), 1055. https://doi.org/10.1186/s40064-016-2435-7 doi: 10.1186/s40064-016-2435-7
[65]
N. K. Thakur, A. Ojha, P. K. Tiwari, R. K. Upadhyay, An investigation of delay induced stability transition in nutrient-plankton systems, Chaos Solitons Fractals, 142 (2021), 110474. https://doi.org/10.1016/j.chaos.2020.110474 doi: 10.1016/j.chaos.2020.110474
[66]
S. R. J. Jang, J. Baglama, J. Rick, Nutrient-phytoplankton-zooplankton models with a toxin, Math. Comput. Modell., 43 (2006), 105–118. https://doi.org/10.1016/j.mcm.2005.09.030 doi: 10.1016/j.mcm.2005.09.030
This article has been cited by:
1.
Xiumei Deng, Qihua Huang,
A toxicant-taxis reaction–diffusion model for the interaction between two species and a toxicant in an aquatic ecosystem,
2024,
183,
09600779,
114930,
10.1016/j.chaos.2024.114930
2.
M. PRIYANKA, P. MUTHUKUMAR,
BIFURCATION ANALYSIS AND CHAOS CONTROL OF THE DISSOLVED OXYGEN-PHYTOPLANKTON DYNAMICAL MODEL,
2024,
32,
0218-3390,
1039,
10.1142/S0218339024500347
3.
Xiumei Deng, Qihua Huang, Hao Wang,
A nonautonomous model for the impact of toxicants on size-structured aquatic populations: Well-posedness and long-term dynamics,
2025,
381,
00255564,
109382,
10.1016/j.mbs.2025.109382
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[J]. Mathematical Biosciences and Engineering, 2023, 20(8): 15496-15523. doi: 10.3934/mbe.2023692
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[J]. Mathematical Biosciences and Engineering, 2023, 20(8): 15496-15523. doi: 10.3934/mbe.2023692
Strength of intraspecific competition among the zooplankton population
L/μg/day
0.1
—
Figure 1. Time evolutions of (a) nutrients, (b) phytoplankton and (c) zooplankton in models (2.1) and (2.2)
Figure 2. Results for model (2.2) with σ1=σ2=σ3=0.05+0.01sin(πt/20). (a) The mean for phytoplankton biomass in the (α1,α2) plane; (b) the mean for zooplankton biomass in the (α1,α2) plane; (c) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=0.1,α2=8; (d) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=8,α2=0.1; (e) the evolution of a single path of the phytoplankton population and zooplankton population in model (2.2) with α1=5,α2=5
Figure 3. Results for σ1=σ2=σ3=0.05+0.01sin(πt/20). (a) The variance and the third-order central moment of phytoplankton population and (b) zooplankton population versus different nutrient pulses with α2=4; (c) the variance and the third order central moment of the phytoplankton population and (d) zooplankton population versus different zooplankton pulses with α1=3
Figure 4. Results for model (2.2) with σ1=σ2=σ3=0.05+0.01sin(πt/20) and α1=α2=exp(1), showing the mean and the variance of phytoplankton biomass and zooplankton biomass with respect to number of pulses per period. The path of plankton populations indicates the changes of plankton biomasses for p=2 and p=5, respectively
Figure 5. Results for model (2.2) with σ1=σ2=σ3=0.1+0.01sin(πt/20) and α1=α2=exp(2). (a) The mean and the variance of phytoplankton and zooplankton biomasses; evolution of a single path of (b) nutrient concentration, (c) phytoplankton population, (e) zooplankton population
Figure 6. Time evolutions of nutrients (first column), phytoplankton (second column) and zooplankton (third column) for model (2.2) and the corresponding deterministic model. Here, we have chosen α1=α2=exp(2) and p=3; σ1=σ2=σ3=0.2+0.01sin(πt/20) in (a−c) and σ1=σ2=σ3=0.5+0.01sin(πt/20) in (d−f)
Catalog
Abstract
1.
Introduction
2.
The mathematical model
3.
The main results
3.1. Preliminaries
3.2. Existence and uniqueness of the global positive solution