Research article

Threshold dynamics of a stochastic SIHR epidemic model of COVID-19 with general population-size dependent contact rate


  • Received: 23 December 2021 Revised: 25 January 2022 Accepted: 03 February 2022 Published: 21 February 2022
  • In this paper, we propose a stochastic SIHR epidemic model of COVID-19. A basic reproduction number Rs0 is defined to determine the extinction or persistence of the disease. If Rs0<1, the disease will be extinct. If Rs0>1, the disease will be strongly stochastically permanent. Based on realistic parameters of COVID-19, we numerically analyze the effect of key parameters such as transmission rate, confirmation rate and noise intensity on the dynamics of disease transmission and obtain sensitivity indices of some parameters on Rs0 by sensitivity analysis. It is found that: 1) The threshold level of deterministic model is overestimated in case of neglecting the effect of environmental noise; 2) The decrease of transmission rate and the increase of confirmed rate are beneficial to control the spread of COVID-19. Moreover, our sensitivity analysis indicates that the parameters β, σ and δ have significantly effects on Rs0.

    Citation: Tianfang Hou, Guijie Lan, Sanling Yuan, Tonghua Zhang. Threshold dynamics of a stochastic SIHR epidemic model of COVID-19 with general population-size dependent contact rate[J]. Mathematical Biosciences and Engineering, 2022, 19(4): 4217-4236. doi: 10.3934/mbe.2022195

    Related Papers:

    [1] Damilola Olabode, Jordan Culp, Allison Fisher, Angela Tower, Dylan Hull-Nye, Xueying Wang . Deterministic and stochastic models for the epidemic dynamics of COVID-19 in Wuhan, China. Mathematical Biosciences and Engineering, 2021, 18(1): 950-967. doi: 10.3934/mbe.2021050
    [2] Xinghua Chang, Maoxing Liu, Zhen Jin, Jianrong Wang . Studying on the impact of media coverage on the spread of COVID-19 in Hubei Province, China. Mathematical Biosciences and Engineering, 2020, 17(4): 3147-3159. doi: 10.3934/mbe.2020178
    [3] Xinyu Bai, Shaojuan Ma . Stochastic dynamical behavior of COVID-19 model based on secondary vaccination. Mathematical Biosciences and Engineering, 2023, 20(2): 2980-2997. doi: 10.3934/mbe.2023141
    [4] Jiying Ma, Wei Lin . Dynamics of a stochastic COVID-19 epidemic model considering asymptomatic and isolated infected individuals. Mathematical Biosciences and Engineering, 2022, 19(5): 5169-5189. doi: 10.3934/mbe.2022242
    [5] Fen-fen Zhang, Zhen Jin . Effect of travel restrictions, contact tracing and vaccination on control of emerging infectious diseases: transmission of COVID-19 as a case study. Mathematical Biosciences and Engineering, 2022, 19(3): 3177-3201. doi: 10.3934/mbe.2022147
    [6] Tingting Xue, Long Zhang, Xiaolin Fan . Dynamic modeling and analysis of Hepatitis B epidemic with general incidence. Mathematical Biosciences and Engineering, 2023, 20(6): 10883-10908. doi: 10.3934/mbe.2023483
    [7] Sarita Bugalia, Jai Prakash Tripathi, Hao Wang . Estimating the time-dependent effective reproduction number and vaccination rate for COVID-19 in the USA and India. Mathematical Biosciences and Engineering, 2023, 20(3): 4673-4689. doi: 10.3934/mbe.2023216
    [8] Hamdy M. Youssef, Najat A. Alghamdi, Magdy A. Ezzat, Alaa A. El-Bary, Ahmed M. Shawky . A new dynamical modeling SEIR with global analysis applied to the real data of spreading COVID-19 in Saudi Arabia. Mathematical Biosciences and Engineering, 2020, 17(6): 7018-7044. doi: 10.3934/mbe.2020362
    [9] Pannathon Kreabkhontho, Watchara Teparos, Thitiya Theparod . Potential for eliminating COVID-19 in Thailand through third-dose vaccination: A modeling approach. Mathematical Biosciences and Engineering, 2024, 21(8): 6807-6828. doi: 10.3934/mbe.2024298
    [10] Quentin Griette, Jacques Demongeot, Pierre Magal . What can we learn from COVID-19 data by using epidemic models with unidentified infectious cases?. Mathematical Biosciences and Engineering, 2022, 19(1): 537-594. doi: 10.3934/mbe.2022025
  • In this paper, we propose a stochastic SIHR epidemic model of COVID-19. A basic reproduction number Rs0 is defined to determine the extinction or persistence of the disease. If Rs0<1, the disease will be extinct. If Rs0>1, the disease will be strongly stochastically permanent. Based on realistic parameters of COVID-19, we numerically analyze the effect of key parameters such as transmission rate, confirmation rate and noise intensity on the dynamics of disease transmission and obtain sensitivity indices of some parameters on Rs0 by sensitivity analysis. It is found that: 1) The threshold level of deterministic model is overestimated in case of neglecting the effect of environmental noise; 2) The decrease of transmission rate and the increase of confirmed rate are beneficial to control the spread of COVID-19. Moreover, our sensitivity analysis indicates that the parameters β, σ and δ have significantly effects on Rs0.



    The COVID-19 has been significantly impacting our lives since the emerge of the first case in early December 2019 in Wuhan, China. As of 14 January 2022, there have been 318,648,834 confirmed cases, including 5,518,343 deaths in the world [1]. Multiple mutant strains of COVID-19 have emerged, such as Alpha, Beta, Gamma, Delta, Omicron and so on. Among these strains, Omicron, the newly discovered strain, is known as the fast speed of transmission and the strong ability of infection. Therefore it is extremely urgent to study and control the transmission of COVID-19.

    To better understand the transmission and develop efficient control strategies, researchers have employed mathematical models to analyze the dynamic behavior and control the outbreak of COVID-19 [2,3,4,5,6,7,8,9,10]. Recently, Allegretti et al. [3] considered a modified SIR model of COVID-19 and found that a high fraction of avoided contacts leads to the stability of the disease free equilibrium. Naik et al. [4] proposed a COVID-19 epidemic model and indicated that reducing transmission rate of the coronavirus is the most essential strategy to prevent the virus further spread. Okuonghae et al. [5] formulated a mathematical model and examined the impact of various non-pharmaceutical control measures on the population dynamics of COVID-19 by using the available data from Lagos, Nigeria. Fatma et al. [6] studied the interactions between COVID-19 and diabetes by using real data from Turkey and numerically visualized the population dynamics of COVID-19. Tang et al. [7] proposed a Filippov SIR model to investigate the impacts of three control strategies (media coverage, vaccination and treatment) and choose the switching policy properly to reduce the infected size. Humphrey et al. [8] developed an SEIRL model and found that testing and tracing asymptomatic individuals frequently can help in controlling new cases. Jin et al. [9] proposed a generalised SEIR model to seek optimal strategies for disease control, finding that reducing the transmission rates and increasing contact tracing are possible to hinder the fast spread of COVID-19. It is found that in the early stages of COVID-19, isolating confirmed cases was considered as a more effective control measure due to the inability to quickly produce highly effective vaccines [11]. Hence, Jiao et al. [11] proposed an SIHR model incorporating confirmed cases with general population-size dependent contact rate as follows:

    {dSdt=Aβf(N)SIμS,dIdt=βf(N)SI(γ+δ+μ+μ1)I,dHdt=δI(m+μ+μ2)H,dRdt=γI+mHμR. (1)

    Here the infectious cases are divided into two sub-populations: non-confirmed cases (I) and confirmed cases (H). The non-confirmed cases are infected individuals who have not been tested by medical institutions. Once the nucleic acid tests are positive, they would become confirmed cases and be isolated. In model (1), I(t) and H(t) are the non-confirmed infected individuals and the confirmed individuals at time t, respectively; and S(t), R(t) and N(t) denote respectively the susceptible individuals, the recovered individuals and the total population. A is the recruitment rate of the population and it is assumed that all the newcomers are susceptible. βf(N)SI is the general population-size dependent incidence, in which the parameter β is the transmission rate from the infectious class to the susceptible class, and f(N) is a function of N and it comes in many forms, see [11,12,13]. μ and γ denote the natural death rate and the natural recovery rate, respectively. δ is the confirmation rate from the infected population to confirmed cases. μ1, μ2 are respectively the extra disease-related death rate constants in compartments I and H. m is the transform rate from the confirmed population to the removed population. All parameters are nonnegative constants. Model (1) always exists the disease-free equilibrium E0=(Aμ,0,0,0), whose stability is determined by the basic reproduction number R0=Aβf(Aμ)μ(δ+γ+μ+μ1). If R0<1, then E0 is globally asymptotically stable; and if R0>1, E0 becomes unstable and an endemic equilibrium E(S,I,H,R) appears and it is locally asymptotically stable, see [11] for details.

    However, deterministic model (1) has certain limitation, it cannot describe the effects of random environment. In fact, there are many stochastic factors that can effect the transmission of disease. For example, Jamshidi et al. [14] investigated the impact of mobility, urban density, population, homestay, and mask-wearing separately on COVID-19 by conducting a multiple regression analysis and found that a higher level of population mobility and traveling can increase the transmission rate. Sabbir Hossain et al. [15] studied the impact of weather on COVID-19 in part of South Asian countries through adopting the Autoregressive Integrated Moving Average with Explanatory Variables (ARIMAX) model and revealed that maximum wind speed had a significant negative effect on the transmission rate in India, whlie rainfall, relative humidity, temperature and maximum air pollutants matter PM 2.5 had different influence on COVID-19 in different areas. We also refer the readers to Habeebullah [16], Baniasad [17] and Damette [18] for learning more about the effect of weather, pollution and mobility to the transmission efficiency of COVID-19. Therefore, the effects of these random factors can be translated to the fluctuations in the transmission rate β [19]. It has been well established in literatures that introducing parameter perturbations can affect the dynamic behavior of population. Gray et al. [20] studied the effect of stochastic parameter perturbation on SIS model and fill previous gap. Li et al. [21] applied similar parameter perturbation to investigate the effect of stochastic environmental variability on inter-pandemic transmission dynamics of influenza A. Cai et al. [22] showed that appropriate parameter perturbation to the system is useful in controlling the spread of the disease. Motivated by these, in this paper we suppose that the transmission rate β fluctuates around an average value due to the continuous fluctuations in the environment by the white noise β+σ˙B(t), and then obtain the following stochastic model:

    {dS=[Aβf(N)SIμS]dtσf(N)SIdB(t),dI=[βf(N)SI(γ+δ+μ+μ1)I]dt+σf(N)SIdB(t),dH=[δI(m+μ+μ2)H]dt,dR=(γI+mHμR)dt, (2)

    where B(t) is a standard Brownian motion, which is defined on the complete probability space (Ω,F,{Ft}t0,P), and σ represents the noise intensity; the function f():R+R+ satisfies the following assumptions:

    f(x)0,(f(x)x)0, (3)

    where f(x) and (f(x)x) respectively denote the derivative of f(x) and (f(x)x).

    The main purpose of this paper is to explore the effect of random variability in the environments on the spread of COVID-19 based on realistic parameters from [11] and [23], and obtain the strict threshold condition of the disease. The main contributions of our study can be summarized as follows.

    ● It has been shown in [24] that uncertainty is certain in the disease transmission rate of COVID-19 and there are large variation in its range. Therefore, it seems necessary and important to consider random factors in the context of COVID-19.

    ● Under the setting of more general population-size dependent contact rate, we obtain the threshold condition of disease extinction and persistence by constructing suitable Lyapunov functions. In this sense, we extend the previous studies such as [25] and [26], where the standard incidence form is used.

    ● We have proved that the basic reproduction Rs0 for our stochastic model completely determines the extinction or persistence of the disease. This is contrasted with the existing literatures [27,28,29,30,31,32], where except for the conditions required for the basic reproduction number, there still have some additional conditions for noise to ensure the extinction or persistence of disease. Therefore our results can be regard as an significant extension of the previous articles and also can provide effective information to the control of COVID-19.

    The organization of this paper is as follows. In Section 2, we prove the existence and uniqueness of the positive solution. In Section 3, using techniques in [33,34,35,36] we derive the threshold Rs0 which completely determines the extinction and strongly permanent of the disease. Finally, in Section 4, numerical simulations are carried out to illustrate our theoretical results by analyzing the effect of key parameters on disease and obtain sensitivity indices of some parameters on Rs0 by sensitivity analysis. We also discuss the impact of some measures (media coverage, government intervention, testing and tracking) on COVID-19 and give a summary.

    In order to investigate the dynamics of stochastic model (2), we first need to show that the model has a unique global positive solution. Denote by Φ(t)=(S(t),I(t),H(t),R(t)) the solution of model (2) and R4+:={(x1,x2,x3,x4)R4:xi>0,i=1,2,3,4}. Moreover, for any function VC2(R4,);R+), define the differential operator L associated with model (2) as

    LV=VΦ(Φ)˜f(Φ)+12˜gT(Φ)VΦΦ(Φ)˜g(Φ),

    where VΦ(Φ) and VΦΦ(Φ) are the gradient and Hessian of V(); ˜f and ˜g are respectively the drift and diffusion coefficients of model (2). By Itˆo's formula,

    dV=LVdt+VΦ(Φ)˜g(Φ)dB(t).

    We have the following theorem.

    Theorem 2.1. For any initial value (S(0),I(0),H(0),R(0))R4+, there is a unique solution (S(t),I(t),H(t),R(t)) of model (2) on t0, and the solution will remain in R4+ with probability 1.

    Proof. Notice that the coefficients of model (2) satisfy the local Lipschitz condition. Then there is a unique local solution (S(t),I(t),H(t),R(t)) on [0,τe) for any initial value (S(0),I(0),H(0),R(0)) in R4+, where τe is the explosion time [37]. In order to show the global existence of the positive solution, we need to prove that τe= almost surely (a.s.).

    By summing all the equations in model (2) we get that

    dN(t)=(AμNμ1Iμ2H)dt, (4)

    where N(t)=S(t)+I(t)+H(t)+R(t). It then follows that for all t<τ0,

    N(t)max{S(0)+I(0)+H(0)+R(0),Aμ}:=C1, (5)

    where

    τ0:=inf{t0:S(t)0 or I(t)0 or H(t)0 or R(t)0}.

    Obviously, τ0τe, a.s. To prove τe=, we only need to prove τ0=, a.s. For this, we set ε0>0 such that S(0)>ε0. I(0)>ε0, H(0)>ε0, R(0)>ε0. For any ε>0 satisfying εε0, define the stopping time

    τε=inf{t[0,τe):S(t)ε or I(t)ε or H(t)ε or R(t)ε}, (6)

    where we let inf= (as usual denotes the empty set). It is easy to see that τε increases to τ0 as ε tends to 0, that is τ0=limε0τε, a.s. Now we prove that τ0=, a.s.

    Suppose that this statement is false, then there exist a pair of constants T>0 and ρ(0,1) such that

    P{τ0T}>ρ.

    Thus there is a positive constant ε1ε0 such that P{τεT}>ρ for any positive εε1. Define a function V:R4+R+ by

    V(Φ(t))=lnSC1lnIC1lnHC1lnRC1.

    Clearly, V is positive definite. Applying Itˆo's formula in Appendix, we obtain

    dV=LVdt+σf(N)(IS)dB(t),

    where

    LV=AS+βf(N)I+μ+12σ2I2f2(N)βf(N)S+(γ+δ+μ+μ1)+12σ2S2f2(N)δIH+(m+μ+μ2)γIRmHR+μ.

    By using models (5) and (3), we can obtain

    LVβf(N)N+12σ2(I2+S2)f2(N)+4μ+γ+δ+μ1+m+μ2βf(C1)C1+12σ2(C1f(C1))2+4μ+γ+δ+μ1+m+μ2:=C2.

    Therefore

    dVC2dt+σf(N)(IS)dB(t).

    Integrating both sides from 0 to τεT and taking expectations, yields

    EV(Φ(τεT))V(Φ(0))+C2T.

    Set Ωε={τεT} for any positive εε1, then we have P(Ωε)>ρ. Note that for every ωΩε, there is at least one of S(τε,ω), I(τε,ω), H(τε,ω) and R(τε,ω) equals ε, then

    V(Φ(τε))lnεC1.

    Consequently,

    V(Φ(0))+C2TE[IΩεV(Φ(τεT))]=P(Ωε)V(Φ(τε))>ρlnεC1,

    where IΩε is the indicator function of Ωε. Letting ε0, we obtain the contradiction

    >V(Φ(0))+C2T=.

    So τ0=, a.s. The proof of Theorem 2.1 is thus completed.

    Denote

    Δ={(S,I,H,R)R4+:Aμ+μ1+μ2S+I+H+RAμ}. (7)

    It is easy to see from model (4) that Δ is the positive invariant set of stochastic model (2). Thus, in the sequel of this paper, we only need to consider the dynamics of model (2) constrained in Δ.

    In this section, we perform the persistence and extinction analysis of stochastic model (2). Define

    Rs0=Aβf(Aμ)μ(δ+γ+μ+μ1+12f2(Aμ)(Aμ)2σ2). (8)

    We can see from below that Rs0 plays the similar role of basic reproduction number of disease as defined in classical deterministic epidemic models, called stochastic basic reproduction number, which completely determines the dynamics of stochastic model (2).

    Denote

    g(S,I,H,R)=12σ2f2(N)S2+βf(N)S(γ+δ+μ+μ1) (9)

    and let

    λ:=g(Aμ,0,0,0)=12σ2f2(Aμ)(Aμ)2+βf(Aμ)Aμ(γ+δ+μ+μ1). (10)

    It is easy to check that Rs0=1 implies λ=0, and moreover Rs0<1 if and only if λ<0.

    Consider the following stochastic differential equation of x(t)Rn:

    dx(t)=a(x(t))dt+b(x(t))d˜B(t), (11)

    where a():RnRn and b():RnRn×d; ˜B(t) is a Ft–adapted Rd–valued standard Brownian motion. Assume that x(t)=0 is the trivial solution of model (11). The following lemma presents a proper adaptation of Theorem 3.1 in Dang and Yin [38], which will be used later to establish the condition for the extinction of disease.

    Lemma 3.1. Let D be a neighborhood of 0Rn and V:DR+ which satisfies that V(x)=0 if and only if x=0 and that V(x) is continuous on D, twice continuously differentiable in D{0}. Then the trivial solution is asymptotically stable in probability provided there exists a negative constant number c such that for any xD{0},

    LV(x)cV(x).

    The following result is about the extinction of disease.

    Theorem 3.2. If Rs0<1, then for any initial value (S(0),I(0),H(0),R(0))Δ, Φ(t)(Aμ,0,0,0) a.s. as t, i.e., the disease will go to extinction. Moreover,

    P{limtlnI(t)t=λ<0}=1. (12)

    Proof. Notice that Rs0<1 implies that λ<0. Thus we can choose a sufficiently small number ξ>0 such that λ+ξ<0. Consider the Lyapunov function

    V(S,I,H,R)=(AμS)2+Ip+H2+qR2, (13)

    where q=μ22m2 and p(0,1) will be determined later. Obviously, V(S,I,H,R)=0 if and only if (S,I,H,R)=(Aμ,0,0,0). Applying Itˆo's formula, it then follows from model (13) that

    LV(S,I,H,R)=2(AμS)(Aβf(N)SIμS)+pIp1[βf(N)SI(γ+δ+μ+μ1)I]+σ2f2(N)S2I2+12p(p1)Ipσ2f2(N)S2+2H[δI(m+μ+μ2)H]+2qR(γI+mHμR)=2μ(AμS)22(m+μ+μ2)H22qμR2+I[2(AμS)βf(N)S+σ2f2(N)S2I+2δH+2qγR]+pIpg(S,I,H,R)+2qmHR+12p2σ2f2(N)S2Ip. (14)

    We first perform some estimates on the items appeared in model (14). Notice that

    2μ(AμS)2(2m+2μ2+μ)H2qμR2p(λ+ξ)[(AμS)2+H2+qR2], (15)

    provided p is sufficiently small such that

    min{2μ,2m+2μ2+μ,qμ}p|λ+ξ|. (16)

    Now denote Uδ1:=(Aμδ1,Aμ]×[0,δ1)3 for δ1(0,Aμ). Notice the continuity of functions g() and f(). We can take δ1 and p sufficiently small such that model (16) holds and for any (S,I,H,R)Uδ1, the following two inequalities hold:

    pIpg(S,I,H,R)p[g(Aμ,0,0,0)+ξ1]Ip=p(λ+ξ1)Ip

    and

    I[2(AμS)βf(N)S+σ2f2(N)S2I+2δH+2qγR]+12p2σ2f2(N)S2Ippξ2Ip,

    where 0<ξ1,ξ2<ξ and ξ1+ξ2=ξ. Consequently, we have

    pIpg(S,I,H,R)+I[2(AμS)βf(N)S+σ2f2(N)S2I+2δH+2qγR]+12p2σ2f2(N)S2Ipp(λ+ξ)Ip. (17)

    Moreover, we can easily check that

    qμR2+2qmHRμH20. (18)

    Combining models (15), (17) and (18), we know that if we take p and δ1 both sufficiently small, it then follows from model (14) that for any (S,I,H,R)Uδ1, we have

    LV(S,I,H,R)p(λ+ξ)V(S,I,H,R).

    According to Lemma 3.1, we know that the disease free equilibrium (Aμ,0,0,0) is asymptotically stable in probability. That is, for any ε>0, there exists a δ2, 0<δ2<δ1 such that

    P{limtΦ(t)=(Aμ,0,0,0)}1ε (19)

    for any (S(0),I(0),H(0),R(0))Uδ2, where Uδ2=(Aμδ2,Aμ]×[0,δ2)3. Now we are in a position to prove that any solution starting in Δ will eventually enter Uδ2.

    Define τδ2=inf{t0:S(t)Aμδ2}. Consider the Lyapunov function V1(Φ(t))=c1(S+1)c2, where c1 and c2 are two positive constants to be specified. By calculating we obtain

    LV1(Φ(t))=c2(S+1)c22[(S+1)(μ(AμS)βf(N)SI)+c212σ2f2(N)S2I2].

    For any S(0,Aμδ2], we have (S+1)μ(AμS)μδ2 and inf{σ2f2(N)}>0, then we can choose a sufficiently large c2 such that

    (S+1)[μ(AμS)βf(N)SI]+c212σ2f2(N)S2I212μδ2.

    Hence

    LV1(Φ(t))12μδ2.

    By Dynkin's formula [39], we obtain

    E[V1(Φ(τδ2t))]=V1(Φ(0))+Eτδ2t0LV1(Φ)dsV1(Φ(0))12μδ2E(τδ2t).

    Letting t and using Fatou's lemma yields that

    E[V1(Φ(τδ2))]V1(Φ(0))12μδ2E(τδ2).

    Due to V1 is bounded on R4+, then E(τδ2)<. By the strong Markov property, from model (19) and E(τδ2)< we have that

    P{limtΦ(t)=(Aμ,0,0,0)}1ε

    for any ε>0 and (S(0),I(0),H(0),R(0))Δ. Therefore

    P{limtΦ(t)=(Aμ,0,0,0)}=1 (20)

    for any (S(0),I(0),H(0),R(0))Δ. Applying the Itˆo's formula, we have

    dlnI=(βf(N)S(γ+δ+μ+μ1)12σ2f2(N)S2)dt+σf(N)SdB(t). (21)

    Integrating both sides of model (21) from 0 to t leads to

    lnI(t)lnI(0)=t0g(Φ(u))du+t0σf(N)SdB(u). (22)

    By the strong law of large numbers for martingales, we obtain from models (9) and (20) that

    limt1tt0g(Φ(u))du=λ

    and

    limt1tt0σf(N)S(u)dB(u)=0, a.s.

    It then follows from model (22) that limtlnI(t)t=λ, which implies model (12). The proof is thus completed.

    In this section, we prove that the disease will be persistent provided Rs0>1. We first present the following useful lemma.

    Lemma 3.3. Let Δ:={(S,I,H,R)Δ:I=0}. Then for any (S(0),I(0),H(0),R(0))Δ, there exists T>0 such that

    ET0g(S(t),I(t),H(t),R(t))dt3λ4T. (23)

    Proof. If I(0)=0, then I(t)=0 for all t0 and model (2) becomes

    {dS=(AμS)dt,dI=0,dH=(m+μ+μ2)Hdt,dR=(mHμR)dt. (24)

    Obviously, for any (S(0),I(0),H(0),R(0))Δ, (S(t),I(t),H(t),R(t)) tends asymptotically to the disease free equilibrium (Aμ,0,0,0). Therefore,

    limt1tt0g(S(s),I(s),H(s),R(s))ds=g(Aμ,0,0,0)=λ.

    Then there exists a T>0 such that model (23) holds.

    The following theorem is about the persistence of the disease.

    Theorem 3.4. If Rs0>1, then for any initial value (S(0),I(0),H(0),R(0))Δ, the disease is strongly stochastically permanent, namely, for any ε>0, there exists a α>0 such that

    lim inftP{I(t)α}>1ε. (25)

    Proof. Notice that Rs0>1 implies that λ>0. Consider the Lyapunov function Vθ=Iθ, where θR is a constant. Applying the Itˆo's formula, we have

    LVθ=θ[βf(N)S(γ+δ+μ+μ1)+θ12σ2f2(N)S2]Iθ. (26)

    Denoting Qθ=sup(S,I,H,R)Δ{θ[βf(N)S(γ+δ+μ+μ1)+θ12σ2f2(N)S2]}, then we have from model (26) that LVθQθIθ for any initial value (S(0),I(0),H(0),R(0))Δ. By integrating both sides of model (26) from 0 to t and taking expectation yields

    E(Iθ(t))=E(Iθ(0))+Et0LVθdsE(Iθ(0))+t0QθE(Iθ(s))ds.

    Using Gronwall inequality, for any t0 and (S(0),I(0),H(0),R(0))Δ, we have

    E(Iθ(t))Iθ(0)exp(Qθt). (27)

    Similarly, we have for tnT,

    E(Iθ(t))E(Iθ(nT))exp[Qθ(tnT)]. (28)

    Denote lnI(0)lnI(t)=W(t). It then follows from model (22) that

    W(t)=t0g(Φ(u))dut0σf(N)S(u)dB(u). (29)

    By Feller property and Lemma 3.3, it then follows from model (29) that there exists a sufficiently small δ3>0 such that for any (S(0),I(0),H(0),R(0))Δ with I(0)<δ3, we have

    E(W(T))=ET0g(Φ(t))dtλ2T. (30)

    Notice also from model (27) that for any fixed t0,

    E(eW(t)+eW(t))=E(I(0)I(t)+I(t)I(0))E(eQ1t+eQ1t)<. (31)

    Using Lemma A.1 in Appendix, we obtain

    lnE(eθW(T))E(θW(T))+ˆQ1θ2,θ[0,0.5],

    where ˆQ1 is a constant which depends on T, Q1 and Q1. For a sufficiently small θ satisfying ˆQ1θ2λθ4T, by model (30), we have

    E(Iθ(0)Iθ(T))=E(eθW(T))exp(λθ2T+ˆQ1θ2)exp(λθ4T).

    Then

    E(Iθ(T))Iθ(0)exp(λθ4T)=qIθ(0) (32)

    for I(0)<δ3, where q=exp(λθ4T).

    Next, by model (27), we obtain that

    E(Iθ(T))δθ3exp(QθT):=C (33)

    for any I(0)>δ3. Then model (33) together with model (32) implies that

    E(Iθ(T))qIθ(0)+C

    for any (S(0),I(0),H(0),R(0))Δ. By the Markov property, we have

    E[Iθ((k+1)T)]qE(Iθ(kT))+C.

    Using this recursively, we obtain

    E[Iθ((k+2)T)]q2E(Iθ(kT))+qC+C

    and

    E(Iθ(nT))qnIθ(0)+C(1qn)1q.

    This together with model (28) leads to

    E(Iθ(t))(qnIθ(0)+C(1qn)1q)exp(QθT),t[nT,(n+1)T].

    Letting n, we obtain

    lim suptE(Iθ(t))C1qexp(QθT):=Z.

    For any ε>0, let α=ε1θZ1θ. By Chebyshev's inequality we obtain

    P{|I(t)|<α}=P{1|I(t)|θ>1αθ}αθE(|Iθ(t)|).

    That is,

    lim suptP{|I(t)|<α}αθZ=ε.

    Therefore

    lim inftP{I(t)α}>1ε.

    The proof is thus completed.

    Remark 1. In this section, we obtain the threshold condition Rs0 which completely determines the extinction and persistence of disease. By contrast, in [40], besides the conditions required on basic reproduction number R0, there is one additional condition μ>σ21σ22σ23σ242 to make the disease extinct or persistent. Thus, our results can be regard as an significant extension of COVID-19 and could provide government effective information to the control of of disease transmission.

    In this section, model (2) with the incidence function f(N)=11+bN+1+2bN [12] is applied to verify/extend our analytical results based on the realistic parameter values of COVID-19 from Asamoah [23] and Jiao [11]. We fix the initial values (S(0),I(0),H(0),R(0))=(50000,5,2,1) except for other specification. We divide our simulations into the following two subsections.

    In the subsection, we numerically simulate the solution of model (2) using MATLAB R2016b to illustrate the theoretical results obtained in Section 3, mainly revealing the effect of σ, β and δ on the dynamics of model (2). The numerical scheme is obtained through Milstein's higher order method [41].

    Case 1. The effect of noise intensity

    We first suppose β=0.1860 and b=0.3900 and other parameter values are shown in Table 1. In this case one can get that R0=2.3081. It then follows from [11] that deterministic model (1) has a unique endemic equilibrium E(12824304.8094,3743.8322,9661.8313,16762296.1178), which is stable. It follows from Rs0=1.9900 that we have σ=0.1000. By Theorem 3.4, we can get that the disease is strongly stochastically permanent. The computer simulations shown in Figure 1(a) clearly support the result. To show how the noise affects the dynamics of disease, now we take σ=0.0100, and other parameters remain unchanged. In this case, we obtain Rs0=2.3045>1. The similar simulation result is shown in Figure 1(b). We observed that the path of I(t) for model (2) is oscillating around the steady state value I=3743.8322. Compared with Figure 1(a), one can get that when Rs0>1, the small noise does not change the stability of the equilibrium state of model (2), but with the intensity of white noise increasing, the volatility of I(t) is getting larger. Finally, we consider σ=0.3000. It is easy to compute that Rs0=0.9494<1, according to Theorem 3.2, the disease will go extinct almost surely as shown in Figure 1(c). However, deterministic model (1) claims the persistence of the disease. This discrepancy highlights the impact of stochastic environmental to the disease dynamics.

    Table 1.  Parameter values used in the simulation.
    Parameters Description Values Sources
    A The recruitment rate 1319.2940 [23]
    μ The natural death rate 0.000042578 [23]
    γ The natural recovery rate 0.0185 [11]
    m The recovery rate of confirmed individuals 0.0667 [11]
    δ The confirmed rate 0.1836 [11]
    μ1 The disease-induced death rate of infected individuals 0.0044 [23]
    μ2 The disease-induced death rate of confirmed cases 0.0044 [23]
    Rs0 The basic reproduction number of COVID-19 1.9900 [23]

     | Show Table
    DownLoad: CSV
    Figure 1.  The evolution of a single path of I(t) of models (1) and (2) is graphed for different values of σ (0.1000, 0.0100, 0.3000). Here we take β=0.1860, b=0.3900 and other parameter values are tabulated in Table 1 (Color figure online).

    Case 2. The effect of transmission rate

    Here we assume b=0.3900, σ=0.1000, and other parameters take values as in Table 1. In this case we consider three different values of β=0.0500,0.1860,0.3500 to see the effect of transmission rate on the spread of infectious disease. The corresponding values of Rs0 are respectively 0.5353, 1.9900 and 3.7473. By Theorems 3.2 and 3.4, we can get that the disease is strongly stochastically permanent when β=0.1860 and β=0.3500, while the disease is extinctive when β=0.0500. The computer simulations are shown in Figure 2(a) which clearly support these results. Figure 2(b) shows the corresponding persistence level of I(t) for various values of β. It is observed that when Rs0>1, the persistence level of I(t) is reduced gradually with the decrease of transmission rate. This indicates that decreasing transmission rate is beneficial to the control of the spread of COVID-19. So we can take some measures to reduce the scale of outbreaks by decreasing transmission rate. For example, transmission rate can be reduced through improving the media response rate to reports on the severity of COVID-19 and encouraging citizens to actively prevent disease. Moreover, government can adopt a series of policies including wearing masks, avoiding farm and wild animals, travel restrictions, stay at home, lockdowns, and so on to decline transmission rate. These measures could effectively reduce the number of infected cases and suppress the outbreak of disease.

    Figure 2.  (a) The evolution of a single path of I(t) of stochastic model (2) is graphed for different values of β (0.3500, 0.1860, 0.0500). (b) The corresponding persistence level of infected individuals of model (2) is graphed for various values of β. Here we take σ=0.1000, b=0.3900 and other parameter values are shown in Table 1 (Color figure online).

    Case 3. The effect of confirmed rate

    Here we assume β=0.1860, b=0.3900, σ=0.1000, and other parameters taking values as in Table 1. In this case we choose three different values of δ0.1000,0.1836,0.6000 to see the effect of confirmed rate on the spread of infectious diseases. The corresponding values of Rs0 are respectively 3.0601, 1.9900, 0.7270. According to Theorems 3.2 and 3.4, the disease persists when δ=0.1000 and δ=0.1836, while the disease will be extinct when δ=0.6000. Figure 3(a) clearly support these results. Figure 3(b) shows that the corresponding persistence level of I(t) for various values of δ. We can see that when Rs0>1, the persistence level of I(t) is reduced gradually with the increase of confirmed rate. This indicates that the increase of confirmed rate is beneficial to control the spread of COVID-19, while blind testing is not desirable, it will cause a huge burden on society. Therefore, in order to enhance confirmation rate, a positive tracking and testing strategy should be carried out to control the spread of disease [8]. Frequently testing in smaller scale populations, such as schools, factories, community, etc., where virus is more easier to spread, and test less frequently parts of the population who are not as exposed. Detecting continually close contacts also leads to the increase of confirmed cases.

    Figure 3.  (a) The evolution of a single path of I(t) of stochastic model (2) is graphed for different values of δ (0.1000, 0.1836, 0.6000). (b) The corresponding persistence level of infected individuals of model (2) is graphed for various values of δ. Here we take β=0.1860, σ=0.1000, b=0.3900 and other parameter values are given in Table 1 (Color figure online).

    Varying parameter values will have different effects on the output of model (2). In order to qualitatively analyze the influence of some parameters on the output of model (2), the sensitivity analysis method is adopted. The normalized forward sensitivity index of a variable, R, that depends on a parameter, h, is defined as:

    ΥRh=Rh×|hR|.

    Using the above formula, we analyze the sensitivity of state variable Rs0 to the following parameters of model (2):

    A=1319.2940,β=0.1860,δ=0.1836,σ=0.1000μ=0.000042578,μ1=0.0044,γ=0.0185. (34)

    From Table 2, parameters with positive sensitivity index, A and β, indicate that the transmission of COVID-19 increases with the increase of these two parameters. Similarly, parameters with negative sensitivity index, μ, γ, δ, σ and μ1, mean that the transmission of COVID-19 decreases with the increase of these parameters. As shown in Figure 4, we observe that β, σ and δ have significant effects on Rs0. This verifies that our analysis of parameters is meaningful in Subsection 4.1.

    Table 2.  Normalized sensitivity index for some parameters for Rs0.
    Parameters Description Sensitivity Index
    A The recruitment rate +0.0001477
    μ The natural death rate -0.0003257
    γ The natural recovery rate -0.0773
    β The transmission rate +1.0000
    δ The confirmed rate -0.7675
    σ The noise intensity -0.2746
    μ1 The disease-induced death rate of infected individuals -0.0184

     | Show Table
    DownLoad: CSV
    Figure 4.  The sensitivity indices of state variable Rs0 with respect to some parameters for model (2).

    In reality, there exist many random environmental factors, like weather, relative humidity, temperature and population mobility, which may have significant effects on the transmission of COVID-19. Therefore, considering stochastic influences into the epidemic model seems necessary and important. In this paper, we propose and investigate a stochastic SIHR epidemic model with the environmental variability in the transmission rate to describe the transmission of COVID-19, and based on which we numerically illustrate the evolution dynamic of COVID-19 using the realistic parameter values from literatures.

    The main contribution of our paper can be summarized as the following two aspects. Mathematically, we prove that the stochastic dynamics of stochastic model (2) is completely determined by the reproduction number Rs0: If Rs0<1, the disease will go to extinction ultimately, and if Rs0>1, the disease stochastically permanent. Epidemiologically, we can conclude that: (ⅰ) The presence of environmental noises can sustain the irregular recurrence of disease and the volatility of infected population increases with the increasing noise intensity if Rs0>1. When the noise increases to a certain level such that Rs0<1, the disease will go to extinction (See Figure 1). And also it can be seen from Figure 1(c) that white noise may reshape the solution behavior of corresponding deterministic model (1). In other words, noise may change the evolution tendency of disease. (ⅱ) The decrease of transmission rate and the increase of confirmed rate are beneficial to the control of COVID-19 spread (See Figures 2 and 3). (ⅲ) Our sensitivity analysis indicates that the transmission rate β, noise intensity σ and confirmed rate δ are the most sensitive parameters to Rs0 (See Figure 4).

    More than two years have lasted since the emergence of COVID-19 in the world. It is well known the transmission of disease will necessarily be affected by other factors such as media coverage, seasonal changes and so on [42,43,44,45,46,47]. Considering the seasonal effect or the switching of environments in model (2) will be an interesting research topic. We leave this for our future investigation.

    Research is supported by the National Natural Science Foundation of China (No. 12071239).

    The authors declare there is no conflict of interest.

    The following lemma is from [33], which is used in the establishment of conditions for the persistence of disease.

    Lemma A.1. Let Y be a random variable, suppose Eexp(Y)+Eexp(Y)K1. Then the log-Laplace transform ϕ(θ)=lnEexp(θY) is twice differentiable on [0,0.5] and dϕdθ(0)=EY, 0d2ϕdθ2(θ)2K2, θ[0,0.5] for some K2>0 depending only on K1. Thus, it follows from Taylor's expansion that

    ϕ(θ)θEY+K2θ2,θ[0,0.5].


    [1] World Health Organization. Available from: https://covid19.who.int.
    [2] A. Din, Y. Li, T. Khan, G. Zaman, Mathematical analysis of spread and control of the novel corona virus (COVID-19) in China, Chaos Solitons Fractals, 141 (2020), 110286. https://doi.org/10.1016/j.chaos.2020.110286 doi: 10.1016/j.chaos.2020.110286
    [3] S. Allegretti, I. Bulai, R. Marino, M. Menandro, K. Parisi, Vaccination effect conjoint to fraction of avoided contacts for a Sars-Cov-2 mathematical model, Math. Model. Num. Sim. Appl., 2 (2021), 56–66. https://doi.org/10.53391/mmnsa.2021.01.006 doi: 10.53391/mmnsa.2021.01.006
    [4] P. Naik, M. Yavuz, S. Qureshi, J. Zu, S. Townley, Modeling and analysis of COVID-19 epidemics with treatment in fractional derivatives using real data from Pakistan, Eur. Phys. J. Plus., 135 (2020), 795. https://doi.org/10.1140/epjp/s13360-020-00819-5 doi: 10.1140/epjp/s13360-020-00819-5
    [5] D. Okuonghae, A. Omame, Analysis of a mathematical model for COVID-19 population dynamics in Lagos, Nigeria, Chaos Solitons Fractals, 139 (2020), 110032. https://doi.org/10.1016/j.chaos.2020.110032 doi: 10.1016/j.chaos.2020.110032
    [6] F. ¨Ozk¨ose, M. Yavuz, Investigation of interactions between COVID-19 and diabetes with hereditary traits using real data: A case study in Turkey, Comput. Bio. Med., 141 (2021), 105044. https://doi.org/10.1016/j.compbiomed.2021.105044 doi: 10.1016/j.compbiomed.2021.105044
    [7] J. Deng, S. Tang, H. Shu, Joint impacts of media, vaccination and treatment on an epidemic Filippov model with application to COVID-19, J. Theor. Biol., 523 (2021), 110698. https://doi.org/10.1016/j.jtbi.2021.110698 doi: 10.1016/j.jtbi.2021.110698
    [8] L. Humphrey, E. Thommes, R. Fields, L. Coudeville, N. Hakim, A. Chit, et al., Large-scale frequent testing and tracing to supplement control of COVID-19 and vaccination rollout constrained by supply, Infect. Dis. Modell., 6 (2021) 955–974. https://doi.org/10.1016/j.idm.2021.06.008 doi: 10.1016/j.idm.2021.06.008
    [9] J. Asamoah, Z. Jin, G. Sun, B. Seidu, E. Yankson, A. Abidemi, et al., Sensitivity assessment and optimal economic evaluation of a new COVID-19 compartmental epidemic model with control interventions, Chaos Solitons Fractals, 146 (2021), 110885. https://doi.org/10.1016/j.chaos.2021.110885 doi: 10.1016/j.chaos.2021.110885
    [10] X. Duan, X. Li, M. Martcheva, S. Yuan, Using an age-structured COVID-19 epidemic model and data to model virulence evolution in Wuhan, China, J. Biol. Dynam., 16 (2022), 14–28. https://doi.org/10.1080/17513758.2021.2020916 doi: 10.1080/17513758.2021.2020916
    [11] S. Jiao, M. Huang, An SIHR epidemic model of the COVID-19 with general population-size dependent contact rate, AIMS Math., 5 (2020), 6714–6725. https://doi.org/10.3934/math.2020431 doi: 10.3934/math.2020431
    [12] J. Zhang, Z. Ma, Global dynamics of an SEIR epidemic model with saturating contact rate, Math. Biosci., 185 (2003), 15–32. https://doi.org/10.1016/S0025-5564(03)00087-7 doi: 10.1016/S0025-5564(03)00087-7
    [13] Q. Liu, D. Jiang, N. Shi, T. Hayat, B. Ahmad, Stationary distribution and extinction of a stochastic SEIR epidemic model with standard incidence, Physica A, 476 (2017), 58–69. https://doi.org/10.1016/j.physa.2017.02.028 doi: 10.1016/j.physa.2017.02.028
    [14] S. Jamshidi, M. Baniasad, D. Niyogi, Global to USA county scale analysis of weather, urban density, mobility, homestay, and mask use on COVID-19, Int. J. Environ. Res. Public Health, 17 (2020), 7847. https://doi.org/10.3390/ijerph17217847 doi: 10.3390/ijerph17217847
    [15] M. S. Hossain, S. Ahmed, M. J. Uddin, Impact of weather on COVID-19 transmission in south Asian countries: An application of the ARIMAX model, Sci. Total Environ., 761 (2021), 143315. https://doi.org/10.1016/j.scitotenv.2020.143315 doi: 10.1016/j.scitotenv.2020.143315
    [16] M. Habeebullah, H. A. Abd El-Rahim, A. Morsy, Impact of outdoor and indoor meteorological conditions on the COVID-19 transmission in the western region of Saudi Arabia, J. Environ. Manage., 288 (2021), 112392. https://doi.org/10.1016/j.jenvman.2021.112392 doi: 10.1016/j.jenvman.2021.112392
    [17] M. Baniasad, G. Mofrad, B. Bahmanabadi, S. Jamshidi, COVID-19 in Asia: Transmission factors, re-opening policies, and vaccination simulation, Environ. Res., 202 (2021), 111657. https://doi.org/10.1016/j.envres.2021.111657 doi: 10.1016/j.envres.2021.111657
    [18] O. Damette, S. Goutte, Weather, pollution and COVID-19 spread: A time series and wavelet reassessment, HAL, 2020. Available from: https://halshs.archives-ouvertes.fr/halshs-02629139.
    [19] M. Keeling, P. Rohani, Modeling infectious diseases in human and animals, Princeton University Press, New Jersey, 2008. https://doi.org/10.2307/j.ctvcm4gk0
    [20] A. Gray, D. Greenhalgh, L. Hu, X. Mao, J. Pan, A stochastic differential equation SIS epidemic model, SIAM J. Appl. Math., 71 (2014), 876–902. https://doi.org/10.2307/23073365 doi: 10.2307/23073365
    [21] D. Li, J. Cui, M. Liu, S. Liu, The evolutionary dynamics of stochastic epidemic model with nonlinear incidence rate, Bull. Math. Biol., 77 (2015), 1705–1743. https://doi.org/10.1007/s11538-015-0101-9 doi: 10.1007/s11538-015-0101-9
    [22] Y. Cai, J. Li, Y. Kang, K. Wang, W. Wang, The fluctuation impact of human mobility on the influenza transmission, J. Franklin Inst., 357 (2020), 8899–8924. https://doi.org/10.1016/j.jfranklin.2020.07.002 doi: 10.1016/j.jfranklin.2020.07.002
    [23] J. Asamoah, M. Owusu, Z. Jin, F. Oduro, A. Abidemi, E. Gyasi, Global stability and cost-effectiveness analysis of COVID-19 considering the impact of the environment: using data from Ghana, Chaos Solitions Fractals, 140 (2020), 110103. https://doi.org/10.1016/j.chaos.2020.110103 doi: 10.1016/j.chaos.2020.110103
    [24] C. Manski, F. Molinari, Estimating the COVID-19 infection rate: anatomy of an inference problem, J. Econom., 220 (2020), 181–192. https://doi.org/10.1016/j.jeconom.2020.04.041 doi: 10.1016/j.jeconom.2020.04.041
    [25] Z. Cao, W. Feng, X. Wen, L. Zu, M. Cheng, Dynamics of a stochastic SIQR epidemic model with standard incidence, Physica A, 527 (2019), 121180. https://doi.org/10.1016/j.physa.2019.121180 doi: 10.1016/j.physa.2019.121180
    [26] D. Adak, A. Majumder, N. Bairagi, Mathematical perspective of COVID-19 pandemic: Disease extinction criteria in deterministic and stochastic models, Chaos Solitions Fractals, 27 (2021), 104472. https://doi.org/10.1016/j.chaos.2020.110381 doi: 10.1016/j.chaos.2020.110381
    [27] M. Fatini, M. Khalifi, R. Gerlach, A. Laaribi, R. Taki, Stationary distribution and threshold dynamics of a stochastic SIRS model with a general incidence, Physica A, 534 (2019), 120696. https://doi.org/10.1016/j.physa.2019.03.061 doi: 10.1016/j.physa.2019.03.061
    [28] C. Ji, D. Jiang, Threshold behaviour of a stochastic SIR model, Appl. Math. Model., 38 (2014), 5067–5079. https://doi.org/10.1016/j.apm.2014.03.037 doi: 10.1016/j.apm.2014.03.037
    [29] Y. Zhou, W. Zhang, S. Yuan, Survival and stationary distribution of a SIR epidemic model with stochastic perturbations, Appl. Math. Comput., 244 (2014), 118–131. https://doi.org/10.1016/j.amc.2014.06.100 doi: 10.1016/j.amc.2014.06.100
    [30] Q. Liu, D. Jiang, Dynamics of a stochastic multigroup SIQR epidemic model with standard incidence rates, J. Franklin Inst., 356 (2019), 2960–2993. https://doi.org/10.1016/j.jfranklin.2019.01.038 doi: 10.1016/j.jfranklin.2019.01.038
    [31] Q. Liu, Stability of SIRS system with random perturbations, Physica A, 388 (2009), 3677–3686. https://doi.org/10.1016/j.physa.2009.05.036 doi: 10.1016/j.physa.2009.05.036
    [32] G. Lan, S. Yuan, B. Song, The impact of hospital resources and environmental perturbations to the dynamics of SIRS model, J. Franklin Inst., 358 (2021), 2405–2433. https://doi.org/https://doi.org/10.1016/j.jfranklin.2021.01.015 doi: 10.1016/j.jfranklin.2021.01.015
    [33] T. Tuong, H. Dang, N. Dieu, K. Tran, Extinction and permanence in a stochastic SIRS model in regime-switching with general incidence rate, Nonlinear Anal. Hybri., 34 (2019), 121–130. https://doi.org/10.1016/j.nahs.2019.05.008 doi: 10.1016/j.nahs.2019.05.008
    [34] M. Benaïm, C. Lobry, Lotka-volterra with randomly fluctuating environments or how switching between beneficial environments can make survival harder, Ann. Appl. Probab., 26 (2016), 3754–3785. https://doi.org/10.1214/16-AAP1192 doi: 10.1214/16-AAP1192
    [35] M. Benaïm, Stochastic persistence, arXiv: 1806.08450.
    [36] A. Hening, H. D. Nguyen, Coexistence and extinction for stochastic kolmogorov systems, Ann. Appl. Probab., 28 (2018), 1893–1942. https://doi.org/10.1214/17-AAP1347 doi: 10.1214/17-AAP1347
    [37] X. Mao, Stochastic Differential Equations and Applications, Horwood Publishing, Chichester, 2007.
    [38] N. Dang, G. Yin, Stability of regime-switching diffusion systems with discrete states belonging to a countable set, SIAM J. Control Optim., 56 (2018), 3893–3917. https://doi.org/10.1137/17M1118476 doi: 10.1137/17M1118476
    [39] F. Klebaner, Introduction to Stochastic Calculus with Applications, Imperial College Press, 2005.
    [40] R. Ikram, A. Khan, M. Zahri, A. Saeed, M. Yavuz, P. Kumam, Extinction and stationary distribution of a stochastic COVID-19 epidemic model with time-delay, Comput. Biol. Medi., 141 (2022), 105115. https://doi.org/10.1016/j.compbiomed.2021.105115 doi: 10.1016/j.compbiomed.2021.105115
    [41] D. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. https://doi.org/10.1137/S0036144500378302 doi: 10.1137/S0036144500378302
    [42] A. Din, Y. Li, Stationary distribution extinction and optimal control for the stochastic hepatitis B epidemic model with partial immunity, Phys. Scripta, 7 (2021), 074005. https://doi.org/10.1088/1402-4896/abfacc doi: 10.1088/1402-4896/abfacc
    [43] A. Din, Y. Li, Stochastic optimal control for norovirus transmission dynamics by contaminated food and water, Chinese Phys. B, 2021. https://doi.org/10.1088/1674-1056/ac2f32 doi: 10.1088/1674-1056/ac2f32
    [44] A. Din, The stochastic bifurcation analysis and stochastic delayed optimal control for epidemic model with general incidence function, Chaos, 31 (2021), 104649. https://doi.org/10.1063/5.0063050 doi: 10.1063/5.0063050
    [45] A. Yang, B. Song, S. Yuan, Noise-induced transitions in a non-smooth SIS epidemic model with media alert, Math. Biosci. Eng., 18 (2021), 745–763. https://doi.org/10.3934/mbe.2021040 doi: 10.3934/mbe.2021040
    [46] J. Jaramillo, J. Ma, P. Driessche, S. Yuan, Host contact structure is important for the recurrence of Influenza A, J. Math. Biol., 77 (2018), 1563–1588. https://doi.org/10.1007/s00285-018-1263-5 doi: 10.1007/s00285-018-1263-5
    [47] Y. Zhao, L. Zhang, S. Yuan, The effect of media coverage on threshold dynamics for a stochastic SIS epidemic model, Physica A, 512 (2018), 248–260. https://doi.org/10.1016/j.physa.2018.08.113 doi: 10.1016/j.physa.2018.08.113
  • This article has been cited by:

    1. Zin Thu Win, Mahmoud A. Eissa, Boping Tian, Stochastic Epidemic Model for COVID-19 Transmission under Intervention Strategies in China, 2022, 10, 2227-7390, 3119, 10.3390/math10173119
    2. Qun Liu, Daqing Jiang, Stationary distribution and probability density for a stochastic SEIR-type model of coronavirus (COVID-19) with asymptomatic carriers, 2023, 169, 09600779, 113256, 10.1016/j.chaos.2023.113256
    3. Junaid Iqbal Khan, Farman Ullah, Sungchang Lee, Attention based parameter estimation and states forecasting of COVID-19 pandemic using modified SIQRD Model, 2022, 165, 09600779, 112818, 10.1016/j.chaos.2022.112818
    4. Tingting Ding, Tongqian Zhang, Asymptotic behavior of the solutions for a stochastic SIRS model with information intervention, 2022, 19, 1551-0018, 6940, 10.3934/mbe.2022327
    5. Guijie Lan, Baojun Song, Sanling Yuan, Epidemic threshold and ergodicity of an SEIR model with vertical transmission under the telegraph noise, 2023, 167, 09600779, 113017, 10.1016/j.chaos.2022.113017
    6. Guijie Lan, Sanling Yuan, Baojun Song, Threshold behavior and exponential ergodicity of an sir epidemic model: the impact of random jamming and hospital capacity, 2024, 88, 0303-6812, 10.1007/s00285-023-02024-1
    7. Xiaojie Mu, Daqing Jiang, A stochastic SIHR epidemic model with general population-size dependent contact rate and Ornstein–Uhlenbeck process: dynamics analysis, 2024, 112, 0924-090X, 10703, 10.1007/s11071-024-09586-9
    8. Gaohui Fan, Ning Li, Application and analysis of a model with environmental transmission in a periodic environment, 2023, 31, 2688-1594, 5815, 10.3934/era.2023296
    9. Jing Yang, Shaojuan Ma, Juan Ma, Jinhua Ran, Xinyu Bai, Stochastic Analysis for the Dual Virus Parallel Transmission Model with Immunity Delay, 2024, 1557-8666, 10.1089/cmb.2024.0662
    10. Muzi Li, Guijie Lan, Chunjin Wei, Threshold dynamics of stochastic H7N9 model with Markov switching and hybrid strategy, 2024, 361, 00160032, 916, 10.1016/j.jfranklin.2023.12.034
    11. Jiacheng Song, Wangyong Lv, Yaling Deng, Zhehao Sun, A double stochastic SIS network epidemic model with nonlinear contact rate and limited medical resources, 2024, 112, 0924-090X, 6743, 10.1007/s11071-024-09291-7
    12. Aojie Huang, M. Md Husin, Analysis of the Profit Model of Internet Money Market Funds Based on the Characteristics of Internet Economy–Taking balance treasure as an example, 2024, 193, 2261-2424, 01031, 10.1051/shsconf/202419301031
  • Reader Comments
  • © 2022 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(3134) PDF downloads(182) Cited by(12)

Figures and Tables

Figures(4)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog