Parameter | Parameter description |
Λ | The per capita constant birth rate |
β | The transmission rate |
μ0 | The natural death rate |
μ1 | The disease induced death rate |
v | The vaccination rate |
γ1 | The constant recovery rate |
Citation: Kangbo Bao, Qimin Zhang, Xining Li. A model of HBV infection with intervention strategies: dynamics analysis and numerical simulations[J]. Mathematical Biosciences and Engineering, 2019, 16(4): 2562-2586. doi: 10.3934/mbe.2019129
[1] | 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 |
[2] | 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 |
[3] | Xiaoqin Wang, Yiping Tan, Yongli Cai, Kaifa Wang, Weiming Wang . Dynamics of a stochastic HBV infection model with cell-to-cell transmission and immune response. Mathematical Biosciences and Engineering, 2021, 18(1): 616-642. doi: 10.3934/mbe.2021034 |
[4] | Ling Xue, Caterina Scoglio . Network-level reproduction number and extinction threshold for vector-borne diseases. Mathematical Biosciences and Engineering, 2015, 12(3): 565-584. doi: 10.3934/mbe.2015.12.565 |
[5] | 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 |
[6] | Fathalla A. Rihan, Hebatallah J. Alsakaji . Analysis of a stochastic HBV infection model with delayed immune response. Mathematical Biosciences and Engineering, 2021, 18(5): 5194-5220. doi: 10.3934/mbe.2021264 |
[7] | Helong Liu, Xinyu Song . Stationary distribution and extinction of a stochastic HIV/AIDS model with nonlinear incidence rate. Mathematical Biosciences and Engineering, 2024, 21(1): 1650-1671. doi: 10.3934/mbe.2024072 |
[8] | Junyoung Jang, Kihoon Jang, Hee-Dae Kwon, Jeehyun Lee . Feedback control of an HBV model based on ensemble kalman filter and differential evolution. Mathematical Biosciences and Engineering, 2018, 15(3): 667-691. doi: 10.3934/mbe.2018030 |
[9] | Yiping Tan, Yongli Cai, Zhihang Peng, Kaifa Wang, Ruoxia Yao, Weiming Wang . Dynamics of a stochastic HBV infection model with drug therapy and immune response. Mathematical Biosciences and Engineering, 2022, 19(8): 7570-7585. doi: 10.3934/mbe.2022356 |
[10] | 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 |
Outbreaks of infectious diseases have caused seriously endanger to human life and property safety, which is also a major global health problem. According to the World Health Organization (WHO) report [1]: an estimated 2 billion people worldwide carry hepatitis B virus, about 2 million people suffer from chronic liver infections, more than 780, 000 people die every year. How to prevent and control the spread of hepatitis B is one of the hot issues that people care about.
When an infectious disease appears and spreads, one of the measures taken by the disease management department is information intervention [2]. As an important non-drug control measure, intervention strategies (e.g., media coverage, health education) has attracted more and more attention [3,4,5]. It plays an important role in helping the government to formulate intervention measures to control diseases and reduce the infection rate of human [6,7]. For example, during the outbreak of H7N9 influenza in 2013 [8] and the outbreak of cholera virus in Tanzania [9], various information intervention strategies were used. These strategies told people the correct knowledge of disease prevention and greatly reduced the number of contacts per unit time, thus reduced the infection rate. Cui et al. [10,11] studied the impact of media coverage on the control of infectious diseases and reached the conclusion that media coverage is essential to eradicate the diseases. Therefore, it is of great significance to consider the influence of information intervention in preventing hepatitis B virus (HBV) transmission.
Mathematical models have been showed to be an important tool that helps understand the spread and control of infectious diseases. A large number of mathematical models have been developed to study the dynamics of hepatitis B [12,13,14]. A mathematical model was developed by Zou et al. [12] to study the transmission dynamics and prevalence of HBV infection in China. They investigated the existence and stability of equilibrium points, sensitivity analysis of the model parameters are also performed. Khan et al. [13] investigated the dynamics of acute and chronic hepatitis B epidemic problem by using a HBV transmission model. They proposed an optimal control strategy to control the spread of hepatitis B. These studies [12,13,14] used deterministic hepatitis B epidemic models.
For human diseases, due to the unpredictability of human-to-human contacts, the natural growth and spread of epidemic are essentially random [15]. Environmental variations are also important for the development of epidemic [16]. Stochastic epidemic model is more suitable for describing the effect of environmental fluctuations on the dynamics of disease [17,18,19,20,21,22,23]. There are very few stochastic hepatitis B epidemic models. Khan et al. [24] discussed the dynamics of disease by proposing a stochastic hepatitis B model with a varying population environment. They investigated the influence of noise intensity on the disease transmission and obtained the sufficient conditions for the extinction and persistence. However, to the best of our knowledge, very few studies, if any, have been done to consider the influence of information intervention into the above-mentioned hepatitis B models [12,13,14,24].
In this paper, we studies the dynamics of the stochastic hepatitis B epidemic model incorporating information intervention under environmental noise. By using the Markov semigroups theory, we find that the stochastic basic reproduction number can be used to govern the hepatitis B extinction or persistence. Our innovation points are as follows:
● The effect of information intervention is taken into account the stochastic hepatitis B model.
● By using the Markov semigroups theory, we show the stochastic hepatitis B model admits a stationary distribution.
The hepatitis B model and preliminaries will be introduced in Section 2. In Section 3, we will give our main results. The proofs of the main results in details will be provided in Section 4. In Section 5, some numerical simulations will be conducted to illustrate the influence of environmental noise and information intervention on the hepatitis B dynamics. Finally, we finish the paper with conclusions and future directions in the last section.
In this section, we will introduce the stochastic hepatitis B epidemic model incorporating information intervention. It is followed by some preliminaries.
In [24], Khan et al. proposed the following hepatitis B epidemic model:
{dS(t)=[Λ−βS(t)I(t)−(μ0+ν)S(t)]dt,dI(t)=[βS(t)I(t)−(μ0+μ1+γ1)I(t)]dt,dR(t)=[γ1I(t)+νS(t)−μ0R(t)]dt, | (2.1) |
where S(t), I(t) and R(t) represent susceptible, infective and recovered population, respectively. All parameters are assumed to be positive and the descriptions are listed in Table 1.
Parameter | Parameter description |
Λ | The per capita constant birth rate |
β | The transmission rate |
μ0 | The natural death rate |
μ1 | The disease induced death rate |
v | The vaccination rate |
γ1 | The constant recovery rate |
In this paper, we consider the effect of intervention strategies into the hepatitis B model (2.1). In order to do this, using a function β=β1−β2f(I) and the function f(I) satisfies the following assumption:
(A1)f(0)=0,f′(I)>0andlimI→∞f(I)=1, |
where β1 is the usual contact rate without considering the infectious individuals, and β2 is the maximum reduced contact rate due to the presence of the infected individuals. Then we obtain the following hepatitis B model with information intervention:
{dS(t)=[Λ−(β1−β2f(I))S(t)I(t)−(μ0+ν)S(t)]dt,dI(t)=[(β1−β2f(I))S(t)I(t)−(μ0+μ1+γ1)I(t)]dt,dR(t)=[γ1I(t)+νS(t)−μ0R(t)]dt. | (2.2) |
To consider the effect of environment noise, we suppose that the contact transmission coefficient β1 is stochastically perturbed, β1→β1+σ˙B(t). The hepatitis B model (2.2) becomes
{dS(t)=[Λ−(β1−β2f(I))S(t)I(t)−(μ0+ν)S(t)]dt−σS(t)I(t)dB(t),dI(t)=[(β1−β2f(I))S(t)I(t)−(μ0+μ1+γ1)I(t)]dt+σS(t)I(t)dB(t),dR(t)=[γ1I(t)+νS(t)−μ0R(t)]dt, | (2.3) |
where B(t) is the standard Brownian motion and σ2 is the intensity of the noise.
In this subsection, we introduce some definitions and results about the Markov semigroup and asymptotic properties [25,26,27,28,29,30,31].
Let (Ω,F,{Ft}t≥0,Prob) be a complete probability space with a filtration {Ft}t≥0 which meet the general conditions (i.e., it is right continuous and increasing while F0 contains all P-null sets).
Let Σ=B(X) be the σ-algebra of Borel subsets of X and m the Lebesgue measure on (X,Σ). D=D(X,Σ,m) denotes the subset of the space L1=L1(X,Σ,m) which contains all densities, i.e.,
D={g∈L1:g≥0,‖g‖=1}, |
where ‖⋅‖ is the norm in L1. If P(D)⊂D, then a linear mapping P:L1→L1 is called a Markov operator.
If there exists a measurable function k:X×X→[0,∞) such that
∫Xk(x,y)m(dx)=1, | (2.4) |
for almost every y∈X, then
Pg(x)=∫Xk(x,y)g(y)m(dy) |
is an integral Markov operator and the function k is called a kernel of the Markov operator P.
A family {P(t)}t≥0 of Markov operators which satisfies conditions:
(a) P(0)=Id;
(b) P(t+s)=P(t)P(s) for s,t≥0;
(c) for every g∈L1 the function t↦P(t)g is continuous.
is called a Markov semigroup.
A Markov semigroup {P(t)}t≥0 is called integral, if for every t>0, the operator P(t) is an integral Markov operator, then there exists a measurable function k:(0,∞)×X×X→[0,∞) such that
P(t)g(x)=∫Xk(t,x,y)g(y)m(dy) |
for every density g.
Lemma 2.1. [25] Let {P(t)}t≥0 be an integral Markov semigroup with a continuous kernel k(t,x,y) for t>0 which satisfies (2.4) for all y∈X. If for every g∈D
∫∞0P(t)g(x)dt>0, |
then this semigroup is asymptotically stable or is sweeping with respect to compact sets.
For any A∈Σ, we denote the transition probability function by P(t,x,y,z,A) for the diffusion process (S(t),I(t),R(t)), i.e.
P(t,x,y,z,A)=Prob{(S(t),I(t),R(t))∈A} |
with the initial condition (S(0),I(0),R(0))=(x,y,z). If (S(t),I(t),R(t)) is a solution of system (2.3) such that the distribution of (S(0),I(0),R(0)) is absolutely continuous with the density v(x,y,z). Then there also exists the density U(t,x,y,z) of (S(t),I(t),R(t)) which satisfies the Fokker-Planck equation [30,31]:
∂U∂t=12σ2(∂2(φU)∂x2−2∂2(φU)∂x∂y+∂2(φU)∂y2)−∂(f1U)∂x−∂(f2U)∂y−∂(f3U)∂z, | (2.5) |
where φ(x,y,z)=x2y2 and
f1(x,y,z)=Λ−(β1−β2f(y))xy−(μ0+ν)x,f2(x,y,z)=(β1−β2f(y))xy−(μ0+μ1+γ1)y,f3(x,y,z)=γ1y+νx−μ0z. | (2.6) |
Let P(t)V(x,y,z)=U(x,y,z,t) for V∈D. Due to the operator P(t) is a contraction on D, it can be extended to a contraction on L1. Then the operators {P(t)}t≥0 form a Markov semigroup. Denote A the infinitesimal generator of semigroup {P(t)}t≥0, i.e.
AV=12σ2(∂2(φV)∂x2−2∂2(φV)∂x∂y+∂2(φV)∂y2)−∂(f1V)∂x−∂(f2V)∂y−∂(f3V)∂z. |
The adjoint operator of A is of the form
A∗V=12σ2φ(∂2V∂x2−2∂2V∂x∂y+∂2V)∂y2)+∂(f1V)∂x+∂(f2V)∂y+∂(f3V)∂z. |
The basic reproduction number R0 is a threshold which represents how many secondary infections result from the introduction of one infected individual into a population of susceptible [32]. We can calculate the basic reproduction number R0 for the deterministic hepatitis B model (2.2), given by
R0=Λβ1(μ0+ν)(μ0+μ1+γ1), |
which can be shown to be a threshold of extinction or persistence of disease for the model (2.2).
In addition, we easily know that there exist two equilibriums for model (2.2): the disease free equilibrium E0=(Λμ0+ν,0,νΛμ0(μ0+ν)) always exists and the unique endemic equilibrium E∗=(S∗,I∗,R∗) exists whenever R0>1, which is a positive solution of the following system:
{Λ−(β1−β2f(I∗))S∗I∗−(μ0+ν)S∗=0,(β1−β2f(I∗))S∗I∗−(μ0+μ1+γ1)I∗=0,γ1I∗+νS∗−μ0R∗=0, |
where S∗,I∗,R∗ satisfy
S∗=μ0+μ1+γ1β1−β2f(I∗),R∗=γ1I∗(β1−β2f(I∗))+ν(μ0+μ1+γ1)μ0(β1−β2f(I∗)), |
and
Λ−(μ0+μ1+γ1)I∗−(μ0+ν)(μ0+μ1+γ1)β1−β2f(I∗)=0. |
Set
F(I):=Λ−(μ0+μ1+γ1)I−(μ0+ν)(μ0+μ1+γ1)β1−β2f(I). |
Since
F(0)=Λ−(μ0+ν)(μ0+μ1+γ1)β1=(μ0+ν)(μ0+μ1+γ1)(R0−1)β1, |
if R0>1, then F(0)>0. It follows from the assumption (A1) that F(I) is a decreasing function. Therefore, F(I)=0 has a unique positive solution I∗ and model (2.2) has a unique endemic equilibrium E∗=(S∗,I∗,R∗) with
S∗=μ0+μ1+γ1β1−β2f(I∗),R∗=γ1I∗(β1−β2f(I∗))+ν(μ0+μ1+γ1)μ0(β1−β2f(I∗)). |
Now, we give the results concerning the existence of unique positive solution, extinction and persistence of the disease for the hepatitis B model (2.3). For simplicity, define a bounded set Γ by
Γ={(S(t),I(t),R(t))∈X:S(t)>0,I(t)>0,R(t)>0,Λμ0+μ1≤S(t)+I(t)+R(t)≤Λμ0}⊂X, |
and denote the stochastic basic reproduction number Rs for the hepatitis B model (2.3) by
Rs=Λβ1(μ0+ν)(μ0+μ1+γ1+σ2Λ22(μ0+ν)2). |
The main results of this paper are given by the following three theorems.
Theorem 3.1. There exists a unique positive solution (S(t),I(t),R(t)) to model (2.3) on t≥0 for any initial value (S(0),I(0),R(0))∈X, and the solution will remain in X with probability one.
Theorem 3.2. Let (S(t),I(t),R(t)) be the solution of the hepatitis B model (2.3) with initial value (S(0),I(0),R(0))∈Γ. If
Rs<1,andσ2≤β1(μ0+ν)Λ, | (3.1) |
then
lim supt→∞logI(t)t<0,a.s. |
namely, I(t) converges to 0 exponentially a.s., and the disease in model (2.3) will die out with probability one. In addition,
limt→∞S(t)=Λμ0+ν=S0,limt→∞R(t)=νΛμ0(μ0+ν)=R0. | (3.2) |
Theorem 3.3. For every t>0, the distribution of (S(t),I(t),R(t)) has a density U(t,x,y,z). If Rs>1 and
θ2<2μ0ν,σ2<2(μ0+ν+12θ2ν)θ1I∗min{1,A1,A2},A1=2μ0+2μ1+γ1(μ0+ν+12θ2ν)S∗2+2μ0+2μ1+γ1,A2=θ2(μ0+12ν)(μ0+ν+12θ2ν)S∗2+θ2(μ0+12ν), | (3.3) |
hold, where θ1=4μ0+2μ1+γ1+νβ1−β2f(I∗), θ2=2μ0+μ1γ1, then there exists a unique density U∗(x,y,z) which is a stationary solution of model (2.3) and
limt→∞∭X|U(t,x,y,z)−U∗(x,y,z)|dxdydz=0. |
In addition, we have
Π≡suppU∗={(x,y,z)∈X:Λμ0+μ1<x+y+z<Λμ0}. | (3.4) |
Remark 3.4. It follows from Theorem 3.2 that if Rs<1, the disease in hepatitis B model (2.3) will die out. The results of Theorem 3.3 mean that the disease is prevalent if Rs>1. Therefore, together with Theorem 3.2 and 3.3, we can clearly see that Rs can be a threshold of disease persistence and extinction. We also conclude that the large random noise can suppress the outbreak of disease.
In this section, we analyze the dynamical behavior of the system (2.3) and give the detailed proofs of our main results.
The aim of this subsection is to show the existence of unique positive global solution of stochastic model (2.3), namely to prove Theorem 3.1.
Proof. The proof of Theorem 3.1 is similar to that in [4,24]. Here we omit it.
The following result shows that the solutions of system (2.3) are bounded.
Lemma 4.1. The unique solution of stochastic hepatitis B epidemic model (2.3) on t≥0 for any initial value (S(0),I(0),R(0))∈X will enter Γ and will remain in Γ with probability one.
Proof. Let the total size of population be N(t)=S(t)+I(t)+R(t). From model (2.3), we have
dN(t)dt=Λ−μ0N(t)−μ1I(t). |
This implies that
Λ−(μ0+μ1)N(t)≤dN(t)dt≤Λ−μ0N(t). |
Letting t→∞, we get
Λμ0+μ1≤lim inft→∞N(t)≤lim supt→∞N(t)≤Λμ0. |
Then the region
Γ={(S(t),I(t),R(t)):S(t)>0,I(t)>0,R(t)>0,Λμ0+μ1≤N(t)≤Λμ0}. |
Therefore, all solution S(t), I(t) and R(t) of model (2.3) are bounded by Λμ0. Thus, we get that Γ is the positively invariant bounded set. In conclusion, the trajectories of all solution initiating anywhere of X will enter Γ and then remain in Γ with probability one.
In this subsection, for simplicity, we begin with introducing the following notation and lemma:
⟨x(t)⟩=1t∫t0x(s)ds. |
Lemma 4.2. [33] Let M={M(t)}t≥0 be a real-valued continuous local martingale vanishing at t=0. Then
limt→∞⟨M,M⟩t=∞a.s.⇒limt→∞M(t)⟨M,M⟩t=0a.s. |
and
lim supt→∞⟨M,M⟩tt<∞a.s.⇒limt→∞M(t)t=0a.s., |
where ⟨M,M⟩t denotes the quadratic variation of M.
Next, we investigate the extinction of disease for the stochastic hepatitis B model (2.3), which means to prove Theorem 3.2.
Proof. An integration of system (2.3) yields
{S(t)−S(0)t=Λ−(β1−β2f(I))⟨S(t)I(t)⟩−(μ0+ν)⟨S(t)⟩−σt∫t0S(s)I(s)dB(s),I(t)−I(0)t=(β1−β2f(I))⟨S(t)I(t)⟩−(μ0+μ1+γ1)⟨I(t)⟩+σt∫t0S(s)I(s)dB(s),R(t)−R(0)t=γ1⟨I(t)⟩+ν⟨S(t)⟩−μ0⟨R(t)⟩. | (4.1) |
According to (4.1), we have
S(t)−S(0)t+I(t)−I(0)t=Λ−(μ0+ν)⟨S(t)⟩−(μ0+μ1+γ1)⟨I(t)⟩. | (4.2) |
We compute that
⟨S(t)⟩=Λμ0+ν−μ0+μ1+γ1μ0+ν⟨I(t)⟩+φ(t), | (4.3) |
where φ(t) is defined by
φ(t)=−1μ0+ν[S(t)−S(0)t+I(t)−I(0)t]. |
Obviously,
limt→∞φ(t)=0.a.s. |
Applying the Itô's formula [16] to system (2.3) leads to
dlogI(t)=[(β1−β2f(I))S(t)−(μ0+μ1+γ1)−12σ2S2(t)]dt+σS(t)dB(t). |
Integrating it from 0 to t and dividing t on both sides, we obtain
logI(t)−logI(0)t=(β1−β2f(I))⟨S(t)⟩−(μ0+μ1+γ1)−12σ2⟨S2(t)⟩+σt∫t0S(s)dB(s)≤β1⟨S(t)⟩−(μ0+μ1+γ1)−12σ2⟨S(t)⟩2+σt∫t0S(s)dB(s). | (4.4) |
Substituting (4.3) into (4.4) yields
logI(t)−logI(0)t≤β1(Λμ0+ν−μ0+μ1+γ1μ0+ν⟨I(t)⟩+φ(t))−(μ0+μ1+γ1)−12σ2(Λμ0+ν−μ0+μ1+γ1μ0+ν⟨I(t)⟩+φ(t))2+σt∫t0S(s)dB(s)≤β1Λμ0+ν−β1(μ0+μ1+γ1)μ0+ν⟨I(t)⟩−(μ0+μ1+γ1)−12σ2Λ2(μ0+ν)2+σ2Λ(μ0+μ1+γ1)(μ0+ν)2⟨I(t)⟩+M(t)t+ψ(t)≤−(μ0+μ1+γ1+12σ2Λ2(μ0+ν)2)(1−Rs)−(μ0+μ1+γ1μ0+ν)(β1−σ2Λμ0+ν)⟨I(t)⟩+M(t)t+ψ(t). | (4.5) |
where
ψ(t)=βφ(t)−12σ2φ2(t)+σ2(μ0+μ1+γ1)μ0+ν⟨I(t)⟩φ(t)−σ2Λφ(t)μ0+ν |
and M(t)=σ∫t0S(s)dB(s), which is a local continuous martingale with M(0)=0. Moreover,
lim supt→∞⟨M,M⟩tt≤σ2Λ2μ20<∞a.s. |
By Lemma 4.2 and limt→∞φ(t)=0, we obtain
limt→∞M(t)t=0andlimt→∞ψ(t)=0a.s. |
If the condition (3.1) is satisfied, it follows from (4.5) that
lim supt→∞logI(t)t≤−(μ0+μ1+γ1+12σ2Λ2(μ0+ν)2)(1−Rs)−(μ0+μ1+γ1μ0+ν)(β1−σ2Λμ0+ν)⟨I(t)⟩<0a.s. |
which implies
limt→∞I(t)=0a.s. | (4.6) |
Next, we prove the assertion (3.2). According to model (2.3), we have
d(S(t)+I(t)+R(t))=[Λ−μ0(S(t)+I(t)+R(t))−μ1I(t)]dt. |
We also have
S(t)+I(t)+R(t)=e−μ0t(S(0)+I(0)+R(0)+∫t0[Λ−μ1I(s)]eμ0sds). |
Applying L'Hospital's rule and (4.6), we get
limt→∞(S(t)+R(t))=limt→∞(S(0)+I(0)+R(0)+∫t0[Λ−μ1I(s)]eμ0sdseμ0t−I(t))=limt→∞Λ−μ1I(t)μ0=Λμ0. | (4.7) |
Thus, we obtain
limt→∞(S(t)+R(t))=Λμ0a.s. |
According to model (2.3), the first equation with limiting system yields
dS(t)=(Λ−(μ0+ν)S(t))dt. |
Then we obtain
limt→∞S(t)=Λμ0+ν=S0a.s. |
Therefore, by (4.7), we have
limt→∞R(t)=νΛμ0(μ0+ν)=R0a.s. |
This finishes the proof.
The aim of this subsection is to investigate that the solutions of model (2.3) are converging to the endemic dynamics when Rs>1 under mild extra conditions. We prove Theorem 3.3 about the existence of stationary distribution for the solution of model (2.3), which implies that the disease is persistent.
In order to prove main result, we give the following lemmas and study the asymptotic stability of the Markov semigroups. First, we check that the semigroup has an invariant density.
Lemma 4.3. For each point (x0,y0,z0)∈X and t>0, the transition probability function P(t,x0,y0,z0,A) has a continuous density k(t,x,y,z;x0,y0,z0) with respect to the Lebesgue measure.
Proof. Let
a0(S,I,R)=(Λ−(β1−β2f(I))SI−(μ0+ν)S(β1−β2f(I))SI−(μ0+μ1+γ1)Iγ1I+νS−μ0R)anda1(S,I,R)=(−σSIσSI0). |
Then Lie bracket [a0,a1] is given by
a2=[a0,a1]=(−σI(Λ−(μ0+μ1+γ1)S+β2f′(I)S2I)σI(Λ−(μ0+ν)S+β2f′(I)S2I)σ(ν−γ1)SI) |
and
a3=[a1,a2]=(−σ2I(ΛI+(ν−μ1−γ1)S2+(β2f″(I)S−β2f′(I))S2I2+β2f′(I)S3I)σ2I2(Λ+(ν−μ1−γ1)S+(β2f″(I)S−β2f′(I))S2I+β2f′(I)S3)σ2(ν−γ1)(S−I)SI). |
Thus, a1,a2,a3 are linearly independent on X. Then for each (S,I,R)∈X, the vector a1(S,I,R), a2(S,I,R), a3(S,I,R) span the space X. According to the Hörmander theorem on the existence of smooth densities for degenerate diffusion process (see [34], Theorem 4.3), the transition probability function P(t,x0,y0,z0,A) has a continuous density k(t,x,y,z;x0,y0,z0) and k∈C∞((0,∞)×X×X).
Using the similar method mentioned in [35], we check the positivity of k.
Fix a point (x0,y0,z0)∈X and a function ϕ∈L2([0,T];R), then consider the following system:
{xϕ(t)=x0+∫t0[f1(xϕ(s),yϕ(s),zϕ(s))−σϕxϕ(s)yϕ(s)]ds,yϕ(t)=y0+∫t0[f2(xϕ(s),yϕ(s),zϕ(s))+σϕxϕ(s)yϕ(s)]ds,zϕ(t)=z0+∫t0f3(xϕ(s),yϕ(s),zϕ(s))ds, | (4.8) |
where f1(x,y,z), f2(x,y,z), f3(x,y,z) are defined as (2.6).
Denote X=(x,y,z)T, X0=(x0,y0,z0)T, let DX0;ϕ be the Fréchet derivative of the function h↦Xϕ+h(T) from L2([0,T];R) to X. Then k(T,x,y,z;x0,y0,z0)>0 for X=Xϕ(T) holds, if the derivative DX0;ϕ has rank 3 for some ϕ∈L2([0,T];R). Let
Ψ(t)=f′(Xϕ(t))+ϕg′(Xϕ(t)), |
where f′, g′ are the Jacobians of
f=(f1(x,y,z)f2(x,y,z)f3(x,y,z))andg=(−σxyσxy0), |
respectively. Let Q(t,t0)(0≤t0≤t≤T) be a matrix function, and Q(t0,t0)=Id, ∂Q(t,t0)∂t=Ψ(t)Q(t,t0), then
DX0;ϕh=∫T0Q(T,s)g(s)h(s)ds. |
Lemma 4.4. There exists T>0 such that k(T,x,y,z;x0,y0,z0)>0 for every (x0,y0,z0)∈Π and (x,y,z)∈Π.
Proof. We consider a continuous control function ϕ and rewrite the system (4.8) as follows:
{x′ϕ(t)=f1(xϕ(t),yϕ(t),zϕ(t))−σϕxϕ(t)yϕ(t),y′ϕ(t)=f2(xϕ(t),yϕ(t),zϕ(t))+σϕxϕ(t)yϕ(t),z′ϕ(t)=f3(xϕ(t),yϕ(t),zϕ(t)). | (4.9) |
Step 1: Let ε∈(0,T) and h(t)=χ[T−ε,T](t)xϕ(t)yϕ(t), t∈[0,T], where χ is the characteristic function. Since
Q(T,s)=Id+Ψ(T)(s−T)+12Ψ2(T)(s−T)2+o((s−T)2). |
Then
DX0;ϕh=εv−12ε2Ψ(T)v+16ε3Ψ2(T)v+o(ε3), |
where
v=(−σσ0). |
Compute
Ψ(T)v=σ(μ0+ν+(σϕ+β1−β2f(y))(y−x)+β2f′(y)xy−(μ0+μ1+γ1)−(σϕ+β1−β2f(y))(y−x)−β2f′(y)xyγ1−ν) |
and
Ψ2(T)v=σ(a11a22a33), |
where
a11=−β22f′2(y)x2y2+(2(β1−β2f(y)+σϕ)(x−y)−(2μ0+μ1+γ1+ν))β2f′(y)xy+(β1−β2f(y)+σϕ)((2μ0+μ1+γ1+ν)x−2(μ0+ν)y)−(β1−β2f(y)+σϕ)2(x−y)2,a22=β22f′2(y)x2y2−2((β1−β2f(y)+σϕ)(x−y)−(μ0+μ1+γ1))β2f′(y)xy−(β1−β2f(y)+σϕ)(2(μ0+μ1+γ1)x−(2μ0+μ1+γ1+ν)y)+(β1−β2f(y)+σϕ)2(x−y)2+(μ0+μ1+γ1)2,a33=−γ1(2μ0+μ1+γ1)+ν(2μ0+ν)+(ν−γ1)(β2f′(y)xy+(β1−β2f(y)+σϕ)(y−x)). |
It is easy to see that v, Ψ(T)v, and Ψ2(T)v are linearly independent. Therefore, the rank of DX0;ϕ is 3.
Step 2: We check that there exist a control function ϕ and T>0 such that Xϕ(0)=X0, Xϕ(T)=X for any two points X0∈Π and X∈Π holds. Let wϕ=xϕ+yϕ+zϕ, then system (4.9) can be replaced by
{x′ϕ(t)=g1(xϕ(t),wϕ(t),zϕ(t))−σϕxϕ(t)(wϕ(t)−xϕ(t)−zϕ(t)),y′ϕ(t)=g2(xϕ(t),wϕ(t),zϕ(t)),z′ϕ(t)=g3(xϕ(t),wϕ(t),zϕ(t)), | (4.10) |
where
g1(x,w,z)=Λ−(β1−β2f(w−x−z))x(w−x−z)−(μ0+ν)x,g2(x,w,z)=Λ+μ1(x+z)−(μ0+μ1)w,g3(x,w,z)=γ1w−(γ1−ν)x−(γ1+μ0)z. | (4.11) |
Let
Π0={(x,w,z)∈X:0<x,z<Λμ0,Λμ0+μ1<w<Λμ0andx,z<w}. |
Now we claim that there exist a control function ϕ and T>0, for any (x0,w0,z0)∈Π0 and (x1,w1,z1)∈Π0, we have (xϕ(0),wϕ(0),zϕ(0))=(x0,w0,z0) and (xϕ(T),wϕ(T),zϕ(T))=(x1,w1,z1).
To construct the function ϕ, first, we find a positive constant T and a differentiable function
wϕ:[0,T]→(Λμ0+μ1,Λμ0), |
such that wϕ(0)=w0, wϕ(T)=w1, w′ϕ(0)=g2(x0,w0,z0)=wd0, w′ϕ(T)=g2(x1,w1,z1)=wdT and
Λ−(μ0+μ1)wϕ(t)<w′ϕ(t)<Λ−μ0wϕ(t),t∈[0,T]. | (4.12) |
Next we separate the construction of the function wϕ on three subintervals [0,ε], [ε,T−ε] and [T−ε,T], where 0<ε<T2. Let
η=12min{w0−Λμ0+μ1,w1−Λμ0+μ1,Λμ0−w0,Λμ0−w1}. |
If wϕ∈(Λμ0+μ1+η,Λμ0−η), then
Λ−(μ0+μ1)wϕ(t)<−(μ0+μ1)η<0,andΛ−μ0wϕ(t)>μ0η>0fort∈[0,T]. | (4.13) |
Therefore, from (4.13) it follows that we can find a C2-function wϕ: [0,ε]→(Λμ0+μ1+η,Λμ0−η) such that
wϕ(0)=w0,w′ϕ(0)=wd0,w′ϕ(ε)=0, |
and for all t∈[0,ε], the differentiable function wϕ satisfies (4.12). The same proof works for t∈[T−ε,T], we also find a C2-function wϕ: [T−ε,T]→(Λμ0+μ1+η,Λμ0−η) such that
wϕ(T)=w1,w′ϕ(T)=wdT,w′ϕ(T−ε)=0, |
and for all t∈[T−ε,T], the differentiable function wϕ satisfies (4.12). We choose T sufficiently large such that
wϕ:[0,ε]∪[T−ε,T]→(Λμ0+μ1+η,Λμ0−η) |
can be extend to a C2-function wϕ which defined on the whole interval [0,T], then we have
Λ−(μ0+μ1)wϕ(t)<−(μ0+μ1)η<w′ϕ(t)<μ0η<Λ−μ0wϕ(t),fort∈[ε,T−ε], |
and the differentiable function wϕ satisfies (4.12) on [0,T].
Thus, two C2-function xϕ and zϕ can be found to satisfy the second and third equation of (4.10). Finally, there exists a continuous control function ϕ which can be determined from the first equation of (4.10). This completes the proof.
Lemma 4.5. If Rs>1, then for every density g and semigroup {P(t)}t≥0,
limt→∞∭ΠP(t)g(x,y,z)dxdydz=1. |
Proof. Let Z(t)=S(t)+I(t)+R(t), then system (2.3) becomes
{dS(t)=g1(S(t),Z(t),R(t))dt−σS(t)(Z(t)−S(t)−R(t))dB(t),dZ(t)=g2(S(t),Z(t),R(t))dt,dR(t)=g3(S(t),Z(t),R(t))dt, | (4.14) |
and the functions g1, g2 and g3 are defined in (4.11). For the positive solution (S(t),I(t),R(t)) of model (2.3), we have
Λ−(μ0+μ1)Z(t)<dZ(t)dt<Λ−μ0Z(t),t∈(0,+∞),a.s. | (4.15) |
Next, we check that for almost each ω∈Ω, there exists t0=t0(ω) such that
Λμ0+μ1<Z(ω,t)<Λμ0,fort>t0. |
Here we have the following three possible situations.
(a) The case Z(0)∈(Λμ0+μ1,Λμ0) is simple to see from (4.15).
(b) Consider the case Z(0)∈(0,Λμ0+μ1). Assume our assertion is false, then for ω∈Ω′, there exists Ω′⊂Ω with Prob(Ω′)>0 such that Z(ω,t)∈(0,Λμ0+μ1). Obviously, from (4.15) we see Z(ω,t) is strictly increasing on [0,∞] for any ω∈Ω′. Thus, we have
limt→∞Z(ω,t)=Λμ0+μ1,ω∈Ω′. |
By the second equation of (4.14), we obtain for any ω∈Ω′
Z(t)=e−(μ0+μ1)t⋅(Z(0)+∫t0e(μ0+μ1)s[Λ+μ1(S(s)+R(s))]ds). |
Then for ω∈Ω′, we have
limt→∞S(ω,t)=limt→∞R(ω,t)=0. |
Therefore, limt→∞I(ω,t)=Λμ0+μ1, ω∈Ω′, which implies that
limt→∞logI(t)−logI(0)t=0,ω∈Ω′. |
Application of the Itô's formula yields
dlogI(t)=[(β1−β2f(I))S(t)−(μ0+μ1+γ1)−12σ2S2(t)]dt+σS(t)dB(t), |
and
logI(t)−logI(0)t=(β1−β2f(I))⟨S(t)⟩−(μ0+μ1+γ1)−12σ2⟨S2(t)⟩+σt∫t0S(s)dB(s). |
According to Lemma 3, we have
limt→∞σt∫t0S(s)dB(s)=0. |
Hence,
limt→∞logI(t)−logI(0)t=limt→∞((β1−β2f(I))⟨S(t)⟩−(μ0+μ1+γ1)−12σ2⟨S2(t)⟩+σt∫t0S(s)dB(s))=−(μ0+μ1+γ1). |
This leads to the contradiction that limt→∞logI(t)−logI(0)t=0, then the assertion holds.
(c) Consider the case Z(0)∈(Λμ0,+∞). The same proof works to the case (b), by contradiction, suppose that there exists ω∈Ω′ with Prob(Ω′)>0 such that
limt→∞Z(ω,t)=Λμ0,ω∈Ω′. |
The second and third equation of (4.14) implies that for any ω∈Ω′
Z(t)=e−(μ0+μ1)t⋅(Z(0)+∫t0e(μ0+μ1)s[Λ+μ1(S(s)+R(s))]ds),R(t)=e−(μ0+γ1)t⋅(R(0)+∫t0e(μ0+γ1)s[γ1Z(s)−(γ1−ν)S(s)]ds). |
It follows that
limt→∞S(ω,t)=Λμ0+ν,limt→∞I(ω,t)=0,limt→∞R(ω,t)=νΛμ0(μ0+ν)forω∈Ω′. |
Thus,
limt→∞logI(t)−logI(0)t=limt→∞((β1−β2f(I))⟨S(t)⟩−(μ0+μ1+γ1)−12σ2⟨S2(t)⟩+σt∫t0S(s)dB(s))=β1Λμ0+ν−σ2Λ22(μ0+ν)2−(μ0+μ1+γ1)=(μ0+μ1+γ1+σ2Λ22(μ0+ν)2)(Rs−1)>0a.s.onΩ′. |
This leads to the contradiction that \lim_{t\rightarrow\infty}I(\omega, t) = 0 a.s. and the assertion holds.
Remark 4.1. The results of Lemma 5 and Lemma 6 mean that if there exists a stationary solution U_* for Fokker-Planck equation (2.5), then \text{supp}\; U_* = \Pi .
Lemma 4.6. If \mathscr{R}_s > 1 , then the semigroup \{P(t)\}_{t\geq0} is asymptotically stable or is sweeping with respect to compact sets.
Proof. The result of Lemma 4 shows that \{P(t)\}_{t\geq0} is an integral Markov semigroup, which has a continuous kernel k(t, x, y, z; x_0, y_0, z_0) for t > 0 . Thus, the distribution of (S(t), I(t), R(t)) has a density U(x, y, z, t) and it satisfies the Fokker-Planck equation (2.5). The Lemma 6 implies that certify the restriction of the semigroup \{P(t)\}_{t\geq0} to the space L^1(\Pi) is sufficient. By Lemma 5, we know for every g\in\mathbb{D}
\begin{equation*} \int_0^\infty P(t)gdt \gt 0, \; \text{a.s.}\; \text{on} \; \Pi. \end{equation*} |
Therefore, in view of Lemma 1 that the semigroup \{P(t)\}_{t\geq0} is asymptotically stable or is sweeping with respect to compact sets. This finishes the proof.
Lemma 4.7. If \mathscr{R}_s > 1 then the semigroup \{P(t)\}_{t\geq0} is asymptotically stable provided following conditions are satisfied:
\begin{equation} \theta_2 \lt \frac{2\mu_0}{\nu}, \; \sigma^2 \lt \frac{2(\mu_0+\nu+\frac{1}{2}\theta_2\nu)}{\theta_1I^*}\min\{1, A_1, A_2\}, \end{equation} | (4.16) |
where \theta_1 = \frac{4\mu_0+2\mu_1+\gamma_1+\nu}{\beta_1-\beta_2f(I^*)} , \theta_2 = \frac{2\mu_0+\mu_1}{\gamma_1} and A_1, A_2 are defined in (3.3).
Proof. The Lemma 4.6 implies that the semigroup \{P(t)\}_{t\geq0} satisfies the Foguel alternative. We construct a nonnegative C^2 -function V and a closed set O\in\Sigma such that
\begin{equation*} \sup\limits_{(S, I, R)\in\mathbb{X}\setminus O}\mathscr{A}^*V \lt 0, \end{equation*} |
where the function V is called Khasminski\u{i} function [29]. Since there exists an endemic equilibrium E^* of system (2.2) when \mathscr{R}_0 > 1 , then we have
\begin{equation*} \Lambda = (\mu_0+\nu)S^*+(\beta_1-\beta_2f(I^*))S^*I^*, \end{equation*} |
\begin{equation*} (\beta_1-\beta_2f(I^*))S^*I^* = (\mu_0+\mu_1+\gamma_1)I^*, \end{equation*} |
\begin{equation*} \gamma_1I^* = \mu_0R^*-\nu S^*. \end{equation*} |
Define a nonnegative C^2 -function V by
\begin{equation*} \begin{split} V(S, I, R) = &\frac{1}{2}(S-S^*+I-I^*+R-R^*)^2+\frac{1}{2}(S-S^*+I-I^*)^2\\ &+\theta_1\left(I-I^*-I^*\log\frac{I}{I^*}\right)+\frac{\theta_2}{2}(R-R^*)^2\\\ : = &V_1+V_2+\theta_1V_3+\theta_2V_4, \end{split} \end{equation*} |
where \theta_1 and \theta_2 are defined in Lemma 4.7. First, we compute
\begin{equation} \begin{split} \mathscr{A}^*V_1 = &(S-S^*+I-I^*+R-R^*)(\Lambda-\mu_0S-(\mu_0+\mu_1)I-\mu_0R)\\ = &(S-S^*+I-I^*+R-R^*)(-\mu_0(S-S^*)-(\mu_0+\mu_1)(I-I^*)-\mu_0(R-R^*))\\ = &-\mu_0(S-S^*)^2-(\mu_0+\mu_1)(I-I^*)^2-\mu_0(R-R^*)^2-(2\mu_0+\mu_1)(S-S^*)(I-I^*)\\ &-2\mu_0(S-S^*)(R-R^*)-(2\mu_0+\mu_1)(I-I^*)(R-R^*). \end{split} \end{equation} | (4.17) |
Next, we have
\begin{equation} \begin{split} \mathscr{A}^*V_2 = &(S-S^*+I-I^*)(\Lambda-(\mu_0+\nu)S-(\mu_0+\mu_1+\gamma_1)I)\\ = &(S-S^*+I-I^*)(-(\mu_0+\nu)(S-S^*)-(\mu_0+\mu_1+\gamma_1)(I-I^*))\\ = &-(\mu_0+\nu)(S-S^*)^2-(\mu_0+\mu_1+\gamma_1)(I-I^*)^2\\ &-(2\mu_0+\mu_1+\gamma_1+\nu)(S-S^*)(I-I^*). \end{split} \end{equation} | (4.18) |
We calculate
\begin{equation} \begin{split} \mathscr{A}^*V_3 = &(I-I^*)((\beta_1-\beta_2f(I))S-(\mu_0+\mu_1+\gamma_1))+\frac{1}{2}I^*\sigma^2S^2\\ = &(I-I^*)((\beta_1-\beta_2f(I))S-(\beta_1-\beta_2f(I^*))S^*)+\frac{1}{2}I^*\sigma^2S^2\\ = &-\beta_2S(f(I)-f(I^*))(I-I^*)+(\beta_1-\beta_2f(I^*))(S-S^*)(I-I^*)+\frac{1}{2}I^*\sigma^2S^2\\ \leq&\; (\beta_1-\beta_2f(I^*))(S-S^*)(I-I^*)+\frac{1}{2}I^*\sigma^2S^2. \end{split} \end{equation} | (4.19) |
At last, for V_4 , we have
\begin{equation} \begin{split} \mathscr{A}^*V_4 = &(R-R^*)(\gamma_1I+\nu S-\mu_0R)\\ = &(R-R^*)(\nu (S-S^*)+\gamma_1(I-I^*)-\mu_0(R-R^*))\\ = &-\mu_0(R-R^*)^2+\nu(S-S^*)(R-R^*)+\gamma_1(I-I^*)(R-R^*). \end{split} \end{equation} | (4.20) |
Combining (4.17)– (4.20), we obtain
\begin{equation*} \begin{split} \mathscr{A}^*V = &\mathscr{A}^*V_1+\mathscr{A}^*V_2+\theta_1\mathscr{A}^*V_3+\theta_2\mathscr{A}^*V_4\\ \leq&-\Big(\mu_0+\nu+\frac{1}{2}\theta_2\nu\Big)(S-S^*)^2-(2\mu_0+2\mu_1+\gamma_1)(I-I^*)^2\\ &-\theta_2\Big(\mu_0+\frac{1}{2}\nu\Big)(R-R^*)^2+\frac{1}{2}\theta_1I^*\sigma^2S^2\\ = &-\Big(\mu_0+\nu+\frac{1}{2}\theta_2\nu-\frac{1}{2}\theta_1I^*\sigma^2\Big)\Big(S-\frac{2\mu_0+2\nu+\theta_2\nu}{2\mu_0+2\nu+\theta_2\nu- \theta_1I^*\sigma^2}S^*\Big)^2 \\ &- (2\mu_0+2\mu_1+\gamma_1)(I-I^*)^2-\theta_2\Big(\mu_0+\frac{1}{2}\nu\Big)(R-R^*)^2\\ &+\frac{\theta_1I^*\Big(\mu_0+\nu+\frac{1}{2}\theta_2\nu\Big)\sigma^2}{2\mu_0+2\nu+\theta_2\nu- \theta_1I^*\sigma^2}S^{*2}\\ : = &-b_1(S-c_1S^*)^2-b_2(I-I^*)^2-b_3(R-R^*)^2+b_4. \end{split} \end{equation*} |
It follows from the condition (4.16) of Lemma 4.7 that
\begin{equation*} \begin{split} &\frac{\theta_1I^*\Big(\mu_0+\nu+\frac{1}{2}\theta_2\nu\Big)\sigma^2}{2\mu_0+2\nu+\theta_2\nu- \theta_1I^*\sigma^2}S^{*2}\\ \lt &\min\Bigg\{\frac{2\Big(\mu_0+\nu+\frac{1}{2}\theta_2\nu\Big)^2S^{*2}}{2\mu_0+2\nu+\theta_2\nu- \theta_1I^*\sigma^2}, 2\mu_0+2\mu_1+\gamma_1, \theta_2\Big(\mu_0+\frac{1}{2}\nu\Big)\Bigg\}. \end{split} \end{equation*} |
Then the ellipsoid
\begin{equation*} -b_1(S-c_1S^*)^2-b_2(I-I^*)^2-b_3(R-R^*)^2+b_4 = 0 \end{equation*} |
lies entirely in \mathbb{X} . Thus, there exist a closed set O\in\Sigma which contains this ellipsoid and constant c > 0 such that
\begin{equation*} \sup\limits_{(S, I, R)\in\mathbb{X}\setminus O}\mathscr{A}^*V\leq-c \lt 0. \end{equation*} |
Remark 4.2. Together with Lemma 4.6 and Lemma 4.7, we get Theorem 3.3.
In this section, we give some numerical simulations to illustrate the effectiveness of our analytical results. Choosing the function f(I) = \frac{I}{H+I} (as in [3,11]), which satisfy the assumption \mathbf{(A1)} clearly. Using the Milstein method for stochastic differential equations [36], we consider the following discretization for system (2.3):
\begin{equation*} \begin{cases} \begin{split} S(k+1) = &S(k)+\Big(\Lambda-\Big(\beta_1-\frac{\beta_2I(k)}{H+I(k)}\Big) S(k)I(k)-(\mu_0+\nu)S(k)\Big)\Delta t \\ &- \sigma S(k)I(k)\sqrt{\Delta t}\xi(k) - \frac{\sigma^2}{2}S(k)I(k)(\xi^2(k)-1)\Delta t, \\ I(k+1) = &I(k)+\Big(\Big(\beta_1-\frac{\beta_2I(k)}{H+I(k)}\Big) S(k)I(k)-(\mu_0+\mu_1+\gamma_1)I(k)\Big)\Delta t \\ &+ \sigma S(k)I(k)\sqrt{\Delta t}\xi(k) + \frac{\sigma^2}{2}S(k)I(k)(\xi^2(k)-1)\Delta t, \\ R(k+1) = &R(k)+(\gamma_1 I(k)+\nu S(k)-\mu_0R(k))\Delta t, \end{split} \end{cases} \end{equation*} |
where \xi(k), k = (1, 2, ..., n) are independent Gaussian random variables N(0, 1) . We take the parameter values in system (2.3) as in Table 2. The initial value of population size is S(0) = 0.9, I(0) = 0.8, R(0) = 0.6 [24]. For system (2.2), with parameter values of Table 2, easy to calculate that the basic reproduction number \mathscr{R}_0 = \frac{\Lambda\beta_1}{(\mu_0+\nu)(\mu_0+\mu_1+\gamma_1)} = 3.7500 > 1 . Thus, for any initial values (S(0), I(0), R(0)) , there exist a unique endemic equilibrium E^* = (0.9273, 0.2347, 3.7207) which is globally stable and a disease-free equilibrium E_0 = (1.2500, 0, 3.7500) . For a clear comparison with the path of stochastic hepatitis B model (2.3), we show the path of S(t) , I(t) , R(t) for deterministic model (2.1) in Figure 1.
Parameter | Parameter description | Value | The source of data |
\Lambda | Birth rate | 0.5 | [24] |
\beta_1 | Transmission rate | 0.6 | [24] |
\beta_2 | Maximum reduced contact rate | 0.3 | Assumed |
\mu_0 | Natural death rate | 0.1 | [24] |
\mu_1 | Disease induced death rate | 0.05 | Assumed |
v | Vaccination rate | 0.3 | [24] |
\gamma_1 | Recovery rate | 0.4 | [24] |
H | 10 | [11]. |
Choosing the intensity of noise \sigma = 0.1 , then \mathscr{R}_s = 1.3445 > 1 , \sigma^2 = 0.0100 < 0.5595 and the condition \theta_2 = 0.6250 < 0.6667 hold. It follows from Theorem 3.3 that the disease will persistent and we give the simulation result in Figure 2(a). Compared with Figure 1, the solution of system (2.3) in Figure 2(a) shows small fluctuations. We increase the intensity of noise \sigma to 0.3 ( \mathscr{R}_s = 1.2091 > 1 , \sigma^2 = 0.0900 < 0.5595 ) and 0.5 ( \mathscr{R}_s = 1.0063 > 1 , \sigma^2 = 0.2500 < 0.5595 ), respectively. We find that the fluctuations become stronger with the increase of the noise intensity and the simulation results are showed in Figure 2(b)–(c).
Moreover, the histograms of the probability density function for I(t) are obtained from 10, 000 simulation runs for three different noise intensities at t = 100 , which is displayed in Figure 2(d)–(f). From Figure 2(d)–(f), we can see that the skewness of the distribution for I(100) is changing with the increase of the noise intensity \sigma . More precisely, the distribution is close to a standard distribution when \sigma = 0.1 (see Figure 2(d)). But if increase \sigma to 0.3 and 0.5, respectively, the distributions are positively skewed (see Figure 2(e)–(f)).
We note that in this paper, the model (2.3) is a continuous time and continuous state space model, the values of I(t) are non-zero quantities increase to 16th decimal during the numerical experiments. Therefore, we assume that 10, 000 individuals are deemed to be 1 unit populations approximately, and assume that the disease is regarded as extinction when the value of I(t) less than 0.0001 [4].
To learn the stochastic disease-free dynamics of model (2.3), we choose \sigma = 0.52 , then \mathscr{R}_s = 0.9852 < 1 and \sigma^2 = 0.2704 < \frac{\beta_1(\mu_0+\nu)}{\Lambda} = 0.4800 . According to Theorem 3.2, the disease goes extinct exponentially almost surely, which is illustrated by Figure 3(a). We increase \sigma to 0.54 ( \mathscr{R}_s = 0.9642 < 1 , \sigma^2 = 0.2916 < 0.4800 ) and to 0.56 ( \mathscr{R}_s = 0.9434 < 1 , \sigma^2 = 0.3136 < 0.4800 ), respectively. We find that the disease goes to extinction with probability one, as shown in Figure 3(b)-(c).
Furthermore, we conduct 10, 000 numerical simulation runs and calculate the mean extinction time of I(t) . We obtain that the mean extinction time for the three different noise intensities \sigma (i.e., 0.52, 0.54, 0.56) is 82.5642, 76.9118, 69.5481, respectively. Then we conclude that the mean extinction time of disease decreases with the increase of noise intensity \sigma .
In the following, we show the influence of information intervention. Therefore, we mainly discuss the effect of different values of \beta_2 on infected population. First, we show the I(t) for deterministic model (2.2) under different values of \beta_2 ( \beta_2 = 0, 0.2, 0.4, 0.6 ) in Figure 4(a). It can be seen from Figure 4(a) that \beta_2 has great influence on I(t) . The number of the infected population decreases with the increase of the \beta_2 .
We next choose noise intensity \sigma = 0.25 , then make 10, 000 numerical simulation runs and calculate mean value. From Figure 4(b), it can be seen that the increase \beta_2 can reduce the value of I(t) . Further, we fix \beta_2 and compare Figure 4(a) with 4(b), we find that noise intensity \sigma can also reduce the number of the infected population. If we increase \sigma to 0.5, then we have a similar conclusion and the simulation result is shown in Figure 4(c). This simulation shows that the information intervention can help reduce the number of the disease outbreak.
Outbreaks of infectious diseases have caused substantial deaths, social and economic losses in the whole world. As a result, a large number of prevention strategies, including information policies, such as media coverage, health education, psychological suggestion, etc. have been used to help to understand the transmission and control of infectious diseases. Environmental noise, an important component in real world, also has greatly affect on the development of infectious diseases. In this paper, we investigated the long term behavior of the stochastic model for the transmission dynamic of hepatitis B with varying population environment. Our major results are as follows:
(ⅰ) Using the Markov semigroups theory, we investigated the dynamics of the stochastic hepatitis B model with information intervention perturbed by environment noises. Our results showed that the dynamics of the stochastic hepatitis B model can be governed by the stochastic basic reproduction number \mathscr{R}_s . There exists a disease-free equilibrium and the disease is predicted to die out with probability one when \mathscr{R}_s < 1 under mild extra condition. If \mathscr{R}_s > 1 under mild extra conditions, the stochastic model has an endemic equilibrium and the disease would persist.
(ⅱ) We showed that environmental noises can play an important role in the long-term dynamics. If the intensity of noises is small enough to imply that \mathscr{R}_s > 1 , which means the disease will prevail and there exists a stationary distribution for the stochastic model. If the intensity of noises is large (leading to \mathscr{R}_s < 1 ), then the disease would be eradicated. Thus, large environmental noises are able to suppress the emergence of the disease outbreak. We further evaluated the influence of information intervention. It leads to the changes in human behavior, which reduces the effective contact rates of susceptible people and provide (temporary) protection from infection. We found that large intensity of information intervention ( \beta_2 ) can lead to the decrease of I(t) (see Figure 4). Therefore, information intervention can also reduce the number of infected population and suppress the outbreak of hepatitis B.
The environmental white noise considered in this paper is a continuous random process. However, population systems may have sudden impact of various factors, such as volcanoes, toxic pollutants, earthquakes, abrupt climate change, etc. Stochastic models with Brownian motion cannot describe these phenomena. It is worth studying epidemic model with a discontinuous random process (Lévy noise, Markov noise) by using Markov semigroups theory. On the other hand, we considered the role of information intervention in the control of hepatitis B. We found that information intervention can actually mitigate the spread of hepatitis B and reduce the number of infected population. But we note that these information intervention policies and vaccination often require more costs. How to minimize the infected populations and the costs of these control strategies by considering the optimal control problem? We will study this issue later.
The research is supported by the Natural Science Foundation of China (Grant numbers 11661064).
The authors declare there is no conflict of interest.
[1] | WHO, Hepatitis B, 2014 (revised Agust 2014). Available from: http://www.who.int/meadiacenter /factsheet/fs204/en/index.html. |
[2] | J.Wu, U.Wang, S. Chang, et al., Impacts of a mass vaccination campaign against pandemic H1N1 2009 influenza in Taiwan: a time-series regression analysis, Int. J. Infect. Dis., 23 (2014), 82–89. |
[3] | Y. Liu and J. Cui, The impact of media coverage on the dynamics of infectious disease, Int. J. Biomath., 1 (2008), 65–74. |
[4] | Y. Cai, Y. Kang, M. Banerjee, et al., A stochastic SIRS epidemic model with infectious force under intervention strategies, J. Differ. Equations, 259 (2015), 7463–7502. |
[5] | Q. Wang, L. Zhao, R. Huang, et al., Interaction of media and disease dynamics and its impact on emerging infection management, Discrete Cont. Dyn.-B, 20 (2014), 215–230. |
[6] | Y. Xiao, T. Zhao and S. Tang, Dynamics of an infectious diseases with media/psychology induced non-smooth incidence, Math. Biosci. Eng., 10 (2013), 445–461. |
[7] | M. Brinn, K. Carson, A. Esterman, et al., Mass media interventions for preventing smoking in young people, Cochrane DB Syst. Rev., 11 (2010), CD001006. |
[8] | Q. Li, L. Zhou, M. Zhou, et al, Epidemiology of human infections with avian influenza A(H7N9) virus in China, New Engl. J. Med., 370 (2014), 520–532. |
[9] | WHO, Cholera-United Republic of Tanzania, 2018. Available from: http://www.who.int/csr/don/ 12-january-2018-cholera-tanzania/en/. |
[10] | J. Cui, Y. Sun and H. Zhu, The impact of media on the control of infectious diseases, J. Dyn. Differ. Equ., 20 (2018), 31–53. |
[11] | J. Cui, X. Tao and H. Zhu, An SIS infection model incorporating media coverage, Rocky MT J. Math., 38 (2008), 1323–1334. |
[12] | L. Zou, W. Zhang and S. Ruan, Modeling the transmission dynamics and control of hepatitis b virus in China, J. Theor. Biol., 262 (2010), 330–338. |
[13] | T. Khan, G. Zaman and M. I. Chohan, The transmission dynamic and optimal control of acute and chronic hepatitis b, J. Biol. Dyn., 11 (2017), 172–189. |
[14] | T. Zhang, K. Wang and X. Zhang, Modeling and analyzing the transmission dynamics of HBV epidemic in Xinjiang, China, PloS One, 10 (2015), e0138765. |
[15] | S. Spencer, Stochastic epidemic models for emerging diseases, PhD thesis, University of Nottingham, 2008. |
[16] | B. Øksendal, Stochastic Differential Equations: An Introduction with Applications, Springer, 2010. |
[17] | Q. Yang and X. Mao, Extinction and recurrence of multi-group SEIR epidemic models with stochastic perturbations, Nonlinear Anal.-Real, 14 (2013), 1434–1456. |
[18] | Q. Liu, D. Jiang, N. Shi, et al., Stationarity and periodicity of positive solutions to stochastic SEIR epidemic models with distributed delay, Discrete Cont. Dyn. - B, 22 (2017), 2479–2500. |
[19] | Y. Cai and W. Wang, Fish-hook bifurcation branch in a spatial heterogeneous epidemic model with cross-diffusion, Nonlinear Anal.- Real, 30 (2016), 99–125. |
[20] | X. Mao, G. Marion and E. Renshaw, Environmental noise suppresses explosion in population dynamics, Stoch. Proc. Appl., 97 (2002), 95–110. |
[21] | Y. Cai, Y. Kang, M. Banerjee, et al., Complex Dynamics of a host-parasite model with both horizontal and vertical transmissions in a spatial heterogeneous environment, Nonlinear Anal.- Real, 40 (2018), 444–465. |
[22] | L. Allen, An introduction to stochastic epidemic models, Mathematical Epidemiology, Lect. Notes Math., 1945 (2008), 81–130. |
[23] | K. Bao and Q. Zhang, Stationary distribution and extinction of a stochastic SIRS epidemic model with information intervention, Adv. Differ. Equ., 352 (2017), doi:10.1186/s13662-017-1406-9. |
[24] | T. Khan, A. Khan and G. Zaman, The extinction and persistence of the stochastic hepatitis B epidemic model, Chaos, Solitons Fract., 108 (2018), 123–128. |
[25] | R. Rudnicki, Long-time behaviour of a stochastic prey-predator model, Stoch. Proc. Appl., 108 (2003), 93–107. |
[26] | R. Rudnicki, K. Pichr and M. Tyrankamiska, Markov semigroups and their applications, Lect. Notes Phys., 597 (2002), 215–238. |
[27] | K. Pichr and R. Rudnicki, Continuous Markov semigroups and stability of transport equations, J. Math. Anal. Appl., 249 (2000), 668–685. |
[28] | R. Rudnicki, On asymptotic stability and sweeping for Markov operators, B. Pol. Acad. Sci. Math., 43 (1995), 245–262. |
[29] | K. Pichr and R. Rudnicki, Stability of Markov semigroups and applications to parabolic systems, J. Math. Anal. Appl., 215 (1997), 56–74. |
[30] | R. Rudnicki, Asymptotic properties of the Fokker-Planck equation, Springer Berlin Heidelberg, (1995), 517–521. |
[31] | P. Martin, Statistical physics: Statics, dynamics, and renormalization, World Scientific, 2000. |
[32] | O. Diekmann and J. Heesterbeek, Mathematical Epidemiology of Infectious Diseases: Model Building, Analysis and Interpretation, Wiley, New York, 2000. |
[33] | R. Lipster, A strong law of large numbers for local martingales, Stochastics, 3 (1980), 217–228. |
[34] | F. Baudoin and M. Hairer, A version of Hörmanders theorem for the fractional Brownian motion, Probab. Theory Rel., 139 (2007), 373–395. |
[35] | G. Arous and R. Léandre, Décroissance exponentielle du noyau de la chaleur sur la diagonale (II), Probab. Theory Rel., 90 (1991), 377–402. |
[36] | D. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. |
1. | Dong-Me Li, Bing Chai, Qi Wang, A model of hepatitis B virus with random interference infection rate, 2021, 18, 1551-0018, 8257, 10.3934/mbe.2021410 |
Parameter | Parameter description |
\Lambda | The per capita constant birth rate |
\beta | The transmission rate |
\mu_0 | The natural death rate |
\mu_1 | The disease induced death rate |
v | The vaccination rate |
\gamma_1 | The constant recovery rate |
Parameter | Parameter description | Value | The source of data |
\Lambda | Birth rate | 0.5 | [24] |
\beta_1 | Transmission rate | 0.6 | [24] |
\beta_2 | Maximum reduced contact rate | 0.3 | Assumed |
\mu_0 | Natural death rate | 0.1 | [24] |
\mu_1 | Disease induced death rate | 0.05 | Assumed |
v | Vaccination rate | 0.3 | [24] |
\gamma_1 | Recovery rate | 0.4 | [24] |
H | 10 | [11]. |
Parameter | Parameter description |
\Lambda | The per capita constant birth rate |
\beta | The transmission rate |
\mu_0 | The natural death rate |
\mu_1 | The disease induced death rate |
v | The vaccination rate |
\gamma_1 | The constant recovery rate |
Parameter | Parameter description | Value | The source of data |
\Lambda | Birth rate | 0.5 | [24] |
\beta_1 | Transmission rate | 0.6 | [24] |
\beta_2 | Maximum reduced contact rate | 0.3 | Assumed |
\mu_0 | Natural death rate | 0.1 | [24] |
\mu_1 | Disease induced death rate | 0.05 | Assumed |
v | Vaccination rate | 0.3 | [24] |
\gamma_1 | Recovery rate | 0.4 | [24] |
H | 10 | [11]. |