Symbol | Value | Unit | Source |
a | 0.01 | g/g/h | Estimated |
f | 0.4 | m3/m3/h | Estimated |
μ | 0.015 | g/g/h | Estimated |
Aup | 1 | g/m3 | [29] |
h | 0.1 | m | [2,30] |
c | 0.1 | m3/g/h | [31,32] |
β | 0.2 | g/g | [33] |
k | 150 | g/m2 | [5] |
In this paper, we develop two stochastic mussel-algae models: one is autonomous and the other is periodic. For the autonomous model, we provide sufficient conditions for the extinction, nonpersistent in the mean and weak persistence, and demonstrate that the model possesses a unique ergodic stationary distribution by constructing some suitable Lyapunov functions. For the periodic model, we testify that it has a periodic solution. The theoretical findings are also applied to practice to dissect the effects of environmental perturbations on the growth of mussel.
Citation: Dengxia Zhou, Meng Liu, Ke Qi, Zhijun Liu. Long-time behaviors of two stochastic mussel-algae models[J]. Mathematical Biosciences and Engineering, 2021, 18(6): 8392-8414. doi: 10.3934/mbe.2021416
[1] | Yan Xie, Zhijun Liu, Ke Qi, Dongchen Shangguan, Qinglong Wang . A stochastic mussel-algae model under regime switching. Mathematical Biosciences and Engineering, 2022, 19(5): 4794-4811. doi: 10.3934/mbe.2022224 |
[2] | 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 |
[3] | Sanling Yuan, Xuehui Ji, Huaiping Zhu . Asymptotic behavior of a delayed stochastic logistic model with impulsive perturbations. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1477-1498. doi: 10.3934/mbe.2017077 |
[4] | Ying He, Junlong Tao, Bo Bi . Stationary distribution for a three-dimensional stochastic viral infection model with general distributed delay. Mathematical Biosciences and Engineering, 2023, 20(10): 18018-18029. doi: 10.3934/mbe.2023800 |
[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] | 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 |
[7] | Ke Qi, Zhijun Liu, Lianwen Wang, Qinglong Wang . Survival and stationary distribution of a stochastic facultative mutualism model with distributed delays and strong kernels. Mathematical Biosciences and Engineering, 2021, 18(4): 3160-3179. doi: 10.3934/mbe.2021157 |
[8] | Haolei Gu, Kedong Yin . Forecasting algae and shellfish carbon sink capability on fractional order accumulation grey model. Mathematical Biosciences and Engineering, 2022, 19(6): 5409-5427. doi: 10.3934/mbe.2022254 |
[9] | Frédéric Mazenc, Michael Malisoff, Patrick D. Leenheer . On the stability of periodic solutions in the perturbed chemostat. Mathematical Biosciences and Engineering, 2007, 4(2): 319-338. doi: 10.3934/mbe.2007.4.319 |
[10] | Katarzyna Pichór, Ryszard Rudnicki . Stochastic models of population growth. Mathematical Biosciences and Engineering, 2025, 22(1): 1-22. doi: 10.3934/mbe.2025001 |
In this paper, we develop two stochastic mussel-algae models: one is autonomous and the other is periodic. For the autonomous model, we provide sufficient conditions for the extinction, nonpersistent in the mean and weak persistence, and demonstrate that the model possesses a unique ergodic stationary distribution by constructing some suitable Lyapunov functions. For the periodic model, we testify that it has a periodic solution. The theoretical findings are also applied to practice to dissect the effects of environmental perturbations on the growth of mussel.
A nail-cap-sized zebra mussel was first discovered in the US waters in 1988 and has a powerful reproductive capacity [1]. The invasion of zebra mussel has caused great inconvenience to people, such as blocking pipes, polluting water sources, crowding out local species and causing serious economic losses. According to an estimation from the Center for Invasive Species Research at UC Riverside [1], the US spends as much as 500 million dollars every year to manage mussel in the Great Lakes. As a result, many biologists, ecologists, and mathematicians have studied the invasion of mussel from different perspectives.
The growth and survival of mussel depend heavily on the availability of food sources for algae. A lot of literatures have revealed that the food supply of algae can limit mussel intake [2,3,4]. In order to uncover the relationships between mussel and algae, Koppel et al. [5] proposed a diffusive mussel-algae model, considering the corresponding nondiffusive form:
{dM(t)dt=βcA(t)M(t)−μkk+M(t)M(t),dA(t)dt=(Aup−A(t))f−chA(t)M(t), | (1.1) |
where M(t) and A(t) respectively denote the size of mussel and algae, β represents the conversion rate of ingested algae to mussel production, c is the consumption constant, μ is the maximal per capita mussel death rate, k stands for the value of M(t) at which mortality is half maximal, Aup denotes the concentration of algae in the upper water layer, f describes the rate of exchange between the lower and upper water layers, h is the height of the lower water layer. All the parameters are positive.
The research on model (1.1) has attracted much attention. For example, based on model (1.1), Koppel et al. [5] analyzed the scale-dependent feedback and regular spatial patterns of young mussel beds, and uncovered that the self-organization patterns would affect the emergent properties of ecosystems in large-scale space. Cangelosi et al. [6] established a mussel-algae model with Turing patterns and carried out a series of stability analyses. Song et al. [7] dissected Turing-Hopf bifurcation of model (1.1) with reaction-diffusion. Similar diffusive models to study the spatial dynamics of mussel-algae can be found in [8,9,10,11]. In addition, quite a few researchers pay attention on the control of mussel-algae. A model describing mussel bed appearance was proposed in [12] to explore the habitat suitability analysis for littoral mussel beds in the Dutch Wadden Sea. The effects of 19 macroalgal species on the settlement and metamorphosis of the mussel were investigated [13].
Considering that the growth of mussel is affected by intraspecific competition, we transform model (1.1) into the following model:
{dM(t)dt=βcA(t)M(t)−aM2(t)−μkk+M(t)M(t),dA(t)dt=(Aup−A(t))f−chA(t)M(t), | (1.2) |
where a is the intraspecific competition strength of mussel and positive. Other parameters are defined in the same as in model (1.1).
Note that the above studies are all deterministic models. However, environmental uncertainties are ubiquitous in aquatic ecosystems, the populations are always inevitably influenced by environmental noises, which is a momentous element in ecosystems [14]. Environmental stochasticity may involve water temperature, noise, salinity, depth and predators, which might affect the growth and evolution of the populations. Accordingly, stochastic models are usually more realistic, and it is essential to bring environmental stochasticity into model (1.2). Quite a few existing literatures focus on this and obtain excellent results, e.g., survival analysis [15], asymptotic stability [16], stationary distribution [17], optimal harvesting [18,19] and so on. However, as we know, a very little bit of work has been done with stochastic mussel-algae models, especially the corresponding stochastic version of model (1.2).
For M(t) and A(t) in model (1.2), given Δt>0 is a fixed step size. Define ΓΔt(pΔt)=(MΔt(pΔt), AΔt(pΔt)), p=0,1,2,…. Let a normal distribution random variable sequence {ΘΔti(p)}∞p=0 satisfy E[ΘΔti(p)]=0, E[ΘΔti(p)]2=σ2iΔt, i=1,2, where the constants σ21 and σ22 reflect the size of the random perturbations. In each time period [pΔt,(p+1)Δt], we hypothesize that ΓΔt grows in the light of the discrete modification of model (1.2) as well as a stochastic amount (MΔt(pΔt)ΘΔt1(p), AΔt(pΔt)ΘΔt2(p)), then we get
{MΔt((p+1)Δt)=MΔt(pΔt)+[βcAΔt(pΔt)MΔt(pΔt)−a(MΔt(pΔt))2MΔt((p+1)Δt)=−μkk+MΔt(pΔt)MΔt(pΔt)]Δt+MΔt(pΔt)ΘΔt1(p),AΔt((p+1)Δt)=AΔt(pΔt)+[(Aup−AΔt(pΔt))f−chAΔt(pΔt)MΔt(pΔt)]Δt+AΔt(pΔt)ΘΔt2(p). |
On the basis of [20] (Theorem 7.1 and Lemma 8.2), as Δt→0, ΓΔt converges weakly to the solution of the following stochastic differential equation:
{dM(t)=[βcA(t)M(t)−aM2(t)−μkk+M(t)M(t)]dt+σ1M(t)dB1(t),dA(t)=[(Aup−A(t))f−chA(t)M(t)]dt+σ2A(t)dB2(t), | (1.3) |
where B1(t) and B2(t) are independent standard Brownian motions defined on a complete probability space (Ω,F,{F}t≥0,P).
The effects of a periodically varying environment are important as populations evolve influenced by external effects, for example, seasonal changes, food supply, living habits and other factors, which changes significantly through the whole life of populations. This idea has found much attention and is incorporated into dynamical models [21,22,23,24,25]. Till date, to investigate whether these models will exist period solutions or not is still worth noting. Keeping given this fact, model (1.3) may need to be extended into the following periodic version:
{dM(t)=[βc(t)A(t)M(t)−a(t)M2(t)−μ(t)k(t)k(t)+M(t)M(t)]dt+σ1(t)M(t)dB1(t),dA(t)=[(Aup(t)−A(t))f(t)−c(t)hA(t)M(t)]dt+σ2(t)A(t)dB2(t), | (1.4) |
where the coefficients c(t), a(t), μ(t), k(t), Aup(t), f(t) are positive continuous T-periodic functions.
It is well known that stability is one of the key topics in mathematical biology. For autonomous stochastic population models, scholars are concerned with the stable "stochastic positive equilibrium"---stationary distribution. For periodic stochastic population models, positive periodic solution is an attractive concept. To the best of our knowledge, however, both the stationary distribution of model (1.3) and the existence of periodic solution of model (1.4) have not been considered. The objectives of this paper are to test these two issues. The rest arrange of this paper is as follows. In the next section, the existence and uniqueness of the global positive solution are testified. In Section 3, the extinction, nonpersistent in the mean and weak persistence of model (1.3) are probed. Section 4 provides the conditions under which model (1.3) possesses a unique ergodic stationary distribution. In Section 5, we explore the existence of T-periodic solution of model (1.4). To illustrate the theoretical findings, some numerical simulations are given in Section 6. A few biological meanings of conditions and results are discussed to end Section 7.
Theorem 2.1. For arbitrary initial data (M(0),A(0))∈R2+, model (1.3) has a unique global positive solution with probability one.
Proof. Recalling model (1.3), assign M(t)=e˜M(t), we obtain
{d˜M(t)=[βcA(t)−σ212−ae˜M(t)−μkk+e˜M(t)]dt+σ1dB1(t),dA(t)=[(Aup−A(t))f−chA(t)e˜M(t)]dt+σ2A(t)dB2(t) | (2.1) |
with (˜M(0),A(0))=(lnM(0),A(0)). One can see that the coefficients of model (2.1) obey the locally Lipschitz continuous conditions, as a result, it possesses a unique solution (˜M(t),A(t)) on [0,τe), where τe≤+∞. Accordingly, model (1.3) possesses a unique positive solution (M(t),A(t))=(e˜M(t),A(t)) on [0,τe). To finish the proof, we only need to testify that τe=+∞ a.s. Choose an integer n0>0 which obeys that 1/n0≤M(0),A(0)≤n0. For every n≥n0, define
τn=inf{t∈[0,τe]:min{M(t),A(t)}≤1/normax{M(t),A(t)}≥n}. |
Set τ∞=limn→+∞τn. As a result, τ∞≤τe. Now we only need to testify that τ∞=+∞. If it is not true, then one can find two constants T>0 and ϵ∈(0,1) such that P{τ∞≤T}>ϵ. As a result, one can set an integer n1≥n0 which satisfies
P{τn≤T}≥ϵ. | (2.2) |
Define
V(M,A)=M+hβA. |
Taking advantage of Itô's formula, one has
dV(M,A)=LV(M,A)dt+Mσ1dB1(t)+hβAσ2dB2(t), | (2.3) |
where
LV(M,A)=(βcAM−aM2−μkk+MM)+hβ[(Aup−A)f−chAM]LV(M,A)=βcAM−aM2−μkk+MM+hβAupf−hβfA−βcAMLV(M,A)≤hβAupf=G. |
Integrating both sides of Eq (2.3) from 0 to τn∧T yields
∫τn∧T0dV(M,A)≤∫τn∧T0Gdt+∫τn∧T0Mσ1dB1(t)+∫τn∧T0hβAσ2dB2(t). |
Taking expectation on both sides results in
EV(M(τn∧T),A(τn∧T))≤V(M(0),A(0))+GE(τn∧T)EV(M(τn∧T),A(τn∧T))≤V(M(0),A(0))+GT. | (2.4) |
Set Ωn={τn≤T} for n≥n1. According to Eq (2.2), P(Ωn)≥ϵ. For any θ∈Ωn, at least one of M(τn,θ),A(τn,θ) equals to n or 1/n. Thus, we derive
V(M(τn,θ),A(τn,θ))≥(n−1−lnn)∧(1n−1−ln1n). |
Therefore, Eq (2.4) implies that
V(M(0),A(0))+GT≥E[1Ωn(θ)V(M(τn,θ),A(τn,θ))]V(M(0),A(0))+GT≥ε[(n−1−lnn)∧(1n−1−ln1n)], |
where 1Ωn denotes the indicator function of Ωn. Letting n→+∞ causes the contradiction:
V(M(0),A(0))+GT>+∞. |
This finishes the proof.
Remark 2.1. Similar to the proof of Theorem 2.1, one can testify that model (1.4) has a unique global positive solution with probability one, and the details are left out.
Lemma 3.1. Given (M(0),A(0))∈R2+, model (1.3) admits lim supt→+∞[M(t)+βhA(t)]<+∞ and
limt→+∞1t∫t0σ1M(s)dB1(s)=0, limt→+∞1t∫t0σ2A(s)dB2(s)=0 a.s. |
Proof. Denote Z(t)=M(t)+βhA(t). From model (1.3), we have
dZ=(βcAM−aM2−μkk+MM+βh[(Aup−A)f−chAM])dtdZ=+σ1M(t)dB1(t)+βhσ2A(t)dB2(t)dZ≤[a(−2M+1)+βhAupf−βhfA]dt+σ1M(t)dB1(t)+βhσ2A(t)dB2(t)dZ≤(a+βhAupf−δZ)dt+σ1M(t)dB1(t)+βhσ2A(t)dB2(t), |
where δ=min{2a,f}>0. Consider
{dY=(a+βhAupf−δY)dt+σ1M(t)dB1(t)+βhσ2A(t)dB2(t),dY(0)=(M(0),A(0)). | (3.1) |
The solution of model (3.1) is
Y(t)=a+βhAupfδ+[Y(0)−a+βhAupfδ]e−δt+N(t), |
where
N(t)=σ1∫t0e−δ(t−s)M(s)dB1(s)+βhσ2∫t0e−δ(t−s)A(s)dB2(s) |
is a local martingale satisfying N(0)=0 a.s. Thus
Y(t)=Y(0)+Q(t)−P(t)+N(t), |
where
Q(t)=a+βhAupfδ(1−e−δt), P(t)=Y(0)(1−e−δt) |
with Q(0)=P(0)=0. Clearly, Q(t) and P(t) are continuous increasing functions. By [26], we have limt→+∞Y(t)<+∞ a.s., then by stochastic comparison theorem, one has lim supt→+∞Z(t)<+∞ a.s.
Let N1=∫t0σ1M(s)dB1(s) and N2=∫t0σ2A(s)dB2(s). Through calculation, we obtain
⟨N1,N1⟩(t)=σ21∫t0M2(s)ds, |
then
limt→+∞∫t0σ21M2(s)ds(1+s)2≤σ21supt≥0{M2(t)}<+∞. |
In light of [27], limt→+∞t−1N1(t)=0 a.s. Similarly, we have limt→+∞t−1N2(t)=0.
Theorem 3.1. If λ0=βcAup−μ−σ21/2<0 and a>μ/k, then M(t) is extinct a.s.
Proof. We deduce from model (1.3) that
d(1hM+βA)=1h(βcAM−aM2−μkk+MM)dt+β[(Aup−A)f−chAM]dtd(1hM+βA)=+σ1hMdB1(t)+βσ2AdB2(t)d(1hM+βA)=[βchAM−ahM2−μkh(k+M)M+β(Aup−A)f−βchAM]dtd(1hM+βA)=+σ1hMdB1(t)+βσ2AdB2(t)d(1hM+βA)=(βAupf−βfA−ahM2−μkh(k+M)M)dt+σ1hMdB1(t)+βσ2AdB2(t), |
which implies that
βAupf−βft∫t0A(s)ds−aht∫t0M2(s)ds−μht∫t0kM(s)k+M(s)ds=φ1(t)t, | (3.2) |
where
φ1(t)=1hM(t)−1hM(0)+βA(t)−βA(0)−1h∫t0σ1M(s)dB1(s)−β∫t0σ2A(s)dB2(s) |
satisfying limt→+∞φ1(t)/t=0. In light of Eq (3.2), we have
1t∫t0A(s)ds=Aup−aβfht∫t0M2(s)ds−μβfht∫t0kM(s)k+M(s)ds−φ1(t)βft. | (3.3) |
By the first equation of model (1.3) and using Itô's formula, we obtain
dlnM(t)=(βcA−aM−μkk+M−12σ21)dt+σ1dB1(t), |
then together with Eq (3.3), one has
1tlnM(t)M(0)=βct∫t0A(s)ds−at∫t0M(s)ds−12σ21−μt∫t0kk+M(s)ds+1t∫t0σ1dB1(s)1tlnM(t)M(0)=βc(Aup−aβfht∫t0M2(s)ds−μβfht∫t0kM(s)k+M(s)ds−φ1(t)βft)1tlnM(t)M(0)=−at∫t0M(s)ds−12σ21−μ+1t∫t0μM(s)k+M(s)ds+1t∫t0σ1dB1(s)1tlnM(t)M(0)=βcAup−μ−12σ21−acfht∫t0M2(s)ds−μcfht∫t0kM(s)k+M(s)ds−cφ1(t)ft1tlnM(t)M(0)=−1t∫t0(a−μk+M(s))M(s)ds+1t∫t0σ1dB1(s). | (3.4) |
Since the strong law of numbers implies that
limt→+∞1t∫t0σ1dB1(s)=0. | (3.5) |
Thus, it follows from Eqs (3.4) and (3.5) that
limt→+∞t−1lnM(t)≤βcAup−μ−12σ21<0, |
which implies the required assertion.
Theorem 3.2. If λ0=0 and a>μ/k, then M(t) is nonpersistent in the mean a.s., namely, limt→+∞t−1∫t0M(s)ds=0 a.s.
Proof. Let ρ>0 be a constant which satisfies ρ<a−μ/k. From Eq (3.4), we have
lnM(t)−lnM(0)=λ0t−acfh∫t0M2(s)ds−μcfh∫t0kM(s)k+M(s)ds−cφ1(t)flnM(t)−lnM(0)=−∫t0(a−μk+M(s))M(s)ds+∫t0σ1dB1(s)lnM(t)−lnM(0)≤−∫t0ρM(s)ds+∫t0σ1dB1(s). | (3.6) |
Note that for any ε>0, there is T>0 such that for t≥T,
t−1lnM(0)≤ε/2, t−1∫t0σ1dB1(s)≤ε/2. | (3.7) |
Substituting Eq (3.7) into Eq (3.6), we have
lnM(t)≤εt−ρ∫t0M(s)ds, t≥T. |
Set ϱ(t)=∫t0M(s)ds, then we get
ln(dϱ(t)/dt)≤εt−ρϱ(t). |
Hence for t>T, we have
eρϱ(t)(dϱ(t)/dt)≤eεt. |
Integrating this inequality from T to t, one can derive that
ρ−1(eρϱ(t)−eρϱ(T))≤ε−1(eεt−eεT). |
That is,
eρϱ(t)≤eρϱ(T)+ρε−1eεt−ρε−1eεT. | (3.8) |
Taking the logarithm of both sides of Eq (3.8) results in
ϱ(t)≤ρ−1ln(eρϱ(T)+ρε−1eεt−ρε−1eεT). |
Note that ϱ(t)=∫t0M(s)ds, then one can obtain that
lim supt→+∞t−1∫t0M(s)ds≤ρ−1lim supt→+∞t−1ln{eρϱ(T)+ρε−1eεt−ρε−1eεT}. |
Applying L'Hospital's rule leads to
lim supt→+∞t−1∫t0M(s)ds≤ε/ρ. |
It then follows from the arbitrariness of ε that lim supt→+∞t−1∫t0M(s)ds≤0. This proof is complete.
Theorem 3.3. If λ0>0, then M(t) is weakly persistent a.s., namely, lim supt→+∞M(t)>0 a.s.
Proof. We first testify that
lim supt→+∞t−1lnM(t)≤0 a.s. | (3.9) |
By Itô's formula,
d(etlnM)=etlnMdt+etdlnMd(etlnM)=et{[lnM+βcA−aM−μkk+M−12σ21]dt+σ1dB1(t)}. |
Integrating the both sides from 0 to t, we have
etlnM(t)−lnM(0)=∫t0es[lnM(s)+βcA(s)−aM(s)−μkk+M(s)−12σ21]ds+W(t), | (3.10) |
where W(t)=∫t0esσ1dB1(s) is a local martingale with the quadratic form
⟨W(t),W(t)⟩=σ21∫t0e2sds. |
By the exponential martingale inequality (see [26] on page 44), for arbitrary positive constants T0, ι and ν, one has
P{sup0≤t≤T0[W(t)−ι2⟨W(t),W(t)⟩]>ν}≤e−ιν. |
Choose T0=ϑr, ι=e−ϑr and ν=ϖeϑrlnr, then we obtain
P{sup0≤t≤ϑr[W(t)−0.5e−ϑr⟨W(t),W(t)⟩]>ϖeϑrlnr}≤r−ϖ, |
where ϖ>1, ϑ>0. By the Borel-Cantalli lemma (see [26] on page 7), for almost all ζ∈Ω, there exists a r0(ζ) such that for r≥r0(ζ),
W(t)≤0.5e−ϑr⟨W(t),W(t)⟩+ϖeϑrlnr, 0≤t≤ϑr. | (3.11) |
Combining Eq (3.10) with Eq (3.11), we obtain
etlnM(t)−lnM(0)≤∫t0es[lnM(s)+βcA(s)−aM(s)−12σ21]dsetlnM(t)−lnM(0)≤+0.5e−ϑrσ21∫t0e2sds+ϖeϑrlnretlnM(t)−lnM(0)=∫t0es[lnM(s)+βcA(s)−aM(s)−12σ21+0.5es−ϑrσ21]dsetlnM(t)−lnM(0)≤+ϖeϑrlnr. | (3.12) |
Since lnM(t)+βcA(t)−aM(t)−12σ21+0.5et−ϑrσ21 is bounded, for any 0≤s≤ϑr, there is a constant C independent of r such that
lnM(t)+βcA(t)−aM(t)−12σ21+0.5et−ϑrσ21≤C. | (3.13) |
Substituting Eq (3.13) into Eq (3.12), we obtain
etlnM(t)−lnM(0)≤C[et−1]+ϖeϑrlnr. | (3.14) |
Dividing the both sides of Eq (3.14) by et leads to
lnM(t)≤e−tlnM(0)+C[1−e−t]+ϖe−teϑrlnr. |
Consequently, if ϑ(r−1)≤t≤ϑr and r≥r0(ζ), then one can observe that
t−1lnM(t)≤e−tt−1lnM(0)+Ct−1[1−e−t]+ϖe−ϑ(r−1)eϑrt−1lnr, |
which is the needed assertion Eq (3.9) by letting r→+∞.
Now let us testify lim supt→+∞M(t)>0 a.s. If not, then we denote S={lim supt→+∞M(t)=0}, P(S)>0. In light of Eq (3.4), one has
1tlnM(t)M(0)=λ0−acfht∫t0M2(s)ds−μcfht∫t0kM(s)k+M(s)ds−cφ1(t)ft1tlnM(t)M(0)=−1t∫t0(a−μk+M(s))M(s)ds+1t∫t0σ1dB1(s). |
For all ζ∈S, we have limt→+∞M(t,ζ)=0, and the law of large numbers for local martingales indicates that limt→+∞1t∫t0σ1dB1(s)=0. Thus we have lim supt→+∞t−1lnM(t,ζ)=λ0>0. By Eq (3.9), a contradiction arises.
Now we dissect the stationary distribution for model (1.3) by taking advantage of Has'minskii's results [28]. Denote by X(t) a time-homogeneous Markov process in Rn which obeys
dX(t)=b(X)dt+m∑r=1σr(X)dBr(t). |
Let I(x)=(aij(x)) be the diffusion matrix of X(t), where
aij(x)=m∑r=1σir(x)σjr(x). |
For any C2- function V1(x), define
LV1=l∑i=1bi(x)∂V1(x)∂xi+12l∑i,j=1aij(x)∂2V1(x)∂xi∂xj. |
Lemma 4.1. If there is a bounded domain U⊂Rd with regular boundary such that ([28])
● there is a positive number Λ which obeys
2∑i,j=1aij(x)ξiξj≥Λ|ξ|2, x∈U, ξ=(ξ1,ξ2)∈Rd, |
● there is a nonnegative C2- function V2 such that LV2(x)<−1 for any x∈Rd∖U,
then X(t) admits a unique ESD.
Define
R0=Aupfβc(μ+12σ21)(f+12σ22). |
Theorem 4.1. If λ0>0 and σ22 is sufficiently small such that
σ22<min{2λ0fμ+12σ21, hβf}, |
then model (1.3) admits a unique ESD.
Proof. Considering the function V3(M,A)=−m1lnM−m2lnA, and m1, m2 are positive constants to be chosen later, we obtain
LV3(M,A)=−m1M(βcAM−aM2−μkk+MM)−m2A[(Aup−A)f−chAM]+m12σ21+m22σ22LV3(M,A)=−m1βcA+am1M+μkm1k+M−AupAm2f+m2f+chm2M+m12σ21+m22σ22LV3(M,A)≤−m1βcA−AupAm2f+am1M+μm1+m2f+chm2M+m12σ21+m22σ22LV3(M,A)≤−2√m1m2βcAupf+(μ+σ21/2)m1+(f+σ22/2)m2+(am1+chm2)MLV3(M,A)=−2√βcAupf(μ+σ21/2)(f+σ22/2)+2+(am1+chm2)MLV3(M,A)=−2(√βcAupf(μ+σ21/2)(f+σ22/2)−1)+(am1+chm2)MLV3(M,A)=−2(√R0−1)+(am1+chm2)MLV3(M,A)=−D1+(am1+chm2)M, |
where
m1=1μ+σ21/2, m2=1f+σ22/2, D1=2(√R0−1)>0. |
Define
V4(M,A)=12(M+hβA)2−lnA, |
we have
L(12(M+hβA)2)=(M+hβA)(βcAM−aM2−μkk+MM+hβ[(Aup−A)f−chAM])L(12(M+hβA)2)=+σ212M2+hβ2σ22A2L(12(M+hβA)2)≤(M+hβA)(−aM2+hβAupf−hβfA)+σ212M2+hβ2σ22A2L(12(M+hβA)2)=−aM3+hβAupfM−hβfAM−ahβAM2+h2β2AupfA−h2β2fA2L(12(M+hβA)2)=+σ212M2+hβ2σ22A2L(12(M+hβA)2)≤−aM3+hβAupfM+h2β2AupfA−h2β2fA2+σ212M2+hβ2σ22A2L(12(M+hβA)2)≤−a2M3−h2β2f2A2+D2. |
Notice that h2β2f−hβσ22>0, hence
D2=sup(M,A)∈R2+{−a2M3+σ212M2+hβAupfM−h2β2f2A2+hβ2σ22A2+h2β2AupfA}<+∞. |
In addition,
L(−lnA)=−1A[(Aup−A)f−chAM]+σ222L(−lnA)=−AupfA+f+chM+σ222L(−lnA)=−AupfA+chM+D3, |
where D3=f+σ22/2.
Therefore,
LV4(M,A)=L(12(M+hβA)2)+L(−lnA)LV4(M,A)≤−a2M3−h2β2f2A2−AupfA+chM+D2+D3. |
Now define V5(M,A)=λV3(M,A)+V4(M,A), where λ>0 is sufficiently large. Hence,
lim infq1→+∞,(M,A)∈R2+∖Uq1V5(M,A)=+∞, |
where Uq1=(1q1,q1)×(1q1,q1), q1 is a sufficiently large number. Notice that V5(M,A) is continuous. Thus V5(M,A) has a minimum point (M0,A0) in R2+. Define
V6(M,A)=V5(M,A)−V5(M0,A0). |
Thus, we can get
LV6(M,A)≤λ(−D1+(am1+chm2)M)+(−a2M3−h2β2f2A2−AupfA+chM+D2+D3)LV6(M,A)≤−λD1−a2M3−h2β2f2A2−AupfA+[λ(am1+chm2)+ch]M+D2+D3. |
Define a bounded close set:
U={ε≤M≤1/ε,ε≤A≤1/ε}, |
where 0<ε<1 is sufficient small. We can split R2+∖U into the following four ranges,
U1={M<ε}, U2={A<ε}, U3={M>1/ε}, U4={A>1/ε}. |
Case 1. If (M,A)∈U1, then we have
LV6(M,A)≤−λD1−h2β2f2A2−a2M3−AupfA+[λ(am1+chm2)+ch]ε+D2+D3LV6(M,A)≤−λD1+[λ(am1+chm2)+ch]ε+D2+D3. | (4.1) |
Case 2. If (M,A)∈U2, then one can see that
LV6(M,A)≤−Aupfε−a2M3+[λ(am1+chm2)+ch]M+D2+D3LV6(M,A)≤−Aupfε+F1+D2+D3, | (4.2) |
where
F1=supM∈R+{−a2M3+[λ(am1+chm2)+ch]M}. |
Case 3. If (M,A)∈U3, then one has
LV6(M,A)≤−a4M3−a4M3+[λ(am1+chm2)+ch]M+D2+D3LV6(M,A)≤−a4ε3+F2+D2+D3, | (4.3) |
where
F2=supM∈R+{−a4M3+[λ(am1+chm2)+ch]M}. |
Case 4. If (M,A)∈U4, then we obtain
LV6(M,A)≤−h2β2f2ε2−a2M3+[λ(am1+chm2)+ch]M+D2+D3LV6(M,A)≤−h2β2f2ε2+F1+D2+D3. | (4.4) |
In R2+∖U, let ε be sufficiently small which satisfies
−λD1+[λ(am1+chm2)+ch]ε+D2+D3<−1,−Aupfε+F1+D2+D3<−1,−a4ε3+F2+D2+D3<−1,−h2β2f2ε2+F1+D2+D3<−1. | (4.5) |
It follows from Eqs (4.1)–(4.5) that
sup(M,A)∈R2+∖ULV6(M,A)<−1. | (4.6) |
The diffusion matrix of model (1.3) has the form
I(M,A)=(σ21M200σ22A2). |
Choosing Λ=min(M,A)∈Uq{σ21M2,σ22A2}>0, we have
2∑i,j=1aij(M,A)ξiξj=σ21M2ξ21+σ22A2ξ22≥Λ|ξ|2, ξ=(ξ1,ξ2)∈R2+. | (4.7) |
According to Eqs (4.6), (4.7) and Lemma 4.1 that we complete the proof.
Consider the stochastic periodic equation
dx(t)=v(t,x(t))dt+g(t,x(t))dB(t), | (5.1) |
where v(t) and g(t) are T-periodic functions in t.
Lemma 5.1. If there exists a function V7(t,x)∈C2 which is T-periodic and satisfies the conditions ([28])
● inf|x|>ΘV7(t,x)→∞ as Θ→∞,
● LV7(t,x)≤−1 on the outside of some compact set,
then there exists a periodic solution to Eq (5.1).
Define
R1=⟨(Aupfβc)12⟩T(⟨μ+σ21/2⟩T⟨f+σ22/2⟩T)12. |
Define ⟨g⟩T=1T∫T0g(s)ds, where g(t)∈[0,∞) is an integrable function.
Define gu=maxt∈[0,+∞)g(t), gl=mint∈[0,+∞)g(t), where g(t)∈[0,+∞) is a bounded function.
Theorem 5.1. If R1>1 and (σ22)u<hβfl, then model (1.4) admits a positive T-periodic solution.
Proof. Define
V8(t,M,A)=−b1lnM−b2lnA, |
and b1, b2 are positive constants to be chosen later. By Itô's formula, we have
LV8(t,M,A)=−b1M(βc(t)AM−a(t)M2−μ(t)k(t)k(t)+MM)−b2A[(Aup(t)−A)f(t)−c(t)hAM]LV8(t,M,A)=+σ21(t)2b1+σ22(t)2b2LV8(t,M,A)=−b1βc(t)A+a(t)b1M+μ(t)k(t)b1k(t)+M−Aup(t)Ab2f(t)+b2f(t)+c(t)hb2MLV8(t,M,A)=+σ21(t)2b1+σ22(t)2b2LV8(t,M,A)≤−b1βc(t)A−Aup(t)Ab2f(t)+(μ(t)+σ21(t)2)b1+(f(t)+σ22(t)2)b2LV8(t,M,A)=+(a(t)b1+c(t)hb2)MLV8(t,M,A)≤−2√b1b2βc(t)Aup(t)f(t)+(μ(t)+σ21(t)2)b1+(f(t)+σ22(t)2)b2LV8(t,M,A)=+(aub1+cuhb2)MLV8(t,M,A)=R(t)+(aub1+cuhlb2)M, | (5.2) |
where
R(t)=−2√b1b2βc(t)Aup(t)f(t)+(μ(t)+σ21(t)2)b1+(f(t)+σ22(t)2)b2,b1=1⟨μ+σ21/2⟩T, b2=1⟨f+σ22/2⟩T. |
Let ˉω(t) be the solution of the following equation
ˉω′(t)=⟨R(t)⟩T−R(t). | (5.3) |
Then ˉω(t) is a T-periodic function. On the basis of Eqs (5.2) and (5.3), we can obtain
L(V8+ˉω(t))≤⟨R(t)⟩T+(aub1+cuhb2)ML(V8+ˉω(t))=−2⟨(Aupfβc)12⟩T(⟨μ+σ21/2⟩T⟨f+σ22/2⟩T)12+2+(aub1+cuhb2)ML(V8+ˉω(t))=−2(R1−1)+(aub1+cuhb2)ML(V8+ˉω(t))=−α1+(aub1+cuhb2)M, | (5.4) |
where α1=2(R1−1)>0.
Define
V9(t,M,A)=12(M+hβA)2−lnA. |
Applying Itô's formula, one has
L(12(M+hβA)2)=(M+hβA)(βc(t)AM−a(t)M2−μ(t)k(t)k(t)+MML(12(M+hβA)2)=+hβ[(Aup(t)−A)f(t)−c(t)hAM])+σ21(t)2M2+hβ2σ22(t)A2L(12(M+hβA)2)≤(M+hβA)(−a(t)M2+hβAup(t)f(t)−hβf(t)A)L(12(M+hβA)2)=+σ21(t)2M2+hβ2σ22(t)A2L(12(M+hβA)2)=−a(t)M3+hβAup(t)f(t)M−hβf(t)AM−a(t)hβAM2L(12(M+hβA)2)=+h2β2Aup(t)f(t)A−h2β2f(t)A2+σ21(t)2M2+hβ2σ22(t)A2L(12(M+hβA)2)≤−a(t)M3+hβAup(t)f(t)M+h2β2Aup(t)f(t)AL(12(M+hβA)2)=−h2β2f(t)A2+σ21(t)2M2+hβ2σ22(t)A2L(12(M+hβA)2)≤−al2M3−h2β2fl2A2+α2, | (5.5) |
where
α2=sup(M,A)∈R2+{−al2M3+(σ21)u2M2+hβAuupfuM−h2β2fl2A2α2=+hβ2(σ22)uA2+h2β2AuupfuA}<+∞. |
Similarly, one deduces
L(−lnA)=−1A[(Aup(t)−A)f(t)−c(t)hAM]+12σ22(t)L(−lnA)=−Aup(t)Af(t)+f(t)+c(t)hM+12σ22(t)L(−lnA)≤−AlupflA+cuhM+fu+12(σ22)u. | (5.6) |
According to Eqs (5.5) and (5.6) one can get
LV9(t,M,A)≤−h2β2fl2A2−al2M3−AlupflA+cuhM+α2+fu+(σ22)u2. | (5.7) |
Define
V10(t,M,A)=H(V8+ˉω)+V9, |
where H is positive constant. Clearly,
lim infq2→+∞, (M,A)∈R2+∖Uq2V10(t,M,A)→+∞, |
where Uq2=(1q2,q2)×(1q2,q2), q2 is a sufficiently large number. Combining with Eqs (5.4) and (5.7), we have
LV10≤−Hα1+(aub1+cuhb2)HM−h2β2fl2A2−al2M3−AlupflA+cuhMLV10≤+α2+fu+(σ22)u2LV10=−Hα1−h2β2fl2A2−al2M3−AlupflA+[H(aub1+cuhb2)+cuh]MLV10≤+α2+fu+(σ22)u2. |
Define a bounded close set:
K={(M,A)∈R2+:ε≤M≤1/ε,ε≤A≤1/ε}, |
where 0<ε<1 is a sufficient small number. We divide R2+∖K into the following four ranges
K1={M<ε}, K2={A<ε}, K3={M>1/ε}, K4={A>1/ε}. |
Case 1'. If (M,A)∈K1, then we get
LV10(t,M,A)≤−Hα1−h2β2fl2A2−al2M3−AlupflA+[H(aub1+cuhb2)+cuh]εLV10(t,M,A)≤+α2+fu+(σ22)u/2LV10(t,M,A)≤−Hα1+[H(aub1+cuhb2)+cuh]ε+α2+fu+(σ22)u/2. | (5.8) |
Case 2'. If (M,A)∈K2, then we have
LV10(t,M,A)≤−Alupflε−al2M3+[H(aub1+cuhb2)+cuh]M+α2+fu+(σ22)u/2LV10(t,M,A)≤−Alupflε+J1+α2+fu+(σ22)u/2, | (5.9) |
where
J1=supM∈R+{−al2M3+[H(aub1+cuhb2)+cuh]M}. |
Case 3'. If (M,A)∈K3, then we derive
LV10(t,M,A)≤−al4M3−al4M3+[H(aub1+cuhb2)+cuh]M+α2+fu+(σ22)u/2LV10(t,M,A)≤−al4ε3+J2+α2+fu+(σ22)u/2, | (5.10) |
where
J2=supM∈R+{−al4M3+[H(aub1+cuhb2)+cuh]M}. |
Case 4'. If (M,A)∈K4, then one has
LV10(t,M,A)≤−h2β2fl2ε2−al2M3+[H(aub1+cuhb2)+cuh]M+α2+fu+(σ22)u/2LV10(t,M,A)≤−h2β2fl2ε2+J1+α2+fu+(σ22)u/2. | (5.11) |
In the set R2+∖K, we choose ε sufficiently small such that
−Hα1+[H(aub1+cuhb2)+cuh]ε+α2+fu+(σ22)u/2<−1, | (5.12) |
−Alupflε+J1+α2+fu+(σ22)u/2<−1, | (5.13) |
−al4ε3+J2+α2+fu+(σ22)u/2<−1, | (5.14) |
−h2β2fl2ε2+J1+α2+fu+(σ22)u/2<−1. | (5.15) |
It then follows from Eqs (5.8)–(5.15) that LV10(t,M,A)<−1 for all (M,A)∈R2+∖K.
In this section, we take advantage of some real data (see Table 1) and the Euler-Maruyama method [34] to illustrate the above results. For model (1.3), we pay attention to the discretization equation:
{Mn+1=Mn+[βcAnMn−aM2n−μkk+MnMn]Δt+σ1Mnζ1n√ΔtMn+1=+12σ21M2n(ζ21n−1)Δt,An+1=An+[(Aup−An)f−chAnMn]Δt+σ2Anζ2n√ΔtMn+1=+12σ22A2n(ζ22n−1)Δt, |
Symbol | Value | Unit | Source |
a | 0.01 | g/g/h | Estimated |
f | 0.4 | m3/m3/h | Estimated |
μ | 0.015 | g/g/h | Estimated |
Aup | 1 | g/m3 | [29] |
h | 0.1 | m | [2,30] |
c | 0.1 | m3/g/h | [31,32] |
β | 0.2 | g/g | [33] |
k | 150 | g/m2 | [5] |
where ζ1n, ζ2n mean independent Gaussian random variable.
From Theorems 3.1–3.3, one can find out that under the assumption a>μ/k, λ0 is sufficient conditions determining the persistence and extinction of the mussel species. More precisely, under the assumption a>μ/k, if λ0<0, then the mussel species dies out; if λ0=0, then the mussel species is nonpersistent in the mean. If λ0>0, then the mussel species is weakly persistent. Note that λ0=βcAup−μ−12σ21, which suggests that white noise can greatly influence the survival of mussel: when the intensity of the noise is large enough, it could make mussel become extinct.
Theorem 4.1 suggests that if λ0>0 and σ22 is sufficiently small such that
σ22<min{2λ0fμ+12σ21, hβf}, |
then model (1.3) possesses a unique ESD on R2+. This ESD can be used to estimate the outbreak possibility of mussel.
Figure 1(a)–(c) characterize the persistence and extinction of the mussel species in model (1.3) with different σ1. We choose σ2=0.3, initial value M(0)=0.1 and A(0)=0.12. Figure 1(a) is with σ1=0.3, which reflects that the mussel species dies out with probability one; Figure 1(b) is with σ1=0.1, which suggests that the mussel species is nonpersistent in the mean; Figure 1(c) is with σ1=0.01, which shows that the mussel species is weakly persistent. Comparing Figure 1(c) with Figure 1(a), one can observe that with the increasing value of σ1, the mussel species tends to go to extinction. In other words, large noise could lead to the extinction of mussel.
Figure 2 plots the probability density function (PDF) of the stationary distribution of model (1.3) with σ1=0.005 and σ2=0.0048.
For model (1.4), we focus on the following discretization equation:
{Mn+1=Mn+[βc(nΔt)AnMn−a(nΔt)M2n−μ(nΔt)k(nΔt)k(nΔt)+MnMn]ΔtMn+1=+σ1(nΔt)Mnζ1n√Δt+12σ21(nΔt)Mn(ζ21n−1)Δt,An+1=An+[(Aup(nΔt)−An)f(nΔt)−c(nΔt)hAnMn]ΔtAn+1=+σ2(nΔt)Anζ2n√Δt+12σ22(nΔt)An(ζ22n−1)Δt, |
where ζ1n, ζ2n mean independent Gaussian random variable.
Theorem 5.1 provides sufficient conditions (i.e., R1>1 and (σ22)u<hβfl) under which model (1.4) admits a positive periodic solution. This offers some insightful understanding on how environmental fluctuations affect the survival of mussel and algae.
The initial values M(0)=0.1 and A(0)=0.12 are kept the same as in Fig. 1, and we choose β=0.2, h=1, Aup(t)=10+0.1sin(πt), a(t)=0.2+0.1sin(πt), f(t)=1+0.1sin(πt), c(t)=0.5+0.1sin(πt), μ(t)=0.1+0.1sin(πt), k(t)=10+0.1sin(πt), σ1(t)=0.03+0.01sin(πt), σ2(t)=0.04+0.01sin(πt). It follows from Theorem 5.1 that model (1.4) has a T-periodic solution, see Figsure 3(a), (c). Moreover, Figsure 3(a), (b) show that M(t) and A(t) fluctuate periodically, that is, mussel and algae will not die out. In particular, the effect of environmental noises can be easily found by comparing Figsure 3(a), (c) with Figsure 3(b), (d).
Understanding the effect of random perturbations on the evolution of the mussel is useful for managing this species. This paper proposed two stochastic mussel-algae models (one is autonomous, and the other is non-autonomous) to test the effect of environmental fluctuations on the evolution of mussel. For the autonomous model, the critical value between extinction and weak persistence was obtained. In addition, sufficient conditions for the existence of an ESD were established. For the non-autonomous model, the existence of a positive periodic solution was examined. Some vital impacts of environmental fluctuations on the evolution of mussel were uncovered.
In comparison with the existing papers, this research has the following contributions:
● Our models consider the environmental fluctuations which are more reasonable. Actually, to the best of our knowledge, this research is the first attempt to dissect the stochastic mussel-algae models.
● We obtain the critical value between extinction and weak persistence for the mussel, and uncover that environmental fluctuations can significantly affect the extinction/persistence of the mussel.
● We give some conditions under which model (1.3) has an ESD. This ESD is useful to estimate the outbreak probability of the mussel.
● We provide sufficient conditions for existence of a positive periodic solution of model (1.4). This positive periodic solution is helpful for the understanding how environmental fluctuations affects the survival of mussel and algae.
Some studies on mussel-algal models are worth further investigations. Actually, Theorem 3.1 and Theorem 3.2 have an assumption a>μ/k, what happens when a<μ/k is still unclear. In addition, Theorem 4.1 testifies that if λ0>0 and
σ22<min{2λ0fμ+12σ21,hβf}, |
then model (1.3) possesses a unique ESD on R2+. It is interesting to relax the restriction on σ22. Finally, one may put forward some more realistic and meaningful models, such as considering the effects of Lévy jump [35,36], impulsive perturbations [37,38], time delay [39,40] or fractional order [41,42]. We will leave these for future works.
The authors thank the editor and referees for their careful reading and valuable comments. This work is supported by the National Natural Science Foundations of China (Nos.11771174, 11871201).
The authors declare there is no conflict of interest.
[1] | Quagga & Zebra Mussels, Available from: https://cisr.ucr.edu/invasive-species/quagga-zebra-mussels. |
[2] |
D. J. Wildish, D. D. Kristmanson, Importance to mussels of the benthic boundary layer, Can. J. Fish. Aquat. Sci., 41 (1984), 1618–1625. doi: 10.1139/f84-200
![]() |
[3] |
P. Dolmer, Algal concentration profiles above mussel beds, J. Sea Res., 43 (2000), 113–119. doi: 10.1016/S1385-1101(00)00005-8
![]() |
[4] |
J. Widdows, J. S. Lucas, M. D. Brinsley, P. N. Salkeld, F. J. Staff, Investigation of the effects of current velocity on mussel feeding and mussel bed stability using an annular flume, Helgol. Mar. Res., 56 (2002), 3–12. doi: 10.1007/s10152-001-0100-0
![]() |
[5] |
J. Koppel, M. Rietkerk, N. Dankers, P. Herman, Scale-dependent feedback and regular spatial patterns in young mussel beds, Am. Nat., 165 (2005), E66–E77. doi: 10.1086/428362
![]() |
[6] |
R. A. Cangelosi, D. J. Wollkind, B. J. Kealy-Dichone, I. Chaiya, Nonlinear stability analyses of Turing patterns for a mussel-algae model, J. Math. Biol., 70 (2015), 1249–1294. doi: 10.1007/s00285-014-0794-7
![]() |
[7] | Y. L. Song, H. P. Jiang, Q. X. Liu, Y. Yuan, Spatiotemporal dynamics of the diffusive Mussel-Algae model near Turing-Hopf bifurcation, SIAM J. Appl. Dyn. Syst., 16 (2018), 2030–2062. |
[8] |
R. A. Cangelosi, D. J. Wollkind, B. J. Kealy-Dichone, I. Chaiya, Nonlinear stability analyses of Turing patterns for a mussel-algae model, J. Math. Biol., 70 (2015), 1249–1294. doi: 10.1007/s00285-014-0794-7
![]() |
[9] |
M. Holzer, N. Popovic, Wavetrain solutions of a reaction-diffusion-advection model of mussel-algae interaction, SIAM. J. Appl. Dyn. Syst., 16 (2017), 431–478. doi: 10.1137/15M1040463
![]() |
[10] |
Z. L. Shen, J. J. Wei, Spatiotemporal patterns in a delayed reaction-diffusion mussel-algae model, Int. J. Bifur. Chaos., 29 (2019), 1950164. doi: 10.1142/S0218127419501645
![]() |
[11] |
Z. L. Shen, J. J. Wei, Stationary pattern of a reaction-diffusion mussel-algae model, Bull. Math. Biol., 82 (2020), 1–31. doi: 10.1007/s11538-019-00680-3
![]() |
[12] |
A. G. Brinkman, N. Dankers, M. van Stralen, An analysis of mussel bed habitats in the Dutch Wadden Sea, Helgol. Mar. Res., 56 (2002), 59–75. doi: 10.1007/s10152-001-0093-8
![]() |
[13] |
J. L. Yang, C. G. Satuito, W. Y. Bao, H. Kitamura, Larval settlement and metamorphosis of the mussel Mytilus galloprovincialis on different macroalgae, Mar. Biol., 152 (2007), 1121–1132. doi: 10.1007/s00227-007-0759-0
![]() |
[14] | R. M. May, Stability and Complexity in Model Ecosystems, Princeton University Press, Princeton, 2001. |
[15] |
C. Lu, L. J. Chen, Y. M. Wang, S. Gao, The threshold of stochastic Gilpin-Ayala model subject to general Lévy jumps, J. Appl. Math. Comput., 60 (2019), 731–747. doi: 10.1007/s12190-018-01234-x
![]() |
[16] |
G. D. Liu, H. K. Qi, Z. B. Chang, X. Z. Meng, Asymptotic stability of a stochastic May mutualism system, Comput. Math. Appl., 79 (2020), 735–745. doi: 10.1016/j.camwa.2019.07.022
![]() |
[17] |
J. Hu, Z. J. Liu, L. W. Wang, R. H. Tan, Extinction and stationary distribution of a competition system with distributed delays and higher order coupled noises, Math. Biosci. Eng, 17 (2020), 3240–3251. doi: 10.3934/mbe.2020184
![]() |
[18] | M. Liu, C. Z. Bai, Optimal harvesting of a stochastic mutualism model with regime-switching, Appl. Math. Comput., 373 (2020), 125040. |
[19] |
Y. Zhao, L. You, D. Burkow, S. L. Yuan, Optimal harvesting strategy of a stochastic inshore-offshore hairtail fishery model driven by Lévy jumps in a polluted environment, Nonlinear Dyn., 95 (2019), 1529–1548. doi: 10.1007/s11071-018-4642-y
![]() |
[20] | R. Durrett, Stochastic Calculus: A Practical Introduction, CRC Press, New York, 1996. |
[21] |
Z. Z. Liu, Z. W. Shen, H. Wang, Z. Jin, Analysis of a local diffusive SIR model with seasonality and nonlocal incidence of infection, SIAM J. Appl. Math., 79 (2019), 2218–2241. doi: 10.1137/18M1231493
![]() |
[22] |
X. N. Liu, Y. Wang, X. Q. Zhao, Dynamics of a climate-based periodic Chikungunya model with incubation period, Appl. Math. Model., 80 (2020), 151–168. doi: 10.1016/j.apm.2019.11.038
![]() |
[23] |
H. K. Qi, X. N. Leng, X. Z. Meng, T. H. Zhang, Periodic solution and ergodic stationary distribution of SEIS dynamical systems with active and latent patients, Qual. Theory Dyn. Syst., 18 (2019), 347–369. doi: 10.1007/s12346-018-0289-9
![]() |
[24] | X. H. Zhang, D. Q. Jiang, Periodic solutions of a stochastic food-limited mutualism model, Mathodol. Comput. Appl., 22 (2020), 267–278. |
[25] | C. Lu, X. H. Ding, Periodic solutions and stationary distribution for a stochastic predator-prey system with impulsive perturbations, Appl. Math. Comput., 350 (2019), 313–322. |
[26] | X. R. Mao, Stochastic Differential Equations and Applications, Horwood Publishing, Chichester, 1997. |
[27] |
R. S. Lipster, A strong law of large numbers for local martingales, Stochastics, 3 (1980), 217–228. doi: 10.1080/17442508008833146
![]() |
[28] | R. Khasminskii, Stochastic Stability of Differential Equations, Springer, Berlin, 2011. |
[29] |
G. C. Cadee, J. Hegeman, Phytoplankton in the Marsdiep at the end of the 20th century: 30 years monitoring biomass, primary production, and Phaeocystis blooms, J. Sea. Res., 48 (2002), 97–110. doi: 10.1016/S1385-1101(02)00161-2
![]() |
[30] |
D. K. Muschenheim, C. R. Newell, Utilization of seston flux over a mussel bed, Mar. Ecol. Prog. Ser., 85 (1992), 131–136. doi: 10.3354/meps085131
![]() |
[31] |
H. Scholten, A. C. Smaal, Responses of Mytilus edulis L. to varying food concentrations: testing EMMY, an ecophysiological model, J. Exp. Mar. Biol. Ecol., 219 (1998), 217–239. doi: 10.1016/S0022-0981(97)00182-2
![]() |
[32] |
H. U. Riisgard, On measurement of filtration rates in bivalves–-the stony road to reliable data: review and interpretation, Mar. Ecol. Prog. Ser., 211 (2001), 275–291. doi: 10.3354/meps211275
![]() |
[33] |
A. A. Sukhotin, D. Abele, H. O. Portner, Growth, metabolism and lipid peroxidation in Mytilus edulis: age and size effects, Mar. Ecol. Prog. Ser., 226 (2002), 223–234. doi: 10.3354/meps226223
![]() |
[34] | D. J. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 5250–546. |
[35] |
D. C. Shangguan, Z. J. Liu, L. W. Wang, R. H. Tan, A stochastic epidemic model with infectivity in incubation period and homestead-isolation on the susceptible, J. Appl. Math. Comput., 67 (2021), 785–805. doi: 10.1007/s12190-021-01504-1
![]() |
[36] |
X. J. Mu, Q. M. Zhang, L. B. Rong, Optimal vaccination strategy for an SIRS model with imprecise parameters and Lévy noise, J. Frankl. Inst., 356 (2019), 11385–11413. doi: 10.1016/j.jfranklin.2019.03.043
![]() |
[37] | X. W. Yu, S. L. Yuan, T. H. Zhang, Survival and ergodicity of a stochastic phytoplankton-zooplankton modelwith toxin-producing phytoplankton in an impulsive polluted environment, Appl. Math. Comput., 347 (2019), 249–264. |
[38] |
D. M. Li, T. Guo, Y. J. Xu, The effects of impulsive toxicant input on a single-species population in a small polluted environment, Math. Biosci. Eng., 16 (2019), 8179–8194. doi: 10.3934/mbe.2019413
![]() |
[39] |
Q. Liu, D. Q. Jiang, T. Hayat, A. Alsaedi, Dynamical behavior of stochastic predator-prey models with distributed delay and general functional response, Stoch. Anal. Appl., 38 (2020), 403–426. doi: 10.1080/07362994.2019.1695628
![]() |
[40] |
L. L. Liu, R. Xu, Z. Jin, Global dynamics of a spatial heterogeneous viral infection model with intracellular delay and nonlocal diffusion, Appl. Math. Model., 82 (2020), 150–167. doi: 10.1016/j.apm.2020.01.035
![]() |
[41] |
P. A. Naik, K. M. Owolabi, M. Yavuz, J. Zu, Chaotic dynamics of a fractional order HIV-1 model involving AIDS-related cancer cells, Chaos Solitons Fractals, 140 (2020), 110272. doi: 10.1016/j.chaos.2020.110272
![]() |
[42] |
M. Yavuz, N. Sene, Stability analysis and numerical computation of the fractional predator-prey model with the harvesting rate, Fractal Fract., 4 (2020), 35. doi: 10.3390/fractalfract4030035
![]() |
1. | Ricardo Vinicius Gonçalves Rosas, Anna Regina Corbo Costa, Claudia Mazza Dias , Charles Henrique Xavier Barreto Barbosa, José Carlos Rubianes Silva, Dayse Haime Pastore, Raquel Medeiros Andrade Figueira, An app for monitoring the population of Golden Mussels, 2022, 43, 1679-0375, 171, 10.5433/1679-0375.2022v43n2p171 | |
2. | Warda Islam, Muhammad Z. Baber, Nauman Ahmed, Ali Akgül, Muhammad Rafiq, Ali Raza, I.S. Yahia, H. Algarni, Wajaree Weera, Investigation the soliton solutions of mussel and algae model leading to concentration, 2023, 70, 11100168, 133, 10.1016/j.aej.2023.01.025 | |
3. | Yan Xie, Zhijun Liu, Ke Qi, Dongchen Shangguan, Qinglong Wang, A stochastic mussel-algae model under regime switching, 2022, 19, 1551-0018, 4794, 10.3934/mbe.2022224 | |
4. | José Carlos Rubianes Silva, Claudia Mazza Dias, Dayse Haime Pastore, Anna Regina Corbo Costa, Raquel Medeiros Andrade Figueira, Humberto Freitas de Medeiros Fortunato, Charles Henrique Xavier Barreto Barbosa, Breylla Campos Carvalho, Population growth of the golden mussel (L. fortunei) in hydroelectric power plants: a study via mathematical and computational modeling, 2022, 27, 2318-0331, 10.1590/2318-0331.272220210124 | |
5. | Peng Zhu, Min Xiao, Xia Huang, Fuchen Zhang, Zhen Wang, Jinde Cao, Spatiotemporal dynamics optimization of a delayed reaction–diffusion mussel–algae model based on PD control strategy, 2023, 173, 09600779, 113751, 10.1016/j.chaos.2023.113751 | |
6. | Charles H.X.B. Barbosa, Claudia M. Dias, Dayse H. Pastore, José C.R. Silva, Anna R.C. Costa, Isaac P. Santos, Ramoni Z.S. Azevedo, Raquel M.A. Figueira, Humberto F.M. Fortunato, Analysis of a mathematical model for golden mussels infestation, 2023, 486, 03043800, 110502, 10.1016/j.ecolmodel.2023.110502 | |
7. | Ramoni Z. S. Azevedo, Charles H. X. B. Barbosa, Isaac P. Santos, José C. R. Silva, Dayse H. Pastore, Anna R. C. Costa, Claudia M. Dias, Raquel M. A. Figueira, Humberto F. M. Fortunato, 2023, Chapter 14, 978-3-031-37128-8, 163, 10.1007/978-3-031-37129-5_14 |