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

Stochastic analysis of survival and sensitivity in a competition model influenced by toxins under a fluctuating environment

  • Received: 15 January 2024 Revised: 12 February 2024 Accepted: 20 February 2024 Published: 27 February 2024
  • MSC : 34F05, 37H10, 60J70, 92B05

  • This paper proposed a stochastic toxin-dependent competition model to investigate the impact of environmental noise on species interaction dynamics. First, a survival analysis was conducted to establish the sufficient conditions for population extinction and persistence. Second, we proved the existence of a unique ergodic stationary distribution. Finally, the spatial arrangement of random states near the deterministic attractor was investigated using the stochastic sensitivity functions technique. This analytical approach facilitates constructing confidence ellipses and estimating critical noise intensity corresponding to the onset of transition. Both theoretical and numerical findings demonstrated that significant levels of noise experienced by one species lead to its extinction while promoting persistence in its competitor; conversely, negligible levels of noise did not alter the original competition outcomes in the deterministic model. However, when both species encounter moderate levels of noise, various modifications can occur in competition outcomes. These findings have significant implications for preserving ecosystem diversity.

    Citation: Yuanlin Ma, Xingwang Yu. Stochastic analysis of survival and sensitivity in a competition model influenced by toxins under a fluctuating environment[J]. AIMS Mathematics, 2024, 9(4): 8230-8249. doi: 10.3934/math.2024400

    Related Papers:

    [1] Xiaoxia Guo, Dehao Ruan . Long-time analysis of a stochastic chemostat model with instantaneous nutrient recycling. AIMS Mathematics, 2023, 8(4): 9331-9351. doi: 10.3934/math.2023469
    [2] 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
    [3] Jawdat Alebraheem . Asymptotic stability of deterministic and stochastic prey-predator models with prey herd immigration. AIMS Mathematics, 2025, 10(3): 4620-4640. doi: 10.3934/math.2025214
    [4] Kai Zhang, Xinzhu Meng, Abdullah Khames Alzahrani . The ACE2 receptor protein-mediated SARS-CoV-2 infection: dynamic properties of a novel delayed stochastic system. AIMS Mathematics, 2024, 9(4): 8104-8133. doi: 10.3934/math.2024394
    [5] T. K. Buvaneshwari, D. Anuradha . Solving bi-objective bi-item solid transportation problem with fuzzy stochastic constraints involving normal distribution. AIMS Mathematics, 2023, 8(9): 21700-21731. doi: 10.3934/math.20231107
    [6] Shabir Ahmad, Saud Fahad Aldosary, Meraj Ali Khan . Stochastic solitons of a short-wave intermediate dispersive variable (SIdV) equation. AIMS Mathematics, 2024, 9(5): 10717-10733. doi: 10.3934/math.2024523
    [7] Yao Fu, Sisi Zhou, Xin Li, Feng Rao . Multi-assets Asian rainbow options pricing with stochastic interest rates obeying the Vasicek model. AIMS Mathematics, 2023, 8(5): 10685-10710. doi: 10.3934/math.2023542
    [8] Chenlu Zheng, Jianping Zhu . Promote sign consistency in cure rate model with Weibull lifetime. AIMS Mathematics, 2022, 7(2): 3186-3202. doi: 10.3934/math.2022176
    [9] Areej A. Almoneef, Abd-Allah Hyder, Mohamed A. Barakat, Abdelrheem M. Aly . Stochastic solutions of the geophysical KdV equation: Numerical simulations and white noise impact. AIMS Mathematics, 2025, 10(3): 5859-5879. doi: 10.3934/math.2025269
    [10] Chunli You, Linxin Shu, Xiao-bao Shu . Approximate controllability of second-order neutral stochastic differential evolution systems with random impulsive effect and state-dependent delay. AIMS Mathematics, 2024, 9(10): 28906-28930. doi: 10.3934/math.20241403
  • This paper proposed a stochastic toxin-dependent competition model to investigate the impact of environmental noise on species interaction dynamics. First, a survival analysis was conducted to establish the sufficient conditions for population extinction and persistence. Second, we proved the existence of a unique ergodic stationary distribution. Finally, the spatial arrangement of random states near the deterministic attractor was investigated using the stochastic sensitivity functions technique. This analytical approach facilitates constructing confidence ellipses and estimating critical noise intensity corresponding to the onset of transition. Both theoretical and numerical findings demonstrated that significant levels of noise experienced by one species lead to its extinction while promoting persistence in its competitor; conversely, negligible levels of noise did not alter the original competition outcomes in the deterministic model. However, when both species encounter moderate levels of noise, various modifications can occur in competition outcomes. These findings have significant implications for preserving ecosystem diversity.



    The earth's ecosystems are extensively affected by chemical contaminations resulting from a multitude of anthropogenic activities and natural phenomena, which significantly impacts the survival of diverse species and human health, particularly accelerating the extinction of endangered species [1]. To safeguard the ecological environment and maintain ecological equilibrium, it is imperative to accurately evaluate the risk posed by environmental toxins on exposed populations. In recent decades, mathematical models have gained widespread recognition as a powerful tool for assessing the risks of chemical contamination. Currently, the majority of population ecological models related to environmental pollution are based on the assumption that population growth rates adhere to the logistic equation [2,3,4,5]. However, these models fail to quantitatively analyze the impact of environmental toxins on population capacity. To address this limitation, Thieme [6] proposed using the Beverton-Holt equation instead of the logistic equation to describe the population growth rate. This approach can effectively differentiate between various toxins in terms of food uptake, conversion efficiency, and biomass acquisition. Taking into account the influence of toxins on population mortality, [7] built a toxin-dependent aquatic ecosystem model based on Thieme's approach and utilized available data to evaluate mercury's effect on rainbow trout.

    The aforementioned models are all single-species models, assuming that populations solely absorb toxins from their environment. To investigate the impact of environmental toxins on species interactions, some scholars have extended these single-species models into multi-species ones [8,9,10,11,12]. For instance, [13] established a prey-predator model considering the simultaneous exposure of both prey and predator to environmental toxins, revealing counterintuitive results suggesting that intermediate toxin concentrations could actually increase prey biomass. Peace et al. [14] successfully captured the phenomenon known as somatic growth dilution by developing a stoichiometric toxicant-mediated predator-prey model. Recently, Shan and Huang [15] proposed a toxin-dependent model to investigate the direct and indirect impacts of environmental toxins on two competing populations:

    dxidt=αimax{0,1βiyi}1+γixixi(kiyi+mi)xicix1x2,dyidt=aiTσiyiαimax{0,1βiyi}1+γixiyi+cixjyi, (1)

    where the subscript i represents species i, with i,j=1,2, and ij; xi denotes the concentration of population biomass; yi denotes the body burden; αi corresponds to the maximum growth rate; βi indicates the impact of toxin on population gain; γi captures the crowding effect; ki reflects the influence of toxin on mortality; mi stands for natural mortality rate; ci is the competition coefficient; ai denotes the uptake coefficient; and σi measures toxin elimination rate, while T refers to environmental toxin concentration.

    Considering the significantly higher rate of population metabolism compared to population growth, model (1) can be approximated as a two-dimensional model. To be specific, we introduce the following dimensionless quantities, ˜xi=γixi, ˜yi=βiyi, ˜t=α1t, ˜ki=kiβiα1, ˜mi=miα1, ˜α=α2α1, ˜T=β1a1Tσ1, ˜ci=ciα1γj, ˜a=β2a2β1a1, ˜σ=σ2σ1,ϵ=α1σ1, then drop the tildes, so that model (1) becomes

    dx1dt={max{0,1y1}1+x1k1y1m1}x1c1x1x2,dx2dt={αmax{0,1y2}1+x2k2y2m2}x1c2x1x2,ϵdy1dt=Ty1ϵmax{0,1y1}1+x1y1+ϵc1x2y1,ϵdy2dt=aTσy2ϵαmax{0,1y2}1+x2y2+ϵc2x1y2. (2)

    Let ϵ0 in (2), then y1=T, y2=aTσ. Substituting them into the first two equations in model (2) leads to the following quasi-steady system:

    dx1dt=x1[1T1+x1c1x2(k1T+m1)],dx2dt=x2[α(1pT)1+x2c2x1(k2pT+m2)]. (3)

    Throughout this paper, we consistently assume that

    m1<1, m2<α and T<min{1,1p},

    where p=aσ. Otherwise, the study would be meaningless. Under these assumptions, the authors of [15] studied the complete dynamics of model (3), revealing that competition outcomes can be influenced in various counterintuitive ways by both the level of toxins and the distinct vulnerabilities exhibited by the two species toward toxins.

    In practice, ecosystems typically exist within a stochastic environment. Intuitively, when subjected to relatively minor perturbations, random trajectories tend to exhibit small-amplitude oscillations around deterministic attractors; however, the presence of significant disturbances can potentially lead to the collapse of the entire system [16,17,18,19,20]. More importantly, intermediate noise intensity can give rise to certain counterintuitive phenomena that lack counterparts in the corresponding deterministic model, such as stochastic resonance [21], noise-induced transition [22], and noise-enhanced stability [23]. The underlying reason for this mainly lies in the multi-stability and high sensitivity of attractors. Research on the effect of noise on multi-stable system can be traced back to [24] and has been followed by many researchers recently. Bashkirtseva's team conducted a comprehensive investigation into noise-induced switching and transformation phenomena in multi-stable systems [25,26,27,28]. Yuan and his collaborators analyzed the stochastic sensitivity of competitive or predator-prey population models [29,30,31]. Spagnolo'team studied stochastic and coherence resonances in ecology and other fields [32,33,34,35].

    Motivated by the above facts, this paper aims to investigate how environmental noise affects the dynamics of toxin-dependent competition model (3). To achieve this objective, we initially utilize the approach proposed in [36,37] to develop a stochastic version of model (3) as follows:

    {dx1=x1f1(x1,x2)dt+σ1x1dB1(t),dx2=x2f2(x1,x2)dt+σ2x2dB2(t), (4)

    where

    f1(x1,x2)=1T1+x1c1x2(k1T+m1),f2(x1,x2)=α(1pT)1+x2c2x1(k2pT+m2),

    and where Bi(t) are mutually independent standard Brownian motions, and σi are the white noise intensities, i=1,2.

    The impact of noise on competition outcomes between two species will be investigated by conducting survival and sensitivity analyses for stochastic model (4). Specifically, a survival analysis will establish sufficient conditions for population extinction and persistence. Furthermore, we will prove the existence of a unique ergodic stationary distribution using Khasminskii's theory in Section 2. Next, we will study the spatial arrangement of random states near the deterministic attractor through applying the stochastic sensitivity function (SSF) technique in Section 3. Lastly, our study concludes with a concise discussion presented in Section 4. Our analysis results show that (ⅰ) large noise is harmful to two competing species and can lead to their extinction; (ⅱ) small noise does not alter the original competition outcomes observed in the model without random disturbance; and (ⅲ) intermediate noise can significantly influence competition outcomes in various ways. In other words, increasing noise intensity may have a positive impact on one species while negatively affecting another. Moreover, increasing noise intensity may enhance coexistence between two species and maintain species diversity by reducing the persistence of dominant species. These findings have significant implications for maintaining ecosystem diversity.

    The initial presentation of fundamental properties is essential for the investigation of the dynamics of model (4).

    Lemma 2.1. The given initial valve (x1(0),x2(0))R2+ ensures that model (4) possesses a unique solution (x1(t),x2(t)) for t0. Moreover, it can be guaranteed with probability one that the solution will always remain in R2+, i.e., (x1(t),x2(t))R2+ for all t0 almost surely.

    Proof. The coefficients of (4) satisfy the local Lipschitz condition but fail to meet the linear growth condition. Consequently, given an initial value (x1(0),x2(0))R2+, a unique local solution (x1(t),x2(t)) exists for t(0,τe], where τe is commonly referred to as the explosion time [38]. To prove τe= a.s., following the approach in [39, Theorem 3.1], it suffices to construct a nonnegative C2-function V1(x1,x2) that satisfies LV1K, where L is a differential operator and K is a constant. To this end, we define the function V1 as follows:

    V1(x1,x2)=x1n1n1lnx1n1+n2(x21lnx2),

    where n1=m2n2c1 and n2=m1c2. Applying Itô's formula, we obtain

    LV1=(x1n1)[1T1+x1c1x2(k1T+m1)]+12σ21+n2(x21)[α(1pT)1+x2c2x1(k2pT+m2)]+12n2σ221m1x1+c1n1x2+(k1T+m1)n1+12σ21+n2[αm2x2+c2x1+(k2pT+m2)]+12n2σ22=1+(k1T+m1)n1+αn2+(k2pT+m2)n2+12σ21+12n2σ22K.

    Thus, the desired outcome is achieved by employing a similar discourse as presented in [39].

    Lemma 2.2. (see [40]) Let φi(t) (i=1,2) be the solution of the stochastic equation

    dφi(t)=(bidiφi(t))dt+σiφi(t)dBi(t) (5)

    with φi(0)=xi(0)R+, where

    b1=1T, b2=α(1pT), d1=k1T+m1, d2=k2pT+m2.

    then, φi(t) satisfies that

    limt1tt0φi(s)ds=bidi,i=1,2.

    Remark 2.1. From model (4), we know that

    dxi(t)(bidixi(t))dt+σixi(t)dBi(t), i=1,2,

    which follows from the stochastic comparison theory that xi(t)φi(t) a.s.

    In this subsection, survival of the two species will be discussed, including extinction and persistence in mean. i.e., the following result is valid based on Lemmas 2.1 and 2.2.

    Theorem 2.1. For model (4), denote

    λ1=1T(k1T+m1+12σ21),λ2=α(1pT)(k2pT+m2+12σ22).

    (i) If maxi=1,2{λi}<0, both species 1 and species 2 go to extinction.

    (ii) If λi<0, λj>0, ij, i,j=1,2, species i is extinct and species j is persistent in mean.

    (iii) If mini=1,2{λiϑi}>0, both species 1 and species 2 are persistent in mean, where ϑ1=c1b2d2, ϑ2=c2b1d1.

    Remark 2.2. (i) Theorem 2.1 implies that sufficiently small noise will not change the survival of species, while sufficiently large noise will force the entire system to go extinct. Ecologically speaking, large noise is not conducive to maintaining biodiversity.

    (ii) Denote

    λd1=1T(k1T+m1), λd2=α(1pT)(k2pT+m2).

    Clearly, λdi<λi (i=1,2), implying the possibility that λdi<0<λi due to the continuity of λi in Ii. Biologically, this means that when species i is persistent in deterministic model (3) (see [15, Theorem 3.8]), it may be extinct with probability one in stochastic model (4) due to the effect of noise. In other words, noise may influence the competition outcomes of two species.

    Proof. By Itô's formula to (4), we have

    1tlnx1(t)x1(0)=(k1T+m1+12σ21)+1tt0(1T1+x1c1x2)ds+M1t, (6)
    1tlnx2(t)x2(0)=(k2pT+m2+12σ22)+1tt0(α(1pT)1+x2c2x1)ds+M2t, (7)

    where Mi=σiBi(t), i=1,2, and limtMit=0 a.s. by strong law of large numbers.

    (i) When λ1<0, Eq (6) implies that

    lim supt1tlnx1(t)x1(0)1T(k1T+m1+12σ21)<0  a.s.

    That is, species 1 converges to 0 at an exponential rate. Similarly, species 2 goes to extinction under the condition λ2<0.

    (ii) Without loss of generality, we consider the case where i=1 and j=2, then, species 1 is an obvious extinction by case (i), i.e., limtx1(t)=0 a.s. Hence, for arbitrary 0<ϵ1<1, there exists a set Ωϵ1Ω with P(Ωϵ1)1ϵ1 and a constant T1=T1(ϵ1) such that 1tt0c2x1ds<ϵ1, for ωΩϵ1 and t>T1. By Eq (7), one gets

    1tlnx2(t)x2(0)(k2pT+m2+12σ22+ϵ1)+1tt0α(1pT)1+x2ds+M2tα(1pT)(k2pT+m2+12σ22+ϵ1)1tt0α(1pT)x2ds+M2t.

    By [41, Lemma 4] and the arbitrariness of ϵ1, we obtain

    lim inft1tt0x2dsα(1pT)(k2pT+m2+12σ22)α(1pT)=λ2α(1pT)>0  a.s.,

    which means that species 2 is persistent in mean.

    (iii) By Lemma 2.2, one has limt1tt0φ2(s)ds=b2d2 a.s. In other words, for any arbitrary 0<ϵ2<1, there exists a set Ωϵ2Ω with P(Ωϵ2)1ϵ2 and a constant T2=T2(ϵ2) such that 1tt0φ2(s)ds<b2d2+ϵ2 holds for all ωΩϵ2 and t>T2. On the other hand, by Eq (6) and Remark 2.1,

    1tlnx1(t)x1(0)1T(k1T+m1+12σ21)1tt0(1T)x1ds1tt0c1x2ds+M1t1T(k1T+m1+12σ21)1tt0(1T)x1ds1tt0c1φ2ds+M1t1T(k1T+m1+12σ21)c1(b2d2+ϵ2)1tt0(1T)x1ds+M1t. (8)

    In conjunction with [41, Lemma 4] and the arbitrariness of ϵ2, this implies that

    lim inft1tt0x1dsλ1ϑ11T>0  a.s.

    Similarly, for species 2, we have lim inft1tt0x2dsλ2ϑ2α(1pT)>0  a.s. Consequently, both species 1 and species 2 are persistent in mean.

    In the following, the existence of ergodic stationary distribution will be investigated, which reflects the long-term stochastic dynamics of model (4). Let us first denote

    ν1=(b2+c1)b2d2, ν2=(b1+c2)b1d1.

    Using [42, Theorem 4.1], one can prove the following theorem.

    Theorem 2.2. Assuming that mini=1,2{λiνi}>0, a unique and ergodic stationary distribution exists for model (4) given any initial value (x1(0),x2(0))R2+.

    Proof. To complete the proof of Theorem 2.2, based on [42, Theorem 4.1], we need to construct a nonnegative C2-function V2 and a bounded closed Uϱ such that LV2C for (x1,x2)R2+Uϱ, where C is a positive constant.

    Let us first define a C2-function ¯V2 as

    ¯V2(x1,x2)=lnx1lnx2+b1+c2d1x1+b2+c1d2x2,

    where bi and di (i=1,2) are defined as in Lemma 2.2. Obviously, ¯V2(x1,x2) is continuous and

    lim infR,(x1,x2)R2+UR¯V2(x1,x2)=+,

    where UR=(1R,R)×(1R,R), which implies that ¯V2(x1,x2) achieves a lower bound at a point (ˉx1,ˉx2) in the interior of R2+. Define the nonnegative C2-function V2:R2+R+ as

    V2(x1,x2)=¯V2(x1,x2)¯V2(ˉx1,ˉx2).

    By Itô's formula, one has

    LV2=1T1+x1+k1T+m1+c1x2+12σ21α(1pT)1+x2+k2PT+m2+c2x1+12σ22+(b1+c2)x1d1[1T1+x1(k1T+m1)c1x2]+(b2+c1)x2d2[α(1pT)1+x2(k2PT+m2)c2x1](1T)+(1T)x1+k1T+m1+c1x2+12σ21α(1pT)+α(1pT)x2+k2PT+m2+c2x1+12σ22+(b1+c2)x1d1[1Tx1(k1T+m1)c1x2]+(b2+c1)x2d2[α(1pT)x2(k2PT+m2)c2x1]=[λ1(b2+c1)b2d2][λ2(b1+c2)b1d1][(b1+c2)c1d1+(b2+c1)c2d2]x1x2Θ(x1,x2).

    Given that mini=1,2{λiνi}>0, then

    limx∣→0Θ(x1,x2)=[λ1(b2+c1)b2d2][λ2(b1+c2)b1d1]<0,

    and limx∣→Θ(x1,x2)=, where x∣=x21+x22. Hence, there exist a positive constant C and a sufficiently small ϱ such that

    LV2(x1,x2)C  for any (x1,x2)R2+Uϱ,

    where Uϱ=[ϱ,1ϱ]×[ϱ,1ϱ]. In other words, for any initial value (x1(0),x2(0))R2+, a unique stationary distribution exists for model (4), which is also ergodic.

    The survival analysis conducted in Section 2 can characterize to some extent how noise affects the dynamics of species. However, these results are crude and do not well reflect the whole effects of noise on species. In this section, some stochastic sensitivity analysis of model (4) is presented, which can fill the gap well.

    As a basis for further discussion, we first consider the case when only one species without interspecific competition presents in a perturbed environment with toxin. Without loss of generality, suppose that only species x1 is present, then model (4) without interspecific competition is governed by

    dx1=x1f1(x1,0)dt+σx1dB1(t), (9)

    where σ=σ1. When σ=0, model (4) exists two equilibria: x01=0 (unstable) and x1=1m1(1+k1)Tm1+k1T (stable), then the formula

    μ=x1(1+x1)22(1T)

    reflects the sensitivity of x1 (see Appendix), as depicted in Figure 1(a). We can see that the sensitivity of x1 decreases as the toxin level T increases, suggesting that a higher concentration of toxins in the environment always has a detrimental impact on species survival.

    Figure 1.  (a) Stochastic sensitivity of equilibrium x1; (b) Critical noise intensity σ. Here, parameters k1=1 and m1=0.49.

    Following 3σ-rule (the fiducial probability P=0.997), the corresponding confidence interval can be expressed as (x1r,x1+r) with r=3σμ. Furthermore, the critical value of noise intensity can be obtained by x1r=0,

    σ=2(1T)x13(1+x1). (10)

    As depicted in Figure 1(b), the critical value of noise intensity σ exhibits a decreasing trend with increasing toxin level T. Meanwhile, this curve σ(T) divides the parameter region into two parts: persistence (below) and extinction (above).

    In Figure 2, the time series of model (9) (blue) and the boundaries of confidence intervals (red dashed line) are plotted for different noise intensity values. Obviously, as noise intensity increases, the confidence interval expands and the distribution of random states is becoming more and more dispersed. Note that here, σ=0.145. Figure 2 further shows that when σ<0.145, the dispersion of random states is located in the interior of the corresponding confidence interval with probability 0.997, which means that the species is persistent; while for σ>0.145, the left boundary of confidence interval is less than zero, suggesting that the survival of species has transformed from persistence to extinction.

    Figure 2.  Stochastic trajectories of model (9) with initial value x1(0)=0.155, parameters k1=1, m1=0.49, and T=0.2 for (a) σ=0.05, (b) σ=0.1, (c) σ=0.15.

    Taking the parameters of model (3) as in Figure 3, it follows from [15] that deterministic model (3) has two types of bistability. As shown in Figure 3(a), exclusion equilibrium E01 and coexistence equilibrium El are asymptotically stable simultaneously. That is, two basins of attraction are separated by the stable manifold (black dashed-dotted) of saddle point Er. Obviously, the trajectory with initial value below the separatrix approaches E01, whereas the trajectory with initial value above the separatrix converges to El. Figure 3(b) shows another type of bistability, that is, both exclusion equilibria E01 and E02 are asymptotically stable simultaneously. In this situation, two basins of attraction are separated by the stable manifold of saddle point Ec.

    Figure 3.  Vector field of deterministic model (3) and separatrix of two attraction domains. Filled (open) circles denote stable (unstable). The black dashed-dotted line denotes the separatrix of two attraction domains. Here, (a) α=0.8, p=1.09, k1=1, k2=1, m1=0.49, m2=0.4, T=0.05, c1=0.6, c2=0.4; (b) α=0.8, p=2.2, k1=1, k2=1, m1=0.49, m2=0.4, T=0.01, c1=0.6, c2=0.4.

    We now focus on the sensitivity analysis of the stable equilibria E01 and El depicted in Figure 3(a). However, for the exclusion equilibria E01 and E02 shown in Figure 3(b), a similar analysis can be conducted but is omitted here. The equilibrium El can be analyzed in two cases.

    (i) σ1=σ2=σ. As depicted in Figure 4, the stochastic trajectory of model (4) with an initial value above the separatrix deviates from the deterministic attractor El, subsequently giving rise to a corresponding stochastic attractor. In the presence of small noise, the states of this stochastic attractor tend to concentrate near El; whereas for relatively large noise levels, qualitative changes in dynamics are observed in model (4). More precisely, the stochastic trajectory escapes from the attraction domain of El, subsequently traverses the separatrix, and ultimately enters the attraction domain of E01.

    Figure 4.  Phase trajectories for model (4) with initial value (x1(0),x2(0))=(0.41,0.23). Here, (a) σ=0.006 (small); (b) σ=0.12 (large), and the other parameters are the same as those in Figure 3(a).

    The above detailed qualitative analysis can be illustrated clearly by use of SSF technique (see Appendix). To this end, denote El=(x1,x2) and

    F=[f11f12f21f22],  G=[g1100g22],

    where

    f11=x1fx1(x1,x2), f12=x1fx2(x1,x2), f21=x2gx1(x1,x2), f22=x2gx2(x1,x2)

    and

    g11=(x1)2, g22=(x2)2.

    The stochastic sensitivity matrix

    W=[w11w12w21w22],

    can be derived from (A.2) in Appendix A, which satisfies the equation

    {2f11w11+f12w12+f12w21=g211,f21w11+(f11+f22)w12+f12w22=0,f21w11+(f11+f22)w21+f12w22=0,f21w12+f21w21+2f22ww22=g222.

    Here, w11 and w22, respectively, describe the sensitivity of E along the x1-axis and x2-axis, and w12=w21 denotes the covariation. The graphs of functions w11(T) and w22(T) are shown in Figure 5. Clearly, w11(T)>w22(T) for each T[0.05,0.07], which shows that the sensitivity of species 1 is large than species 2. Most importantly, the two functions tend to infinity as T approaches 0.05. This signalizes the possible about noise-induced transitions, which can be visualized by use of confidence ellipses.

    Figure 5.  The sensitivity of equilibrium E. Here, the values of all parameters except for T are identical to those in Figure 3(a).

    The confidence ellipse is a geometrical model describing the configurational arrangement of random states near stable equilibrium. Next, we will illustrate how confidence ellipse can be used to explain the noise-induced transition from persistence to extinction. By Eq (A.4) in Appendix, the corresponding confidence ellipse equation of El can be expressed by

    (x1x1,x2x2)T,W1(x1x1,x2x2)T=2σ2K2, (11)

    where K2=ln(1P), and P is a fiducial probability. In Figure 6, the random states (blue) of model (4) for σ=0.006 are plotted, along with the confidence ellipse (green) representing a fiducial probability P=0.95. It can be observed that the random states are distributed around El and fall within the interior of confidence ellipse with a probability P=0.95. This observation, in conjunction with Figure 5, further implies that the dispersion of the random states near El are influenced by both noise intensity and sensitivity.

    Figure 6.  Random states (blue) of stochastic model (4) and confidence ellipse (green) are shown for σ=0.006, with the other parameters being the same as those in Figure 3(a).

    The confidence ellipses (solid) in Figure 7 represent the fixed fiducial probability P=0.95 for three different noise intensity values: σ=0.004 (small), σ=0.0069 (middle), and σ=0.01 (large). It can be observed that when the noise intensity is small, the entire confidence ellipse falls within the attraction domain of El. Consequently, in this scenario, the corresponding random orbit is concentrated near El as depicted in Figure 4(a). As we increase the noise intensity to a critical value, the confidence ellipse starts expanding and eventually crosses over to occupy the attraction domain of E01 after crossing the separatrix (dashed-dotted line). In this case, with a high probability, the corresponding random orbit may leave the attraction domain of El and form a stochastic attractor near E01 as shown in Figure 4(b). The intersection point between the confidence ellipse and separatrix represents an approximate value for this critical intensity denoted by σ. Notably, here we have approximately estimated that σ0.0069. By comparing Figures 7 with 4, it can be concluded that these quantitative analysis results are consistent with direct numerical simulations conducted earlier on this topic. Biologically speaking, noise is not conducive to the coexistence of two species; however, it does confer benefits on the survival of species 1.

    Figure 7.  Separatix (dashed-dotted) and confidence ellipses (solid) are shown for σ=0.004 (small), σ=0.0069 (middle), and σ=0.01 (large), with the other parameters being the same as those in Figure 3(a).

    (ii) σ1σ2. Noise-induced transitions under σ1σ2 can be analyzed by means of numerical simulations. In Figure 8, phase trajectories of model (4) with initial value near El are plotted for two sets of different noise intensity:

    (a) σ1=0.18, σ2=0.006; (b) σ1=0.7, σ2=0.9.
    Figure 8.  Phase trajectories for model (4) with initial value (x1(0),x2(0))=(0.41,0.23). Here, (a) σ1=0.18, σ2=0.004; (b) σ1=0.7, σ2=0.9, and the other parameters are the same as those in Figure 3(a).

    Comparing Figures 8(a) with 4(a), we can see that increasing the noise experienced by species 1 but still at a low level can force the state of the system to shift from near El to near E02. This means that noise is not conducive to the coexistence of two species, but species 2 actually benefits from large noise. Similarly, it follows from comparison of Figures 8(b) with 4(a) that further increasing the noise experienced by two species will destroy their coexistence and make both species extinct simultaneously. It seems that the large noise experienced by two species is a disaster for the persistence of the competition system. The appropriate noise intensity can also force the system state to shift from near El to near Er, but its corresponding phase trajectory is not drawn due to the similarity to Figure 8.

    In the remaining space of this subsection, we discuss the species 1-only equilibrium E01. It is noted that the stochastic sensitivity of E01 cannot be analyzed by constructing the corresponding confidence ellipses as in case (ⅰ). Therefore, here we numerically explore the noise-induced phenomenon. Figure 9 shows all possible phase portraits of model (4) with initial value near E01 for different noise intensities except the following case: the noise-induced transition from near E01 to near Er. It can be seen from Figure 9 that the system state can switch from near E01 to near any other equilibrium. It is noteworthy that these state transitions are only related to noise intensity, but not to the stability of the equilibrium, which is different from case (ⅰ). It can be seen from these numerical simulations that intermediate noise intensity can promote the coexistence of two competing species by reducing the persistence level of dominant species.

    Figure 9.  Phase trajectories for model (4) with initial value (x1(0),x2(0))=(0.77,0.01). Here, (a) σ1=σ2=0.01; (b) σ1=0.08, σ2=0.01; (c) σ1=0.19, σ2=0.01; (d) σ1=1, σ2=0.76, and the other parameters are the same as those in Figure 3(a).

    To predict the impact of toxic pollutants on aquatic ecosystems, ecologists and mathematicians have recently established a large number of biological mathematical models from different perspectives [8,9,13,14,15]. These models explicitly incorporate the effects of toxic substances on the survival of aquatic species while not accounting for stochastic fluctuations in the external environment. The ubiquity of environmental noises in aquatic ecosystems arises from the inherent unpredictability of weather, temperature, and various other physical factors that are intricately intertwined within these ecosystems. In this study, we proposed a stochastic toxin-dependent competition model (4) based on the findings of [15]. We initially conducted survival analysis and successfully proved the existence of a unique ergodic stationary distribution. From Theorems 2.1 and 2.2, the relationships between the noise intensity and the survival of two species are summarized below:

    ● When σ2i>ηi, i=1,2, two species are extinct simultaneously (see Figures 8(b) and 9(d));

    ● When σ2i>ηi, σ2j<ηj, ij, i,j=1,2, species i is extinct and species j is persistent in mean (see Figures 4(b), 8(a), 9(a), and 9(c));

    ● When σ2i<ηi2ϑi, i=1,2, two species are persistent in mean simultaneously (see Figures 4(a) and 9(b));

    ● When σ2i<ηi2νi, i=1,2, model (4) exists a unique ergodic stationary distribution,

    where

    η1=2(1T(k1T+m1)), η2=2(α(1pT)(k2pT+m2)).

    The foregoing theoretical analysis shows the impact of noise on two species and estimates the corresponding noise intensity. Due to the limitations of Lyapunov's method, we do not know what happens when ηi2ϑi<σ2i<ηi, i=1,2. For this reason, we further discussed the stochastic sensitivity of model (4) with the help of the SSF technique and numerical method.

    For the single species model (9), we constructed the confidence interval and estimated the critical noise intensity σ. As shown in Figure 2, the critical noise intensity determines the survival of species with probability P=0.997. For the two species competition model (4), the noise-induced phenomena were studied in detail by different ways. When σ1=σ2, we constructed the confidence ellipse by use of the SSF technique and plotted the sensitivity of attractor to toxin concentration based on stochastic sensitivity matrix, as shown in Figures 57. It allows us to see clearly the configurational arrangement of random states near stable equilibrium. Meanwhile, the critical noise intensity was estimated numerically; see Figure 7. Information about the critical noise intensity enables us to predict the conditions of the noise-induced transition. In this case, large noise but still at a low level can force the system state to transform from the coexistence to the species 1-only. When σ1σ2, we numerically explored the noise-induced phenomena; see Figures 8 and 9. In this case, the state transitions are only related to the noise intensities σ1 and σ2, but not to the stability of equilibrium, which is different from the case σ1=σ2.

    After theoretical analysis and numerical simulations, it is suggested that noise can significantly influence competition outcomes between two species. Specifically, (ⅰ) high levels of noise are detrimental to the competing species and can lead to their extinction, (ⅱ) low levels of noise do not alter the original competition outcomes in the deterministic model; in this scenario, species survival depends on its initial population size, and (ⅲ) intermediate levels of noise can induce various changes in competition outcomes. Increasing noise intensity may have a positive impact on one species while negatively affecting another due to reduced persistence level and increased resource availability for its competitor. Moreover, higher noise intensity may promote coexistence and maintain species diversity by reducing the persistence level of dominant species. In summary, increasing noise intensity has the potential to reverse competition outcomes between two species. These findings are intriguing and hold significant theoretical implications for ecological resource management.

    The model proposed in this paper assumes that environmental noise follows a Gaussian white noise distribution and is directly proportional to the variables. However, in reality, the natural growth of species often encounters abrupt environmental disturbances such as harvesting and earthquakes. Additionally, the growth rates of certain populations exhibit significant variations between summer and winter seasons. These phenomena cannot be adequately captured by white noise alone. Some scholars have proposed alternative types of environmental noise, such as Lévy noise [43] and telephone noise [44], to better describe these complex dynamics. We acknowledge the importance of investigating these alternatives in future research.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The work was supported by the National Natural Science Foundation of China (No. 12171441), the Scientific Research Team Plan of Zhengzhou University of Aeronautics (No. 23ZHTD01013), the Key Scientific Research Project of Colleges and Universities of Henan Province (No. 24A110012), and the Basic Research Projects of Key Scientific Research Projects Plan in Henan Higher Education Institutions (No. 24ZX008).

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Let B(t)=(B1(t),,Bl(t))T, t0 be an l-dimensional Brownian motion defined on the complete probability space (Ω,F,P) adapted to the filtration {Ft}t0.

    Consider the following stochastic system:

    dx=f(x)dt+σg(x)dB(t), (A.1)

    where x is an n-dimensional vector, f(x) is an n-dimensional vector function, g(x) is an n×l-matrix function, and σ is a nonnegative scalar parameter that denotes the noise intensity. Suppose that the corresponding deterministic system of (A.1) has an asymptotically stable equilibrium x. The stochastic trajectories of (A.1) with initial values near x may leave x and form a corresponding stochastic attractor with probabilistic density function ρ(x,σ), which is governed by the corresponding Kolmogorov-Fokker-Planck (K-P) equation. For one-dimensional model, ρ(x,σ) can be solved directly by the K-P equation, but for the multidimensional model, it is difficult to solve it. In this case, it follows from Refs. [45,46] that we can write an approximation of this probability distribution as follows:

    ρ(x,σ)1(2πσ2)ndetWexp(<xx,W1(xx)>2σ2),

    where <,> represents Euclidean scalar product, matrix W denotes the stochastic sensitivity function of equilibrium x, σ2W is the covariance matrix, and W is a unique solution of the following equation:

    FW+WFT=S, (A.2)

    where F=fx(x) and S=g(x)g(x)T. With the help of matrix W, the corresponding confidence ellipsoid can be constructed as follows (see [47]):

    <xx,W1(xx)>=σ2K(P), (A.3)

    where P is a fiducial probability. The function K(P) is an inverse function of P(K), where

    P(K)=Ψn(K)Ψn(), Ψn(K)=K0exp(t22)tn1dt.

    For the case n=1, P(K)=erf(K2), erf(x)=2πx0exp(t2)dt, then the corresponding confidence interval is (xr,x+r) with r=σ2μerf1(P), and the stochastic sensitivity μ can be determined by μ=g2(x)2f(x). It follows from 3σ-rule [48] that r=3σμ.

    For the case n=2, P(K)=1exp(K2), K(P)=2ln(1P), then the corresponding confidence ellipsoid is

    <xx,W1(xx)>=2σ2K2, (A.4)

    where K2=ln11P. Similarly, for n=3, P(K)=erf(K2)2KπeK2; and for n=4, P(K)=1exp(K2)(1+12K).



    [1] C. H. Walker, R. M. Sibly, S. P. Hopkin, D. B. Peakall, Principles of ecotoxicology, Boca Raton: CRC Press, 2012. https://doi.org/10.1201/b11767
    [2] H. I. Freedman, J. B. Shukla, Models for the effect of toxicant in single-species and predator-prey systems, J. Math. Biol., 30 (1991), 15–30. https://doi.org/10.1007/BF00168004 doi: 10.1007/BF00168004
    [3] T. G. Hallam, C. E. Clark, R. R. Lassiter, Effect of toxicants on populations: a qualitative approach Ⅰ. Equilibrium environmental exposure, Ecol. Model., 18 (1983), 291–304. https://doi.org/10.1016/0304-3800(83)90019-4 doi: 10.1016/0304-3800(83)90019-4
    [4] M. Liu, K. Wang, Persistence and extinction of a stochastic single-specie model under regime switching in a polluted environment Ⅱ, J. Theor. Biol., 267 (2010), 283–291. https://doi.org/10.1016/j.jtbi.2010.08.030 doi: 10.1016/j.jtbi.2010.08.030
    [5] D. M. Thomas, T. W. Snell, S. M. Jaffar, A control problem in a polluted environment, Math. Biosci., 133 (1996), 139–163. https://doi.org/10.1016/0025-5564(95)00091-7 doi: 10.1016/0025-5564(95)00091-7
    [6] H. R. Thieme, Mathematics in population biology, Princeton University Press, 2003.
    [7] Q. Huang, L. Parshotam, H. Wang, C. Bampfylde, M. A. Lewis, A model for the impact of contaminants on fish population dynamics, J. Theor. Biol., 334 (2013), 71–79. https://doi.org/10.1016/j.jtbi.2013.05.018 doi: 10.1016/j.jtbi.2013.05.018
    [8] W. Wang, Biodynamic understanding of mercury accumulation in marine and freshwater fish, Adv. Environ. Res., 1 (2012), 15–35. https://doi.org/10.12989/aer.2012.1.1.015 doi: 10.12989/aer.2012.1.1.015
    [9] W. Wang, P. S. Rainbow, Comparative approaches to understand metal bioaccumulation in aquatic animals, Comp. Biochem. Phys. C, 148 (2008), 315–323. https://doi.org/10.1016/j.cbpc.2008.04.003 doi: 10.1016/j.cbpc.2008.04.003
    [10] Y. Zhao, S. Yuan, J. Ma, Survival and stationary distribution analysis of a stochastic competitive model of three species in a polluted environment, Bull. Math. Biol., 77 (2015), 1285–1326. https://doi.org/10.1007/s11538-015-0086-4 doi: 10.1007/s11538-015-0086-4
    [11] P. Zhou, Q. Huang, A spatiotemporal model for the effects of toxicants on populations in a polluted river, SIAM J. Appl. Math., 82 (2022), 95–118. https://doi.org/10.1137/21M1405629 doi: 10.1137/21M1405629
    [12] A. Q. Khan, S. S. Kazmi, T. D. Alharbi, Bifurcations of a three-species prey-predator system with scavenger, Ain Shams. Eng. J., 14 (2023), 102514. https://doi.org/10.1016/j.asej.2023.102514 doi: 10.1016/j.asej.2023.102514
    [13] Q. Huang, H. Wang, M. A. Lewis, The impact of environmental toxins on predator-prey dynamics, J. Theor. Biol., 378 (2015), 12–30. https://doi.org/10.1016/j.jtbi.2015.04.019 doi: 10.1016/j.jtbi.2015.04.019
    [14] A. Peace, M. D. Poteat, H. Wang, Somatic growth dilution of a toxicant in a predator-prey model under stoichiometric constraints, J. Theor. Biol., 407 (2016), 198–211. https://doi.org/10.1016/j.jtbi.2016.07.036 doi: 10.1016/j.jtbi.2016.07.036
    [15] C. Shan, Q. Huang, Direct and indirect effects of toxins on competition dynamics of species in an aquatic environment, J. Math. Biol., 78 (2019), 739–766. https://doi.org/10.1007/s00285-018-1290-2 doi: 10.1007/s00285-018-1290-2
    [16] D. Li, S. Liu, J. Cui, Threshold dynamics and ergodicity of an SIRS epidemic model with Markovian switching, J. Differ. Equations, 263 (2017), 8873–8915. https://doi.org/10.1016/j.jde.2017.08.066 doi: 10.1016/j.jde.2017.08.066
    [17] X. Chen, X. Li, Y. Ma, C. Yuan, The threshold of stochastic tumor-immune model with regime switching, J. Math. Anal. Appl., 522 (2023), 126956. https://doi.org/10.1016/j.jmaa.2022.126956 doi: 10.1016/j.jmaa.2022.126956
    [18] X. Zhang, Q. Yang, D. Jiang, A stochastic predator-prey model with Ornstein-Uhlenbeck process: characterization of stationary distribution, extinction and probability density function, Commun. Nonlinear. Sci., 122 (2023), 107259. https://doi.org/10.1016/j.cnsns.2023.107259 doi: 10.1016/j.cnsns.2023.107259
    [19] Q. Liu, D. Jiang, Analysis of a stochastic inshore-offshore hairtail fishery model with Ornstein-Uhlenbeck process, Chaos Soliton. Fract., 172 (2023), 113525. https://doi.org/10.1016/j.chaos.2023.113525 doi: 10.1016/j.chaos.2023.113525
    [20] M. N. Srinivas, K. S. Reddy, A. Sabarmathi, Optimal harvesting strategy and stochastic analysis for a two species commensaling system, Ain Shams. Eng. J., 5 (2014), 515–523. https://doi.org/10.1016/j.asej.2013.10.003 doi: 10.1016/j.asej.2013.10.003
    [21] R. Benzi, A. Sutera, A. Vulpiani, The mechanism of stochastic resonance, J. Phys. A. Math. Gen., 14 (1981), 4531981. https://doi.org/10.1088/0305-4470/14/11/006 doi: 10.1088/0305-4470/14/11/006
    [22] D. O. Filatov, D. V. Vrzheshch, O. V. Tabakov, A. S. Novikov, A. I. Belov, I. N. Antonov, et al., Noise-induced resistive switching in a memristor based on ZrO2(Y)/Ta2O5 stack, J. Stat. Mech., 2019 (2019), 124026. https://doi.org/10.1088/1742-5468/ab5704 doi: 10.1088/1742-5468/ab5704
    [23] N. V. Agudov, A. V. Safonov, A. V. Krichigin, A. A. Kharcheva, A. A. Dubkov, D. Valenti, et al., Nonstationary distributions and relaxation times in a stochastic model of memristor, J. Stat. Mech., 2020 (2020), 024003. https://doi.org/10.1088/1742-5468/ab684a doi: 10.1088/1742-5468/ab684a
    [24] L. S. Pontryagin, A. A. Andronov, A. A. Witt, On statistical analysis of dynamical systems, Zh. Eksp. Teor. Fiz., 3 (1933), 165.
    [25] I. Bashkirtseva, L. Ryashko, Constructive analysis of noise-induced transitions for coexisting periodic attractors of the Lorenz model, Phys. Rev. E, 79 (2009), 041106. https://doi.org/10.1103/PhysRevE.79.041106 doi: 10.1103/PhysRevE.79.041106
    [26] I. Bashkirtseva, L. Ryashko, T. Ryazanova, Stochastic sensitivity technique in a persistence analysis of randomly forced population systems with multiple trophic levels, Math. Biosci., 293 (2017), 38–45. https://doi.org/10.1016/j.mbs.2017.08.007 doi: 10.1016/j.mbs.2017.08.007
    [27] L. Ryashko, T. Perevalova, I. Bashkirtseva, Stochastic bifurcations and multistage order-chaos transitions in a 4D eco-epidemiological model, Int. J. Bifurcat. Chaos, 33 (2023), 2350112. https://doi.org/10.1142/S0218127423501122 doi: 10.1142/S0218127423501122
    [28] I. Bashkirtseva, L. Ryashko, How noise induces multi-stage transformations of oscillatory regimes in a thermochemical model, Phys. Lett. A, 476 (2023), 128884. https://doi.org/10.1016/j.physleta.2023.128884 doi: 10.1016/j.physleta.2023.128884
    [29] C. Xu, S. Yuan, T. Zhang, Stochastic sensitivity analysis for a competitive turbidostat model with inhibitory nutrients, Int. J. Bifurcat. Chaos, 26 (2016), 1650173. https://doi.org/10.1142/S021812741650173X doi: 10.1142/S021812741650173X
    [30] D. Wu, H. Wang, S. Yuan, Stochastic sensitivity analysis of noise-induced transitions in a predator-prey model with environmental toxins, Math. Biosci. Eng., 16 (2019), 2141–2153. https://doi.org/10.3934/mbe.2019104 doi: 10.3934/mbe.2019104
    [31] S. Yuan, D. Wu, G. Lan, H. Wang, Noise-induced transitions in a nonsmooth producer-grazer model with stoichiometric constraints, Bull. Math. Biol., 82 (2020), 55. https://doi.org/10.1007/s11538-020-00733-y doi: 10.1007/s11538-020-00733-y
    [32] D. Valenti, G. Fazio, B. Spagnolo, Stabilizing effect of volatility in financial markets, Phys. Rev. E, 97 (2018), 062307. https://doi.org/10.1103/PhysRevE.97.062307 doi: 10.1103/PhysRevE.97.062307
    [33] I. A. Surazhevsky, V. A. Demin, A. I. Ilyasov, A. V. Emelyanov, K. E. Nikiruy, V. V. Rylkov, et al., Noise-assisted persistence and recovery of memory state in a memristive spiking neuromorphic network, Chaos Soliton. Fract., 146 (2021), 110890. https://doi.org/10.1016/j.chaos.2021.110890 doi: 10.1016/j.chaos.2021.110890
    [34] A. V. Yakimov, D. O. Filatov, O. N. Gorshkov, D. A. Antonov, D. A. Liskin, I. N. Antonov, et al., Measurement of the activation energies of oxygen ion diffusion in yttria stabilized zirconia by flicker noise spectroscopy, Appl. Phys. Lett., 114 (2019), 253506. https://doi.org/10.1063/1.5098066 doi: 10.1063/1.5098066
    [35] B. Spagnolo, C. Guarcello, L. Magazzˊu, A. Carollo, D. P. Adorno, D. Valenti, Nonlinear relaxation phenomena in metastable condensed matter systems, Entropy, 19 (2017), 20. https://doi.org/10.3390/e19010020 doi: 10.3390/e19010020
    [36] C. Xu, S. Yuan, Competition in the chemostat: a stochastic multi-species model and its asymptotic behavior, Math. Biosci., 280 (2016), 1–9. https://doi.org/10.1016/j.mbs.2016.07.008 doi: 10.1016/j.mbs.2016.07.008
    [37] X. Yu, S. Yuan, T. Zhang, The effects of toxin-producing phytoplankton and environmental fluctuations on the planktonic blooms, Nonlinear Dyn., 91 (2018), 1653–1668. https://doi.org/10.1007/s11071-017-3971-6 doi: 10.1007/s11071-017-3971-6
    [38] X. Mao, Stochastic differential equations and spplications, Chichester: Horwood Publishing Limited, 1997.
    [39] A. Gray, D. Greenhalgh, L. Hu, X. Mao, J. Pan, A stochastic differential equation SIS epidemic model, SIAM J. Appl. Math., 71 (2011), 876–902. https://doi.org/10.1137/10081856X doi: 10.1137/10081856X
    [40] Q. Yang, D. Jiang, N. Shi, C. Ji, The ergodicity and extinction of stochastically perturbed SIR and SEIR epidemic models with saturated incidence, J. Math. Anal. Appl., 388 (2012), 248–271. https://doi.org/10.1016/j.jmaa.2011.11.072 doi: 10.1016/j.jmaa.2011.11.072
    [41] M. Liu, C. Bai, Analysis of a stochastic tri-trophic food-chain model with harvesting, J. Math. Biol., 73 (2016), 597–625. https://doi.org/10.1007/s00285-016-0970-z doi: 10.1007/s00285-016-0970-z
    [42] R. Khasminskii, Stochastic stability of differential equations, Berlin: Springer, 2012. https://doi.org/10.1007/978-3-642-23280-0
    [43] Y. Zhao, S. Yuan, Q. Zhang, The effect of Lévy noise on the survival of a stochastic competitive model in an impulsive polluted environment, Appl. Math. Model., 40 (2016), 7583–7600. https://doi.org/10.1016/j.apm.2016.01.056 doi: 10.1016/j.apm.2016.01.056
    [44] S. Zhang, X. Meng, T. Feng, T. Zhang, Dynamics analysis and numerical simulations of a stochastic non-autonomous predator-prey system with impulsive effects, Nonlinear Anal.-Hybri., 26 (2017), 19–37. https://doi.org/10.1016/j.nahs.2017.04.003 doi: 10.1016/j.nahs.2017.04.003
    [45] A. Dembo, O. Zeitouni, Large deviations techniques and applications, Boston: Jones and Bartlett Publishers, 1995.
    [46] M. I. Freidlin, A. D. Wentzell, Random perturbations of dynamical systems, New York: Springer, 1984. https://doi.org/10.1007/978-1-4684-0176-9
    [47] R. C. Smith, P. Cheeseman, On the representation and estimation of spatial uncertainty, Int. J. Rob. Res., 5 (1986), 56–68. https://doi.org/10.1177/027836498600500404 doi: 10.1177/027836498600500404
    [48] A. Hastings, T. Powell, Chaos in a three-species food chain, Ecology, 72 (1991), 896–903. https://doi.org/10.2307/1940591 doi: 10.2307/1940591
  • Reader Comments
  • © 2024 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(888) PDF downloads(61) Cited by(0)

Figures and Tables

Figures(9)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog