Loading [MathJax]/jax/output/SVG/jax.js
 

Asymptotic behavior of a delayed stochastic logistic model with impulsive perturbations

  • Received: 23 June 2016 Accepted: 27 September 2016 Published: 01 October 2017
  • MSC : Primary: 65C20, 65C30; Secondary: 92D25

  • In this paper, we investigate the dynamics of a delayed logistic model with both impulsive and stochastic perturbations. The impulse is introduced at fixed moments and the stochastic perturbation is of white noise type which is assumed to be proportional to the population density. We start with the existence and uniqueness of the positive solution of the model, then establish sufficient conditions ensuring its global attractivity. By using the theory of integral Markov semigroups, we further derive sufficient conditions for the existence of the stationary distribution of the system. Finally, we perform the extinction analysis of the model. Numerical simulations illustrate the obtained theoretical results.

    Citation: Sanling Yuan, Xuehui Ji, Huaiping Zhu. Asymptotic behavior of a delayed stochastic logistic model with impulsive perturbations[J]. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1477-1498. doi: 10.3934/mbe.2017077

    Related Papers:

    [1] Chun Lu, Bing Li, Limei Zhou, Liwei Zhang . Survival analysis of an impulsive stochastic delay logistic model with Lévy jumps. Mathematical Biosciences and Engineering, 2019, 16(5): 3251-3271. doi: 10.3934/mbe.2019162
    [2] Yan Zhang, Shujing Gao, Shihua Chen . Modelling and analysis of a stochastic nonautonomous predator-prey model with impulsive effects and nonlinear functional response. Mathematical Biosciences and Engineering, 2021, 18(2): 1485-1512. doi: 10.3934/mbe.2021077
    [3] 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
    [4] Yan Wang, Tingting Zhao, Jun Liu . Viral dynamics of an HIV stochastic model with cell-to-cell infection, CTL immune response and distributed delays. Mathematical Biosciences and Engineering, 2019, 16(6): 7126-7154. doi: 10.3934/mbe.2019358
    [5] Cristina Anton, Alan Yong . Stochastic dynamics and survival analysis of a cell population model with random perturbations. Mathematical Biosciences and Engineering, 2018, 15(5): 1077-1098. doi: 10.3934/mbe.2018048
    [6] Biwen Li, Qiaoping Huang . Synchronization of time-delay systems with impulsive delay via an average impulsive estimation approach. Mathematical Biosciences and Engineering, 2024, 21(3): 4501-4520. doi: 10.3934/mbe.2024199
    [7] 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
    [8] Zeyan Yue, Sheng Wang . Dynamics of a stochastic hybrid delay food chain model with jumps in an impulsive polluted environment. Mathematical Biosciences and Engineering, 2024, 21(1): 186-213. doi: 10.3934/mbe.2024009
    [9] Qi Zhou, Huaimin Yuan, Qimin Zhang . Dynamics and approximation of positive solution of the stochastic SIS model affected by air pollutants. Mathematical Biosciences and Engineering, 2022, 19(5): 4481-4505. doi: 10.3934/mbe.2022207
    [10] Yansong Pei, Bing Liu, Haokun Qi . Extinction and stationary distribution of stochastic predator-prey model with group defense behavior. Mathematical Biosciences and Engineering, 2022, 19(12): 13062-13078. doi: 10.3934/mbe.2022610
  • In this paper, we investigate the dynamics of a delayed logistic model with both impulsive and stochastic perturbations. The impulse is introduced at fixed moments and the stochastic perturbation is of white noise type which is assumed to be proportional to the population density. We start with the existence and uniqueness of the positive solution of the model, then establish sufficient conditions ensuring its global attractivity. By using the theory of integral Markov semigroups, we further derive sufficient conditions for the existence of the stationary distribution of the system. Finally, we perform the extinction analysis of the model. Numerical simulations illustrate the obtained theoretical results.


    1. Introduction

    This paper considers the long-term behavior of a system that results from impulsive and stochastic perturbations of a deterministic dynamical system with delay. This type of dynamics occurs naturally for example in the modeling of biological systems such as a growing bacterial colony, modeled deterministically, from which samples of random size are drawn regularly in an experiment or a marine ecosystem from which fish are harvested with a net [2], or the modeling of stochasticity in gene regulatory networks with feedback through a cell's signaling system [19], etc.

    The associated mathematical model with stochastic perturbations

    dx(t)=x(t)(r(t)a(t)x(t))dt+σ(t)x(t)dB(t), (1)

    where x(t) is the population size, B(t) is a standard Brownian motion, r(t), a(t) and σ(t) are continuous bounded functions on [0,), has been considered by many authors [3,6,11,12,15,16]. Recently, Liu and Wang [17] studied a stochastic logistic system with impulsive effects which takes the following form

    {dx(t)=x(t)(r(t)a(t)x(t))dt+σ(t)x(t)dB(t), ttk,x(t+k)x(tk)=bkx(tk), t=tk,kN, (2)

    where N denotes the set of positive integers. The time sequence {tk} is strictly increasing such that limktk=+. The term bk is impulsive perturbations at the moments of time tk, and satisfies that 1+bk>0 for all kN. The authors obtained the sufficient conditions for extinction, non-persistence in the mean, weak persistence, persistence in the mean and stochastic permanence of the system.

    As is well known, time delay always exists in the evolutionary processes of the population (e.g., resource regeneration times, maturation periods, feeding times, reaction times, etc) and it can cause the oscillatory or even unstable phenomena (see monographs [8,14]). Therefore, it seems more realistic to consider model (2) with delay. As far as we know, few investigations have been made on such type system. Based on model (2), we propose the following model:

    {dx(t)=x(t)(rax(t)c0f(s)x(ts)ds)dt+σ(t)x(t)dB(t), ttk,x(t+k)x(tk)=bkx(tk), t=tk,kN, (3)

    where c stands for the effect power of the past history. f(s), called the delay kernel, is a weighting factor which indicates how much emphasis should be given to the size of the population at earlier times to determine the present effect on resource availability[24]. Usually, we use the Gamma distribution delay kernel

    fnα(s)=αnsn1eαs(n1)!,

    where α>0 is a constant, n an integer, with the average delay T=n/α. The corresponding version of model (3) with a Gamma distribution delay kernel can be written as

    {dx(t)=x(t)[rax(t)c0fnα(s)x(ts)ds]dt+σx(t)dB(t), ttk,x(t+k)x(tk)=bkx(tk), t=tk, kN. (4)

    To perform a thorough analysis on model (3) with a general delay kernel function is very difficult. So in this paper we mainly devote our attention to the investigation on the dynamics of its special case model (4).

    Notice that in the absence of impulsive and random perturbations, it is obvious that model (4) has a stable positive equilibrium x=ra+c. However, under impulsive and random perturbations, the equilibrium does not exist. That is, the steady state of model (4) can no longer be represented by a single point. Therefore, we turn our attention to the study of the stationary distribution of the system. To the best of our knowledge, though some significant progress has been made in the techniques and methods of determining the existence of the stationary distribution and the stability of the density for stochastic differential equations [20,9], unfortunately, few works have been performed for the corresponding problems of impulsive stochastic differential equations up to now. So, our work in this paper may be considered as a first attempt to the investigation on the stationary distribution of the impulsive stochastic differential equations.

    The organization of this paper is as follows. In the next section, we show the existence and uniqueness of a global positive solution of model (4). Then, in section 3, we present our main results: we first carry out the global attractivity analysis of the model; then using the Hasminskii's method and constructing Lyapunov function, we prove the existence of the stationary distribution of the model; we also perform an extinction analysis of the model. Finally, some discussions and numerical simulations are presented in section 4.


    2. Existence and uniqueness of a global positive solution

    In this section, we show the uniqueness and global existence of a positive solution of model (4) for any positive initial value. Using the linear chain trick, if we define

    yn(t)=0fnα(s)x(ts)ds,

    then model (4) is equivalent to the following system

    {dx(t)=x(t)(rax(t)cyn(t))dt+σx(t)dB(t),dy1(t)=α(y1(t)x(t))dt,dyi(t)=α(yi(t)yi1(t))dt, i=2,3,,n,} ttk,x(t+k)x(tk)=bkx(tk), t=tk,kN. (5)

    Therefore, to study the dynamics of model (4), we need only to consider system (5).

    To begin with, we consider a general d-dimensional impulsive stochastic differential equation:

    {dX(t)=F(t,X(t))dt+G(t,X(t))dB(t), ttk, kN,X(t+k)X(tk)=bkX(tk), kN (6)

    with initial condition X(0)=X0Rd. B(t) denotes m-dimensional standard Brownian motions defined on the probability space (Ω,F,{Ft}t0,P), which is a complete probability space with a filtration {Ft}t0 satisfying the usual conditions (i.e. it is right continuous and F0 contains all P-null sets). Denote the differential operator L associated with Eq. (6) by

    L=t+di=1Fi(t,X)Xi+12di,j=1[GT(t,X)G(t,X)]ij2XiXj (7)

    for t[tk1,tk). Let PC1,2([tk1,tk)×Rd;R+) denote the family of all nonnegative functions V(t,X) on [tk1,tk)×Rd which are continuous once differentiable in t and twice differentiable in X. If L acts on such V(t,X), then

    LV(t,X)=Vt(t,X)+VX(t,X)F(t,X)+12trace[GT(t,X)VXX(t,X)G(t,X)].

    Definition 2.1. (See [17]) A stochastic process X(t)=(X1(t),X2(t)...,Xd(t))T,tR+, is said to be a solution of Eq. (6) if

    (ⅰ) X(t) is Ft-adapted and is continuous on (0,t1) and each interval (tk,tk+1)R+,kN; F(t,X(t))L1(R+;Rd),G(t,X(t))L2(R+;Rd), where Lk(R+;Rd) is all Rn valued measurable {Ft}-adapted processes f(t) satisfying T0|f(t)|kdt< almost surely for every T>0;

    (ⅱ) for each tk,kN,X(t+k)=limtt+kX(t) and X(tk)=limttkX(t) exist and X(tk)=X(tk) with probability one;

    (ⅲ) for almost all t[0,t1], X(t) obeys the integral equation

    X(t)=X(0)+t0F(s,X(s))ds+t0G(s,X(s))dB(s)

    and for almost all t(tk,tk+1],kN,X(t) obeys the integral equation

    X(t)=X(t+k)+t0F(s,X(s))ds+t0G(s,X(s))dB(s).

    Moreover, X(t) satisfies the impulsive conditions at each t=tk,kN with probability one.

    We now establish a fundamental lemma to reduce the dynamics of nonlinear stochastic differential system under impulsive perturbation to the corresponding problems of a nonlinear stochastic differential system without impulses.

    Lemma 2.1. Consider the following stochastic differential equation:

    dY(t)=f(t,Y(t))dt+g(t,Y(t))dB(t), (8)

    where

    f(t,Y(t))=0<tk<t(1+bk)1F(t,0<tk<t(1+bk)Y(t)),g(t,Y(t))=0<tk<t(1+bk)1G(t,0<tk<t(1+bk)Y(t))

    with the initial Y(0)=X0.

    ( i ) If Y(t) is a solution of Eq. (8), then X(t)=0<tk<t(1+bk)Y(t) is a solution of Eq. (6) on [0,);

    ( ii ) if X(t) is a solution of Eq. (6), then Y(t)=0<tk<t(1+bk)1X(t) is a solution of Eq. (8) on [0,).

    Proof. The proof is inspired by the method of Yan [26]. Fist, we prove (i). Let Y(t) be a possible solution to Eq. (8), it is easy to see that X(t)=0<tk<t(1+bk)Y(t) is continuous on each interval (tk,tk+1)[0,),kN, and for any ttk,

    dX(t)=d(0<tk<t(1+bk)Y(t))=0<tk<t(1+bk)f(t,Y(t))dt+0<tk<t(1+bk)g(t,Y(t))dB(t)=F(t,0<tk<t(1+bk)Y(t))dt+G(t,0<tk<t(1+bk)Y(t))dB(t)=F(t,X(t))dt+G(t,X(t))dB(t).

    So, X(t) satisfies Eq. (6) for almost everywhere on [0,)/{tk},k=1,2,...

    On the other hand, for each kN and tk[0,),

    X(t+k)=limtt+k0<tj<t(1+bj)Y(t)=0<tjtk(1+bj)Y(t+k)=(1+bk)0<tj<tk(1+bj)Y(tk)=(1+bk)X(tk)

    and

    X(tk)=limttk0<tj<t(1+bj)Y(t)=0<tj<tk(1+bj)Y(tk)=0<tj<tk(1+bj)Y(tk)=X(tk),

    which implies that X(t) is the solution of Eq. (6).

    We now prove (ⅱ). Since X(t) is a solution of Eq. (6) and is continuous on each interval (tk,tk+1)[0,+), thus Y(t)=0<tk<t(1+bk)1X(t) is continuous on (tk,tk+1)[0,+). What's more, we have that, for any kN and tk[0,+),

    Y(t+k)=limtt+k0<tj<t(1+bj)1X(t)=0<tjtk(1+bj)1X(t+k)=0<tj<tk(1+bj)1X(tk)=Y(tk)

    and

    Y(tk)=limttk0<tj<t(1+bj)1X(t)=0<tj<tk(1+bj)1X(tk)=Y(tk).

    Therefore, Y(t) is continuous on [0,+). Similar to the proof in the case (i), we can easily check that Y(t) is the solution of Eq. (8) on [0,+). This completes the proof of the Lemma.

    It should be emphasized that we consider Eqs. (6) and (8) on the same probability space. For Eq. (8), the associated differential operator L is the same as that defined in (7) for t[0,) rather than for t[tk1,tk).

    Now we turn to show the existence and uniqueness of the solution of system (5) by employing Lemma 2.1. Let x(t)=0<tk<t(1+bk)u(t) and yi(t)=vi(t), i=1,2,,n, then system (5) becomes

    {du(t)=u(t)(ra0<tk<t(1+bk)u(t)cvn(t))dt+σu(t)dB(t),dv1(t)=α(v1(t)0<tk<t(1+bk)u(t))dt,dvi(t)=α(vi(t)vi1(t))dt, i=2,3,,n. (9)

    By Lemma 2.1, if (u(t),v1(t),,vn(t)) is a solution of system (9), then (x(t),y1(t), ,yn(t)) is a solution of system (5). So, in the following, we only need to show the global existence of the positive solution of system (9).

    Lemma 2.2. There is a unique positive global solution (u(t),v1(t),,vn(t)) of system (9) a.s. for any initial value (u0,v10,,vn0)Rn+1+.

    Proof. Consider the system

    {dˉu(t)=(r12σ2a0<tk<t(1+bk)eˉu(t)ceˉvn(t))dt+σdB(t),dˉv1(t)=α(10<tk<t(1+bk)eˉu(t)eˉv1(t))dt,dˉvi(t)=α(1eˉvi1(t)eˉvi(t))dt, i=2,3,,n (10)

    with initial value ˉu(0)=lnu0, ˉvi(0)=lnvi0, i=1,2,,n. Obviously, the coefficients of system (10) satisfy the local Lipschitz condition, then there is a unique local solution (ˉu(t),ˉv1(t),,ˉvn(t)) on t[0,τe), where τe is the explosion time. By Itô's formula, it is easy to see

    (u(t),v1(t),,vn(t))=(eˉu(t),eˉv1(t),,eˉvn(t))

    is the unique positive local solution of system (9) with initial value (u0,v10,,vn0)Rn+1+.

    We now show the solution of system (9) is global, i.e., τe=. Since the solution is positive for t[0,τe), we have

    du(t)u(t)(ra0<tk<t(1+bk)u(t))dt+σu(t)dB(t).

    Let

    ϕ(t)=e(rσ22)t+σB(t)1u0+a0<tk<t(1+bk)t0e(rσ22)s+σB(s)ds,

    then ϕ(t) is the unique solution of the equation

    {dϕ(t)=ϕ(t)(ra0<tk<t(1+bk)ϕ(t))dt+σu(t)dB(t),ϕ(0)=u0,

    and by the comparison theorem for the stochastic equation, yields

    u(t)ϕ(t),t[0,τe), a.s.

    Besides, we can get

    dv1(t)(αv1(t)+α0<tk<t(1+bk)ϕ(t))dt.

    Obviously,

    φ1(t)=v10eαt+α0<tk<t(1+bk)t0ϕ(s)eα(ts)ds

    is the solution to the equation

    {dφ1(t)=(αφ1(t)+α0<tk<t(1+bk)ϕ(t))dt,φ1(0)=v10

    and v1(t)φ1(t),t[0,τe), a.s. Similarly, let φi, i=2,3,,n be the solution to the equation

    {dφi(t)=α(φi(t)φi1(t))dt,φi(0)=vi0.

    Then we have vi(t)φi(t), where φi(t)=vi0eαt+αt0φi1(s)eα(ts)ds.

    On the other hand,

    du(t)u(t)(ra0<tk<t(1+bk)u(t)cφn(t))dt+σu(t)dB(t).

    It follows that

    u(t)˜ϕ(t), t[0,τe), a.s.

    where

    ˜ϕ(t)=e(rσ22)tct0φn(s)ds+σB(t)1u0+a0<tk<t(1+bk)t0e(rσ22)scs0φn(τ)dτ+σB(s)ds.

    Then one has

    dv1(t)(αv1(t)+α0<tk<t(1+bk)˜ϕ(t))dt.

    Arguing as above, we can get

    v1(t)v10eαt+α0<tk<t(1+bk)t0˜ϕ(s)eα(ts)ds˜φ1(t),t[0,τe),a.s.

    Similarly, we have vi(t)˜φi(t), i=2,3,,n, where

    ˜φi(t)=vi0eαt+αt0˜φi1(s)eα(ts)ds.

    To sum up, we have that

    ˜ϕ(t)u(t)ϕ(t),  ˜φi(t)v(t)φi(t), i=1,2,,n, t(0,), a.s. (11)

    This completes the proof of the lemma.

    By Lemma 2.2, we have the following result on the global existence of the positive solution to system (5).

    Theorem 2.1. There is a unique positive global solution (x(t),y1(t),,yn(t)) to system (5) a.s. for any initial value (x0,y10,,yn0)Rn+1+.


    3. Asymptotic behavior of system (5)

    In this section, we will first investigate the global attractivity of system (5), then prove the existence of its stationary distribution, and finally, we perform an extinction analysis of the system. We first give the following fundamental assumptions on the impulsive perturbations bk.

    Assumption 3.1. 1+bk>0 for all kN.

    Assumption 3.2. There are two positive constants θ1 and θ2 such that for all t>0,

    θ10<tk<t(1+bk)θ2. (12)

    3.1. Global attractivity of system (5)

    Before proving the global attractivity of system (5), we first prepare some lemmas.

    Lemma 3.1. Let (u(t),v1(t),,vn(t)) be a solution of system (9) with any initial value (u(0),v1(0),,vn(0))Rn+1+. Then there are q>1, K1(q) and K2(q) such that

    lim suptE(uq(t))K1(q)   and   lim suptE(vqi(t))K2(q), i=1,2,,n.

    Proof. Applying Itô's formula to system (9), we compute

    E(etuq(t))=uq(0)+Et0esV(s)ds,

    where

    V(t)=aq0<tk<t(1+bk)uq+1(t)+(rq+12q(q1)σ2+1)uq(t)cquq(t)vn(t)aqθ1uq+1(t)+(rq+12q(q1)σ2+1)uq(t)K1(q).

    Therefore,

    E(etuq(t))uq(0)+K1(q)et.

    So

    lim suptE(uq(t))K1(q). (13)

    On the other hand, one can compute that

    dvq1(t)=qvq11(t)(αv1(t)+α0<tk<t(1+bk)u(t))dt.

    It follows that

    dE(vq1(t))dt=qαE[vq11(t)(v1(t)+0<tk<t(1+bk)u(t))]qαE(vq1(t))+qαθ2E(u(t)vq11(t)).

    Using Hölder inequality, one can then derive that

    dE(vq1(t))dtqαE(vq1(t))[1+θ2(Euq(t))1q(E(vq1(t)))1q]. (14)

    From (13) we know that for any ε>0, there exists a T>0 such that when t>T,

    E(uq(t))K1(q)+ε.

    It then follows from (14) that

    dE(vq1(t))dtqαE(vq1(t))[1+θ2(K1(q)+ε)1q(E(vq1(t)))1q].

    Using comparison theorem and noting also that ε is arbitrary, we obtain

    lim supt+E(vq1(t))θq2K1(q)K2(q). (15)

    Similarly, for i=2,3,,n, we compute that

    dE(vqi(t))dt=qα[E(vqi(t))+E(vq1i(t)vi1(t))]qαE(vqi(t))[1+(Evqi1(t))1q(E(vqi(t)))1q].

    Then we can deduce that

    lim supt+E(vqi(t))K2(q).

    This completes the proof of Lemma 3.1.

    Definition 3.1. (See [18]) Let X(t),ˉX(t) be two arbitrary solutions of Eq. (5) with initial values X(0)>0 and ˉX(0)>0, respectively. If

    limt+|X(t)ˉX(t)|=0 a.s.,

    then we say Eq. (5) is globally attractive.

    Lemma 3.2. (See [13]) Suppose that a stochastic process X(t) on t0 satisfies the condition

    E|X(t)X(s)|α1c1|ts|1+β, 0s, t<

    for some positive constants α1,β and c1. Then there exists a continuous modification ˜X(t) of X(t) which has the property that for every γ(0,β/α1), there is a positive random variable h(ω) such that

    P{ω:sup0<|ts|<h(ω),0s,t<|˜X(t,ω)˜X(s,ω)||ts|γ212γ}=1.

    In other words, almost every sample path of ˜X(t) is locally but uniformly Hölder continuous with exponent γ.

    Lemma 3.3. Let (u(t),v1(t),,vn(t)) be a solution of system (9) for any positive initial value (u(0),v1(0),,vn(0)), then almost every sample path of (u(t),v1(t), ,vn(t)) is uniformly continuous for t0.

    Proof. From (13) and the continuity of E(uq(t)) we have that there is a K1(q) such that for t0,

    E(uq(t))K1(q).

    Similarly, by (15) and the continuity of E(vqi(t)), there is a K2(q)>0 such that for t0,

    E(vqi(t))K2(q).

    The first equation of system (9) is equivalent to the following stochastic integral equation

    u(t)=u0+t0u(s)[ra0<tk<s(1+bk)u(s)cvn(s)]ds+t0σu(s)dB(s).

    Then

    E|u(t)[ra0<tk<t(1+bk)u(t)cvn(t)]|q0.5E|(u(t))|2q+0.5E|ra0<tk<t(1+bk)u(t)cvn(t)|2q0.5E|u(t)|2q+0.5×32q1[r2q+(a0<tk<t(1+bk))2qE|u(t)|2q+c2qE|vn(t)|2q]0.5K1(2q)+0.5×32q1[r2q+(aθ2)2qK1(2q)+c2qK2(2q)]L1(q).

    By the moment inequality for stochastic integrals, we know that for 0t1t2 and q>2,

    E|t2t1σu(s)dB(s)|qσq[q(q1)2]q2(t2t1)q21t2t1E|u(s)|qds.

    For 0<t1<t2<, t2t11, and 1p+1q=1, we have

    E(|u(t2)u(t1)|q)=E|t2t1u(s)[ra0<tk<s(1+bk)u(s)cvn(s)]ds+t2t1σu(s)dB(s)|q2q1E|t2t1u(s)[ra0<tk<s(1+bk)u(s)cvn(s)]ds|q+2q1E|t2t1σu(s)dB(s)|q2q1(t2t1)q1t2t1E|u(s)[ra0<tk<s(1+bk)u(s)cvn(s)]|qds+2q1σq[q(q1)2]q2(t2t1)q21t2t1E|u(s)|qds2q1(t2t1)qL1(q)+2q1σq[q(q1)2]q2(t2t1)q2K1(q)2q1(t2t1)q2[(t2t1)q2L1(q)+σq[q(q1)2]q2K1(q)](t2t1)q2L2(q),

    where L2(q)=2q1[L1(q)+σq[q(q1)2]q2K1(q)]. Then it follows from Lemma 3.2 that almost every sample path of u(t) is locally but uniformly Hölder-continuous with exponent γ1 for every γ1(0,q22q).

    From the second equation of system (9), we have

    v1(t)=v10+t0(αv1(s)+α0<tk<s(1+bk)u(s))ds.

    Then for 0<t1<t2< and t2t11, we compute that

    E|v1(t2)v1(t1)|q=E|t2t1(αv1(s)+α0<tk<s(1+bk)u(s))ds|q(t2t1)q1t2t1E|αv1(s)+α0<tk<s(1+bk)u(s)|qds2q1(t2t1)q1t2t1[αE|v1(s)|q+α0<tk<s(1+bk)E|u(s)|q]ds2q1(t2t1)q[αK2(q)+α0<tk<t(1+bk)K1(q)]2q1(t2t1)qL3(q).

    Similarly, for i=2,,n, we can compute that

    E|vi(t2)vi(t1)|q=E|t2t1(αvi(s)+αvi1(s))ds|q2q1(t2t1)q1t2t1[αE|vi(s)|q+αE|vi1(s)|q]ds2q1(t2t1)qαK2(q).

    In view of Lemma 3.2, almost every sample path of vi(t)(i=1,2,,n) is locally but uniformly Hölder-continuous with exponent γ2 for every γ2(0,q1q).

    Therefore, almost every sample path of (u(t),v1(t),,vn(t)) is uniformly continuous on t0.

    Lemma 3.4. (See [4]) Let f be a non-negative function defined on R+=[0,+) such that f is integrable on R+, and is uniformly continuous on R+, then

    limtf(t)=0.

    Now, we are in a position to prove the global attractivity of system (5). We have the following theorem.

    Theorem 3.1. Under Assumptions 3.1 and 3.2, if c<a, then system (5) is globally attractive.

    Proof. Let (x(t),y1(t),,yn(t)) and (ˉx(t),ˉy1(t),,ˉyn(t)) be two arbitrary solutions of system (5) with positive initial values, and suppose x(t)=0<tk<t(1+bk)u(t), yi=vi(t), ˉx(t)=0<tk<t(1+bk)ˉu(t), ˉyi=ˉvi(t). Then for V1(t)=|lnu(t)lnˉu(t)|, the right differential d+V1(t) of V1(t) is

    d+V1(t)=sgn(u(t)ˉu(t))d(lnu(t)lnˉu(t))=sgn(u(t)ˉu(t))[a0<tk<t(1+bk)(u(t)ˉu(t))c(vn(t)ˉvn(t))]dt[a0<tk<t(1+bk)|u(t)ˉu(t)|+c|vn(t)ˉvn(t)|]dt.

    Define V2=ni=1|viˉvi|, by directly calculating the right differential D+V2(t) of V2(t) and then making use of Itô's formula, we have

    d+V2(t)=ni=1sgn(vi(t)ˉvi(t))d(vi(t)ˉvi(t))[α|v1(t)ˉv1(t)|+α0<tk<t(1+bk)|u(t)ˉu(t)|]dt+[αni=2|vi(t)ˉvi(t)|+αni=2|vi1(t)ˉvi1(t)|]dt=[α|vn(t)ˉvn(t)|+α0<tk<t(1+bk)|u(t)ˉu(t)|]dt

    Then for V(t)=V1(t)+λV2(t), λ(c/α,a/α), one has

    d+V(t)[(aλα)0<tk<t(1+bk)|u(t)ˉu(t)|(αλc)|vn(t)ˉvn(t)|]dt. (16)

    Integrating both sides of (16) from 0 to t yields

    V(t)+t0[(aλα)0<tk<s(1+bk)|u(s)ˉu(s)|+(αλc)|vn(s)ˉvn(s)|]dsV(0)<,

    which leads to

    |u(t)ˉu(t)|L1[0,) and |vi(t)ˉvi(t)|L1[0,), i=1,2,,n.

    It follows from Lemmas 3.3 and 3.4 that

    limt|u(t)ˉu(t)|=limt|vi(t)ˉvi(t)|=0.

    Consequently, by Assumption 3.2, we obtain

    limt|x(t)ˉx(t)|=limt|yi(t)ˉyi(t)|=0, i=1,2,,n.

    3.2. Stationary distribution of system (5)

    In this subsection, we will prove the existence of a stationary distribution of system (5). Due to the complexity of system (5) and the lake of effective mathematical techniques available, we only consider the special case for Gamma distribution delay kernel with n=1, i.e., f(s)=αeαs, which is called weak delay kernel, indicating that the maximum weighted response of the growth rate is due to current population density while past densities have (exponentially) decreasing influence [24].

    For convenience, we first study system (9) since systems (5) and (9) are equivalent when we set x(t)=0<tk<t(1+bk)u(t) and y(t)=v(t). Furthermore, we assume that

    limt0<tk<t(1+bk)=θ (17)

    holds and let (˜u(t),˜v(t)) be the solution of

    {d˜u(t)=˜u(t)(raθ˜u(t)c˜v(t))dt+σ˜u(t)dB(t),d˜v(t)=(α˜v(t)+αθ˜u(t))dt. (18)

    Notice that system (18) is the limit system of (9), then by the global attractivity of (9), we only need to study the existence of stationary distribution of system (18).

    Let ˜u(t)=e˜ξ(t) and ˜v(t)=e˜η(t), then system (18) becomes

    {d˜ξ(t)=(rσ22aθe˜ξ(t)ce˜η(t))dt+σdB(t),d˜η(t)=(α+αθe˜ξ(t)˜η(t))dt. (19)

    We now study the existence of stationary distribution of the equivalent system (19) of (18). Let X=R2, Σ be the σalgebra of Borel subsets of X, and m be the Lebesgue measure on (X,Σ). We denote by P(t,˜x,˜y,A) the transition probability function for the diffusion process (˜ξ,˜η), i.e. P(t,˜x,˜y,A) =Prob((˜ξ,˜η)A) and (˜ξ,˜η) is the solution of system (19) with the initial condition (˜ξ(0),˜η(0))=(˜x,˜y). If the distribution of (˜ξ,˜η) is absolutely continuous with respect to the Lebesgue measure with the density U(t,˜x,˜y), then U(t,˜x,˜y) satisfies the following Fokker-Planck equation:

    Ut=12σ22U˜x2(f1(˜x,˜y)U)˜x(f2(˜x,˜y)U)˜y, (20)

    where

    f1(˜x,˜y)=rσ22aθe˜xce˜y,f2(˜x,˜y)=α+αθe˜x˜y. (21)

    Then for the asymptotical stability of system (19), we have the following theorem.

    Theorem 3.2. Let (˜ξ,˜η) be the solution of system (19). Then for every t(>0) the distribution of (˜ξ,˜η) has the density U(t,˜x,˜y) satisfying (20). Furthermore, if r>12σ2, then there exists a unique density U(˜x,˜y) such that

    limtR2|U(t,˜x,˜y)U(˜x,˜y)|d˜xd˜y=0 (22)

    and

    suppU=R2.

    Remark 1. By the support of a measurable function f we simply mean the set

     suppf={(˜x,˜y)R2:f(˜x,˜y)0}.

    Besides, note that the Fokker-Planck equation corresponding to system (19) is of a degenerate type, thus the asymptotic stability of the system can't follow directly from the known results from Hasminskii [9]. We will show it by using the theory of integral Markov semigroups (see Appendix A). Now we introduce an integral Markov semigroup connected with system (19).

    Denote by k(t,˜x,˜y;˜x0,˜y0) the density of P(t,˜x0,˜y0,). Then

    P(t)f(˜x,˜y)=R2k(t,˜x,˜y;u,v)f(u,v)dudv

    and consequently {P(t)}t0 is an integral Markov semigroup. Thus the asymptotic stability of the semigroup {P(t)}t0 implies that all the densities of the process (˜ξ(t),˜η(t)) convergence to an invariant density in L1. Therefore, for Theorem 3.2, we only need to show the asymptotic stability of the semigroup {P(t)}t0. The outline of our proof is as follows:

    (ⅰ) First, using the Hörmander condition [21], we show that the transition function of the process (˜ξ(t),˜η(t)) is absolutely continuous;

    (ⅱ) Then according to support theorems [5,1], we find a set E on which the density of the transition function is positive;

    (ⅲ) Next we show that the set E is an attractor and the semigroup satisfies the Foguel alternative [25,22,23];

    (ⅳ) Finally, we exclude sweeping by showing that there exists a Khasminski˘i function.

    In the following, we give the proof of Theorem 3.2 through four lemmas in succession, which correspond respectively to (ⅰ)-(ⅳ) above.

    Lemma 3.5. The transition probability function P(t,˜x0,˜y0,A) has a continuous density k(t,˜x,˜y; ˜x0,˜y0).

    Proof. The proof is based on the Hörmander theorem for the existence of smooth densities of the transition probability function for degenerate diffusion processes. If a(x) and b(x) are vector fields on Rd, then the Lie bracket [a,b] is a vector field given by

    [a,b]j(X)=dk=1(akbjxk(X)bkajxk(X))T,   j=1,2,,d.

    Let

    a(˜x,˜y)=(rσ22aθe˜xce˜y,α+αθe˜x˜y)T

    and

    b(˜x,˜y)=(σ,0)T.

    Then

    [a,b](˜x,˜y)=(σaθe˜x,σαθe˜x˜y)T.

    It follows that

    |σ0σaθe˜xσθe˜x˜y|=σ2αθe˜x˜y.

    Thus for every (˜x,˜y)R2, vectors b(˜x,˜y) and [a,b](˜x,˜y) span the space R2. This implies that the transition probability function P(t,˜x0,˜y0,A) has a density k(t,˜x,˜y;˜x0,˜y0) and kC((0,)×R2×R2). This completes the proof of lemma 3.5.

    Lemma 3.6. Let E=R2. Then for each (˜x0,˜y0)E and (˜x,˜y)E, there exists T>0 such that k(T,˜x,˜y;˜x0,˜y0)>0.

    Proof. We now use support theorems to check that the kernel k is positive. Fix a point (˜x0,˜y0)R2 and a continuous function ϕL2([0,T];R). Consider the following system

    ˜xϕ(t)=˜x0+t0[f1(˜xϕ(s),˜yϕ(s))+σϕ]ds, (23)
    ˜yϕ(t)=˜y0+t0f2(˜xϕ(s),˜yϕ(s))ds, (24)

    where f1(˜xϕ,˜yϕ)=rσ22aθe˜xϕce˜yϕ and f2(˜xϕ,˜yϕ)=α+αθe˜xϕ˜yϕ.

    Let D˜x0,˜y0;ϕ be the Frechét derivative of the function hXϕ+h(T) with Xϕ+h=[xϕ+h,yϕ+h]T. If for some ϕ the derivative D˜x0,˜y0;ϕ has rank 2, then we get k(T,˜x,˜y;˜x0,˜y0)>0 for x=˜xϕ(T) and y=˜yϕ(T). The derivative Dx0,˜y0;ϕ can be found by means of the perturbation method for ordinary differential equations. In other words, let Γ(t)=f(˜xϕ(t),˜yϕ(t)), where f is the Jacobians of f=[f1(˜x,˜y),f2(˜x,˜y)]T. Let Q(t,t0) be a matrix function such that Q(t0,t0)=I, Q(t,t0)t=Γ(t)Q(t,t0) for Ttt00 and v=(σ,0)T. Then

    D˜x0,˜y0;ϕh=t0Q(T,s)vh(s)ds.

    We first check that the rank of D˜x0,˜y0;ϕ is 2. Let ε(0,T) and h=1[Tε,T](t), where t[0.T] and 1[Tε,T] is the characteristic function of interval [Tε,T]. Since Q(T,s)=I+Γ(T)(Ts)+o(Ts), we obtain

    D˜x0,˜y0;ϕh=εv+12ε2Γ(T)v+o(ε2).

    Then, we have

    Γ(T)v=[aθe˜xce˜yαθe˜x˜yαθe˜x˜y][σ0]=σe˜x[aθαθe˜y].

    Hence, vectors v and Γ(T)v are linearly independent. Thus D˜x0,˜y0;ϕ has rank 2.

    Next, we prove that for any two points (˜x0,˜y0)E and (xT,yT)E, there exists a control function ϕ and T>0 such that ˜xϕ(0)=˜x0, ˜yϕ(0)=˜y0, ˜xϕ(T)=xT and ˜yϕ(T)=yT. Taking derivatives of systems (23) and (24) yield

    ˜xϕ(t)=f1(˜xϕ(t),˜yϕ(t))+σϕ, (25)
    ˜yϕ(t)=f2(˜xϕ(t),˜yϕ(t)). (26)

    We construct the function ϕ in the following way. First, we find a positive constant T and a differential function ˜xϕ:[0,T]R+ such that ˜yϕ(0)=˜y0, ˜yϕ(T)=˜yT,

    ˜yϕ(0)=α+αθe˜x0˜y0,  ˜yϕ(T)=α+αθe˜xT˜yT (27)

    and

    ˜yϕ(t)+α=αθe˜xt˜yt>0  for  t[0,T]. (28)

    Denote constants ˜yϕ(0) and ˜yϕ(T) by a0 and aT. We split the construction of the function ˜yϕ on three intervals [0,ε], [ε,Tε] and [Tε,T], where 0<ε<T/2. By (27), we can construct a C2 function ˜yϕ:[0,ε]R such that ˜yϕ(0)=˜y0, ˜yϕ(0)=a0, ˜yϕ(ε)=0 and ˜yϕ satisfies inequality (28) for t[0,ε]. Similarly, we construct a function ˜yϕ:[Tε,T]R such that ˜yϕ(T)=˜yT, ˜yϕ(T)=aT, ˜yϕ(Tε)=0 and ˜yϕ satisfies inequality (28) for t[Tε,T]. Let T be sufficiently large, we can extend the function ˜yϕ:[0,ε][Tε,T]R to a C2 function ˜xϕ defined on the whole interval [0,T] such that ˜yϕ satisfies inequality (28). From (28), we can find a C1 function ˜xϕ which satisfies (26) and finally we can determine a continuous function ϕ from (25). This completes the proof of Lemma 3.6.

    Notice that for every density f, one has

    limtR2P(t)f(˜x,˜y)d˜xd˜y=1.

    Lemma 3.7. The semigroup {P(t)}t0 is asymptotically stable or is sweeping with respect to compact sets.

    Proof. By Lemma 3.5, the semigroup {P(t)}t0 is an integral Markov semigroup with a continuous kernel k(t,˜x,˜y) for t>0. Let E=R2, then it is sufficient to investigate the restriction of the semigroup {P(t)}t0 to the space L1(ˉE), where ˉE denotes the closure of the set E. In view of Lemma 3.6, for every fD, we have

    0P(t)fdt>0  a.e.  on E,

    where D is defined in Appendix A. Therefore, according to Lemma A.1 in Appendix A, it follows that the semigroup {P(t)}t0 is asymptotically stable or is sweeping with respect to compact sets. This completes the proof of Lemma 3.7.

    Lemma 3.8. The semigroup {P(t)}t0 is asymptotically stable.

    Proof. In order to exclude sweeping, we now construct a nonnegative C2function V and a closed set ΓΣ such that

    sup˜x,˜yΓAV(˜x,˜y)<0,

    where A is the adjoint operator of the infinitesimal generator A of the semigroup {P(t)}t0, which is of the form

    AV=12σ22V˜x2+f1(˜x,˜y)V˜x+f2(˜x,˜y)V˜y, (29)

    where fi(˜x,˜y), i=1,2 are defined in (21). Define

    V(˜x,˜y)=θ(e˜xe˜xe˜x(˜x˜x))+12c(e˜ye˜y)2,

    where

    ˜x=lnraθ+cθ, ˜y=lnra+c.

    By (29), one has

    AV(˜x,˜y)=θ(e˜xe˜y)(rσ22aθe˜xce˜y)+12σ2θe˜x+c(e˜ye˜y)(αe˜y+αθe˜x)=aθ2(e˜xe˜x)2αc(e˜ye˜y)2+12σ2θe˜x

    Thus, the ellipsoid

    aθ2(e˜xe˜x)2+αc(e˜ye˜y)2=12σ2θe˜x

    lies entirely in R2. Thus, there exist a closed set ΓΣ which contains this ellipsoid and c>0 such that

    sup˜x,˜yΓAV(˜x,˜y)c<0.

    The function V is called a Khasminski˘i function. By means of standard arguments similar to those in [22], one can check that the semigroup is not sweeping from the set Γ due to the existence of a Khasminski˘i function. Therefore, according to Lemma A.1 in Appendix A, we conclude that the semigroup {P(t)}t0 is asymptotically stable.

    In view of Theorem 3.2, for system (18) there exists a unique positive invariant density with the support set R2+ since it is equivalent to system (19). Notice that system (18) is the limit system of (9) as limt0<tk<t(1+bk)=θ, then by the global attractivity of (9), we conclude that there exists a unique positive invariant density for system (9). Further notice that system (5) is equivalent to system (9). Therefore, we have the following result for system (5).

    Theorem 3.3. Assume that condition (17) holds. If r>12σ2 and c<a, then for system (5) there exists a unique positive invariant density with the support set R2+.


    3.3. Extinction of system (5)

    In this subsection, we show that large noise can lead to the extinction of system (5).

    Theorem 3.4. Let (x(t),y(t)) be a solution of system (5). Then

    limtlnx(t)tr12σ2.

    In particular, if r<12σ2, then limtx(t)=0 a.s.

    Proof. Applying Itô's formula to system (9), one has

    dlnu(t)=(r12σ2a0<tk<t(1+bk)u(t)cv(t))dt+σdB(t).

    Integrating both sides from 0 to t, we have

    lnu(t)lnu(0)+(r12σ2)t+M1(t), (30)

    where M1(t)=t0σdB(s). Note that M1(t) is a local martingale, whose quadratic variation is M1(t),M1(t)=σ2t. Making use of the strong law of large numbers for local martingales leads to

    limtM1(t)/t=0, a.s.

    From (30), we know

    lnx(t)=0<tk<tln(1+bk)+lnu(t)lnu(0)+lnθ+(r12σ2)t+M1(t). (31)

    It then follows from (31) that

    limtlnx(t)tr12σ2.

    Remark 2. From Theorems 3.3 and 3.4, one can see that if the intensity of the noise is small, there exists an invariant and asymptotically stable density of the system, while large noise will make the population extinct eventually. Noting further that if c<a and condition (17) hold, then r12σ2 can be considered as a threshold determining the asymptotical stability and the extinction of system.


    4. Simulations and discussions

    Impulsive and uncertain variability together with time delay are always present in a natural system, which should be accounted for in its mathematical model. The research performed in this paper is an attempt in this direction, using an impulsive stochastic model with delay. More specifically, the impulse is introduced at fixed moments, the stochastic perturbation is of white noise type and is assumed to be proportional to the population density, and the delay takes the distributed type with a weak delay kernel. To perform a detailed analysis on the dynamics of model (4), we first transform it to an equivalent stochastic system (5) with impulsive effects using the linear chain trick. Then based on Lemma 2.1, it can be further reduced to the problem of a nonlinear stochastic differential system without impulses, i.e., system (9).

    For system (5), we first carry out the analysis of its global attractivity. Theorem 3.1 shows that the system is globally attractive provided that c<a, Assumptions 3.1 and 3.2 hold. Note that Assumptions 3.1 and 3.2 are only relative to the impulse, so we see that the noise and the delay do not affect the global attractivity of the system. Then we study the existence of the stationary distribution of system (5). It is difficult to directly study its distribution, so we turn to study its equivalent system (9). Notice that (9) is non-autonomous and it has been showed to be globally attractive in Theorem 3.1. So we only need to study the limit system (18) of (9). Due to the Fokker-Planck equation corresponding to system (18) is of a degenerate type, thus the existence of the stationary distribution can't follow directly from the known result from Hasminskii [9]. We show it based on the theory of integral Markov semigroups. Theorem 3.3 shows that for system (5) there exists a unique invariant density with the support R2+ provided that r>12σ2, c<a and condition (17) hold. Moreover, it is shown in Theorem 3.4 that if r<12σ2, the population will be extinct eventually. Thus, under c<a and condition (17), r12σ2 can be considered as a threshold determining the asymptotical stability and the extinction of system (5).

    To illustrate the results obtained above, some numerical simulations are carried out by using Milstein scheme [10]. Consider the discretization of model (5) for t=0,Δt,2Δt,,nΔt:

    {xi+1=xi+xi(raxicyi)Δt+σxiΔtξi,yi+1=yi+(αyi+αxi)Δt,} ttk,x(t+k)x(tk)=bkx(tk), t=tk,kN.

    where time increment Δt>0 and ξi are N(0,1)distributed independent random variables which can be generated numerically by pseudorandom number generators.

    First, we demonstrate system (5) is asymptotical stability. For this purpose, let r=0.8, α=0.6, a=0.4, c=0.33, tk=k (kN), initial value (x(0),y(0))=(1,1), bk vary in the range [0.3/k,0.3/k] and σ vary in the range [0,0.03]. Simple computations show that σ2<2r=1.6 and limt0<tk<t(1+bk) exists.

    By Theorem 3.3, system (5) is asymptotical stability. Our simulation supports this conclusion as shown in Fig. 1, where we show the effect from different noise intensities σ and different impulse intensities bk, respectively.

    Figure 1. Trajectories of impulsive stochastic system (5) with different σ and bk. Here (S(0),x(0))=(1,1) and E(x,y)=(1.0909,1.8182). System (5) is asymptotically stable.

    It is seen in Fig. 1 (a) that the steady state of the system is point E(x,y)=(1.0909,1.8182) in the absence of impulsive and random perturbations (i.e. σ=bk=0). Increasing the value of σ and keeping bk the same, we can see that the steady state of the system can no longer be represented by a single point; instead, it is represented by a region around E. Shown in Fig. 1 (a)-(c), the larger the noise intensity is, the more diffusive of the system state is. When bk changes and σ stays 0, we can see in Fig. 1 (d)-(e) that there is an attractor for the system, but E is not included in the attraction, which is above the attraction in the case of bk<0 and is below the attraction in the case of bk>0. When we choose σ=0.03 and bk=0.3/k, the system is still stable, please see Fig. 1 (f).

    Fig. 1 shows the trajectories of system (5) under different values of parameters. However, we should point out that the trajectory in each subgraph is drawn for a single sample, which is stochastic for each sample under the same parameters, that is the outcome for a single trajectory is not predictable. But, the probability distribution of all possible outcomes can be determined. Our simulation supports this conclusion as shown in Fig. 2, where σ=0.03, bk=0.3/k and the densities are drawn based on 10000 sample pathes, computed with differential initial value and different iterative times, respectively. We can see from the figure the distribution is stationary.

    Figure 2. Probability densities of x(t) for system (5) based on 10000 sample pathes, computed with the noise intensity σ=0.03, and for (a) differential initial value and (b) different iterative times. There exists stationary distribution and the density is stable.

    Next, we show system (5) is extinct. To this end, we set initial value (x(0),y(0))=(1,1), r=0.3, a=0.4, c=0.33, α=0.6,  bk=0,  tk=5k and σ=0.8. Since r<12σ2=0.32, by Theorem 3.4, the population is extinct eventually. Our simulation supports this conclusion as shown in Figs. 3 (a) and (b). Decreasing the value of σ from 0.8 to 0.1, we obtain r>12σ2=0.005. Then the population is persistent and system (5) is asymptotically stable, please see Figs. 3 (c) and (d). Thus, as we mentioned before, under condition (17), r12σ2 can be considered as a threshold determining the asymptotical stability and the extinction of system (5).

    Figure 3. The dynamics of the system (5) with different σ. (a) The population x(t) is extinct with σ=0.8. (c) The population x(t) is persistent with σ=0.1.

    To sum up, this paper presents an investigation on the dynamics of a impulsive stochastic system with delay. Our findings are useful for better understanding of the effects of impulses, stochastic perturbations and delay on the dynamics of a system. We should point out there are still some other interesting topics meriting further investigation, for example, the long term behavior of multi-population system with impulsive and stochastic perturbations. We leave these for future considerations.


    Appendix A.

    Let the triple (X,Σ,m) be a σfinite measure space. Denote by D be the subset of the space L1 which contains all densities, i.e.

    D={fL1:f0,f=1}.

    A linear mapping P:L1L1 is called a Markov operator if P(D)D.

    The Markov operator P is called an integral or kernel operator if there exists a measurable function k:X×X[0,) such that

    Xk(˜x,˜y)m(d˜x)=1 (A.1)

    for all ˜yX and

    Pf(x)=Xk(˜x,˜y)f(˜y)m(d˜y)

    for every density f.

    A family {P(t)}t0 of Markov operators which satisfies conditions:

    (ⅰ) P(0)=Id,

    (ⅱ) P(t+s)=P(t)P(s) for s,t0,

    (ⅲ) for each fL1 the function tP(t)f is continuous with respect to the L1 norm is called a Markov semigroup. A Markov semigroup {P(t)}t0 is called integral, if for each t>0, the operator P(t) is an integral Markov operator.

    A density f is called invariant if P(t)f=f for each t>0. The Markov semigroup {P(t)}t0 is called asymptotically stable if there is an invariant density f such that

    limtP(t)ff=0  for  fD.

    A Markov semigroup {P(t)}t0 is called sweeping with respect to a set AΣ if for every fD

    limtAP(t)f(˜x)m(d˜x)=0.

    We need some result concerning asymptotic stability and sweeping which can be found in [7].

    Lemma A.1. Let X be a metric space and Σ be the σalgebra of Borel sets. Let {P(t)}t0 be an integral Markov semigroup with a continuous kernel k(t,˜x,˜y) for t>0, which satisfies (A.1) for all ˜yX. We assume that for every fD, we have

    0P(t)fdt>0  a.e.

    Then this semigroup is asymptotically stable or is sweeping with respect to compact sets.

    The property that a Markov semigroup {P(t)}t0 is asymptotically stable or sweeping for a sufficiently large family of sets (e.g. for all compact sets) is called the Foguel alternative.


    Acknowledgments

    We would like to thank anonymous reviewers for helpful suggestions which greatly improved this paper.


    [1] [ S. Aida,S. Kusuoka,D. Strook, On the support of Wiener functionals, Longman Scient. Tech., 284 (1993): 3-34.
    [2] [ T. Alkurdi,S. Hille,O. Gaans, Ergodicity and stability of a dynamical system perturbed by impulsive random interventions, J. Math. Anal. Appl., 407 (2013): 480-494.
    [3] [ L. Arnold, Stochastic Differential Equations: Theory and Applications, Wiley, New York-London-Sydney, 1974.
    [4] [ I. Barbalat, Systems dequations differentielles d'osci d'oscillations nonlineaires, Rev. Roumaine Math. Pures Appl., 4 (1959): 267-270.
    [5] [ G. Ben Arous,R. Léandre, Décroissance exponentielle du noyau de la chaleur sur la diagonale (Ⅱ), Probab. Theory Related Fields, 90 (1991): 377-402.
    [6] [ A. Freedman, Stochastic differential equations and their applications, Stochastic Differential Equations, 77 (1976): 75-148.
    [7] [ S. Foguel, Harris operators, Israel J. Math., 33 (1979): 281-309.
    [8] [ K. Gopalsamy, Stability and Oscillations in Delay Differential Equations of Population Dynamics, Springer-Verlag, New York, 1992.
    [9] [ R. Z. Has'minskii, Stochastic Stability of Differential Equations, Sijthoof & Noordhoof, Alphen aan den Rijn, The Netherlands, 1980.
    [10] [ D. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Review, 43 (2001): 525-546.
    [11] [ D. Jiang,N. Shi, A note on nonautonomous logistic equation with random perturbation, J. Math. Anal. Appl., 303 (2005): 164-172.
    [12] [ D. Jiang,N. Shi,X. Li, Global stability and stochastic permanence of a non-autonomous logistic equation with random perturbation, J. Math. Anal. Appl., 340 (2008): 588-597.
    [13] [ I. Karatzas and S. Shreve, Brownian Motion and Stochastic Calculus, Springer Verlag, Berlin, 1991.
    [14] [ Y. Kuang, Delay differential equations with applications in population dynamics, in Mathematics in Science and Engineering, Academic Press, New York, 1993.
    [15] [ X. Li,X. Mao, Population dynamical behavior of non-autonomous Lotka-Volterra competitive system with random perturbation, Discrete Contin. Dyn. Syst., 24 (2009): 523-545.
    [16] [ M. Liu,K. Wang, Persistence and extinction in stochastic non-autonomous logistic systems, J. Math. Anal. Appl., 375 (2011): 443-457.
    [17] [ M. Liu,K. Wang, On a stochastic logistic equation with impulsive perturbations, Comput. Math. Appl., 63 (2012): 871-886.
    [18] [ Z. Ma and Y. Zhou, Qualitative and Stability Method of Ordinary Differential Equation, Science Press, Beijing, 2001.
    [19] [ M. Mackey,M. Kamińska,R. Yvinec, Molecular distributions in gene regulatory dynamics, J. Theoret. Biol., 247 (2011): 84-96.
    [20] [ X. Mao, Stochastic Differential Equations and their Applications, Horwood publishing, Chichester, England, 1997.
    [21] [ J. Norris, Simplified Malliavin calculus, in SLeminaire de probabilitiLes XX, Lecture Notes in Mathematics, Springer, New York, 1024 (1986), 101–130.
    [22] [ K. Pichór,R. Rudnicki, Stability of Markov semigroups and applications to parabolic systems, J. Math. Anal. Appl., 215 (1997): 56-74.
    [23] [ K. Pichór,R. Rudnicki, Continuous Markov semigroups and stability of transport equations, J. Math. Anal. Appl., 249 (2000): 668-685.
    [24] [ S. Ruan, Delay differential equations in single species dynamics, in Delay Differential Equations and Applications, Springer, Berlin, 205 (2006), 477–517.
    [25] [ R. Rudnicki, On asymptotic stability and sweeping for Markov operators, Bull. Polish Acad. Math., 43 (1995): 245-262.
    [26] [ J. Yan, On the oscillation of impulsive neutral delay differential equations, Chinese Ann. Math., 21A (2000): 755-762.
  • This article has been cited by:

    1. Li Sun, Haitao Zhu, Yanhui Ding, Impulsive control for persistence and periodicity of logistic systems, 2020, 171, 03784754, 294, 10.1016/j.matcom.2019.06.006
    2. Xuejin Lv, Xinzhu Meng, Xinzeng Wang, Extinction and stationary distribution of an impulsive stochastic chemostat model with nonlinear perturbation, 2018, 110, 09600779, 273, 10.1016/j.chaos.2018.03.038
    3. Chaoqun Xu, Sanling Yuan, Richards Growth Model Driven by Multiplicative and Additive Colored Noises: Steady-State Analysis, 2020, 19, 0219-4775, 2050032, 10.1142/S0219477520500327
    4. Xinguo Sun, Wenjie Zuo, Daqing Jiang, Tasawar Hayat, Unique stationary distribution and ergodicity of a stochastic Logistic model with distributed delay, 2018, 512, 03784371, 864, 10.1016/j.physa.2018.08.048
    5. Zaowang Xiao, Xiangdong Xie, Yalong Xue, Stability and bifurcation in a Holling type II predator–prey model with Allee effect and time delay, 2018, 2018, 1687-1847, 10.1186/s13662-018-1742-4
    6. Jing Hu, Zhijun Liu, Incorporating two coupling noises into a nonlinear competitive system with saturation effect, 2020, 13, 1793-5245, 2050012, 10.1142/S1793524520500126
    7. Siyu Chen, Zhijun Liu, Ronghua Tan, Lianwen Wang, Survival analysis of an impulsive stochastic facultative mutualism system with saturation effect, 2021, 14, 1793-5245, 2150009, 10.1142/S1793524521500091
    8. Ming Liu, Dongpo Hu, Fanwei Meng, Stability and bifurcation analysis in a delay-induced predator-prey model with Michaelis-Menten type predator harvesting, 2020, 0, 1937-1179, 0, 10.3934/dcdss.2020259
  • Reader Comments
  • © 2017 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(3372) PDF downloads(501) Cited by(8)

Article outline

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog