
In this paper, we study the attracting sets for two plankton models perturbed by bounded noises which are modeled by the Ornstein-Uhlenbeck process. Specifically, we prove the existence and uniqueness of the solutions for these random models, as well as the existence of the attracting sets for the random dynamical systems generated by the solutions. In order to further reveal the survival of plankton species in a fluctuating environment, we analyze the internal structure of the attracting sets and give sufficient conditions for the persistence and extinction of the plankton species. Some numerical simulations are shown to support our theoretical results.
Citation: Zhihao Ke, Chaoqun Xu. Structure analysis of the attracting sets for plankton models driven by bounded noises[J]. Mathematical Biosciences and Engineering, 2023, 20(4): 6400-6421. doi: 10.3934/mbe.2023277
[1] | Tomás Caraballo, Renato Colucci, Javier López-de-la-Cruz, Alain Rapaport . Study of the chemostat model with non-monotonic growth under random disturbances on the removal rate. Mathematical Biosciences and Engineering, 2020, 17(6): 7480-7501. doi: 10.3934/mbe.2020382 |
[2] | Xin Zhao, Lijun Wang, Pankaj Kumar Tiwari, He Liu, Yi Wang, Jianbing Li, Min Zhao, Chuanjun Dai, Qing Guo . Investigation of a nutrient-plankton model with stochastic fluctuation and impulsive control. Mathematical Biosciences and Engineering, 2023, 20(8): 15496-15523. doi: 10.3934/mbe.2023692 |
[3] | Meng Gao, Xiaohui Ai . A stochastic Gilpin-Ayala mutualism model driven by mean-reverting OU process with Lévy jumps. Mathematical Biosciences and Engineering, 2024, 21(3): 4117-4141. doi: 10.3934/mbe.2024182 |
[4] | Giacomo Ascione, Enrica Pirozzi . On a stochastic neuronal model integrating correlated inputs. Mathematical Biosciences and Engineering, 2019, 16(5): 5206-5225. doi: 10.3934/mbe.2019260 |
[5] | 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 |
[6] | Virginia Giorno, Serena Spina . On the return process with refractoriness for a non-homogeneous Ornstein-Uhlenbeck neuronal model. Mathematical Biosciences and Engineering, 2014, 11(2): 285-302. doi: 10.3934/mbe.2014.11.285 |
[7] | Yuanpei Xia, Weisong Zhou, Zhichun Yang . Global analysis and optimal harvesting for a hybrid stochastic phytoplankton-zooplankton-fish model with distributed delays. Mathematical Biosciences and Engineering, 2020, 17(5): 6149-6180. doi: 10.3934/mbe.2020326 |
[8] | Alexander P. Krishchenko, Konstantin E. Starkov . The four-dimensional Kirschner-Panetta type cancer model: How to obtain tumor eradication?. Mathematical Biosciences and Engineering, 2018, 15(5): 1243-1254. doi: 10.3934/mbe.2018057 |
[9] | Hong Yang . Global dynamics of a diffusive phytoplankton-zooplankton model with toxic substances effect and delay. Mathematical Biosciences and Engineering, 2022, 19(7): 6712-6730. doi: 10.3934/mbe.2022316 |
[10] | Giuseppina Albano . Detecting time-changes in PM10 during Covid pandemic by means of an Ornstein Uhlenbeck type process. Mathematical Biosciences and Engineering, 2021, 18(1): 888-903. doi: 10.3934/mbe.2021047 |
In this paper, we study the attracting sets for two plankton models perturbed by bounded noises which are modeled by the Ornstein-Uhlenbeck process. Specifically, we prove the existence and uniqueness of the solutions for these random models, as well as the existence of the attracting sets for the random dynamical systems generated by the solutions. In order to further reveal the survival of plankton species in a fluctuating environment, we analyze the internal structure of the attracting sets and give sufficient conditions for the persistence and extinction of the plankton species. Some numerical simulations are shown to support our theoretical results.
Plankton are located in the first trophic layer of the aquatic food chain and are the base of the aquatic ecosystem [1]. They not only generate organic compounds by absorbing carbon dioxide dissolved in the surrounding environment but also perform photosynthesis, which has an important impact on large-scale global processes such as the global carbon cycle, climate change and ocean-atmosphere dynamics [2].
Toxin producing phytoplankton (TPP) are a kind of harmful plankton with the ability to release toxic chemicals into the environment. The toxic chemicals may inhibit predation pressure from phytoplankton and other predator populations in planktonic systems [3] and then contribute to the formation of harmful algal blooms (HABs) [4]. For example, some experimental observations in [5] have indicated that the toxic dinoflagellate Alexandrium fundyense can negatively affect the growth rate of the copepod Acartia hudsonica, and the toxic effects may have profound implications on the ability of grazers to control the HABs. Over the past few decades, research on the complex dynamics of planktonic systems has attracted great interests of researchers; see previous studies [6,7,8,9,10,11,12] and the references therein.
For example, for a nutrient-phytoplankton system with TPP, Chakraborty et al. [1] established the following nonlinear mathematical model:
{dNdt=a−bNP−hN+kP,dPdt=cNP−dP−θP2μ2+P2, | (1.1) |
where N(t) and P(t) are the concentrations of nutrient and TPP at time t, respectively. Parameter a is the external nutrient inflow rate, b is the nutrient uptake rate of phytoplankton, c(c<b) is the conversion rate of nutrients into phytoplankton, h is the loss rate of nutrients, d is the mortality rate of phytoplankton, k(k<d) is the nutrient recycle rate due to the death of phytoplankton, μ is the half saturation constant, and θ represents the rate of release of toxic chemicals by the TPP population. All the parameters are assumed to be positive. The authors showed that, for a certain range of θ, model (1.1) exhibits periodic solutions. They also observed that toxin produced by the TPP may act as a biological control in the termination of the planktonic bloom, which is in good agreement with some earlier findings.
Because the real environment is full of stochasticity, and every ecosystem is inevitably affected by environmental noise, it seems more appropriate to develop some stochastic ecological models by considering the influence of environmental noise [13,14,15,16,17]. For example, Ji et al. [13] established a stochastic Lotka-Volterra predator-prey model with white noise, and obtained some criteria for persistence and extinction of the species. Zhang et al. [14] proposed and studied a stochastic non-autonomous prey-predator model with impulsive effects. They showed that the stochastic noise and impulsive perturbations have crucial effects on the persistence and extinction of each species.
In fact, plankton systems are more susceptible to environmental fluctuations such as light, water temperature and water pH [18,19,20]. Therefore, based on deterministic model (1.1), Yu et al. [18] constructed the following plankton model with white noise:
{dN=(a−bNP−hN+kP)dt+α1NdW1(t),dP=(cNP−dP−θP2μ2+P2)dt+α2PdW2(t), | (1.2) |
where Wi(t) are standard Wiener processes with intensities αi, i=1,2, and Wi(t) are defined on a complete probability space (Ω,F,P) with filtration {Ft}t≥0. For stochastic model (1.2), the authors gave sufficient criteria for the existence of ergodic stationary distribution and investigated the extinction and persistence of the phytoplankton species. They also showed that the TPP and environmental fluctuations may have great influence on planktonic blooms.
Different from the standard Wiener process, the Ornstein-Uhlenbeck (O-U) process [21] can be used to model the bounded environmental fluctuation in a real ecosystem. The ecological model driven by O-U process is closer to reality, as stated by Caraballo et al. [22]: "The most common stochastic process that is considered when modeling disturbances in real life is the well-known standard Wiener process. Nevertheless, this stochastic process has the property of having continuous but not bounded variation paths, which does not suit to the idea of modeling real situations since, in most of cases, the real life is subjected to fluctuations which are known to be bounded." So, ecological models driven by O-U process have been proposed and analyzed by some scholars; see [22,23,24,25,26,27,28] and the references therein. For example, Caraballo et al. [22] used the O-U process to model the bounded noise perturbations in a logistic system and competitive Lotka-Volterra system. They present an example testing the theoretical result with real data and verified that this method is a realistic one. A general eco-epidemiological system, in which the birth rate of prey population is driven by O-U process, was considered in [24]. The authors proved the existence of a global random attractor and the persistence of susceptible prey, and provided some conditions for the simultaneous extinction of infective preys and predators. In [25], López-de-la-Cruz derived a random chemostat model driven by O-U process and investigated the existence and internal structure of the attracting set (or attractor) for the random model. In reality, the internal structure of the attracting set can reflect the survival of species in ecosystem [25,29]. For other systems with O-U process we refer also to [30,31,32,33].
In view of the latest research and the advantages of O-U process, for the nutrient-phytoplankton system with TPP, we consider the environmental fluctuations to be bounded and model the bounded noise by using suitable O-U process in this paper. Based on deterministic model (1.1) and stochastic model (1.2), respectively, we first construct two random plankton models and then investigate the existence and internal structures of the attracting sets for these models.
This paper is organized as follows: In Section 2, we analyze a random plankton model corresponding to deterministic model (1.1) in which the external nutrient inflow rate a is driven by an O-U process. In Section 3, we use the O-U process to transform stochastic plankton model (1.2) into a random one and investigate the attracting set for the random model. A simple discussion is given in Section 4. For completeness, some mathematical backgrounds of the O-U process and random dynamical system is given in the Appendix.
In this section, we will consider a suitable O-U process to perturb the external nutrient inflow rate a in a deterministic plankton system in the same way as in [25,34]. Particularly, we are interested in replacing a by the random term a+σz∗(θtω) in deterministic model (1.1), where z∗(θtω) denotes the O-U process which will be introduced in the Appendix, and σ>0 represents the intensity of perturbation. In such a way, the resulting random model is given by the following system of random differential equations:
{dNdt=(a+σz∗(θtω))−bNP−hN+kP,dPdt=cNP−dP−θP2μ2+P2. | (2.1) |
We would also like to note that, thanks to the property limλ→∞z∗(θtω)=0 shown in Proposition A.1, for every fixed ω∈Ω, it is possible to take λ large enough such that a+σz∗(θtω)∈(a1,a2) for every t∈R+, where a1 and a2 are positive constants.
We will introduce the main results for random model (2.1) in the following three subsections, including existence and uniqueness of the global positive solution and existence and internal structure of the attracting set.
Theorem 2.1. For any initial value S(0):=(N(0),P(0))∈R2+ and any ω∈Ω, model (2.1) possesses a unique global positive solution
S(t;0,ω,S(0))=(N(t;0,ω,N(0)),P(t;0,ω,P(0)))∈C1(R+,R2+) |
with S(0;0,ω,S(0))=S(0).
Proof. Let S(t;0,ω,S(0))=(N(t;0,ω,N(0)),P(t;0,ω,P(0))), then model (2.1) can be rewritten as
dSdt=L(θtω)⋅S+F(S,θtω), |
where
L(θtω)=[−hk0−d], |
and F:R2+×R+→R2 is given by
F(η,θtω)=[a+σz∗(θtω)−bη1η2cη1η2−θη22μ2+η22], |
where η=(η1,η2)∈R2+.
One can find that F(η,θtω) is locally Lipschitz with respect to (η1,η2), then model (2.1) possesses a unique local solution. To prove the local solution is a global one, we define the new variable U=N+P. It is easy to see that U satisfies the following equation:
dUdt=a+σz∗(θtω)−(b−c)NP−hN−(d−k)P−θP2μ2+P2. | (2.2) |
Notice that a+σz∗(θtω)∈(a1,a2), and from Eq (2.2), we know
dUdt≤a2−m1U, | (2.3) |
where m1=min{h,d−k}. It is straightforward to check that U does not blow up at any finite time, and the same happens to N and P. Therefore, the unique local solution can be extended to a global one.
Moreover, from Eq (2.1), we know that
dNdt|N=0=a+σz∗(θtω)+kP>0 |
for all P≥0, and
dPdt|P=0=0 |
for all N≥0. Thus, the unique global solution S(t;0,ω,S(0)) of random model (2.1) remains in the positive quadrant R2+ for every initial value S(0)∈R2+.
Remark 2.1. Define a mapping φS:R+×Ω×R2+→R2+ given by
φS(t,ω)S(0)=S(t;ω,S(0)),for allt∈R+,ω∈Ω,S(0)∈R2+. |
Since the function F is continuous in (S,t) and is measurable in ω, we obtain the (B(R+)×F×B(R2+),B(R2+))-measurability of this mapping, which defines a random dynamical system generated by the solution mapping of model (2.1).
Theorem 2.2. There exists a deterministic compact attracting set
B0={(N,P)∈R2+:U1≤N+P≤U2,N_1≤N} |
for the solution of model (2.1), where U1=a1M1, U2=a2m1, N_1=a1bU2+h, and
M1=max{(b−c)U2+h,d−k+θ2μ}. |
Proof. According to inequality (2.3), we can obtain
limt→∞U(t)≤a2m1=U2. | (2.4) |
On the other hand, it follows from equation (2.2) that
dUdt≥a1−(b−c)NP−hN−(d−k)P−θ2μP≥a1−[(b−c)U2+h]N−(d−k+θ2μ)P. |
By setting M1=max{(b−c)U2+h,d−k+θ2μ}, we get
dUdt≥a1−M1U, |
and then
limt→∞U(t)≥a1M1=U1. | (2.5) |
According to inequality (2.4), we know that, for every initial value S(0)∈R2+ and any given ε>0, there exists some time T(ω,S(0),ε)>0 such that U(t)≤U2+ε for all t≥T(ω,S(0),ε). Therefore, we know N(t)+P(t)≤U2 holds for every time t large enough. It then follows from a+σz∗(θtω)>a1 and P≤U2 that
dNdt=(a+σz∗(θtω))−bNP−hN+kP≥a1−(bU2+h)N, |
and
limt→∞N(t)≥a1bU2+h=N_1. | (2.6) |
Thus, from inequalities (2.4), (2.5) and (2.6), we can obtain that
B0={(N,P)∈R2+:U1≤N+P≤U2,N_1≤N} |
is a deterministic attracting set for the solution of model (2.1).
Remark 2.2. The existence of attracting set B0 indicates that the inequalities
U1≤N(t)+P(t)≤U2andN_1≤N(t) |
hold for every time t large enough, where S(t)=(N(t),P(t)) is the solution of model (2.1). In what follows, we always believe that these inequalities are true, because the purpose of this paper is to explore the long-time behavior of the plankton species.
Taking parameters a1=1.3, a2=2.7, b=0.8, c=0.7, k=0.1, h=0.4, d=0.5, θ=0.1 and μ=2 in model (2.1), we can calculate U1=1.21, U2=6.75 and N_1=0.22, and the simulation of the attracting set is shown in Figure 1.
Theorem 2.3. For model (2.1), if the condition
cU2−d<0 |
holds, then the attracting set B0 is reduced to a line segment on the coordinate axis. More precisely, it is
B0={(N,P)∈R2+:N_2≤N≤¯N,P=0}, |
where N_2=a1h and ¯N=a2h.
Proof. According to Remark (2.2), we know N<U2. It then follows from
dPdt=cNP−dP−θP2μ2+P2 |
that
dPdt≤(cU2−d)P. |
If cU2−d<0, we know that
limt→∞P=0, |
and then, for every time t large enough, the first equation of model (2.1) can be written as
dNdt=(a+σz∗(θtω))−hN. |
It follows from a+σz∗(θtω)∈(a1,a2) that
a1−hN≤dNdt≤a2−hN, |
and then
N_2=a1h≤limt→∞N(t)≤a2h=¯N. |
Therefore, the attracting set of model (2.1) will become
B0={(N,P)∈R2+:N_2≤N≤¯N,P=0}, |
which is a line segment on the coordinate axis.
Taking parameters a=2, a1=1.3, a2=2.7, b=0.8, c=0.1, k=0.1, h=0.6, d=0.6, θ=0.1, λ=20, μ=2 and σ=0.5 in model (2.1), we can calculate N_2=2.16 and ¯N=4.5. The simulation of the attracting set and three trajectories with different initial values is shown in Figure 2. One can see that the trajectory of model (2.1) eventually enters the line segment B0 on the coordinate axis, which indicates that the phytoplankton species will go extinct, and only the nutrient can be persistent.
Theorem 2.4. For model (2.1), if the condition
cU1−(d+θ2μ)>0 |
holds, then the attracting set B0 is reduced to a plane region in the first quadrant. More precisely, it is
B0={(N,P)∈R2+:U1≤N+P≤U2,N_3≤N,P1≤P}, |
where N_3=a1+kP1bU2+h, P1=cU1−(d+θ2μ)c.
Proof. From Remark 2.2, we know U1−P≤N. It follows from
dPdt=cNP−dP−θP2μ2+P2 |
that
dPdt≥c(U1−P)P−dP−θ2μP=[cU1−(d+θ2μ)−cP]P. |
If cU1−(d+θ2μ)>0, we can conclude
limt→∞P(t)≥cU1−(d+θ2μ)c=P1. |
For every time t large enough, it follows from a+σz∗(θtω)>a1 and P1≤P≤U2 that
dNdt=(a+σz∗(θtω))−bNP−hN+kP≥a1+kP1−(bU2+h)N, |
and then
limt→∞N(t)≥a1+kP1bU2+h=N_3. |
Therefore, the attracting set of model (2.1) will become
B0={(N,P)∈R2+:U1≤N+P≤U2,N_3≤N,P1≤P}. |
At this time, the attracting set lies completely in the first quadrant plane.
Taking parameters a=2, a1=1.3, a2=2.7, b=0.8, c=0.7, k=0.1, h=0.4, d=0.5, θ=0.1, λ=20, μ=2 and σ=0.5 in model (2.1), we can calculate N_3=0.23, P1=0.45, U1=1.2 and U2=6.75. The simulation of the attracting set and three trajectories with different initial values is shown in Figure 3. One can see that the trajectory of model (2.1) eventually enters the plane region B0 in the first quadrant, which indicates that the phytoplankton species and nutrient can be simultaneously persistent.
In this section, we assume that the nutrient and phytoplankton species in plankton system are affected by the same white noise, and then model (1.2) is reduced to the following stochastic model in Itô's sense:
{dN=(a−bNP−hN+kP)dt+αNdW(t),dP=(cNP−dP−θP2μ2+P2)dt+αPdW(t). | (3.1) |
Due to the properties of Stratonovich integrals following the classical rules in calculus, with the help of the well-known conversion between Itô's and Stratonovich's senses, we further rewrite model (3.1) as the following stochastic model in Stratonovich's sense:
{dN=(a−bNP−hN+kP−α22N)dt+αN∘dW(t),dP=(cNP−dP−θP2μ2+P2−α22P)dt+αP∘dW(t). | (3.2) |
In what follows, we use the O-U process to transform stochastic model (3.2) into a random one. To this end, we first define two new variables x(t) and y(t) as follows:
x(t)=N(t)e−αz∗(θtω), y(t)=P(t)e−αz∗(θtω). |
For the sake of simplicity, we will write z∗ instead of z∗(θtω), x instead of x(t), and y instead of y(t). From Eq (3.2) and the Langevin equation shown in the Appendix, we know that variables x and y satisfy the following equations:
{dxdt=ae−αz∗−bxyeαz∗−(h+α22−αλz∗)x+ky,dydt=cxyeαz∗−(d+α22−αλz∗)y−θy2eαz∗μ2+y2e2αz∗. | (3.3) |
According to the property limλ→0λz∗(θtω)=0 shown in Proposition A.1, for every fixed ω∈Ω, it is possible to choose a suitable λ such that α22−αλz∗∈(l1,l2) for every t∈R+, where l1<l2<∞, and so that both h+l1 and d−k+l1 are positive.
We will introduce the main results for random model (3.3) in the following three subsections, including existence and uniqueness of the global positive solution and existence and internal structure of the attracting set.
Theorem 3.1. For any initial value X(0):=(x(0),y(0))∈R2+ and any ω∈Ω, model (3.3) possesses a unique global positive solution
X(t;0,ω,X(0)):=(x(t;0,ω,x(0)),y(t;0,ω,y(0)))∈C1(R+,R2+) |
with X(0;0,ω,X(0))=X(0).
Proof. Let X(t;0,ω,X(0)):=(x(t;0,ω,x(0)),y(t;0,ω,y(0))). Then, model (3.3) can be rewritten as
dX=L(θtω)⋅X+F(X,θtω), |
where
L(θtω)=[−(h+α22−αλz∗)k0−(d+α22−αλz∗)], |
and F:R2+×R+→R2 is given by
F(η,θtω)=[ae−αz∗−bη1η2eαz∗cη1η2eαz∗−θη22eαz∗μ2+η22e2αz∗], |
where η=(η1,η2)∈R2+.
We can find that F(η,θtω) is locally Lipschitz with respect to η=(η1,η2), and then model (3.3) possesses a unique local solution. To prove the local solution is a global one, we define the new state variable V=x+y. It is easy to see that V satisfies the following equation:
dVdt=ae−αz∗−(b−c)xyeαz∗−(h+α22−αλz∗)x−(d−k+α22−αλz∗)y−θy2eαz∗μ2+y2e2αz∗. | (3.4) |
Notice that α22−αλz∗∈(l1,l2), and from Eq (3.4), we know
dVdt≤ael2−α22λ−m2V, | (3.5) |
where m2=min{h+l1,d−k+l1}. It is straightforward to check that V does not blow up at any finite time, and the same happens to x and y. Therefore, the unique local solution can be extended to a global one.
Moreover, from Eq (3.3), we know that
dxdt|x=0=ae−αz∗+ky≥0 |
for all y≥0, and
dydt|y=0=0 |
for all x≥0. Thus, the unique global solution X(t;0,ω,X(0)) of random model (3.3) remains in the positive cone R2+ for every initial value X(0)∈R2+.
Remark 3.1. Define a mapping φX:R+×Ω×R2+→R2+ given by
φX(t,ω)X(0):=X(t;ω,X(0)),for allt∈R+,ω∈Ω,X(0)∈R2+. |
Since the function F is continuous in (X,t) and is measurable in ω, we obtain the (B(R+)×F×B(R2+),B(R2+))-measurability of the mapping, which defines a random dynamical system generated by the solution mapping of model (3.3).
Theorem 3.2. There exists a deterministic compact attracting set
B0={(x,y)∈R2+:V1≤x+y≤V2,x_1≤x} |
for the solution of model (3.3), where V1=aM2el1−α22λ, V2=am2el2−α22λ, x_1=ael1−α22λbV2eα22−l1λ+h+l2 and
M2=max{(b−c)V2eα22−l1λ+h+l2,d−k+θ2μ+l2}. |
Proof. According to inequality (3.5), we can obtain
limt→∞V(t)≤am2el2−α22λ=V2. | (3.6) |
Also, from equation (3.4) and α22−αλz∗∈(l1,l2), we can obtain
dVdt≥ae−αz∗−(b−c)V2eαz∗x−(h+α22−αλz∗)x−(d−k+α22−αλz∗)y−θ2μy≥ael1−α22λ−(b−c)V2eα22−l1λx−(h+l2)x−(d−k+l2)y−θ2μy=ael1−α22λ−[(b−c)V2eα22−l1λ+h+l2]x−(d−k+l2+θ2μ)y. |
By setting M2=max{(b−c)V2eα22−l1λ+h+l2,d−k+θ2μ+l2}, we can get
dVdt≥ael1−α22λ−M2V, |
and find
limt→∞V(t)≥aM2el1−α22λ=V1. | (3.7) |
According to inequalities (3.6), we know that, for every initial value X(0)∈R2+ and any given ε>0, there exists some time T(ω,X(0),ε)>0 such that V(t)≤V2+ε for all t≥T(ω,X(0),ε). Therefore, we know x+y≤V2 holds for every time t large enough. It then follows from α22−αλz∗∈(l1,l2) and y≤V2 that
dxdt=ae−αz∗−bxyeαz∗−(h+α22−αλz∗)x+ky≥ael1−α22λ−(bV2eα22−l1λ+h+l2)x, |
and
limt→∞x≥ael1−α22λbV2eα22−l1λ+h+l2=x_1. |
Therefore,
B0={(x,y)∈R2+:V1≤x+y≤V2,x_1≤x} |
is a deterministic attracting set for the solution of model (3.3).
Remark 3.2. The existence of attracting set B0 indicates that the inequalities
V1≤x(t)+y(t)≤V2andx_1≤x(t) |
hold for every time t large enough, where X(t)=(x(t),y(t)) is the solution of model (3.3). In what follows, we always believe that these inequalities are true, due to the purpose of this paper is to explore the long-time behavior of the plankton species.
Taking parameters a=2, b=1.5, c=1.4, k=0.05, h=0.7, d=0.6, θ=0.1, μ=0.8, α=0.1, λ=0.5, l1=−0.13 and l2=0.13 in model (3.3), we can calculate V1=0.93, V2=6.11 and x_1=0.118, and the simulation of the attracting set is shown in Figure 4.
Theorem 3.3. For model (3.3), if the condition
cV2eα22−l1λ−(d+l1)<0, |
holds, then the attracting set B0 is reduced to a line segment on the coordinate axis. More precisely, it is
B0={(x,y)∈R2+:x_2≤x≤¯x,y=0}, |
where x_2=ah+l2el1−α22λ and ¯x=ah+l1el2−α22λ.
Proof. According to Remark 3.2, we know that x<V2. Then, it follows from
dydt=cxyeαz∗−(d+α22−αλz∗)y−θy2eαz∗μ2+y2e2αz∗ |
and l1≤α22−αλz∗ that
dydt≤[cV2eα22−l1λ−(d+l1)]y. |
If cV2eα22−l1λ−(d+l1)<0, we know that
limt→+∞y=0, |
and then, for every time t large enough, the first equation of model (3.3) can be rewritten as
dxdt=ae−αz∗−(h+α22−αλz∗)x. |
It follows from α22−αλz∗∈(l1,l2) that
ael1−α22λ−(h+l2)x≤dxdt≤ael2−α22λ−(h+l1)x, |
and then
x_2=ah+l2el1−α22λ≤limt→∞x≤ah+l1el2−α22λ=¯x. |
Therefore, the attracting set of model (3.3) will become
B0={(x,y)∈R2+:x_2≤x≤¯x,y=0}, |
which is a line segment on the coordinate axis.
Taking parameters a=0.9, b=1.5, c=0.1, k=0.05, h=0.7, d=0.6, θ=0.1, λ=0.5, α=0.1, μ=0.8, l1=−0.13 and l2=0.13 in model (3.3), we can calculate x_2=0.82 and ¯x=2.02. The simulation of the attracting set and three trajectories with different initial values is shown in Figure 5. One can see that the trajectory of model (3.3) eventually enters the line segment B0 on the coordinate axis, which indicates that the phytoplankton species will go to extinct, and only the nutrient can be persistent.
Theorem 3.4. For model (3.3), if the condition
cV1eα22−l2λ−(d+l2+θ2μ)>0, |
holds, then the attracting set B0 is reduced to a plane region in the first quadrant. More precisely, it is
B0={(x,y)∈R2+:V1≤x+y≤V2,x_3≤x,y1≤y}, |
where x_3=ael1−α22λ+ky1bV2eα22−l1λ+h+l2 and y1=cV1eα22−l2λ−(d+l2+θ2μ)ceα22−l2λ.
Proof. From Remark 3.2, we know that V1−y≤x. Then, it follows from
dydt=cxyeαz∗−(d+α22−αλz∗)y−θy2eαz∗μ2+y2e2αz∗ |
and α22−αλz∗≤l2 that
dydt≥c(V1−y)eα22−l2λy−(d+l2)y−θ2μy=[cV1eα22−l2λ−(d+l2+θ2μ)−ceα22−l2λy]y. |
If cV1eα22−l2λ−(d+l2+θ2μ)>0, then
limt→∞y(t)≥cV1eα22−l2λ−(d+l2+θ2μ)ceα22−l2λ=y1. |
For every time t large enough, it follows from α22−αλz∗∈(l1,l2) and y1≤y≤V2 that
dxdt=ae−αz∗−bxyeαz∗−(h+α22−αλz∗)x+ky≥ael1−α22λ+ky1−(bV2eα22−l1λ+h+l2)x, |
and then
limt→∞x≥ael1−α22λ+ky1bV2eα22−l1λ+h+l2=x_3. |
Therefore, the attracting set of model (3.3) will become
B0={(x,y)∈R2+:V1≤x+y≤V2,x_3≤x,y1≤y}. |
In that case, the attracting set lies completely in the first quadrant plane.
Taking parameters a=2, b=1.5, c=1.4, k=0.05, h=0.7, d=0.6, θ=0.1, μ=0.8, α=0.1, l1=−0.13, l2=0.13 and λ=0.5 in model (3.3), we can calculate x_3=0.12, y1=0.2, V1=0.93 and V2=6.11. The simulation of the attracting set and three trajectories with different initial values is shown in Figure 6. One can see that the trajectory of model (3.3) eventually enters the plane region B0 in the first quadrant, which indicates that the phytoplankton species and nutrient can be simultaneously persistent.
Remark 3.3. From the expressions V1 and V2 shown in Theorem 3.2 and the expressions x_3 and y1 shown in Theorem 3.4, we can find that the values of V1, V2, x_3 and y1 will decrease with the increase of α. That is to say, when the perturbation intensity α increases, the attracting set B0 will move towards the origin of coordinates. Biologically speaking, the perturbation is adverse to the survival of the plankton system.
We have considered two random plankton models for the plankton systems driven by bounded noise. To this end, we make use of the O-U process to ensure the random perturbations are bounded in some interval. The first random model (i.e., model (2.1)) is related to deterministic system (1.1) in which the external nutrient inflow rate a is perturbed by the O-U process. The second one (i.e., model (3.3)) is related to stochastic system (1.2), which can be achieved by appropriate variable substitution associated with the O-U process.
We first proved, respectively, in Theorem 2.1 and Theorem 3.1 that the random models possess unique global solutions for any positive initial conditions. Then, we proved, respectively, in Theorem 2.2 and Theorem 3.2 the existence of attracting sets for the solutions of random model (2.1) and random model (3.3). In order to have more detailed information about the long-time behavior of the plankton species, we further investigated the internal structures of the attracting sets. Specifically, Theorem 2.3 and Theorem 3.3 state some conditions under which the attracting set is reduced to a line segment on the coordinate axis (biologically speaking, the phytoplankton species will go to extinct). Theorem 2.4 and Theorem 3.4 state some conditions under which the attracting set is reduced to a plane region in the first quadrant (biologically speaking, the phytoplankton species can be persistent).
It is important to point out that the attracting sets for the solutions of model (2.1) (show in Theorems 2.2, 2.3 and 2.4) do not depend on the intensity of the perturbation, but the attracting sets for the solutions of model (3.3) (show in Theorems 3.2, 3.3 and 3.4) will move towards the origin of coordinates when the perturbation intensity α increases. In Figure 7, by taking initial value (4, 2); parameters a=2, b=1.5, c=1.4, k=0.05, h=0.7, d=0.6, θ=0.1, μ=0.8, l1=−0.13, l2=0.13, λ=0.5; and different noise intensities α=0.1, α=0.4 and α=0.7, we show trajectories of model (3.3). One can see from Figure 7 that the region that the trajectory finally enters will move towards the origin of coordinates when the perturbation intensity α increases.
The results in the present paper seem to be able to help us better understand the dynamics of the plankton system in a stochastic sense. One can further use the O-U process to model the real bounded fluctuations existing in other ecological systems.
This work was jointly supported by grants from the National Natural Science Foundation of China (62173161, 12161005, 11801224) and Natural Science Foundation of Jiangsu Province (BK20180856).
The authors declare there is no conflict of interest.
In this section, we will recall briefly some useful definitions and results about the O-U process and random dynamical systems to make our presentation as complete as possible.
Let W be a two sided Wiener process. Kolmogorov's theorem ensures that W has a continuous version, which we will denote by ω, whose canonical interpretation is as follows: Let Ω be defined by
Ω={ω∈C(R,R):ω(0)=0}, |
F be the Borel σ-algebra on Ω generated by the compact open topology [35] and P be the corresponding Wiener measure on F. We consider the Wiener shift flow given by
θtω(⋅)=ω(⋅+t)−ω(t),t∈R. |
Then, (Ω,F,P,{θt}t∈R) is a metric dynamical system [35].
Now, let us introduce the following O-U process, defined on (Ω,F,P,{θt}t∈R) as the random variable given by
z∗(θtω)=−λ∫0−∞eλsθtω(s)ds,t∈R,ω∈Ω,λ>0, |
which solves the Langevin equation [35,36]
dz∗=−λz∗dt+dω(t),t∈R, |
where λ>0 is a mean reversion constant that represents the strength with which the process is attracted by the mean or, in other words, how strongly our system reacts under some perturbation. There are some important properties [28,35,36,37] of the O-U process:
Proposition A.1. If there exists a θt-invariant set ˜Ω∈F of Ω of full P-measure, then
● for a.e. ω∈˜Ω and every λ>0,
limt→∞1t|z∗(θtω)|=0, |
limt→∞1t∫t0z∗(θsω)ds=0, |
limt→∞1t∫t0|z∗(θsω)|ds=E[z∗(θtω)]<∞; |
● for a.e. ω∈˜Ω and all t∈R,
limλ→∞z∗(θtω)=0, |
limλ→0λz∗(θtω)=0. |
Let (Ω,F,P) be a probability space and (X,‖⋅‖X) be a separable Banach space. The following definitions about the RDS can be found in [35,38].
Definition A.1. An RDS on X consists of two ingredients: (a) a metric dynamical system (Ω,F,P,{θt}t∈R) with a family of mappings θ:Ω→Ω such that
● θ0=dΩ,
● θt+s=θt∘θs for all t,s∈R,
● the mapping (t,ω)↦θtω is measurable, and
● the probability measure P is preserved by θt, i.e., θtP=P;
and (b) a mapping ψ:[0,+∞)×Ω×X→X which is (B([0,+∞))×F×B(X),B(X)) -measurable, such that for a.e. ω∈Ω,
● the mapping φ(t,ω):X→X,x↦φ(t,ω)x is continuous for every t≥0,
● φ(0,ω) is the identity operator on X, and
● φ(t+s,ω)=φ(t,θsω)φ(s,ω) for all t,s≥0.
Definition A.2. A random set K is a measurable subset of X×Ω with respect to the product σ-algebra B(X)×F. Moreover, K will be called a closed or a compact random set if K(ω)={x:(x,ω)∈K}, ω∈Ω, is closed or compact for P-almost all ω∈Ω, respectively.
Definition A.3. A bounded random set K(ω)⊂X is said to be tempered with respect to {θt}t∈R if for a.e. ω∈Ω and all λ>0,
limt→∞e−βtsupx∈K(θ−tω)‖x‖X=0. |
Definition A.4. A random set B(ω)⊂X is called a random absorbing set in E(X), if for any E∈E(X) and a.e. ω∈Ω, there exists TE(ω)>0 such that for all t≥TE(ω),
φ(t,θ−tω)E(θ−tω))∈B(ω). |
Definition A.5. Let {φ(t,ω)}t≥0,ω∈Ω be an RDS over (Ω,F,P,{θt}t∈R) with state space X, and let A(ω) be a random set. Then, A={A(ω)}ω∈Ω is called a global random E-attractor (or pullback E-attractor) for {φ(t,ω)}t≥0,ω∈Ω if
● A(ω) is a compact set of X for a.e. ω∈Ω;
● φ(t,ω)A(ω)=A(θtω) holds for a.e. ω∈Ω and all t≥0;
● for a.e. ω∈Ω and any E∈E(X),
limt→∞distX(φ(t,θ−tω)E(θ−tω),A(ω))=0, |
where dist(G,H)X=supg∈Ginfh∈H‖g−h‖X is the Hausdorff semi-metric for G,H⊆X.
Proposition A.2. [39,40] Let B∈E(X) be an absorbing set for the continuous RDS {φ(t,ω)}t≥0,ω∈Ω which is closed and satisfies the asymptotic compactness condition for a.e. ω∈Ω, i.e., each sequence xn∈φ(tn,θ−tnω)B(θ−tnω) has a convergent subsequence in X when tn→∞. Then, φ has a unique global random attractor A={A(ω)}ω∈Ω with component subsets
A(ω)=∩τ≥TB(ω)¯⋃t≥τφ(t,θ−tω)B(θ−tω). |
Proposition A.3. [21] Let φu be an RDS on X. Suppose that the mapping T:Ω×X→X possesses the following properties:
● for fixed ω∈Ω, the mapping T(ω,⋅) is a homeomorphism on X;
● for fixed x∈X, the mappings T(⋅,x) and T−1(⋅,x) are measurable.
Then, the mapping
(t,ω,x)→φ(t,ω)x:=T−1(θtω,φ(t,ω)T(ω,x)) |
is a conjugated RDS.
[1] |
S. Chakraborty, S. Chatterjee, E. Venturino, J. Chattopadhyay, Recurring plankton ploom dynamics modeled via toxin-producing phytoplankton, J. Biol. Phys., 33 (2007), 271–290. https://doi.org/10.1007/s10867-008-9066-3 doi: 10.1007/s10867-008-9066-3
![]() |
[2] |
C. Subhendu, P. K. Tiwari, A. K. Misra, J. Chattopadhyay, Spatial dynamics of a nutrient-phytoplankton system with toxic effect on phytoplankton, Math. Biosci., 264 (2015), 94–100. https://doi.org/10.1016/j.mbs.2015.03.010 doi: 10.1016/j.mbs.2015.03.010
![]() |
[3] |
S. Zhao, S. Yuan, T. Zhang, The impact of environmental fluctuations on a plankton model with toxin-producing phytoplankton and patchy agglomeration, Chaos Soliton. Fract., 162 (2022), 112426. https://doi.org/10.1016/j.chaos.2022.112426 doi: 10.1016/j.chaos.2022.112426
![]() |
[4] |
E. J. Philips, S. Badylak, S. Youn, K. Kelley, The occurrence of potentially toxic dinoflagellates and diatoms in a subtropical lagoon, the Indian river lagoon, Florida, USA, Harmful Algae, 3 (2004), 39–49. https://doi.org/10.1016/j.hal.2003.08.003 doi: 10.1016/j.hal.2003.08.003
![]() |
[5] |
S. P. Colin, H. G. Dam, Effects of the toxic dinoflagellate Alexandrium fundyense on the copepod Acartia hudsonica: A test of the mechanisms that reduce ingestion rates, Mar. Ecol. Prog. Ser., 248 (2003), 55–65. https://doi.org/10.3354/meps248055 doi: 10.3354/meps248055
![]() |
[6] |
Y. Peng, Y. Li, T. Zhang, Global bifurcation in a toxin producing phytoplankton-zooplankton system with prey-taxis, Nonlinear Anal.-Real., 61 (2021), 103326. https://doi.org/10.1016/j.nonrwa.2021.103326 doi: 10.1016/j.nonrwa.2021.103326
![]() |
[7] |
L. E. Schmidt, P. J. Hansen, Allelopathy in the prymnesiophyte Chrysochromulina polylepis: Effect of cell concentration, growth phase and pH, Mar. Ecol.: Prog. Ser., 216 (2001), 67–81. https://doi.org/10.3354/meps216067 doi: 10.3354/meps216067
![]() |
[8] |
F. Rao, Spatiotemporal dynamics in a reaction-diffusion toxic-phytoplankton-zooplankton model, J. Stat. Mech.: Theory Exp., 2013 (2013), 08014. https://doi.org/10.1088/1742-5468/2013/08/P08014 doi: 10.1088/1742-5468/2013/08/P08014
![]() |
[9] |
J. Chattopadhyay, E. Venturino, S. Chatterjee, Aggregation of toxin-producing phytoplankton acts as a defencemechanism-a model-based study, Math. Comput. Model. Dyn., 19 (2013), 159–174. https://doi.org/10.1080/13873954.2012.708876 doi: 10.1080/13873954.2012.708876
![]() |
[10] |
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
![]() |
[11] |
T. Scotti, M. Mimura, J. Y. Wakano, Avoiding toxic prey may promote harmful algal blooms, Ecol. Complex, 21 (2015), 157–165. https://doi.org/10.1016/j.ecocom.2014.07.004 doi: 10.1016/j.ecocom.2014.07.004
![]() |
[12] |
S. Jang, J. Baglama, L. Wu, Dynamics of phytoplankton-zooplankton systems with toxin producing phytoplankton, Appl. Math. Comput., 227 (2014), 717–740. https://doi.org/10.1016/j.amc.2013.11.051 doi: 10.1016/j.amc.2013.11.051
![]() |
[13] |
C. Ji, D. Jiang, N. Shi, Analysis of a predator-prey model with modified Leslie-Gower and Holling-type Ⅱ schemes with stochastic perturbation, J. Math. Anal. Appl., 359 (2009), 482–498. https://doi.org/10.1016/j.jmaa.2009.05.039 doi: 10.1016/j.jmaa.2009.05.039
![]() |
[14] |
S. Zhang, X. Meng, T. Feng, T. Zhang, Dynamics analysis and numerical simulations of a stochastic non-autonomous predator-prey system with impulsive effects, Nonlinear Anal.-Hybrid Syst., 26 (2017), 19–37. https://doi.org/10.1016/j.nahs.2017.04.003 doi: 10.1016/j.nahs.2017.04.003
![]() |
[15] |
Q. Luo, X. Mao, Stochastic population dynamics under regime switching, J. Math. Anal. Appl., 334 (2007), 69–84. https://doi.org/10.1016/j.jmaa.2006.12.032 doi: 10.1016/j.jmaa.2006.12.032
![]() |
[16] |
M. Liu, K. Wang, Persistence and extinction of a stochastic single-specie model under regime switching in a polluted environment Ⅱ, J. Theor. Biol., 267 (2010), 283–291. https://doi.org/10.1016/j.jtbi.2010.08.030 doi: 10.1016/j.jtbi.2010.08.030
![]() |
[17] |
M. Liu, K. Wang, Persistence, extinction and global asymptotical stability of a non-autonomous predator-prey model with random perturbation, Appl. Math. Modell., 36 (2012), 5344–5353. https://doi.org/10.1016/j.apm.2011.12.057 doi: 10.1016/j.apm.2011.12.057
![]() |
[18] |
X. Yu, S. Yuan, T. Zhang, The effects of toxin-producing phytoplankton and environmental fluctuations on the planktonic blooms, Nonlinear Dyn., 91 (2018), 1653–1668. https://doi.org/10.1007/s11071-017-3971-6 doi: 10.1007/s11071-017-3971-6
![]() |
[19] |
X. Yu, S. Yuan, T. 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
![]() |
[20] |
Q. Zhao, S. Liu, X. Niu, Stationary distribution and extinction of a stochastic nutrien-phytoplankton-zooplankton model with cell size, Math. Methods Appl. Sci., 43 (2020), 3886–3902. https://doi.org/10.1002/mma.6114 doi: 10.1002/mma.6114
![]() |
[21] | T. Caraballo, X. Han, Applied Nonautonomous and Random Dynamical Systems: Applied Dynamical Systems, Springer, Berlin, 2016. |
[22] |
T. Caraballo, R. Colucci, J. López-De-La-Cruz, A. Rapaport, A way to model stochastic perturbations in population dynamics models with bounded realizations, Commun. Nonlinear Sci., 77 (2019), 239–257. https://doi.org/10.1016/j.cnsns.2019.04.019 doi: 10.1016/j.cnsns.2019.04.019
![]() |
[23] |
X. Zhang, R. Yuan, Pullback attractor for random chemostat model driven by colored noise, Appl. Math. Lett., 112 (2021), 106833. https://doi.org/10.1016/j.aml.2020.106833 doi: 10.1016/j.aml.2020.106833
![]() |
[24] |
L. F. de Jesus, C. M. Silva, H. Vilarinho, Random perturbations of an eco-epidemiological model, Discrete Contin. Dyn.-Ser. B, 27 (2022), 257–275. https://doi.org/10.3934/dcdsb.2021040 doi: 10.3934/dcdsb.2021040
![]() |
[25] |
J. López-de-la-Cruz, Random and stochastic disturbances on the input flow in chemostat models with wall growth, Stoch. Anal. Appl., 37 (2019), 668–698. https://doi.org/10.1080/07362994.2019.1605911 doi: 10.1080/07362994.2019.1605911
![]() |
[26] |
T. Caraballo, R. Colucci, X. Han, Predation with indirect effects in fluctuating environments, Nonlinear Dyn., 84 (2016), 115–126, https://doi.org/10.1007/s11071-015-2238-3 doi: 10.1007/s11071-015-2238-3
![]() |
[27] |
T. Caraballo, M. J. Garrido-Atienza, J. López-de-la-Cruz, Dynamics of some stochastic chemostat models with multiplicative noise, Commun. Pur. Appl. Anal., 16 (2017), 1893–1914. https://doi.org/10.3934/cpaa.2017092 doi: 10.3934/cpaa.2017092
![]() |
[28] |
X. Zhang, R. Yuan, Forward attractor for stochastic chemostat model with multiplicative noise, Chaos, Solitons Fractals, 153 (2021), 111585. https://doi.org/10.1016/j.chaos.2021.111585 doi: 10.1016/j.chaos.2021.111585
![]() |
[29] |
D. Wu, H. Wang, S. Yuan, Stochastic sensitivity analysis of noise-induced transitions in a predatorprey model with environmental toxins, Math. Biosci. Eng., 16 (2019), 2141–2153. https://doi.org/10.3934/mbe.2019104 doi: 10.3934/mbe.2019104
![]() |
[30] |
O. E. Barndorff-Nielsen, N. Shephard, Non-Gaussian Ornstein-Uhlenbeck-based models and some of their uses in financial economics, R. Stat. Soc., 63 (2001), 167–241. https://doi.org/10.1111/1467-9868.00282 doi: 10.1111/1467-9868.00282
![]() |
[31] |
X. Mu, D. Jiang, T. Hayat, A. Alsaedi, Y. Liao, A stochastic turbidostat model with Ornstein-Uhlenbeck process: dynamics analysis and numerical simulations, Nonlinear Dyn., 107 (2022), 2805–2817. https://doi.org/10.1007/s11071-021-07093-9 doi: 10.1007/s11071-021-07093-9
![]() |
[32] |
B. Zhou, D. Jiang, T. Hayat, Analysis of a stochastic population model with mean-reverting Ornstein-Uhlenbeck process and Allee effects, Commun. Nonlinear Sci. Numer. Simul., 111 (2022), 106450. https://doi.org/10.1016/j.cnsns.2022.106450 doi: 10.1016/j.cnsns.2022.106450
![]() |
[33] |
Q. Liu, D. Jiang, Analysis of a stochastic logistic model with diffusion and Ornstein-Uhlenbeck process, J. Math. Phys., 63 (2022), 053505. https://doi.org/10.1063/5.0082036 doi: 10.1063/5.0082036
![]() |
[34] |
T. Caraballo, M. Garrido-Atienza, J. López-De-La-Cruz, A. Rapaport, Modeling and analysis of random and stochastic input flows in the chemostat model, Discrete Contin. Dyn.–Ser. B, 24 (2019), 3591–3614. https://doi.org/10.3934/dcdsb.2018280 doi: 10.3934/dcdsb.2018280
![]() |
[35] | L. Arnold, Random Dynamical Systems, Springer, Berlin, 1998. |
[36] |
T. Caraballo, P. E. Kloeden, B. Schmalfuss, Exponentially stable stationary solutions for stochastic evolution equations and their perturbation, Appl. Math. Optim., 50 (2004), 183–207. https://doi.org/10.1007/s00245-004-0802-1 doi: 10.1007/s00245-004-0802-1
![]() |
[37] |
S. Al-Azzawi, J. Liu, X. Liu, Convergence rate of synchronization of systems with additive noise, Discrete Contin. Dyn.-Ser. B, 22 (2017), 227–245. https://doi.org/10.3934/dcdsb.2017012 doi: 10.3934/dcdsb.2017012
![]() |
[38] |
H. Crauel, P. Kloeden, Nonautonomous and random attractors, Jahresber. Deutsch. Math., 117 (2015), 173–206. https://doi.org/10.1365/s13291-015-0115-0 doi: 10.1365/s13291-015-0115-0
![]() |
[39] |
T. Caraballo, G. Lukaszewicz, J. Real, Pullback attractors for asymptotically compact non-autonomous dynamical systems, Nonlinear Anal.-Theory, Methods Appl., 64 (2006), 484–498. https://doi.org/10.1016/j.na.2005.03.111 doi: 10.1016/j.na.2005.03.111
![]() |
[40] |
F. Flandoli, B. Schmalfuss, Random attractors for the 3D stochastic navier-stokes equation with multiplicative white noise, Stoch. Stoch. Rep., 59 (1996), 21–45. https://doi.org/10.1080/17442509608834083 doi: 10.1080/17442509608834083
![]() |
1. | Francisco Rodríguez, Fatima Z. Lachachi, Ikram Medjahdi, M. Ángeles Castro, Juan Carlos Cortés, A note on the application of the RVT method to general classes of single-species population models formulated by random differential equations, 2024, 43, 2238-3603, 10.1007/s40314-024-02809-0 |