Processing math: 100%
Research article Special Issues

Dynamic modeling and analysis of Hepatitis B epidemic with general incidence


  • New stochastic and deterministic Hepatitis B epidemic models with general incidence are established to study the dynamics of Hepatitis B virus (HBV) epidemic transmission. Optimal control strategies are developed to control the spread of HBV in the population. In this regard, we first calculate the basic reproduction number and the equilibrium points of the deterministic Hepatitis B model. And then the local asymptotic stability at the equilibrium point is studied. Secondly, the basic reproduction number of the stochastic Hepatitis B model is calculated. Appropriate Lyapunov functions are constructed, and the unique global positive solution of the stochastic model is verified by Itô formula. By applying a series of stochastic inequalities and strong number theorems, the moment exponential stability, the extinction and persistence of HBV at the equilibrium point are obtained. Finally, using the optimal control theory, the optimal control strategy to eliminate the spread of HBV is developed. To reduce Hepatitis B infection rates and to promote vaccination rates, three control variables are used, for instance, isolation of patients, treatment of patients, and vaccine inoculation. For the purpose of verifying the rationality of our main theoretical conclusions, the Runge-Kutta method is applied to numerical simulation.

    Citation: Tingting Xue, Long Zhang, Xiaolin Fan. Dynamic modeling and analysis of Hepatitis B epidemic with general incidence[J]. Mathematical Biosciences and Engineering, 2023, 20(6): 10883-10908. doi: 10.3934/mbe.2023483

    Related Papers:

    [1] Kangbo Bao, Qimin Zhang, Xining Li . A model of HBV infection with intervention strategies: dynamics analysis and numerical simulations. Mathematical Biosciences and Engineering, 2019, 16(4): 2562-2586. doi: 10.3934/mbe.2019129
    [2] Tahir Khan, Fathalla A. Rihan, Muhammad Ibrahim, Shuo Li, Atif M. Alamri, Salman A. AlQahtani . Modeling different infectious phases of hepatitis B with generalized saturated incidence: An analysis and control. Mathematical Biosciences and Engineering, 2024, 21(4): 5207-5226. doi: 10.3934/mbe.2024230
    [3] Jiying Ma, Shasha Ma . Dynamics of a stochastic hepatitis B virus transmission model with media coverage and a case study of China. Mathematical Biosciences and Engineering, 2023, 20(2): 3070-3098. doi: 10.3934/mbe.2023145
    [4] Tailei Zhang, Hui Li, Na Xie, Wenhui Fu, Kai Wang, Xiongjie Ding . Mathematical analysis and simulation of a Hepatitis B model with time delay: A case study for Xinjiang, China. Mathematical Biosciences and Engineering, 2020, 17(2): 1757-1775. doi: 10.3934/mbe.2020092
    [5] Dong-Me Li, Bing Chai, Qi Wang . A model of hepatitis B virus with random interference infection rate. Mathematical Biosciences and Engineering, 2021, 18(6): 8257-8297. doi: 10.3934/mbe.2021410
    [6] Yanmei Wang, Guirong Liu . Dynamics analysis of a stochastic SIRS epidemic model with nonlinear incidence rate and transfer from infectious to susceptible. Mathematical Biosciences and Engineering, 2019, 16(5): 6047-6070. doi: 10.3934/mbe.2019303
    [7] Maysaa Al Qurashi, Saima Rashid, Fahd Jarad . A computational study of a stochastic fractal-fractional hepatitis B virus infection incorporating delayed immune reactions via the exponential decay. Mathematical Biosciences and Engineering, 2022, 19(12): 12950-12980. doi: 10.3934/mbe.2022605
    [8] Tingting Xue, Xiaolin Fan, Zhiguo Chang . Dynamics of a stochastic SIRS epidemic model with standard incidence and vaccination. Mathematical Biosciences and Engineering, 2022, 19(10): 10618-10636. doi: 10.3934/mbe.2022496
    [9] Suqi Ma . Low viral persistence of an immunological model. Mathematical Biosciences and Engineering, 2012, 9(4): 809-817. doi: 10.3934/mbe.2012.9.809
    [10] Zhongwei Cao, Jian Zhang, Huishuang Su, Li Zu . Threshold dynamics of a stochastic general SIRS epidemic model with migration. Mathematical Biosciences and Engineering, 2023, 20(6): 11212-11237. doi: 10.3934/mbe.2023497
  • New stochastic and deterministic Hepatitis B epidemic models with general incidence are established to study the dynamics of Hepatitis B virus (HBV) epidemic transmission. Optimal control strategies are developed to control the spread of HBV in the population. In this regard, we first calculate the basic reproduction number and the equilibrium points of the deterministic Hepatitis B model. And then the local asymptotic stability at the equilibrium point is studied. Secondly, the basic reproduction number of the stochastic Hepatitis B model is calculated. Appropriate Lyapunov functions are constructed, and the unique global positive solution of the stochastic model is verified by Itô formula. By applying a series of stochastic inequalities and strong number theorems, the moment exponential stability, the extinction and persistence of HBV at the equilibrium point are obtained. Finally, using the optimal control theory, the optimal control strategy to eliminate the spread of HBV is developed. To reduce Hepatitis B infection rates and to promote vaccination rates, three control variables are used, for instance, isolation of patients, treatment of patients, and vaccine inoculation. For the purpose of verifying the rationality of our main theoretical conclusions, the Runge-Kutta method is applied to numerical simulation.



    Hepatitis B is a global health problem with a high incidence rate in developing countries. According to statistics, asymptomatic HBV carriers in the world exceed 280 million, and China accounts for about 130 million. HBV infects liver cells when it enters the body. Most Hepatitis infections are caused by viruses, infections, germs or addiction to ethyl alcohol and medicines. The transmission of HBV can take place in a variety of ways, for instance, transmission of blood, bodily fluid transmission and mother-to-child vertical transmission, and so on. Vaccination against Hepatitis B is the most basic measure for preventing and controlling the disease. Many mathematicians and biologists have studied the Hepatitis B epidemic. They built different mathematical models to analyze the dynamic behavior of HBV. Then stability theory, bifurcation phenomenon, analysis of sensitivity and optimal control strategies of the infectious disease models are studied. This not only helps to achieve a reduction in Hepatitis B transmission, but also helps to prevent Hepatitis B in daily life. For example, in [1,2], the authors established the epidemic models with bilinear incidence for Hepatitis B. Hepatitis B can be effectively controlled through optimal control strategies. In [3], a Hepatitis B transmission model is established as follows:

    {dSdt=Λα1SI1+α2I(μ0+ν)S,dIdt=α1SI1+α2I(μ0+μ1+γ)I,dRdt=γI+νSμ0R,S(0)0,I(0)0,R(0)0, (1.1)

    where α1 represents the transmission rate of HBV, α2 represents the saturation rate. The kinetic behavior of HBV model is discussed with stability theory. In [4,5], the authors established the fractional HBV model. By means of the fractional Routh-Hurwitz stability criterion, the global dynamic behavior of fractional Hepatitis B model is studied.

    Epidemics are strongly influenced by environmental changes. In the case of human diseases, since one person's contact with another is unpredictable, the prevalence and spread of infectious diseases is random. Therefore, it is necessary to incorporate random effects into mathematical models [6,7]. By doing so, we are able to develop a more reasonable model. Modeling using stochastic differential equations is a very suitable method in the study of epidemic dynamics. Many researchers use stochastic infectious disease models to investigate different diseases. Here we focus on the stochastic HBV models. Khan et al.[8] proposed a stochastic HBV epidemic model with bilinear incidence as follows:

    {dSdt=[ΛαSI(μ0+ν)S]dtηSIdB(t),dIdt=[αSI(μ0+μ1+γ)I]dt+ηSIdB(t),dRdt=(γI+νSμ0R)dt,S(0)0,I(0)0,R(0)0, (1.2)

    where α represents the bilinear incidence rate, η2>0 represents the intensity of white noise, B(t) represents the Brownian motion. Based on the theory of stochastic Lyapunov function, the dynamic behavior of Hepatitis B stochastic model is studied. Anwarud Din et al. [9,10,11] proposed the stochastic models of Hepatitis B with standard incidence. Wu et al. [12] established a stochastic delay model of Hepatitis B with bilinear incidence. Anwarud Din et al. [13] built a stochastic time-delay model of Hepatitis B with standard incidence. In [14], a new stochastic Hepatitis B epidemic model that includes white noise, Markov switching, and vaccination control was developed. The above studies concluded that high noise can guarantee the extinction of Hepatitis B.

    To sum up, most of the current research focus on HBV models with bilinear incidence, standard incidence and saturated incidence, and few studies on HBV models with environmental noise disturbance and general incidence. The general incidence rate is more realistic than the bilinear incidence rate, standard incidence rate and saturated incidence rate. This is the research motivation of this paper. New stochastic and deterministic models of Hepatitis B epidemic with general incidence rate are established. The dynamic behavior of HBV model is studied, and the influence of environmental white noise on the epidemic of HBV is analyzed, and the optimal control strategy for eliminating HBV is developed. The research work in this paper is an extension of the work on [3,8,9,10,11]. The meanings of parameters in the models studied in this paper are as follows:

    Λ: the birth rate;

    β: infection rate from susceptible population to Hepatitis B;

    μ0: the natural mortality rate;

    μ1: the mortality rate from HBV;

    γ: the recovery rate of HBV;

    ν: the vaccination rate of HBV.

    Here is a breakdown of this article's organization. In Section 2, the new models of Hepatitis B are established. In Section 3, we obtain the basic reproduction number and the equilibrium points for the deterministic epidemic model of Hepatitis B. Lyapunov function is used to prove the local asymptotic stability. In Section 4, we verify that the stochastic model has one and only one global positive solution. The extinction, persistence and moment exponential stability of stochastic model are studied by means of stochastic Lyapunov function theory. In section 5, the optimal control strategy to eliminate HBV is developed by using the optimal control theory. To reduce Hepatitis B infection rates and to promote vaccination rates, three control variables are used, for instance, isolation of patients, treatment of patients, and vaccine inoculation. In Section 6, Runge-Kutta method is used for numerical simulation to support our main theoretical conclusions. Section 7 provides a brief summary and outlook of the main findings.

    New deterministic and stochastic mathematical models of HBV transmission are established. We make the following assumptions about the models:

    (A1). N(t) represents the total population at time t, which is divided into three parts: susceptible persons S(t), infected persons I(t) and convalescent patients R(t). Namely, N(t)=S(t)+I(t)+R(t).

    (A2). All parameter values of the models are non-negative.

    (A3). The incidence is set as nonlinear incidence rate.

    (A4). Once successfully vaccinated or cured by treatment, immunity is considered permanent.

    The supposed conditions (A1)(A4) lead to Hepatitis B epidemic model as below:

    {dSdt=ΛβSIf(S,I)(μ0+ν)S,dIdt=βSIf(S,I)(μ0+μ1+γ)I,dRdt=γI+νSμ0R, (2.1)

    with S(0)=S00,I(0)=I00,R(0)=R00, where β represents the transmission rate of Hepatitis B, f(S,I)=1+a1S+a2I+a3SI. For convenience, we define g(S,I)=βSIf(S,I)=βSI1+a1S+a2I+a3SI. Here g(S,I) is the incidence rate, where a1,a2,a30. Morbidity is the number of new infections per population in a given time period. To simulate the spread of disease, some scholars use bilinear incidence (see [15,16,17]), standard incidence (see [18,19,20]) and saturated incidence (see [21,22]). However, the nonlinear morbidity g(S,I) takes many forms, each of which has its own advantages, as below:

    (i) When a1=a2=a3=0, g(S,I) represents the bilinear incidence;

    (ii) When a1=a3=0 or a2=a3=0, g(S,I) is the saturation incidence[3];

    (iii) When a3=0, g(S,I) is the Beddington-DeAngelis functional response as shown in [23];

    (iv) When a3=a1a2, g(S,I) is the Crowley-Martin functional response as shown in [24].

    On the other hand, the following stochastic epidemic model of Hepatitis B is studied by incorporating environmental noise into the above model (2.1):

    {dS=(ΛβSIf(S,I)(μ0+ν)S)dtσSIf(S,I)dB(t),dI=(βSIf(S,I)(μ0+μ1+γ)I)dt+σSIf(S,I)dB(t),dR=(γI+νSμ0R)dt, (2.2)

    with S(0)=S00,I(0)=I00,R(0)=R00, where B(t) represents the Brownian motion, σ>0 represents the white noise intensity.

    Let dSdt=0, dIdt=0, dRdt=0, then E0=(S0,I0,R0)=(Λμ0+ν,0,νΛ(μ0+ν)μ0), for I0=0. The basic reproduction number of Hepatitis B model (2.1) is given below. Let X=(S,I,R), (2.1) is written as dXdt=F(X)V(X), where

    F(X)=(ΛβSIf(S,I)γI+νS),V(X)=(βSIf(S,I)+(μ0+ν)S(μ0+μ1+γ)Iμ0R). (3.1)

    The Jacobian of equation (3.1) around E0. Rd0 can be obtained by calculating the spectral radius of FV1. Then, Rd0=βΛ(μ0+ν+a1Λ)(μ0+μ1+γ).

    Theorem 3.1 E0 is locally asymptotically stable if Rd0<1, otherwise E0 is unstable.

    Proof. The Jacobian matrix of Hepatitis B epidemic model (2.1) at E0 can be calculated as follows:

    J|E0=((μ0+ν)βΛμ0+ν+a1Λ00βΛμ0+ν+a1Λ(μ0+μ1+γ)0νγμ0).

    After calculation, J|E0 has three eigenvalues:

    λ1=(μ0+ν)<0,λ2=μ0<0,λ3=(μ0+μ1+γ)(1Rd0).

    Obviously, the sign of λ3 depends on Rd0. If Rd0<1, then the eigenvalues of J|E0 are all negative. Thus, HBV model (2.1) is locally asymptotically stable at E0. On the contrary, if Rd0>1, then λ3 is positive, so E0 is unstable.

    By simple calculation, we have E=(S,I,R), for I0, where t1=μ0+μ1+γ,t2=μ0+ν,

    S=Λt1It2,I=Δ1+Δ214t21a3Δ22t21a3,R=γI+νSμ0,Δ1=t21a1Λt1a3t1t2a2βt1,Δ2=βΛt1t2Λt1a1.

    Lemma 3.1 If Rd0>1, then E exists, otherwise it does not exist.

    Proof. If Rd0>1, we have

    Δ1=t21a1Λt1a3t1t2a2βt1<(μ0+μ1+γ)[(μ0+ν)(μ0+μ1+γΛ+a2)Λa3]<0,Δ2=βΛt1t2Λt1a1=(μ0+ν+a1Λ)(μ0+μ1+γ)(Rd01)>0.

    And by calculation, we get Δ214t21a3Δ2>0. Thus, I>0. Because 2Λt1a3+Δ1>Δ214t21a3Δ2, so ΛΔ1+Δ214t21a3Δ22t1a3>0. Then, Λt1I>0. Hence, we have S>0, thereby R>0. Consequently, E exists, if Rd0>1.

    Theorem 3.2 E is locally asymptotically stable if Rd0>1, otherwise E is unstable.

    Proof. The Jacobian matrix of Hepatitis B epidemic model (2.1) at E can be calculated as follows:

    J|E=((μ0+ν)βI(1+a2I)f2(S,I)βS(1+a1S)f2(S,I)0βI(1+a2I)f2(S,I)βS(1+a1S)f2(S,I)(μ0+μ1+γ)0νγμ0).

    By calculation, the first eigenvalue of J|E is λ1=μ0<0. Take

    A=((μ0+ν)βI(1+a2I)f2(S,I)βS(1+a1S)f2(S,I)βI(1+a2I)f2(S,I)βS(1+a1S)f2(S,I)(μ0+μ1+γ)).

    By dIdt=0, one has βSf(S,I)=(μ0+μ1+γ). Thus, βS=(μ0+μ1+γ)f(S,I). By the definition of f, we get βS(1+a1S)<(μ0+μ1+γ)f2(S,I). So,

    βS(1+a1S)f2(S,I)<μ0+μ1+γ. (3.2)

    Then, by (3.2), we obtain

    trac(A)=[(μ0+ν)βI(1+a2I)f2(S,I)]+[βS(1+a1S)f2(S,I)(μ0+μ1+γ)]<0

    and

    det(A)=((μ0+ν)+βI(1+a2I)f2(S,I))((μ0+μ1+γ)βS(1+a1S)f2(S,I))+βS(1+a1S)f2(S,I)βI(1+a2I)f2(S,I)>0.

    Thus, trac(A)<0,det(A)>0 if Rd0>1. According to Routh-Hurwitz criterion, the two eigenvalues of the matrix A are negative. It can therefore be concluded that E is locally asymptotically stable when Rd0>1.

    Example 3.1 In model (2.1), let Λ=0.8,β=0.01,a1=0.1,a2=0.2,a3=1.2,μ0=0.02,μ1=0.1,ν=0.01,γ=0.01, S(0)=200, I(0)=100, R(0)=100. After the calculation, we get E0=(27,0,13) and Rd0=0.56<1. By observing Figure 1(a), it can be concluded that E0 is locally asymptotically stable, which verifies the rationality of Theorem 3.1.

    Figure 1.  Simulations of (S(t),I(t),R(t)) in the deterministic model (2.1).

    Example 3.2 The values of the parameters except β=0.9 are the same as those in Example 3.1. After calculation, we get E=(5,5,10) and Rd0=50>1. By observing Figure 1(b), it can be concluded that E is locally asymptotically stable, which verifies the rationality of Theorem 3.2.

    Let (Ω,F,P) be one complete probability space whose filtration {Ft}t0 satisfies the usual conditions (i.e., {Ft}t0 is monotonically increasing and right-continuous while F0 contains all P-null sets). We think over the following n-dimensional stochastic differential equation:

    dx(t)=f(x(t),t)dt+g(x(t),t)dB(t),tt0, (4.1)

    with x(0)=x0Rn, where f(x(t),t):Rn×[0,+]Rn and g(x(t),t):Rn×[0,+]Rn×m are locally Lipschitz functions in x. B(t) represents the n-dimensional Brownian motion defined on (Ω,F,{Ft}t0,P). C2,1(Rn×[0,+],R+) is a family of all nonnegative functions V(x,t) defined on Rn×[0,+], making them continuously differentiable twice in x and once in t. The differential operator L associated with (4.1) is defined [25] by:

    L=t+ni=1fi(x,t)xi+12ni,j=1[gT(x,t)g(x,t)]ij2xixj. (4.2)

    If L acts on VC2,1(Rn×[t0,],R+), then

    LV(x,t)=Vt(x,t)+Vx(x,t)f(x,t)+12trace[gT(x,t)Vxx(x,t)g(x,t)]. (4.3)

    If x(t)Rn, by Itô's formula, one has

    dV(x,t)=LV(x,t)dt+Vx(x,t)g(x,t)dB(t). (4.4)

    Definition 4.1 [26] The equilibrium point x=0 of (4.1) is considered to be pth moment exponentially stable, if there are C1,C2>0 so that

    E(|x(x0,t)|p)C1|x0|peC2t,x0Rn,t0. (4.5)

    A bounded set Δ is defined as below:

    Δ:={x=(x1,x2,x3):x1>0,x2>0,x3>0,x1+x2+x3<Λμ0a.s.}.

    Theorem 4.1 For (S(0),I(0),R(0))Δ, model (2.2) possesses one unique positive solution (S(t),I(t),R(t)) on t0. The solution is still in R3+ with probability 1, namely, (S(t),I(t),R(t))R3+ on t0 a.s.

    Proof. See Appendix A.

    Corollary 4.1 The set Δ is positively invariant; in other words, if (S(0),I(0),R(0))Δ, then P((S(t),I(t),R(t))Δ)=1, t0.

    Lemma 4.1 [26] If there is the function V(t,x)C1,2(R×Rn) so that

    K1|x|pV(t,x)K2|x|p,LV(t,x)K3|x|p,t0,p>0,Ki>0,i=1,2,3, (4.6)

    then the equilibrium point of the Eq (4.1) is pth moment exponentially stable. This implies that the number of infected people goes extinct at an exponential rate. When p=2, it usually means that it is exponentially stable at the mean square, and global asymptotically stable at the equilibrium point x=0.

    Lemma 4.2 [27] Let p2, ε,x,y>0, then the following two inequalities are true

    xp1y(p1)εpxp+1pεp1yp,xp2y2(p2)εpxp+2pε(p2)/2yp. (4.7)

    Theorem 4.2 Let p2. If Rd0<1 and

    σ2<2[(μ0+μ1+γ)(μ0+a1Λ)βΛ](μ0+a1Λ)(p1)Λ2, (4.8)

    then E0 is pth moment exponentially stable in Δ.

    Proof. See Appendix B.

    Corollary 4.2 If Rd0<1 and

    σ2<2[(μ0+μ1+γ)(μ0+a1Λ)βΛ](μ0+a1Λ)Λ2, (4.9)

    then E0 is globally asymptotically stable in Δ.

    In studying Hepatitis B model, we are interested in when the disease becomes extinct and when it becomes persistent in the population. This section establishes sufficient conditions for disease elimination in model (2.2). First, we need to give the basic symbols and lemmas associated with this problem. For the integrable function x(t) defined on (0,), is denoted as x(t)=1tt0x(s)ds. The threshold Rs0 of model (2.2) is as follows

    Rs0=βΛμ0+a1Λσ2Λ22(μ0+a1Λ)2μ0+μ1+γ=μ0+ν+a1Λμ0+a1ΛRd0σ2Λ22(μ0+a1Λ)2(μ0+μ1+γ).

    Theorem 4.3 Let (S(t),I(t),R(t)) be the solution of model (2.2) with (S(0),I(0),R(0))Δ. Suppose (i) σ2>β22(μ0+μ1+γ); or (ii) Rs0<1, σ22β(μ0+a1Λ)Λ. Then

    limsuptlnI(t)tβ22σ2(μ0+μ1+γ)<0a.s.if(i)holds, (4.10)
    limsuptlnI(t)t(Rs01)(μ0+μ1+γ)<0a.s.if(ii)holds. (4.11)

    In other words, I(t) approaches zero at an exponential rate a.s., namely, extinction is a certainty.

    Proof. See Appendix C.

    The conditions that allow the disease to persist are discussed in this section. Now, let's start with the relevant knowledge.

    Definition 4.2 If liminftI(t)>0a.s., then the disease is persistence in the mean.

    Lemma 4.3 [28] Let FC([0,)×Ω,R), fC([0,)×Ω,(0,)) such that limtF(t)t=0a.s. Assume there are λ0,λ>0 so that for all t0,

    lnf(t)λtλ0t0f(τ)dτ+F(t)a.s.

    Thus,

    lim inftf(t)λλ0a.s.

    Theorem 4.4 If Rs0>1, then I(t) is persistence in the mean, i.e.,

    lim inftI(t)(μ0+a1Λ)(μ0+μ1+γ)(Rs01)β[μ0+μ1+Λ(a2+a3Λμ0)]>0a.s.

    Proof. See Appendix D.

    Example 4.1 In model (2.2), let β=0.01, σ=0.4, S(0)=50, I(0)=20, R(0)=20. For other parameter values, see Example 3.1. Through calculation, we can get σ2=0.16<2[(μ0+μ1+γ)(μ0+a1Λ)βΛ](μ0+a1Λ)Λ2=0.38125,Rd0=0.56<1. Then the conditions of Corollary 4.2 are verified. Figure 2(a) illustrates that E0 is globally asymptotically stable, which verifies Corollary 4.2.

    Figure 2.  Simulations of (S(t),I(t),R(t)) for HBV stochastic model (2.2).

    Example 4.2 In model (2.2), let β=0.4, σ=0.8. For other parameter values, see Example 3.1. Through calculation, we can get σ2=0.64>β22(μ0+μ1+γ)0.6154. Then the condition (i) of Theorem 4.3 is verified. By Theorem 4.3, one has limsuptlnI(t)tβ22σ2(μ0+μ1+γ)=0.005<0a.s. As a result, I(t) approaches 0 at an exponential rate with probability 1. In other words, the disease has disappeared. As a confirmation of our findings, we present simulations based on the Euler-Maruyama (EM) method, as shown in Figure 2(b).

    Example 4.3 In model (2.2), let β=0.1, σ=0.15. For other parameter values, see Example 3.1. Through calculation, we can get Rs00.615<1, σ2=0.02252β(μ0+a1Λ)Λ=0.025. Then the condition (ii) of Theorem 4.3 is verified. By Theorem 4.3, one has limsuptlnI(t)t(Rs01)(μ0+μ1+γ)0.05005<0a.s. As a result, I(t) approaches 0 at an exponential rate with probability 1. In other words, the disease has disappeared. The simulation result is shown in Figure 2(c).

    Example 4.4 In model (2.2), let β=0.8, σ=0.4. For other parameter values, see Example 3.1. Through calculation, we can get Rs09.8>1. Then Theorem 4.4 is verified. By Theorem 4.4, one has

    lim inftI(t)(μ0+a1Λ)(μ0+μ1+γ)(Rs01)β[μ0+μ1+Λ(a2+a3Λμ0)]0.003697>0a.s.

    Thus, the disease persists. The simulation result is shown in Figure 2(d).

    In order to control the spread of HBV, this section adopts the optimal control theory[29,30,31]. Our aim is to seek an effective control strategy to reduce HBV infection in the population. In order to reduce Hepatitis B infection rates and to promote vaccination rates, three control variables are used, for instance, u1(t), u2(t) and u3(t). The specific explanation is as below:

    1) u1(t) is the isolation rate. Through this control variable, infected persons are isolated to avoid contact between infected persons and susceptible persons;

    2) u2(t) represents the cure rate. Through this control variable, the number of patients can be decreased by using effective drugs to treat the infected persons;

    3) u3(t) represents vaccination rate. The spread of Hepatitis B can be reduced through vaccination.

    To design a control strategy to eliminate Hepatitis B, we will consider the optimal strategy of deterministic model (2.1). The control strategy is achieved by minimizing the target function as below

    J(u1,u2,u3)=T0[ω1I(t)+12(ω2u21(t)+ω3u22(t)+ω4u23(t))]dt, (5.1)

    subjugate to the control model

    dSdt=ΛβSIf(S,I)(1u1(t))(μ0+ν+u3(t))S,dIdt=βSIf(S,I)(1u1(t))(μ0+μ1+γ+u2(t)+u3(t))I,dRdt=(γ+u2(t)+u3(t))I+(ν+u3(t))Sμ0R, (5.2)

    with

    S(0)0,I(0)0,R(0)0. (5.3)

    In Eq (5.1), ω1,ω2,ω3,ω4>0. In the target function, ω1 represents the weight constant of Hepatitis B infection I(t). ω2, ω3 and ω4 represent the weight constants of quarantine of infected persons and susceptible persons, infected persons' treatment and vaccination, respectively. 12ω2u21(t), 12ω3u22(t) and 12ω4u23(t) represent costs related to segregation, treatment and vaccine inoculation, respectively. The objective of this section is to seek u1,u2,u3 so that

    J(u1,u2,u3)=min{J(u1,u2,u3),u1,u2,u3U} (5.4)

    subordinate to problems (5.2) and (5.3), where

    U:={(u1,u2,u3)|0ui1,ui(t)isLebesguemeasurableon[0,T],i=1,2,3}. (5.5)

    As a result of this part, it will be demonstrated that the control problems (5.2) and (5.3) have a solution. Let

    dXdt=L1X+L2(X), (5.6)

    where

    X=(S(t)I(t)R(t)),L1=((μ0+ν+u3(t))000(μ0+μ1+γ+u2(t)+u3(t))0ν+u3(t)γ+u2(t)+u3(t)μ0),L2(X)=(ΛβSIf(S,I)(1u1(t))βSIf(S,I)(1u1(t))0). (5.7)

    It is obvious that Eq (5.6) is a nonlinear system with bounded coefficients. Let

    F(X)=L1X+L3(X), (5.8)

    which satisfies

    |L3(X1)L3(X2)|k1|S1(t)S2(t)|+k2|I1(t)I2(t)|+k3|R1(t)R2(t)|K1(|S1(t)S2(t)|+|I1(t)I2(t)|+|R1(t)R2(t)|), (5.9)

    where K1=max{ki},i=1,2,3 is not affected by state parameters in system (5.2). And we could write it the same way for this

    |F(X1)F(X2)|K2|X1X2|, (5.10)

    where K2=max{K1,L1}<, which implies that F is continuous and uniformly Lipschitz. It goes without saying that state variables and control variables can't be negative. Thus, the solution of model (5.2) exists. As a next step, we will determine the control variables for minimizing the objective function.

    Theorem 5.1 There is an optimal control u=(u1,u2,u3)U, so that

    J(u1,u2,u3)=minJ(u1,u2,u3) (5.11)

    subjugate to systems (5.2) and (5.3).

    Proof. Our method of demonstrating optimal control is the one proposed in [31,32]. Because the state parameters and the control variables are both positive. Therefore, in the minimization problem, the necessary convexity in u1(t), u2(t) and u3(t) of the objective functional defined in Eq (5.1) is satisfied. As defined, u1, u2, u3U are specified as enclosed and convex variables. In order for optimal control to exist, the optimal system (5.2) must have a bound, which ensures its compactness. In addition, the function is composed of control variables and state variables, which shows the convexity of the objective function. Therefore, the problem under consideration satisfies all assumptions, so (u1,u2,u3) exists.

    Define u=(u1,u2,u3), x=(S,I,R). Then the definition of Lagrangian L is as below:

    L(x,u)=ω1I(t)+12(ω2u21(t)+ω3u22(t)+ω4u23(t)). (5.12)

    And the related Hamiltonian H is defined as below:

    H(x,u,λ)=λg(x,u)+L(x,u), (5.13)

    where

    g(x,u)=(g1(x,u),g2(x,u),g3(x,u)),λ=(λ1,λ2,λ3),

    with

    g1(x,u)=ΛβSIf(S,I)(1u1(t))(μ0+ν+u3(t))S,g2(x,u)=βSIf(S,I)(1u1(t))(μ0+μ1+γ+u2(t)+u3(t))I,g3(x,u)=(γ+u2(t)+u3(t))I+(ν+u3(t))Sμ0R. (5.14)

    Secondly, the main research tool for optimal solution of control problem is the standard Pontryagin maximum principle. Suppose that (x,u) is the optimal solution of (5.1)–(5.3), then there is one nontrivial vector function λ, so that

    {dx(t)dt=Hλ(x,u,λ),0=Hu(x,u,λ),dλ(t)dt=Hx(x,u,λ), (5.15)

    with

    H(x,u,λ)=maxu[0,1]H(x(t),u(t),λ(t)), (5.16)

    and the transversal condition

    λ(T)=0. (5.17)

    Theorem 5.2 For the optimal control problems (5.1) and (5.2), S,I and R are the optimal state solutions of the optimal control variables (u1,u2,u3). Then there are adjoint variables λ1(t),λ2(t) and λ3(t), so that

    λ1(t)=(λ1(t)λ2(t))βI(1+a2I)(1u1(t))f2(S,I)+(λ1(t)λ3(t))u3(t)+λ1(t)(μ0+ν)λ3(t)ν,λ2(t)=ω1+(λ1(t)λ2(t))βS(1+a1S)(1u1(t))f2(S,I)+(λ2(t)λ3(t))(u2(t)+u3(t))+λ2(t)(μ0+μ1+γ)λ3(t)γ,λ3(t)=λ3(t)μ0 (5.18)

    with transversality conditions

    λi(T)=0,i=1,2,3. (5.19)

    And the specific forms of u1(t),u2(t) and u3(t) are as below

    u1(t)=max{min{(λ2(t)λ1(t))βSIf(S,I)ω2,1},0}, (5.20)
    u2(t)=max{min{(λ2(t)λ3(t))Iω3,1},0}, (5.21)
    u3(t)=max{min{(λ1(t)λ3(t))S+(λ2(t)λ3(t))Iω4,1},0}. (5.22)

    Proof. Taking the partial derivatives of S(t), I(t), R(t) in (5.13) yield the adjoint variables (5.18). In addition, to calculate u1,u2,u3, we take partial derivatives of u1, u2, u3 in (5.13). Then we obtain the optimal control variables (5.20)–(5.22).

    By giving the equation of the adjoint variables (5.18) and its related conditions (5.3) and (5.19) and the optimal control parameters (5.20)–(5.22), control variables and state variables of the control problem are solved. The next step is to investigate the control theory of stochastic systems when the same control measures are applied.

    As mentioned in Section 5.1, in this section, we study the stochastic optimal control of model (2.2), and the following stochastic control system can be obtained

    dS=(ΛβSIf(S,I)(1u1(t))(μ0+ν+u3(t))S)dtσSIf(S,I)dB(t),dI=(βSIf(S,I)(1u1(t))(μ0+μ1+γ+u2(t)+u3(t))I)dt+σSIf(S,I)dB(t),dR=[(γ+u2(t)+u3(t))I+(ν+u3(t))Sμ0R]dt (5.23)

    with

    S(0)0,I(0)0,R(0)0. (5.24)

    For easy reading, the vectors are given as below:

    x(t)=(S(t),I(t),R(t)),u(t)=(u1(t),u2(t),u3(t)),g(x(t))=(g1(t),g2(t),g3(t))f(x(t),u(t))=(f1(x,u),f2(x,u),f3(x,u)), (5.25)

    and

    dx(t)=f(x(t),u(t))dt+g(x(t))dB(t) (5.26)

    with x(0)=(S(0),I(0),R(0))=x0, where

    f1(x(t),u(t))=ΛβSIf(S,I)(1u1(t))(μ0+ν+u3(t))S,f2(x(t),u(t))=βSIf(S,I)(1u1(t))(μ0+μ1+γ+u2(t)+u3(t))I,f3(x(t),u(t))=(γ+u2(t)+u3(t))I+(ν+u3(t))Sμ0R,g1(t)=σSIf(S,I),g2(t)=σSIf(S,I),g3(t)=0. (5.27)

    The quadratic function is considered as below:

    J(x,u)=12E{tf0[C1I+12(C2u21+C3u22+C4u23)]dt+C52S2+C62I2+C72R2}, (5.28)

    where Ci>0,i=¯1,7. The objective of this section is to obtain the control vector u(t)=(u1(t),u2(t),u3(t)) so that

    J(u)J(u),uU, (5.29)

    where the specific form of U is as below:

    U={ui(t):ui(t)[0,umaxi],uiL2[0,tf],t(0,tf],i=1,2,3}, (5.30)

    where umaxi>0. The next step is to apply the stochastic maximum principle to define the Hamiltonian H(x,u,p,q), which has the following form:

    H(x,u,p,q)=l(x,u)+g(x),q+f(x,u),p, (5.31)

    where q=(q1,q2,q3),p=(p1,p2,p3) are two different adjoint vectors, and , denotes Euclidean inner product space. As a result of applying the maximum principle, one has

    dx(t)=H(x,u,p,q)pdt+g(x(t))dB(t), (5.32)
    dp(t)=H(x,u,p,q)xdt+q(t)dB(t), (5.33)
    Hm(x,u,p,q)=minuUHm(x,u,p,q). (5.34)

    In this case, x(t) is one optimal path of x(t). It is observed that, at both initial and terminal conditions of Eqs (5.32) and (5.33) are

    x(0)=x0, (5.35)
    p(tf)=h(x(tf))x. (5.36)

    As shown in Eq (5.34), the optimal control x(t) is an operator of q(t), p(t) and x(t). Thus, it means that

    u(t)=Φ(x,p,q), (5.37)

    In this case, Φ can be computed by (5.34). Therefore, Equations (5.32) and (5.33) are written as the following equations

    dx(t)=H(x,u,p,q)pdt+g(x(t))dB(t), (5.38)
    dp(t)=H(x,u,p,q)xdt+q(t)dB(t). (5.39)

    So,

    H=C1I+12(C2u21+C3u22+C4u23)+C52S2+C62I2+C72R2+p1[ΛβSIf(S,I)(1u1(t))(μ0+ν+u3(t))S]+p2[βSIf(S,I)(1u1(t))(μ0+μ1+γ+u2(t)+u3(t))I]+p3[(γ+u2(t)+u3(t))I+(ν+u3(t))Sμ0R]σSIf(S,I)q1+σSIf(S,I)q2. (5.40)

    According to random maximum principle,

    dp(t)=H(x,u,p,q)xdt+q(t)dB(t). (5.41)

    We have

    p1(t)=(p1(t)p2(t))βI(1+a2I)(1u1(t))f2(S,I)+(p1(t)p3(t))u3(t)+p1(t)(μ0+ν)p3(t)ν+σI(1+a2I)f2(S,I)(q1q2),p2(t)=C1+(p1(t)p2(t))βS(1+a1S)(1u1(t))f2(S,I)+(p2(t)p3(t))(u2(t)+u3(t))+p2(t)(μ0+μ1+γ)p3(t)γ+σS(1+a1S)f2(S,I)(q1q2),p3(t)=p3(t)μ0. (5.42)

    An auxiliary initial conditions and end conditions are granted as below

    S(0)=˜S,I(0)=˜I,R(0)=˜R,p(tf)=h(x(tf))x, (5.43)
    h(S,I,R)=k12S2+k22I2+k32R2. (5.44)

    In this case, p1(tf)=k1S,p2(tf)=k2I,p3(tf)=k3R. In the Hamiltonian equation, by taking the derivatives of u1, u2, u3, we can figure out that

    u1(t)=max{min{(p2(t)p1(t))βSIf(S,I)C2,1},0}, (5.45)
    u2(t)=max{min{(p2(t)p3(t))IC3,1},0}, (5.46)
    u3(t)=max{min{(p1(t)p3(t))S+(p2(t)p3(t))IC4,1},0}. (5.47)

    In control theory, the desired objective is achieved by adjusting control variables. By substituting the control parameters into the system, the optimal goal of the power system can be obtained. The limits of control can be set according to Eq (5.27). We will establish the objective function by referring to the method of Eq (5.28). As a result of objective functions, there is a direct relationship between optimality and optimality. So special attention should be paid when selecting the objective function. Whenever the control objective function has multiple factors, the more important items should be given weight. Before applying the Pontryagin Maximum Rules[33], the existence and compactness of an optimal control are tested. The optimal control makes the goal function reach a maximum or minimum value at a certain point. Differential equations can be optimized to Hamiltonian at a certain point. Hamiltonian is defined as below:

    Hamiltonian=(integrandofthegoalfunctional)+(adjoint)(RHSofDifferentialsystem).

    Optimal control involves finding the necessary point u to maximize the Hamiltonian equation. We are able to obtain the adjoint system (5.43) by taking the derivative of H w.r.t. with respect to the state variable and substituting it with the final condition.

    Our analysis results are supported by approximate simulations of Hepatitis B models (2.1) and (2.2), respectively. Simulations can be performed from qualitative aspects. To test the rationality of the results, we use the stochastic Runge-Kutta method is adopted to simulate model (2.2), and the calculation model is obtained as below:

    Sk+1=Sk+(ΛβSkIkf(Sk,Ik)(μ0+ν)Sk)ΔtσSkIkf(Sk,Ik)Δtξk+σ2SkIk2f(Sk,Ik)(ξ2k1)Δt,Ik+1=Ik+(βSkIkf(Sk,Ik)(μ0+μ1+γ)Ik)Δt+σSkIkf(Sk,Ik)Δtξk+σ2SkIk2f(Sk,Ik)(ξ2k1)Δt,Rk+1=Rk+(γIk+νSkμ0Rk)Δt, (6.1)

    where σ>0 is the white noise value, ξk(k=¯1,n) is a standalone Gaussian stochastic variable with N(0,1) and the step length Δt>0.

    Next, the qualitative characteristics of deterministic and stochastic optimal controls are modeled. Firstly, Runge-Kutta iterative technique is used to simulate the deterministic model. The time interval is set as [0,100] in the positive direction, and the state system (5.2) and transverse condition (5.19) are solved by the prescribed method. Then the adjoint equation (5.18) with the same time interval is simulated by Runge-Kutta iterative method in the backward direction supported by transverse condition (5.19). Note that the values of other parameters except β are shown in Example 3.1. The results are shown in Figure 3(a). Figure 3(a) shows the dynamic curves of susceptible population, infected person and recovered person in HBV deterministic model (2.1) with and without optimal control value. A clear difference can be observed between the two conditions with and without control. The simulation results show that with the implementation of control measures, the number of the infected and susceptible population tend to decrease, while the number of recovered patients tend to increase.

    Figure 3.  Simulations of (S(t),I(t),R(t)) with and without controls for the deterministic and stochastic HBV models.

    Next, we simulate optimal control techniques for stochastic models. The stochastic Runge-Kutta iterative technique is applied to simulate the optimal control system. Considering the transverse condition, the optimal control strategy is realized by approximating the state and adjoint model. First, we apply the stochastic Runge-Kutta iterative method to calculate the state system (5.42). Secondly, under the transverse condition (5.43), the corresponding adjoint equation (5.42) of the system is obtained by using the reverse technique and the iterative technique of the state equation. The control is then modified by applying the convex combination of the control and the values from the characterizations (5.45)–(5.47). The algorithm is repeated over and over again, and the iteration is performed until the difference between the values obtained in two successive iterations is very small. Note that the values of other parameters except β,σ are shown in Example 3.1. The results are shown in Figure 3(b). Figure 3(b) respectively shows the dynamic curves of susceptible population, infected persons and recovered persons in the stochastic model (2.2) with and without optimal control values. Under the optimal control parameter values, the dynamic behavior before and after the optimal control is significantly different. The images show that through isolation, treatment and vaccination, the number of susceptible and infected persons is decreasing, but the number of recovered people is increasing.

    This paper investigates the transmission dynamics of HBV. The optimal control strategy is developed to control the transmission of HBV in the population. To this end, we first establish new HBV models with general incidence rate. We calculate the basic reproduction number, equilibrium points of deterministic Hepatitis B model to study the local asymptotic stability under certain conditions. Secondly, we calculate the random threshold. The random Lyapunov function theory is applied to verify that the model has one unique global positive solution. The extinction, persistence and stability of stochastic Hepatitis B model are given. These conditions are expressed as expressions containing the stochastic system parameters and the intensity of the noise term. It is clear that noise intensity has an important effect on disease transmission. To control the transmission of HBV, optimal control strategies are used to eliminate the transmission of HBV. In order to reduce Hepatitis B infection rates and to promote vaccination rates, three control variables are used, for instance, isolation of patients, treatment of patients, and vaccine inoculation. Runge-Kutta method is used for numerical simulations to support the theoretical results. It can be found that when the white noise is stronger, the extinction rate of the disease is higher. Disease is more persistent when white noise is lower in intensity. Virus dynamics-based stochastic epidemic models perform better in our study. A broad range of biomedical applications can be made from this theory, as it provides a solid foundation for studying similar diseases. An infection dynamics model based on stochastic delayed infection can, for example, be considered for studying the effects of incubation periods. Additionally, our research can be applied to analyze other epidemics, such as COIVD-19, tuberculosis, HIV and so on.

    This research is funded by the Natural Science Foundation of Xinjiang Uygur Autonomous Region (Grant No. 2022D01A246, 2021D01B35, 2021D01A65), Natural Science Foundation of Colleges and Universities in Xinjiang Uygur Autonomous Region (Grant No. XJEDU2021Y048) and Doctoral Initiation Fund of Xinjiang Institute of Engineering (Grant No. 2020xgy012302).

    The authors declare there is no conflict of interest.

    Let (S(0),I(0),R(0))Δ, N(t)=S(t)+I(t)+R(t). By model (2.2), one has

    dN(t)=[Λμ0N(t)μ1I(t)]dt. (7.1)

    Then,

    dN(τ)<[Λμ0N(τ)]dτ,0τta.s. (7.2)

    By integration, we get

    N(τ)<Λμ0+(N(0)Λμ0)eμ0τ,0τta.s. (7.3)

    So N(τ)<Λμ0,

    S(τ),I(τ),R(τ)(0,Λμ0),0τta.s. (7.4)

    There is no doubt that model (2.2) meets the local Lipschitz condition. For (S(0),I(0),R(0))R3+, in this model (2.2), there is a unique local solution (S(t),I(t),R(t)), t[0,τe). In this case, τe refers to the duration of the explosion. In order to prove that τe= a.s., we must do the following. Make k0>0 large enough so that (S(0),I(0),R(0))>k0. For each integer kk0, the stopping time is defined as below

    τk=inf{t[0,τe):S(t)korI(t)korR(t)k},
    τ0=limk0τk=inf{t[0,τe):S(t)0orI(t)0orR(t)0}.

    A C2-function V:R3+R+ is defined as below

    V(S,I,R)=ln(SΛμ0)ln(IΛμ0)ln(RΛμ0)=lnSIR+3lnΛμ0.

    By applying the formula Itô, we can get

    dV(S,I,R)=[ΛS(τ)+βI(τ)f(S(τ),I(τ))+(μ0+ν)+σ2I2(τ)2f2(S(τ),I(τ))]dτ+[βS(τ)f(S(τ),I(τ))+(μ0+μ1+γ)+σ2S2(τ)2f2(S(τ),I(τ))]dτ+[γI(τ)R(τ)νS(τ)R(τ)+μ0]dτ+σ(I(τ)S(τ))f(S(τ),I(τ))dB(τ)[3μ0+μ1+γ+ν+βI(τ)f(S(τ),I(τ))+σ2(S2(τ)+I2(τ))2f2(S(τ),I(τ))]dτ+σ(I(τ)S(τ))f(S(τ),I(τ))dB(τ),τ[0,tτk]. (7.5)

    For τ[0,tτk], one has

    S(τ)f(S(τ),I(τ))S(τ)1+a1S(τ)Λμ01+a1Λμ0Λμ0+a1Λ,I(τ)f(S(τ),I(τ))I(τ)1+a1S(τ)Λμ01+a1Λμ0Λμ0+a1Λ. (7.6)

    Thus,

    dV(S,I,R)Kdτ+σ(I(τ)S(τ))f(S(τ),I(τ))dB(τ)a.s., (7.7)

    where K=3μ0+μ1+γ+ν+βΛμ0+a1Λ+σ2Λ2(μ0+a1Λ)2. Integrate the above inequality from 0 to τkt, and then taking the expectation, according to the properties of Brownian motion, we get

    EV(S(τkt),I(τkt),R(τkt))V(S(0),I(0),R(0))+Eτkt0KdtV(S(0),I(0),R(0))+Kt<.

    Because V(S(τkt),I(τkt),R(τkt))>0, so

    EV(S(τkt),I(τkt),R(τkt))=E[1{τkt}V(S(τkt),I(τkt),R(τkt))]+E[1{τk>t}V(S(τkt),I(τkt),R(τkt))]E[1{τkt}V(S(τkt),I(τkt),R(τkt))].

    For τk, some component of S(τk),I(τk),R(τk) is equal to k. Thus, V(S(τk),I(τk),R(τk))ln(kμ0Λ). So,

    EV(S(τkt),I(τkt),R(τkt))E[1{τkt}V(S(τkt),I(τkt),R(τkt))]ln(kμ0Λ)P(τkt). (7.8)

    By (7.8), we obtain

    P(τkt)V(S(0),I(0),R(0))+Ktln(kμ0Λ). (7.9)

    Extending k to 0, one has P(τ0t)=0, t>0. Therefore, P(τ0=)=1. Consequently, τ0=τe=a.s.

    Let p2. The Lyapunov function is considered as below

    V=τ1(Λμ0+νS)p+Ipp, (7.10)

    Here τ1>0 will be determined later. The first inequality in (4.6) is easily proved to be true. Then,

    LV=pτ1(Λμ0+νS)p(μ0+ν)+(Λμ0+νS)p1pτ1βSIf(S,I)+[βSf(S,I)(μ0+μ1+γ)]Ip+τ1p(p1)σ2S2I22f2(S,I)×(Λμ0+νS)p2+(p1)σ2S2Ip2f2(S,I). (7.11)

    In Δ, we get

    LVpτ1(Λμ0+νS)p(μ0+ν)+I(Λμ0+νS)p1pτ1βΛμ0+a1Λ[(μ0+μ1+γ)βΛμ0+a1Λ(p1)σ2Λ22(μ0+a1Λ)2]Ip+τ1p(p1)σ2Λ22(μ0+a1Λ)2I2(Λμ0+νS)p2 (7.12)

    According to Lemma 4.2, we can get

    I(Λμ0+νS)p1(p1)εp(Λμ0+νS)p+1pεp1Ip,I2(Λμ0+νS)p2(p2)εp(Λμ0+νS)p+2pε(p2)/2Ip. (7.13)

    Then

    LV(Λμ0+νS)p×[pτ1(μ0+ν)+βΛτ1ε(p1)μ0+a1Λ+(p2)(p1)ετ1σ2Λ22(μ0+a1Λ)2]+Ip×[τ1βΛ(μ0+a1Λ)εp1+τ1(p1)σ2Λ2(μ0+a1Λ)2ε(p2)/2]+Ip×[(μ0+μ1+γ)+βΛμ0+a1Λ+(p1)σ2Λ22(μ0+a1Λ)2]. (7.14)

    Select ε small enough to make the coefficient of (Λμ0+νS)p negative. By (4.8), we have (μ0+μ1+γ)+βΛμ0+a1Λ+(p1)σ2Λ22(μ0+a1Λ)2<0. When (μ0+μ1+γ)+βΛμ0+a1Λ+(p1)σ2Λ22(μ0+a1Λ)2<0, select τ1 is positive, so that the coefficient of Ip is negative.

    Applying Itô formula for the second equation of model (2.2), one has

    dlnI(t)=[βSf(S,I)(μ0+μ1+γ)σ2S22f2(S,I)]dt+σSf(S,I)dB(t). (7.15)

    Integrating the above equation from 0 to t, and dividing both sides by t at the same time, then we get

    lnI(t)=lnI(0)+t0[βS(τ)f(S(τ),I(τ))(μ0+μ1+γ)σ2S2(τ)2f2(S(τ),I(τ))]dτ+M(t), (7.16)

    where M(t):=t0σS(τ)f(S(τ),I(τ))dB(τ). Then

    lnI(t)t=lnI(0)t+1tt0[σ22(S(τ)f(S(τ),I(τ))βσ2)2(μ0+μ1+γ)+β22σ2]dτ+M(t)t. (7.17)

    The following formula can be obtained from the martingale theorem of large numbers

    lim suptM(t)t=0a.s. (7.18)

    If condition (i) is met, then by (7.17), (7.18), we have

    limsuptlnI(t)tβ22σ2(μ0+μ1+γ)<0a.s. (7.19)

    From the definition of f(S,I), we get

    Sf(S,I)S1+a1SΛμ0+a1Λ. (7.20)

    If condition (ii) is met, then by (7.20), one has

    lnI(t)t=lnI(0)t+1tt0[βS(τ)f(S(τ),I(τ))(μ0+μ1+γ)σ2S2(τ)2f2(S(τ),I(τ))]dτ+M(t)tlnI(0)t+1tt0[βS(τ)1+a1S(τ)(μ0+μ1+γ)σ2S2(τ)2(1+a1S(τ))2]dτ+M(t)tlnI(0)t+βΛμ0+a1Λ(μ0+μ1+γ)σ2Λ22(μ0+a1Λ)2+M(t)t(Rs01)(μ0+μ1+γ)+lnI(0)t+M(t)t. (7.21)

    If condition (ii) is met, then by (7.21), one has

    limsuptlnI(t)t(Rs01)(μ0+μ1+γ)<0a.s. (7.22)

    The above inequality indicates that

    limtI(t)=0a.s. (7.23)

    By Corollary 4.1, we have S(t)<Λμ0a.s. Then,

    βSf(S,I)=βΛμ0+a1Λβμ0μ0+a1ΛΛμ0Sf(S,I)βΛa2I(μ0+a1Λ)f(S,I)βΛa3SI(μ0+a1Λ)f(S,I)βΛμ0+a1Λβμ0μ0+a1Λ(Λμ0S)βΛ(μ0+a1Λ)(a2+a3Λμ0)I=βμ0Sμ0+a1ΛβΛ(μ0+a1Λ)(a2+a3Λμ0)I. (7.24)

    By model (2.2), one has

    d(S+I+R)=[Λμ0(S+I+R)μ1I]dt. (7.25)

    Integrating the above equation from 0 to t, and dividing both sides by t at the same time, then

    S(t)+I(t)+R(t)(S(0)+I(0)+R(0))t=Λμ0S(t)(μ0+μ1)I(t)μ0R(t). (7.26)

    Then

    S(t)=Λμ0μ0+μ1μ0I(t)R(t)φ(t), (7.27)

    here φ(t)=[S(t)+I(t)+R(t)(S(0)+I(0)+R(0))]/(μ0t). In this situation,

    limtφ(t)=0a.s. (7.28)

    Applying Itô formula for the second equation of model (2.2) and combining with (7.20), (7.24), we get

    dlnI(t)=[βSf(S,I)(μ0+μ1+γ)σ2S22f2(S,I)]dt+σSf(S,I)dB(t)[βμ0Sμ0+a1ΛβΛμ0+a1Λ(a2+a3Λμ0)I(μ0+μ1+γ)σ2Λ22(μ0+a1Λ)2]dt+σSf(S,I)dB(t). (7.29)

    Integrating the above equation from 0 to t, and combining with (7.27), then

    lnI(t)lnI(0)+βμ0tμ0+a1Λ(Λμ0μ0+μ1μ0I(t)R(t)φ(t))βΛμ0+a1Λ(a2+a3Λμ0)×t0I(τ)dτ[(μ0+μ1+γ)+σ2Λ22(μ0+a1Λ)2]t+t0σS(τ)f(S(τ),I(τ))dB(τ)[βΛμ0+a1Λ(μ0+μ1+γ)σ2Λ22(μ0+a1Λ)2]tβμ0+a1Λ[μ0+μ1+Λ(a2+a3Λμ0)]×t0I(τ)dτβμ0μ0+a1Λ(R(t)+φ(t))t+t0σS(τ)f(S(τ),I(τ))dB(τ)+lnI(0)(μ0+μ1+γ)(Rs01)tβμ0+a1Λ[μ0+μ1+Λ(a2+a3Λμ0)]×t0I(τ)dτ+Φ(t), (7.30)

    here Φ(t)=lnI(0)+t0σS(τ)f(S(τ),I(τ))dB(τ)βμ0μ0+a1Λ(R(t)+φ(t))t. By combining the martingale theorem of large numbers with (7.28), we can get

    limtΦ(t)t=0a.s. (7.31)

    Lemma 4.3 implies that

    liminftI(t)(μ0+μ1+γ)(μ0+a1Λ)(Rs01)β[μ0+μ1+Λ(a2+a3Λμ0)]>0a.s. (7.32)


    [1] T. Khan, G. Zaman, M. I. Chohan, The transmission dynamic and optimal control of acute and chronic Hepatitis B, J. Biol. Dyn., 11 (2017), 172–189. https://doi.org/10.1080/17513758.2016.1256441 doi: 10.1080/17513758.2016.1256441
    [2] A. Din, Y. J. Li, Q. Liu, Viral dynamics and control of Hepatitis B virus (HBV) using an epidemic model, Alex. Eng. J., 59 (2020), 667–679. https://doi.org/10.1016/j.aej.2020.01.034 doi: 10.1016/j.aej.2020.01.034
    [3] T. Khan, Z. Ullah, Z. Ali, G. Zaman, Modeling and control of the Hepatitis B virus spreading using an epidemic model, Chaos, Solitons Fractals, 124 (2019), 1–9. https://doi.org/10.1016/j.chaos.2019.04.033 doi: 10.1016/j.chaos.2019.04.033
    [4] P. T. Karaji, N. Nyamoradi, Analysis of a fractional SIR model with general incidence function, Appl. Math. Lett., 108 (2020), 106499. https://doi.org/10.1016/j.aml.2020.106499 doi: 10.1016/j.aml.2020.106499
    [5] S. M. Simelane, P. G. Dlamini, A fractional order differential equation model for Hepatitis B virus with saturated incidence, Results Phys., 24 (2021), 104114. https://doi.org/10.1016/j.rinp.2021.104114 doi: 10.1016/j.rinp.2021.104114
    [6] F. Huang, J. L. Li, Exponential ultimate boundedness and stability of stochastic differential equations with impulese, Asian J. Control, 25 (2023), 88–100. https://doi.org/10.1002/asjc.2786 doi: 10.1002/asjc.2786
    [7] F. Huang, J. L. Li, Exponential ultimate boundedness and stability of impulsive stochastic functional differential equations, Int. J. Control, 96 (2023), 568–576. https://doi.org/10.1080/00207179.2021.2005259 doi: 10.1080/00207179.2021.2005259
    [8] T. Khan, A. Khan, G. Zaman, The extinction and persistence of the stochastic Hepatitis B epidemic model, Chaos, Solitons Fractals, 108 (2018), 123–128. https://doi.org/10.1016/j.chaos.2018.01.036 doi: 10.1016/j.chaos.2018.01.036
    [9] A. Din, Y. J. Li, T. Khan, K. Anwar, G. Zaman, Stochastic dynamics of Hepatitis B epidemics, Results Phys., 20 (2021), 103730. https://doi.org/10.1016/j.rinp.2020.103730 doi: 10.1016/j.rinp.2020.103730
    [10] P. J. Liu, A. Din, L. F. Huang, A. Yusuf, Stochastic optimal control analysis for the Hepatitis B epidemic model, Results Phys., 26 (2021), 104372. https://doi.org/10.1016/j.rinp.2021.104372 doi: 10.1016/j.rinp.2021.104372
    [11] A. Din, Y. J. Li, Stationary distribution extinction and optimal control for the stochastic Hepatitis B epidemic model with partial immunity, Phys. Scr., 96 (2021), 74005. https://doi.org/10.1088/1402-4896/abfacc doi: 10.1088/1402-4896/abfacc
    [12] B. Wu, J. W. Jia, Asymptotic behavior of a stochastic delayed model for chronic Hepatitis B infection, Complexity, 2020 (2020), 1875475. https://doi.org/10.1155/2020/1875475 doi: 10.1155/2020/1875475
    [13] A. Din, Y. J. Li, A. Yusuf, Delayed Hepatitis B epidemic model with stochastic analysis, Chaos, Solitons Fractals, 146 (2021), 110839. https://doi.org/10.1016/j.chaos.2021.110839 doi: 10.1016/j.chaos.2021.110839
    [14] A. Din, Y. J. Li, Stochastic optimal analysis for the Hepatitis B epidemic model with Markovian switching, Math. Meth. Appl. Sci., 2022 (2022), 1–26. https://doi.org/10.1002/mma.8218 doi: 10.1002/mma.8218
    [15] 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
    [16] I. A. Baba, E. Hincal, Global stability analysis of two-strain epidemic model with bilinear and non-monotone incidence rates, Eur. Phys. J. Plus, 132 (2017), 208. https://doi.org/10.1140/epjp/i2017-11476-x doi: 10.1140/epjp/i2017-11476-x
    [17] J. J. Wang, J. Z. Zhang, Z. Jin, Analysis of an SIR model with bilinear incidence rate, Nonlinear Anal. Real World Appl., 11 (2010), 2390–2402. https://doi.org/10.1016/j.nonrwa.2009.07.012 doi: 10.1016/j.nonrwa.2009.07.012
    [18] T. T. Xue, X. L. Fan, J. Zhu, A class of deterministic and stochastic fractional epidemic models with vaccination, Comput. Math. Methods Med., 2022 (2022), 1–22. https://doi.org/10.1155/2022/1797258 doi: 10.1155/2022/1797258
    [19] Q. Liu, D. Q. Jiang, N. Z. Shi, Threshold behavior in a stochastic SIQR epidemic model with standard incidence and regime switching, Appl. Math. Comput., 316 (2018), 310–325. https://doi.org/10.1016/j.amc.2017.08.042 doi: 10.1016/j.amc.2017.08.042
    [20] T. T. Xue, X. L. Fan, Z. G. Chang, Dynamics of a stochastic SIRS epidemic model with standard incidence and vaccination, Math. Biosci. Eng., 19 (2022), 10618–10636. https://doi.org/10.3934/mbe.2022496 doi: 10.3934/mbe.2022496
    [21] Q. S. Yang, D. Q. Jiang, N. Z. Shi, C. Y. 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
    [22] Y. N. Zhao, D. Q. Jiang, The threshold of a stochastic SIRS epidemic model with saturated incidence, Appl. Math. Lett., 34 (2014), 90–93. https://doi.org/10.1016/j.aml.2013.11.002 doi: 10.1016/j.aml.2013.11.002
    [23] J. R. Beddington, Mutual interference between parasites or predat ors and its efect on searching efciency, J. Anim. Ecol., 44 (1975), 331–340.
    [24] P. Crowley, E. Martin, Functional responses and interference within and between year classes of a dragonfy population. J. N. Am. Benthol. Soc., 8 (1989), 211–221.
    [25] X. Mao, Stochastic Differential Equations and Applications, Horwood, Chichester, 2007.
    [26] V. N. Afanas'ev, V. B. Kolmanovskii, V. R. Nosov, Mathematical Teory of Control Systems Design, Springer, Dorderecht, The Netherlands, 1996.
    [27] D. Kiouach, Y. Sabbar, Stability and threshold of a stochastic SIRS epidemic model with vertical transmission and transfer from infectious to susceptible individuals, Discrete Dyn. Nat. Soc., 2018 (2018), 7570296. https://doi.org/10.1155/2018/7570296 doi: 10.1155/2018/7570296
    [28] C. Y. Ji, D. Q. Jiang, Threshold behavior 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] G. Zaman, Y. H. Kang, I. H. Jung, Optimal treatment of an SIR epidemic model with time delay, Biosystems, 98 (2009), 43–50. https://doi.org/10.1016/j.biosystems.2009.05.006 doi: 10.1016/j.biosystems.2009.05.006
    [30] M. T. Xia, L. Bottcher, T. Chou, Controlling epidemics through optimal allocation of test kits and vaccine doses across networks, IEEE Trans. Network Sci. Eng., 9 (2022), 1422–1436. https://doi.org/10.1109/TNSE.2022.3144624 doi: 10.1109/TNSE.2022.3144624
    [31] G. Zaman, Y. H. Kang, I. H. Jung, Stability analysis and optimal vaccination of an SIR epidemic model, Biosystems, 93 (2008), 240–249. https://doi.org/10.1016/j.biosystems.2008.05.004 doi: 10.1016/j.biosystems.2008.05.004
    [32] A. V. Kamyad, R. Akbari, A. A. Heydari, A. Heydari, Mathematical modeling of transmission dynamics and optimal control of vaccination and treatment for Hepatitis B virus, Comput. Math. Methods Med., 2014 (2014), 475451. https://doi.org/10.1155/2014/475451 doi: 10.1155/2014/475451
    [33] L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelize, E. F. Mishchenko, The Mathematical Theory of Optimal Processes, Wiley, New York, 1962.
  • This article has been cited by:

    1. Tingting Xue, Xiaolin Fan, Yan Xu, Kinetic Behavior and Optimal Control of a Fractional-Order Hepatitis B Model, 2023, 11, 2227-7390, 3642, 10.3390/math11173642
    2. Tahir Khan, Fathalla A. Rihan, Muhammad Ibrahim, Shuo Li, Atif M. Alamri, Salman A. AlQahtani, Modeling different infectious phases of hepatitis B with generalized saturated incidence: An analysis and control, 2024, 21, 1551-0018, 5207, 10.3934/mbe.2024230
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(2076) PDF downloads(155) Cited by(2)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog