Research article Special Issues

Long-time analysis of a stochastic chemostat model with instantaneous nutrient recycling

  • This paper presents long-time analysis of a stochastic chemostat model with instantaneous nutrient recycling. We focus on the investigation of the sufficient and almost necessary conditions of the exponential extinction and persistence for the model. The convergence to the invariant measure is also established under total variation norm. Our work generalizes and improves many existing results. One of the interesting findings is that random disturbance can suppress microorganism growth, which can provide us some useful control strategies to microbiological cultivation. Finally, some numerical simulations partly based on the stochastic sensitive function technique are given to illustrate theoretical results.

    Citation: Xiaoxia Guo, Dehao Ruan. Long-time analysis of a stochastic chemostat model with instantaneous nutrient recycling[J]. AIMS Mathematics, 2023, 8(4): 9331-9351. doi: 10.3934/math.2023469

    Related Papers:

    [1] Yue Dong, Xinzhu Meng . Stochastic dynamic analysis of a chemostat model of intestinal microbes with migratory effect. AIMS Mathematics, 2023, 8(3): 6356-6374. doi: 10.3934/math.2023321
    [2] Ying He, Bo Bi . Conditions for extinction and ergodicity of a stochastic Mycobacterium tuberculosis model with Markov switching. AIMS Mathematics, 2024, 9(11): 30686-30709. doi: 10.3934/math.20241482
    [3] Yassine Sabbar, José Luis Diaz Palencia, Mouhcine Tilioua, Abraham Otero, Anwar Zeb, Salih Djilali . A general chemostat model with second-order Poisson jumps: asymptotic properties and application to industrial waste-water treatment. AIMS Mathematics, 2023, 8(6): 13024-13049. doi: 10.3934/math.2023656
    [4] Xiaodong Wang, Kai Wang, Zhidong Teng . Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence. AIMS Mathematics, 2023, 8(2): 2829-2855. doi: 10.3934/math.2023148
    [5] Yassine Sabbar, Aeshah A. Raezah . Threshold analysis of an algae-zooplankton model incorporating general interaction rates and nonlinear independent stochastic components. AIMS Mathematics, 2024, 9(7): 18211-18235. doi: 10.3934/math.2024889
    [6] Jingen Yang, Zhong Zhao, Xinyu Song . Statistical property analysis for a stochastic chemostat model with degenerate diffusion. AIMS Mathematics, 2023, 8(1): 1757-1769. doi: 10.3934/math.2023090
    [7] Hong Qiu, Yanzhang Huo, Tianhui Ma . Dynamical analysis of a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps. AIMS Mathematics, 2022, 7(8): 14492-14512. doi: 10.3934/math.2022799
    [8] Chuangliang Qin, Jinji Du, Yuanxian Hui . Dynamical behavior of a stochastic predator-prey model with Holling-type III functional response and infectious predator. AIMS Mathematics, 2022, 7(5): 7403-7418. doi: 10.3934/math.2022413
    [9] Tahir Khan, Fathalla A. Rihan, Muhammad Bilal Riaz, Mohamed Altanji, Abdullah A. Zaagan, Hijaz Ahmad . Stochastic epidemic model for the dynamics of novel coronavirus transmission. AIMS Mathematics, 2024, 9(5): 12433-12457. doi: 10.3934/math.2024608
    [10] Jinji Du, Chuangliang Qin, Yuanxian Hui . Optimal control and analysis of a stochastic SEIR epidemic model with nonlinear incidence and treatment. AIMS Mathematics, 2024, 9(12): 33532-33550. doi: 10.3934/math.20241600
  • This paper presents long-time analysis of a stochastic chemostat model with instantaneous nutrient recycling. We focus on the investigation of the sufficient and almost necessary conditions of the exponential extinction and persistence for the model. The convergence to the invariant measure is also established under total variation norm. Our work generalizes and improves many existing results. One of the interesting findings is that random disturbance can suppress microorganism growth, which can provide us some useful control strategies to microbiological cultivation. Finally, some numerical simulations partly based on the stochastic sensitive function technique are given to illustrate theoretical results.



    The chemostat is a laboratory apparatus used for the continuous culture of microorganisms. It plays an important role in biotechnology, microbial ecology, and population biology, and is the most simple idealization of a biological system for population studies [1]. In recent decades, chemostat modeling has drawn great attention from mathematics and ecology, see [2,3,4,5] and reference therein. One classic chemostat with single species and single substrate is expressed in the form of

    {dS(t)dt=D(S0S(t))1δp(S(t))x(t),dx(t)dt=x(t)(p(S(t))D), (1.1)

    where S(t) and x(t) represent the substrate concentration and the microbial concentration, respectively; S0 is the original input concentration of the nutrient; D is the dilution rate; δ is the growth yield constant reflect the conversion of nutrient to organism, obviously, 1δ1, see literature [6] for details; p(S) is the per-capita growth rate of the species. The nutrient uptake function p is generally assumed to satisfy: p is continuously differentiable; p(0)=0, p(S)>0 for S>0. Some kinds of the nutrient uptake function are often found in literature are as follows:

    ● Lotka-Volterra (see e.g., [7]): p(S)=mS.

    ● Michaelis-Menten (Monod) (see e.g., [8]): p(S)=mSa+S, here m is the maximal growth rate of the microbial species and a is the half-saturation constant. The test result of the relationship between microbial proliferation rate and substrate concentration obtained by Monod is consistent with the form of Michaelis-Menten equation. So, function p(S)=mSa+S is also known as the Monod type function.

    ● Monod-Haldane (see e.g., [9]): p(S)=mSa+S+bS2, here the term bS2 is an inhibition.

    Butler et al. [10] established a multiple competing species chemostat model with a general class of functions describing nutrient uptake. If there is only one microbial species, we can obtain the dynamics of system (1.1) from [10].

    The deterministic models can neglect the stochastic effects only at the macroscopic scale, thanks to the law of large numbers. However, ecosystem is inevitably affected by environmental stochasticity which cannot be neglected at microscopic scale. Also the accumulation of small perturbations in the context of multi-species could not be neglected (see Campillo et al. [11] for more details on this respect). Moreover, stochastic noises may change the behavior in a substantial manner by directly acting on the population densities or affecting the parameter values of ecosystems. In recent years, much research has considered the stochastic influence [12,13,14]. Persistence and extinction are the central question in ecosystems, including microorganism culture [15,16,17,18]. For example, Meng et al. [15] established sufficient criteria for extinction and weak persistence of a nonlinear impulsive stochastic chemostat system in a polluted environment. By using Lyapunov functions method, Lv et al. [17] obtained sufficient conditions for the existence of a unique ergodic stationary distribution of an impulsive stochastic chemostat model similar to that in [15]. While Nguyen [18] proved sufficient and almost necessary condition to determine the persistence and extinction of the model by defining a new threshold parameter. In addition to the research on the central question above, Yang et al. [19] constructed a stochastic chemostat model with degenerate diffusion using a discrete Markov chain. By solving the corresponding Fokker-Planck equation, they derived the explicit expression of the stationary joint probability density and investigated the effect of white noise on the variance and skewness of the concentration of microorganisms. Baratti et al. studied bioreactor models under the assumption of constant volume, and described the probability density function evolution by using the Fokker-Planck theory in [20,21]. Besides, Zhang and Yuan [22] investigated the existence of ergodic stationary distribution of a stochastic delayed chemostat model by using the stochastic Lyapunov analysis method. Subsequently, Zhang and Yuan [23] considered a stochastic chemostat model with a distributed delay and proved the sufficient condition for the extinction and the ergodicity of the solution. For other work related to the delay stochastic chemostat model, we can refer to [24,25,26].

    In industrial applications and most ecological situations an undisturbed dilution rate cannot be expected. There are different possible approaches to introduce noise into stochastic differential equations, both from a biological and from a mathematical perspective. One traditional approach is analogous to that of Stephanopoulos et al. [27] who superimposed a one-dimensional white noise process on the dilution rate. In particular, the authors in [28] replaced the dilution rate D by D+σidBi(t). Then, model (1.1) becomes the following system

    {dS(t)=(D(S0S(t))1δp(S(t))x(t))dt+σ1S(t)dB1(t),dx(t)=x(t)(p(S(t))D)dt+σ2x(t)dB2(t), (1.2)

    where Bi(t), i=1,2 are two mutually independent Brownian motions. Under the following assumptions:

    (ⅰ) pC2([0,),[0,)), p(S)cS for any S(0,+);

    (ⅱ) p(S)S3m0, for any S(0,+),

    the authors in [28] showed that system (1.2) admits a stationary distribution which is ergodic if

    λs=p(S0)D12σ2212c1(σ1S0)2>0, where, c1=max{0,m02DS0}.

    Obviously, Michaelis-Menten response function p(S)=mSa+S satisfies the above assumptions.

    Moreover, nutrient recycling is an important factor among the many processes which influence ecosystem dynamics. Many researchers have studied the effect of nutrient recycling on ecosystem stability and persistence, see [29,30]. Usually, nutrient recycling is regarded as an instantaneous term [31]. So in this work, we consider a stochastic chemostat model with instantaneous nutrient recycling, and for convenience, we take p(S)=mSa+S. Then, the following model is obtained

    {dS(t)=(D(S0S(t))1δmS(t)a+S(t)x(t)+bγx(t))dt+σ1S(t)dB1(t),dx(t)=x(t)(mS(t)a+S(t)Dγ)dt+σ2x(t)dB2(t), (1.3)

    where, γ is mortality rate of microorganism population, b(0,1) is the fraction of the nutrient recycled by bacterial decomposition of the dead microorganism. Obviously, if γ=0, then system (1.3) becomes model (1.2) with the case p(S)=mSa+S.

    There exist some important questions concerning the system: (ⅰ) Under what conditions will microorganism population become extinct? (ⅱ) When do microorganism population is persistent? (ⅲ) How is dynamic behavior affected by the noise intensity? In the existing literature, one of the traditional method analogous to that in [28] to solve the above central question is to obtain the sufficient but not necessary conditions for the existence of ergodic stationary distribution of the system by using Khasmiskii theory. A different path taken by Imhofa and Walcherb [12] is to establish appropriate Lyapunov functions to analyze the long-term dynamic behavior of systems, including stochastic stability. In our paper, we develop a new treatment to deal with system (1.3) to prove the sufficient and almost necessary conditions of the extinction and persistence for the model and analyze the effect of noise intensity on system dynamics. Our results essentially improve the corresponding research in [28]. One of the main difficulties in dealing with this model stems from the fact that the stochastic comparison arguments do not work well for system (1.3) due to the nutrient cycling term. Therefore, some new techniques must require here. The rest of the paper is organised as follows. Section 2 presents some preliminary results and defines the threshold λ which determines the persistence and extinction of the system. In Section 3, we focus on the condition for the extinction and persistence of system (1.3). Finally, some numerical examples and brief discussions are given in Sections 4 and 5.

    Let (Ω,F,P) be a complete probability space with a filtration {Ft}t0 satisfying the usual conditions (i.e., it is increasing and right continuous while F0 contains all P-null sets). R2+:={(x,y):x0,y0} and R2,+={(x,y):x>0,y>0}. Denote ab=min{a,b}, ab=max{a,b}.

    Consider the deterministic system corresponding to (1.3)

    {dS(t)dt=D(S0S(t))1δmS(t)a+S(t)x(t)+bγx(t),dx(t)dt=x(t)(mS(t)a+S(t)Dγ). (2.1)

    We can derive from the monotonicity of the function p(S)=mSa+S that there exists a uniquely defined positive real number λ such that p(S)<D for 0<S<λ; p(S)>D for λ<S; p(S)=D for λ=S. Here, λ represents the break-even concentration of the substrate for the species x(t).

    Define λd=mS0a+S0Dγ. Similar to [10], when p(S0)<D, i.e., λd<0, there exists a boundary equilibrium E0=(S0,0) for system (2.1) which is asymptotically stable, while when p(S0)>D, i.e., λd>0, there exists an internal equilibrium E=(s,x)=(λ,Dδ(S0λ)D+(1bδ)γ) which is asymptotically stable.

    The following theorem concerns the existence and uniqueness of positive solutions. The proof is standard, so we only present it without proof.

    Theorem 2.1. For any given initial value (S(0),x(0))=(u,v)R2+, system (1.3) has a unique global solution (Su,v(t),xu,v(t)) on t0, and the solution will always remain in R2+ with probability 1.

    Lemma 2.1. For any initial value (S(0),x(0))R2+, the solution of system (1.3) is stochastically ultimately bounded.

    Proof. Let V(t)=S(t)+1δx(t), by making use of Itô formula, we have

    dV(t)=dS(t)+1δdx(t)+σ1S(t)dB1(t)+σ21δx(t)dB2(t)=LV(t)+σ1S(t)dB1(t)+σ21δx(t)dB2(t),

    where

    LV(t)=S0DDS(t)(D+γδbγ)1δx(t)S0DDS(t)(D+(1δb)γ)1δx(t)S0DDV(t).

    Then, applying the argument of the proof of [23,Theorem 3.1], one can prove the theorem easily, we omit the details here and refer the reader to [23,Theorem 3.1].

    Remark 2.1. Lemma 2.1 implies that for any T>0, there exists positive H0 such that for any t>T, S(t)H0, x(t)H0. $

    Lemma 2.2. The following assertions hold:

    (ⅰ) Let p1 be a positive solution of

    σ21σ222p2+(σ21σ222D)pD=0.

    Then for any 0q<p1 and p<p1, there exists a constants M such that,

    lim suptEu,v{[S(t)+x(t)]1+p+[S(t)]q}:=M<,(u,v)R2+. (2.2)

    (ⅱ) For any ϵ(0,1), T>0 and H>1, there is ˉH=ˉH(ε,T,H)>1 such that for any initial value (u,v)[H1,H]×[0,H],

    Pu,v{ˉH1S(t)ˉH, 0<x(t)ˉH, t[0,T]}1ϵ. (2.3)

    Proof. Consider the Lyapunov function V(S,x)=(S+1δx)1+p+Sq. By directly calculating the differential operator LV(S,x), we obtain

    LV(S,x)=(1+p)(S+1δx)p(DS0DS+bγxmSa+Sxδ+mSa+Sxδ(D+γ)1δx)+p(1+p)2(S+1δx)p1(σ21S2+σ22δ2x2)+q(1+q)2σ21S2qSq1(DS0DS+bγxmSa+S1δx)(1+p)(S+1δx)p(DS0DS(D+(1δb)γ)1δx)+p(1+p)2(σ21σ22)(S+1δx)p+1qDS0Sq1+q(D+q+12σ21)Sq+qmaSq1δx.

    Let q<ˉp<p, Young's inequality says that

    Sqxδˉp1+ˉpSq(1+ˉp)ˉp+11+ˉp(xδ)1+ˉpˉp1+ˉpSq(1+ˉp)ˉp+11+ˉp(S+1δx)1+ˉp.

    Therefore,

    LV(S,x)(1+p)DS0(S+1δx)p+[p(1+p)2(σ21σ22)(p+1)D](S+1δx)1+p+qma(1+ˉp)(S+1δx)1+ˉpqDS0Sq1+q(D+q+12σ21)Sq+qmˉpa(1+ˉp)Sq(1+ˉp)ˉp.

    We choose 0<H2<(1+p)Dp(1+p)2(σ21σ22). Since 1+ˉp<1+p and q(1+ˉp)ˉp<1+q, we have

    H1=sup(S,x)R2+{LV(S,x)+H2V(S,x)}<.

    Hence,

    LV(S,x)H1H2V(S,x). (2.4)

    Define the stopping time τn=inf{t0: V(S(t),x(t))n}. By Itô formula and (2.4) yield that

    Eu,v[eH2(tτn)V(S(tτn),x(tτn))]V(u,v)+Etτn0eH2s(LV(S(s),x(s))+H2V(S(s),x(s)))dsV(u,v)+H1(eH2t1)H2.

    By letting n, applying Fatou's lemma obtains the part (ⅰ) of Lemma 2.1,

    Eu,v[eH2tV(S(t),x(t)]V(u,v)+H1(eH2t1)H2.

    The proof for the second part is standard, we omit the details here and refer the reader to [32,Lemma 2.1].

    Consider the first equation of (1.3) on boundary when microorganism population is absent, i.e.,

    d˜S(t)=D(S0˜S(t))dt+σ1˜S(t)dB1(t),˜S(0)0. (2.5)

    We write ˜Su(t) for the solution of the Eq (2.5) with initial value ˜S(0)=u. By solving the Fokker-Planck equation, the Eq (2.4) has an ergodic stationary distribution ν0 with density function

    f(s)=βαΓ(α)sα1eβs,s>0,

    where α=2Dσ21+1, β=2DS0σ21. Then we have

    E[˜S(t)]=0sν0(ds)=0sf(s)ds=S0.

    Lemma 2.3. [33] The following assertions hold:

    (ⅰ) limtln˜S(t)t=0, a.s.

    (ⅱ) The unique stationary distribution ν0 of the Markov process ˜S(t) satisfies

    limt1tt0˜S1(s)ν0(ds)=D+12σ21<,a.s. (2.6)

    Define

    λ:=0msa+sν0(ds)Dγ12σ22.

    The inequality msa+smas implies that

    0msa+sν0(ds)ma0sf(s)ds=maS0.

    As a result, λ is well-defined.

    Remark 2.2. The solution ˜S(t) of the boundary equation (2.5) has a stationary distribution ν0. Let S be a random variable with distribution ν0. Then, for the nonlinear convex response function p(S)=mSa+S, there are parameters such that Eν0[p(S)]>D+12σ22>D>p(S0) for γ=0, namely, λ>0>λd; further suppose that σ2=0, then there must be λ>λd (see Figure 1).

    Figure 1.  The graph of function p(S)=mSa+S. Here, Eν0[p(S)]0.3218>D+12σ220.2025>D=0.2>p(S0)0.1286 with parameters S0=0.4, D=0.2, m=0.45, a=1, σ1=0.07 and σ2=0.08.

    Furthermore, we consider the perturbed equation of model (2.5)

    d˜Sθ(t)=(DS0+bγθD˜Sθ(t))dt+σ1˜Sθ(t)dB1(t),˜Sθ(0)0. (2.7)

    According to the comparison theorem, ˜S(t)˜Sθ(t). Similar to (2.5), there exists a unique invariant probability measure νθ to the solution ˜Sθ(t) of (2.7). The following key lemma reflects the relationship between the perturbation equation of (2.5) and itself.

    Lemma 2.4. [34] We have

    limθ0|0msa+sνθ(ds)0msa+sν0(ds)|=0.

    This section analyzes the long term properties of the system by using the parameter λ. Roughly speaking, if λ<0, microorganism population will die out. If λ>0, the system has an invariant probability measure and the transition probability of the solution process converges to the invariant measure and we refer to the cases as the system being persistent.

    Lemma 3.1. Assume that λ<0, for any ϵ>0, H>1, ρ>0, there exists δ0=δ0(ϵ,H)(0,H) such that for any initial value (u,v)[H1,H]×(0,δ0],

    Pu,v{limt|lnx(t)tλ|<ϵ}17ε. (3.1)

    Proof. By the ergidicity of ˜Sθ(t), we obtain

    limt1tt0m˜Sθ(s)a+˜Sθ(s)ds=0msa+sνθ(ds) a.s.. 

    Consequently, for any ϵ>0, there exists a T1=T1(ϵ,H)>0 such that PH(Ω1)1ϵ, where

    Ω1={ω:1tt0m˜Sθ(s)a+˜Sθ(s)ds0msa+sνθ(ds)+ϵ,tT1}1ϵ. (3.2)

    According to the uniqueness of solution, we have ˜Sθu(s)˜SθH(s), t0, where the subscript of ˜Sθu(s) indicates the initial value ˜Sθ(0)=u, which implies that Pu(Ω1)1ε, u[H1,H]. From Lemma 2.4, we obtain that there is a θ0[0,1] such that Pu(Ω2)1ϵ, where

    Ω2={ω:|0msa+sνθ(ds)0msa+sν0(ds)|<ϵ2, θ[0,θ0]}. (3.3)

    The strong law of large numbers for Brownian motions

    limtBi(t)t=0 a.s., for  k=1,2, (3.4)

    implies that there is a T2>0 such that Pu,v(Ω3)1ϵ, where

    Ω3={ω:|σ2B2(t)t|ϵ2, tT2}. (3.5)

    Let T:=T1T2. Using (2.3), there is ˉH>1 such that (u,v)[H1,H]×(0,H] and H>1, Pu,v(Ω4)1ε, where

    Ω4={ω:ˉH1S(t)ˉH,0<I(t)ˉH,t[0,T]}. (3.6)

    To proceed, we will establish the difference of (˜Sθ(t)S(t)). It follows from (1.3) and (2.7) that

    d(˜Sθ(t)S(t))=(D(˜Sθ(t)S(t))+mS(t)a+S(t)x(t)+bγ(θx(t)))dt+σ1(˜Sθ(t)S(t))dB1(t).

    Denoting

    ϑ(t)=exp{(D+12σ21)tσ1B1(t)}.

    Then, by constant variation formula we have

    ˜Sθ(t)S(t)=ϑ1(t)t0ϑ(s)(mS(s)a+S(s)x(s)+bγ(θx(s)))ds.

    Choosing positive constants ρ and θ satisfying

    ρ<min{ϵ,a2mϵ} and θ<min{θ0,ρ2,ρ2ϕ1(ϵ)(mˉHaˉH+bγ)},

    such that for (u,v)[H1,H]×(0,θ], Pu,v(Ω5)1ϵ where

    Ω5={|˜Sθ(t)S(t)|<ρ, t[0,Tτθ]},τθ=inf{t0:x(t)θ}. (3.7)

    Indeed, let ϱ=τθinf{t:S(t)ˉH}, we can obtain by the inequality (ni=1ai)22nni=1a2i, Young's inequality and Burkholder-Davis-Gundy inequality that for any t[0,T],

    Eu,v[supst|˜Sθ(sϱ)S(sϱ)|2]M1θ2+M2t0Eu,v[supsu|˜Sθ(sϱ)S(sϱ)|2]du,

    where M1=16(m2+b2γ2)T2,M2=16(D2T+4σ21). Then further using Gronwall's inequality we have

    Eu,v[supsT|˜Sθ(sϱ)S(sϱ)|2]M1eM2Tθ2.

    Hence, for arbitrarily small θ as defined above,

    Pu,v{supsT|˜Sθ(sϱ)S(sϱ)|2ρ2}M1eM2Tθ2ρ2<ϵ.

    By virtue of

    Pu,v{sϱ=sτθ,s[0,T]}Pu,v{supsTS(s)ˉH}1ϵ,

    we obtain

    Pu,v{|˜Sθ(t)S(t)|<ρ,t[0,Tτθ]}1ϵ.

    Define a stoping time

    τρ=inf{t0:|˜Sθ(t)S(t)|>ρ},τ=τθτρ.

    Applying exponential martingale inequality, there is δ1(0,θ) such that Pu,v(Ω6)1ϵ, where

    Ω6={ω:τθT}. (3.8)

    As a consequence, for ω6i=1Ωi, we have τθT. Applying comparison argument, one has S(t)˜Sθ(t) for ω6i=1Ωi and tτθ. Combining (3.2), (3.3) and (3.5), one has that for (u,v)[H1,H]×(0,δ1], ω6i=1Ωi and tT,

    lnx(t)=lnv+t0(mS(s)a+S(s)Dγ12σ22)ds+σ2B2(t)=lnv+t0(m˜Sθ(s)a+˜Sθ(s)Dγ12σ22)ds+t0(mS(s)a+S(s)m˜Sθ(s)a+˜Sθ(s))ds+σ2B2(t)lnv+t0(m˜Sθ(s)a+˜Sθ(s)Dγ12σ22)ds+σ2B2(t)lnv+(λ+2ϵ)t.

    Therefore, for (u,v)[H1,H]×(0,δ1],ω6i=1Ωi,t[T,τ],

    x(t)ve(λ+2ϵ)t. (3.9)

    In addition, by the law of iterated logarithm, there exits qϵ such that Pu,v(Ω7)1ϵ where

    Ω7={|σ1B1(t)|qϵt(|lnt|+1),t0}.

    Denote qϵ(t)=qϵt(|lnt|+1). Hence, for all (u,v)[H1,H]×(0,δ1],ω7i=1Ωi and tT, one has

    |˜Sθ(tτ)S(tτ)|=ϑ1(tτ)tτ0ϑ(s)(mS(s)a+S(s)+bγ)x(s)ds=e(D+12σ21)(tτ)+σ1B1(tτ)T0e(D+12σ21)sσ1B1(s)(mS(s)a+S(s)+bγ)x(s)ds+e(D+12σ21)(tτ)+σ1B1(tτ)tτTe(D+12σ21)sσ1B1(s)(mS(s)a+S(s)+bγ)x(s)ds(mˉHa+ˉH+bγ)θe(D+12σ21)(tτ)+qϵ(tτ)T0e(D+12σ21)s+qϵ(s)ds+(m+bγ)ve(D+12σ21)(tτ)+qϵ(tτ)tτTe(D+12σ21)s+qϵ(s)eλ+3ϵds(mˉHa+ˉH+bγ)θϕ1(ϵ)+(m+bγ)vϕ2(ϵ),

    where

    ϕ1(ϵ)=supt0e(D+12σ21)(tτ)+qϵ(tτ)T0e(D+12σ21)s+qϵ(s)ds<,
    ϕ2(ϵ)=suptTe(D+12σ21)(tτ)+qϵ(tτ)tτTe(D+12σ21)s+qϵ(s)eλ+3ϵds<.

    Let δ0(0,δ1) satisfying

    δ0e(λ+2ε)T<θ,δ0(m+bγ)ϕ2(ε)<ρ2.

    Then, for all (u,v)[H1,H]×(0,δ0], ω7i=1Ωi and tT, we obtain

    |˜Sθ(tτ)S(tτ)|<ρ.

    It follows that tττρ,tT. Therefore,

    7i=1Ωi{ττρ}.

    Since τ=τθτρ, we have

    7i=1Ωi{τθτρ}.

    Hence, for all (u,v)[H1,H]×(0,δ0],ω7i=1Ωi, tT,

    x(tτθ)δ0e(λ+2ε)T<θ. (3.10)

    This means tτθ<τθ, tT. As a result

    τθ=τρ=.

    Hence, for any (u,v)[H1,H]×(0,δ0] and ω7i=1Ωi,

    lim supt|lnx(t)tλ|lim suptlnvt+lim supt1tt0|mS(s)a+S(s)m˜Sθ(s)a+˜Sθ(s)|ds+lim supt1tt0|m˜Sθ(s)a+˜Sθ(s)m˜S(s)a+˜S(s)|ds+lim supt1tt0σ2(s)dB2(s)maρ+ϵ2<ϵ,

    where Pu,v{ω7i=1Ωi}>17ε.

    Remark 3.1. 3.1 shows that when λ<0, the component x(t) of the solution (S(t),x(t)) starting from any initial value in the designated area [H1,H]×(0,δ0] finally enters the interval (0,δ0] and tends to zero exponentially. Biologically, it implies that microorganism population is eventually extinct.

    Theorem 3.1. If λ<0, for any initial value (u,v)R2,+, we have P{limtlnxu,v(t)t=λ}=1, i.e., the number of microorganism population in the chemostat will tend to zero with exponential rate.

    Proof. We shall use the manners established in [34] to prove this theorem. By virtue of Lemma 2.2 and Cheybshev's inequality, for any initial value (u,v)R2,+, one has

    lim suptPu,v{(S(t),x(t))M}1ε, (3.11)

    where

    M={(s,u):(s,u)[H1,H]×(0,H]}.

    In view of Lemma 3.1, the process (S(t),x(t)) is not recurrent in R2,+. Since Bi(t), i=1,2,3 are independent, (S(t),x(t)) must be transient [35]. Denote

    M1={(s,u):(s,u)[H1,H]×(δ0,H]},

    which is a compact subset of R2,+. It follows from the definition of transience that

    limtPu,v{(S(t),x(t))M1}=0. (3.12)

    Combining (3.11) and (3.12), one has

    lim suptPu,v{(S(t),x(t))MM1}1ϵ.

    Therefore, there exits T>0 such that

    Pu,v{(S(T),x(T))[H1,H]×(0,δ0]}12ϵ. (3.13)

    The Markov property of the process, (3.1) and (3.13) deduce that

    Pu,v{lim supt|lnx(t)tλ|ϵ}19ϵ. (3.14)

    Since ϵ is arbitrary, one has

    Pu,v{lim suptlnx(t)t=λ}=1.

    Theorem 3.2. If λ>0, for any initial value (u,v)R2,+, the solution of the system (1.3) has a unique invariant probability measure π with support R2,+. Moreover, for any π-integrable f(x,y):R2,+R, we have

    limt1tt0f(S(s),x(s))ds=R2,+f(x,y)π(dx,dy)a.s.,

    and

    limtP(t,(u,v),)π()=0,

    where P(t,(u,v),) is the transition probability of (S(t),x(t)) and is the total variation norm.

    Proof. Let (S(0),x(0))=(u,v)R2,0+. It follows from Lemma 2.1 that

    lim suptlnS(t)t0, (3.15)

    and

    lim suptlnx(t)t0. (3.16)

    From the first equation of system (1.3), we have

    lnS(t)=lnS(0)+t0(DS0S(s)D12σ21mx(s)a+S(s)+bγx(s)S(s))ds+σ1B1(t). (3.17)

    In view of (3.15), (3.17) and large numbers for martingales, we have

    lim supt1tt0(DS0˜S(s)D12σ21+DS0S(s)DS0˜S(s)mx(s)a+S(s)+bγx(s)S(s))ds0,

    which follows from (2.6) of Lemma 2.3 that

    lim supt1tt0(DS0S(s)DS0˜S(s)mx(s)a+S(s)+bγx(s)S(s))ds0.

    Then,

    DS0lim inft1tt0(1S(s)1˜S(s))ds+malim inft1tt0x(s)ds0. (3.18)

    From the second equation of system (1.3), we have

    lnx(t)=lnx(0)+t0(mS(s)a+S(s)Dγ12σ22)ds+σ2B2(t). (3.19)

    By (3.16), (3.19) and the strong large numbers for martingales, we have

    lim supt1tt0(m˜S(s)a+˜S(s)Dγ12σ22+mS(s)a+S(s)m˜S(s)a+˜S(s))ds0.

    Hence,

    λ+lim supt1tt0(mS(s)a+S(s)m˜S(s)a+˜S(s))ds0.

    Therefore,

    lim inft1tt0(1S(s)1˜S(s))dsλma. (3.20)

    By combing (3.18) and (3.20) we obtain

    lim inft1tt0x(s)dsDS0m2λ:=Θ. (3.21)

    For 0<η<Θ<κ<, Hölder's inequality yields that

    1tt0Eu,v[1{x(s)η}x(s)]ds(1tt0Eu,v[1{x(s)η}]ds)p1+p(1tt0Eu,v[x(s)]1+pds)11+p.

    By Fatou's lemma, we get

    lim inft1tt0Eu,v[1{x(s)η}]ds(lim inft1tt0Eu,v[1{x(s)η}x(s)]ds)1+pp(lim supt1tt0Eu,v[x(s)]1+pds)1pM1p(lim inft1tt0Eu,v[x(s)]dsη)1+ppM1p(Θη)1+pp>0. (3.22)

    Applying (2.2), we obtain

    lim supt1tt0Eu,v[1{S(s)+x(s)κ}]ds1κ1+plim supt1tt0Eu,v[S(s)+x(s)]1+pdsMκ1+p. (3.23)

    Choosing sufficiently small η and sufficiently large κ. It follows from (3.22) and (3.23) that

    lim inft1tt0Eu,v[1{S(s)+x(s)B}]ds(Θη)1+ppM1pMκ1+p>0, (3.24)

    where B={(x,y):yη,x+yκ}. By the invariance of M={x0,y>0} under system (1.3), we consider the process (S(t),x(t)) on M. Obviously, (S(t),x(t)) is a Feller Markov process. Thus, (3.24) and compactness of B in M implies that there is a invariant probability measure π on M. Then, by similar arguments of [33,Theothem 2.4] we can derive that π is a unique invariant probability measure of (S(t),x(t)) on R2,+ and the strong of large numbers holds. The proof is complete.

    Remark 3.2. Theorem 3.2 shows that when λ>0, the positive solution of system (1.3) has a ergodic stationary distribution. Biologically, it implies that microorganism population is persistent. $

    In this section, we will introduce several numerical examples to illustrate our theoretical results. For the numerical simulation, we use Milstein's higher order method [36] to obtain the discretization equations of system (1.3)

    {Sk+1=Sk+(D(S0Sk)1δmSka+Skxk+bγxk)t+σ1Sktξk+12σ21Sk(ξ2k1)t,xk+1=xk+xk(mSka+SkDγ)t+σ2xktξk+12σ22xk(ξ2k1)t,

    where the time increment t>0, ξk are the independent Gaussian random variables which follow the distribution N(0,1) for k=1,2,,n.

    Example 4.1. Consider the small noises. Without loss of generality, let σ1=σ2=0.08. Set S0=1.4, D=0.2, m=0.4, a=0.45, δ=0.3, γ=0.03, bγ=0.02. By calculation, we have λ0.0135>0, which satisfies the condition of Theorem 3.2. Figure 2a, b show time series diagrams of the stochastic model (1.3) and the corresponding deterministic model. Figure 2c, d depict histograms of S and x, respectively; Figure 2e shows the phase diagram of the stochastic model (1.3). Biologically, Figure 2 shows that the microorganism population will be persistent with lower noise intensity.

    Figure 2.  (a) and (b) Time series diagrams of system (1.3); (c) and (d) Histograms of S and x, respectively; (e) Two-dimensional diagram of S and x; (f) Random states and the confidence ellipse of system (1.3) with fiducial probability P=0.95, and the vector field of the corresponding deterministic model (2.1).

    Although the theorem shows the existence of the stationary distribution π of system (1.3) with small noise, its analytic formula determined by Kramers-Fokker-Planck (KFP) equation appears to be technically a very difficult problem. This means that it is not easy to obtain the threshold value of noise intensity containing (or causing) microbial persistence (or extinction). Under this circumstance, if σ1=σ2, an asymptotic of the the stationary distribution in Gaussian form can be proposed to estimate the noise intensity [37]:

    π(z,σ1)Kexp(zE,W1(zE)2σ21),

    where z=(S,x), W is the stochastic sensitive function (SSF) of deterministic equilibrium E=(0.6088,0.2119) determined by the matrix equation:

    FW+WFT+G=0,

    where

    F=(D1δmax(a+s)21δmsa+s+bγma(a+s)2mSa+SDγ),   G=(s200x2).

    Hence, the confidence ellipse of system (1.3) can be presented by SSF technique [37]:

    zE,W1(zE)=2σ21ln11P, (4.1)

    where P is a fiducial probability. By calculation, we obtain that

    W=(0.92440.13980.13980.2575),   W1=(1.17850.63980.63984.2309),

    and the corresponding confidence ellipse equation is

    1.1785(S0.6088)2+1.2796(S0.6088)(x0.2119)+4.2309(x0.2119)2=2×0.082×ln11P.

    Further if the fiducial probability P=0.95, then the confidence ellipse for model (1.3) is shown in Figure 2e. One can see from Figure 2e that the random states of system are distributed around the corresponding deterministic equilibrium, and belong to the interior of the confidence ellipse with large probability 0.95. $

    Example 4.2. Consider the large noise σ2. Let σ2=0.55 and keep other parameters the same as in Example 4.1. Then λ0.1353<0, the condition of Theorem 3.2 is satisfied. Figure 3 depicts the solutions S(t) and x(t) of system (1.3) and the corresponding deterministic model. We can see that microorganism population go to extinction in the stochastic system (1.3) while it is still persistent in the deterministic model. Thus, larger white noise σ2 is harmful to microorganism population.

    Figure 3.  Time series diagrams of system (1.3) and the corresponding deterministic model (2.1).

    Moreover, it can be seen from the expression of λ that σ1 also has an important effect on microorganism population. From Figure 4 we can see that the peak of the density for ν0 becomes smaller with the increase of the noise σ1, Resonating with the convexity of p(S)=mSa+S, λ as a function of σ1 is not a monotonic function, while λs is a decreasing function about σ1. Remark 2.2 points out for the convex function p(S)=mSa+S and γ=0, Eν0[p(S)]>p(S0). Hence, there are parameter sets such that λ>0>λs. If σ2=0, σ10, then λ>λs. $

    Figure 4.  The density of the stationary distribution ν0 in differential cases: σ1=0.45, 0.5, 0.55 and 0.6, respectively.

    Example 4.3. Consider the effect of the instantaneous nutrient recycling on microorganism. Let S0=2, D=0.1, m=0.6, a=0.1, δ=0.3, γ=0.4, σ1=0.08 and σ2=0.07. Take b=0, 0.75. Although the nutrient recovery rate b is not a key factor affecting the persistence and extinction of microorganism population, numerical simulations show that b may affect the concentration of microorganism. From Figure 5 we can see that the concentration of microorganism population will become larger with the increase of b.

    Figure 5.  (a) Time series diagrams of microorganism population x(t) for system (1.3) with b=0 and b=0.75, respectively. (b) Phase diagrams of system (1.3) with b=0 and b=0.75, respectively.

    In this paper, we proposed a stochastic chemostat model with instantaneous nutrient recycling. Then, we constructed a key threshold value λ which enabled us to obtain the asymptotic behavior of the system. Concretely, if λ<0, then P{limtlnx(t)t=λ}=1, biologically, microorganism population will go extinct with exponential rate; if λ>0, there exists an invariant probability measure such that the distribution of (S(t),x(t)) approximate to this invariant probability at infinite time, biologically, microorganism population is persistent. Although it failed to give an explicit analytic formula of the invariant probability, we constructed the stochastic sensitivity matrix by using the SSF technique to describe the scatter of trajectories in the neighborhood of stable equilibrium. It should be emphasized that if γ=0, then system (1.3) becomes the (1.2) with the case p(S)=mSa+S. We propose a new method to define the threshold value λ instead of λs in [28]. Moreover, different from [28], the condition in Theorem 3.1 (Theorem 3.2) is sufficient and almost necessary condition of the exponential extinction (persistence) for the model. However, the case of λ=0 remains unsolved. We will continue this interesting work with the concluding of this paper.

    This work was supported by Basic Research Project (Free Exploration) of Shanxi Province (No.20210302124256), Shanxi Province Department of Finace under Grant No. Z24179 and Talent Special Project of Research Project of Guangdong Polytechnic Normal University under Grant No.2021SDKYA053. We thank the three referees for the careful reading of our paper and all of the insightful suggestions and comments that greatly improved the presentation of the paper.

    The authors declare that they have no conflict of interest.



    [1] L. Becks, F. M. Hilker, H. Malchow, K. Jürgens, H. Arndt, Experimental demonstration of chaos in a microbial food web, Nature, 435 (2005), 1226–1229. https://doi.org/10.1038/nature03627 doi: 10.1038/nature03627
    [2] A. Novick, L. Szilard, Description of the chemostat, Science, 112 (1950), 715–716. https://doi.org/10.1126/science.112.2920.715 doi: 10.1126/science.112.2920.715
    [3] A. W. Bush, A. E. Cool, The effect of time delay and growth rate inhibition in the bacterial treatment of wastewater, J. Theor. Biol., 63 (1976), 385–395. https://doi.org/10.1016/0022-5193(76)90041-2 doi: 10.1016/0022-5193(76)90041-2
    [4] S. Ruan, S. K. Wolkowicz, Bifurcation analysis of a chemostat model with a distributed delay, J. Math. Anal. Appl., 204 (1996), 786–812. https://doi.org/10.1006/jmaa.1996.0468 doi: 10.1006/jmaa.1996.0468
    [5] X. Meng, Q. Zhao, L. Chen, Global qualitative analysis of new Monod type chemostat model with delayed growth response and pulsed input in polluted environment, Appl. Math. Mech., 29 (2008), 75–87. https://doi.org/10.1007/s10483-008-0110-x doi: 10.1007/s10483-008-0110-x
    [6] M. Rehim, L. L. Sun, X. Abdurahman, Z. D. Teng, Study of chemostat model with impulsive input and nutrient recycling in a environment, Commun. Nonlinear Sci., 16 (2011), 2563–2574. https://doi.org/10.1016/j.cnsns.2010.09.030 doi: 10.1016/j.cnsns.2010.09.030
    [7] G. Pang, F. Wang, L. Chen, Study of Lotka-volterra food chain chemostat with periodically varying dilution rate, J. Math. Chem., 43 (2008), 901–913. https://doi.org/10.1007/s10910-007-9263-5 doi: 10.1007/s10910-007-9263-5
    [8] T. Wang, L. Chen, Global analysis of a three-dimensional delayed Michaelis-Menten chemostat-type models with pulsed input, J. Appl. Math. Comput., 35 (2011), 211–227. https://doi.org/10.1007/s12190-009-0352-4 doi: 10.1007/s12190-009-0352-4
    [9] G. Pang, F. Wang, L. Chen, Analysis of a Monod-Haldene type food chain chemostat with periodically varying substrate, Chaos Soliton. Fract., 38 (2008), 731–742. https://doi.org/10.1016/j.chaos.2007.01.018 doi: 10.1016/j.chaos.2007.01.018
    [10] G. J. Butler, G. S. K. Wolkowicz, A mathematical model of the chemostat with a general class of functions describing nutrient uptake, SIAM J. Appl. Math., 45 (1985), 138–151. https://doi.org/10.1137/0145006 doi: 10.1137/0145006
    [11] F. Campillo, M. Joannides, I. Larramendy-Valverde, Stochastic modeling of the chemostat, Ecol. Model., 222 (2011), 2676–2689. https://doi.org/10.1016/j.ecolmodel.2011.04.027 doi: 10.1016/j.ecolmodel.2011.04.027
    [12] L. Imhof, S. Walcher, Exclusion and persistence in deterministic and stochastic chemostat models, J. Differ. Equ., 217 (2005), 26–53. https://doi.org/10.1016/j.jde.2005.06.017 doi: 10.1016/j.jde.2005.06.017
    [13] H. J. Alsakaji, F. A. Rihan, A. Hashish, Dynamics of a tsochastic epidemic model with vaccination and multiple time-delays for COVID-19 in the UAE, Complexity, 2022 (2022), 4247800. https://doi.org/10.1155/2022/4247800 doi: 10.1155/2022/4247800
    [14] H. J. Alsakaji, F. A. Rihan, Stochastic delay differential equations of three-species prey-predator system with cooperation among prey species, Discrete Contin. Dyn. Syst., 15 (2022), 245–263. https://doi.org/10.3934/dcdss.2020468 doi: 10.3934/dcdss.2020468
    [15] X. Meng, L. Wang, T. Zhang, Global dynamics analysis of a nonlinear impulsive stochastic chemostat system in a polluted environment, J. Appl. Anal. Comput., 6 (2016), 865–875. https://doi.org/10.11948/2016055 doi: 10.11948/2016055
    [16] S. Sun, Y. Sun, G. Zhang, X. Liu, Dynamical behavior of a stochastic two-species Monod competition chemostat model, Appl. Math. Comput., 298 (2017), 153–170. https://doi.org/10.1016/j.amc.2016.11.005 doi: 10.1016/j.amc.2016.11.005
    [17] X. Lv, X. Meng, X. Wang, Extinction and stationary distribution of an impulsive stochastic chemostat model with nonlinear perturbation, Chaos Soliton. Fract., 110 (2018), 273–279. https://doi.org/10.1016/j.chaos.2018.03.038 doi: 10.1016/j.chaos.2018.03.038
    [18] D. H. Nguyen, N. Nguyen, G. Yin, General nonlinear stochastic systems motivated by chemostat models: Complete characterization of long-time behavior, optimal controls, and applications to wastewater treatment, Stoch. Proc. Appl., 130 (2020), 4608–4642. https://doi.org/10.1016/j.spa.2020.01.010 doi: 10.1016/j.spa.2020.01.010
    [19] J. Yang, Z. Zhao, X. Song, Statistical property analysis for a stochastic chemostat model with degenerate diffusion, AIMS Mathematics, 8 (2023), 1757–1769. https://doi.org/10.3934/math.2023090 doi: 10.3934/math.2023090
    [20] R. Baratti, J. Alvarez, S. Tronci, M. Grosso, A. Schaum, Characterization with Fokker-Planck theory of the nonlinear stochastic dynamics of a class of two-state continuous bioreactors, J. Process Contr., 102 (2021), 66–84. https://doi.org/10.1016/j.jprocont.2021.04.004 doi: 10.1016/j.jprocont.2021.04.004
    [21] A. Schaum, S. Tronci, R. Baratti, J. Alvarez, On the dynamics and robustness of the chemostat with multiplicative noise, IFAC-PapersOnLine, 54 (2021), 342–347. https://doi.org/10.1016/j.ifacol.2021.08.265 doi: 10.1016/j.ifacol.2021.08.265
    [22] X. Zhang, R. Yuan, The existence of stationary distribution of a stochastic delayed chemostat model, Appl. Math. Lett., 93 (2019), 15–21. https://doi.org/10.1016/j.aml.2019.01.034 doi: 10.1016/j.aml.2019.01.034
    [23] X. Zhang, R. Yuan, Stochastic properties of solution for a chemostat model with a distributed delay and random disturbance, Int. J. Biomath., 13 (2020), 2050065. https://doi.org/10.1142/s1793524520500667 doi: 10.1142/s1793524520500667
    [24] S. Sun, X. Zhang, Asymptotic behavior of a stochastic delayed chemostat model with nutrient storage, J. Biol. Syst., 26 (2018), 225–246. https://doi.org/10.1142/S0218339018500110 doi: 10.1142/S0218339018500110
    [25] F. Mazenc, S. L. Niculescu, G. Robledo, Stability analysis of mathematical model of competition in a chain of chemostats in series with delay, Appl. Math. Model., 76 (2019), 311–329. https://doi.org/10.1016/j.apm.2019.06.006 doi: 10.1016/j.apm.2019.06.006
    [26] W. Wang, W. Chen, Persistence and extinction of Markov switched stochastic Nicholson's blowies delayed differential equation, Int. J. Biomath., 13 (2020), 2050015. https://doi.org/10.1142/S1793524520500151 doi: 10.1142/S1793524520500151
    [27] G. Stephanopoulos, R. Aris, A. G. Fredrickson, A stochastic analysis of the growth of competing microbial populations in a continuous biochemical reactor, Math. Biosci., 45 (1979), 99–135. https://doi.org/10.1016/0025-5564(79)90098-1 doi: 10.1016/0025-5564(79)90098-1
    [28] L. Wang, D. Jiang, A note on the stationary distribution of the stochastic chemostat model with general response functions, Appl. Math. Lett., 73 (2017), 22–28. https://doi.org/10.1016/j.aml.2017.04.029 doi: 10.1016/j.aml.2017.04.029
    [29] R. M. Nisbet, W. Gurney, Model of material cycling in a closed ecosystem, Nature, 264 (1976), 633–634. https://doi.org/10.1038/264633a0 doi: 10.1038/264633a0
    [30] S. Ruan, Persistence and coexistence in zooplankton-phytoplankton-nutrient models with instantaneous nutrient recycling, J. Math. Biol., 31 (1993), 633–654. https://doi.org/10.1007/BF00161202 doi: 10.1007/BF00161202
    [31] Y. M. Svirezhev, D. O. Logofet, Stability of Biological Communities, Mir Publishers, 1983.
    [32] N. H. Du, N. H. Dang, N. T. Dieu, On stability in distribution of stochastic diferential delay equations with markovian switching, Syst. Control Lett., 65 (2014), 43–49. https://doi.org/10.1016/j.sysconle.2013.12.006 doi: 10.1016/j.sysconle.2013.12.006
    [33] N. T. Dieu, T. Fugo, N. H. Du, Asymptotic behaviors of stochastic epidemic models with jump-diffusion, Appl. Math. Model., 86 (2020), 259–270. https://doi.org/10.1016/j.apm.2020.05.003 doi: 10.1016/j.apm.2020.05.003
    [34] N. H. Du, N. N. Nhu, Permanence and extinction for the stochastic SIR epidemic model, J. Differ. Equ., 269 (2020), 9619–9652. https://doi.org/10.1016/j.jde.2020.06.049 doi: 10.1016/j.jde.2020.06.049
    [35] W. Kliemann, Recurrence and invariant measures for degenerate diffusions, Ann. Probab., 15 (1987), 690–707. https://doi.org/10.1214/aop/1176992166 doi: 10.1214/aop/1176992166
    [36] D. J. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. https://doi.org/10.1137/s0036144500378302 doi: 10.1137/s0036144500378302
    [37] L. B. Ryashko, I. A. Bashkirtseva, On control of stochastic sensitivity, Automat. Rem. Contr., 69 (2008), 1171–1180. https://doi.org/10.1134/S0005117908070084 doi: 10.1134/S0005117908070084
  • This article has been cited by:

    1. Xin Xu, Baodan Tian, Xingzhi Chen, Yanhong Qiu, Dynamics of a stochastic food chain chemostat model with Monod–Haldane functional response and Ornstein–Uhlenbeck process, 2024, 225, 03784754, 495, 10.1016/j.matcom.2024.05.014
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1491) PDF downloads(71) Cited by(1)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog