
Citation: Yanyan Du, Ting Kang, Qimin Zhang. Asymptotic behavior of a stochastic delayed avian influenza model with saturated incidence rate[J]. Mathematical Biosciences and Engineering, 2020, 17(5): 5341-5368. doi: 10.3934/mbe.2020289
[1] | Yoichi Enatsu, Yukihiko Nakata . Stability and bifurcation analysis of epidemic models with saturated incidence rates: An application to a nonmonotone incidence rate. Mathematical Biosciences and Engineering, 2014, 11(4): 785-805. doi: 10.3934/mbe.2014.11.785 |
[2] | Ning Bai, Juan Zhang , Li Li, Zhen Jin . Evaluating the effect of virus mutation on the transmission of avian influenza H7N9 virus in China based on dynamical model. Mathematical Biosciences and Engineering, 2019, 16(5): 3393-3410. doi: 10.3934/mbe.2019170 |
[3] | Tingting Yu, Sanling Yuan . Dynamics of a stochastic turbidostat model with sampled and delayed measurements. Mathematical Biosciences and Engineering, 2023, 20(4): 6215-6236. doi: 10.3934/mbe.2023268 |
[4] | Guo Lin, Shuxia Pan, Xiang-Ping Yan . Spreading speeds of epidemic models with nonlocal delays. Mathematical Biosciences and Engineering, 2019, 16(6): 7562-7588. doi: 10.3934/mbe.2019380 |
[5] | Yan’e Wang , Zhiguo Wang, Chengxia Lei . Asymptotic profile of endemic equilibrium to a diffusive epidemic model with saturated incidence rate. Mathematical Biosciences and Engineering, 2019, 16(5): 3885-3913. doi: 10.3934/mbe.2019192 |
[6] | Keguo Ren, Xining Li, Qimin Zhang . Near-optimal control and threshold behavior of an avian influenza model with spatial diffusion on complex networks. Mathematical Biosciences and Engineering, 2021, 18(5): 6452-6483. doi: 10.3934/mbe.2021321 |
[7] | Yajuan Guo, Zijian Liu, Yuanshun Tan, Yawei Liu . Modeling and analysis of a stochastic giving-up-smoking model with quit smoking duration. Mathematical Biosciences and Engineering, 2023, 20(12): 20576-20598. doi: 10.3934/mbe.2023910 |
[8] | Ke Qi, Zhijun Liu, Lianwen Wang, Qinglong Wang . Survival and stationary distribution of a stochastic facultative mutualism model with distributed delays and strong kernels. Mathematical Biosciences and Engineering, 2021, 18(4): 3160-3179. doi: 10.3934/mbe.2021157 |
[9] | Edoardo Beretta, Dimitri Breda . An SEIR epidemic model with constant latency time and infectious period. Mathematical Biosciences and Engineering, 2011, 8(4): 931-952. doi: 10.3934/mbe.2011.8.931 |
[10] | Tingting Ding, Tongqian Zhang . Asymptotic behavior of the solutions for a stochastic SIRS model with information intervention. Mathematical Biosciences and Engineering, 2022, 19(7): 6940-6961. doi: 10.3934/mbe.2022327 |
Avian influenza is an animal infectious disease caused by the transmission of influenza A viruses. Influenza A viruses are divided into subtypes according to two proteins on the surface of the virus: Hemagglutinin (HA) and neuraminidase (NA) [1]. Most avian influenza viruses infect only certain species and do not infect humans. However, a few avian influenza viruses have crossed the species barrier to infect humans and even kill them, such as H5N1, H7N1, H7N2, H7N3, H7N7, H9N2 and H7N9. Among them, H5N1 is a highly pathogenic avian influenza virus, which was first detected in human in Hong Kong in 1997. After that, humans infection with avian influenza have occurred from time to time. As of December 2019, the global cumulative number of cases of human infection with H5N1 avian influenza arrives 861, with 455 deaths. Unlike H5N1, H7N9 is classified as a low pathogenicity avian influenza virus [2]. In March 2013, there was the first case of human infection with the H7N9 avian influenza virus in Shanghai, China. In the following weeks, this virus spread to several provinces and municipalities in mainland China. As of May 2017, H7N9 has resulted in 1263 human cases in China, of whom 459 died, with a mortality rate of nearly 37%. The frequent outbreak of avian influenza in the world not only brings a serious threat to human health, but also causes psychological panic and huge social impact, and brings a huge blow to the national economy. Therefore, it has been important to understand the dynamical behavior of avian influenza and to predict what may occur. Mathematical modeling has been a useful tool to describe the dynamical behavior of avian influenza and to obtain a better understanding of transmission mechanisms. Recently, many avian influenza models have been built from different perspectives (see [2,3,4,5,6,7,8,9,10,11,12] and references therein).
As we all know, there exist time delays during the spread of avian influenza, which can be used to describe not only the infection period of avian influenza virus in poultry (human) population, but also the incubation period of avian influenza in poultry (human) population and the immune period of recovered human to avian influenza. Therefore, the time delays should be considered such that the avian influenza models are more realistic. Generally speaking, delayed differential equations exhibit more complex dynamical behavior than differential equations without delay because time delay can make a stable equilibrium position to be unstable [13,14,15,16]. Consequently, it is of great interest to describe the transmission mechanism of avian influenza by introducing time delay into the models. For example, Liu et al. [7] and Kang et al. [12] established avian influenza models with different time delays in the poultry and human populations by considering the incubation periods of avian influenza virus and the survival probabilities of infected poultry and humans. By considering the existence of intracellular delay between initial infection of a cell and the release of new virus particles, Samanta [17] established a non-autonomous ordinary differential equation with distributed delay to characterize the spread of avian influenza between poultry and humans. These surveys imply that the research of time delay on avian influenza is a meaningful issue and is still open for study.
On the other hand, many existing literatures only focus on the deterministic avian influenza models that do not consider the impact of environmental noise. However, in the real world, the spread of avian influenza is often affected by the variations of environmental factors, such as humidity, temperature and so on [18,19]. Due to the fluctuations in the environment, an actual avian influenza system would not remain in a stable state, which would interfere with this stable state by acting directly on the density or indirectly affecting the parameter values. Therefore, it is of great significance to reveal the impact of environmental noise on avian influenza model by using stochastic model, so as to obtain more real benefits and accurately predict the future dynamics of avian influenza. To better understand the transmission dynamics of avian influenza, some authors have introduced stochastic perturbations into the deterministic models [20,21,22]. Zhang et al. [20] constructed a stochastic avian-human influenza model with logistic growth for avian population, and discussed the dynamical behavior of this model. Further, Zhang et al. [21] investigated a stochastic avian-human influenza epidemic model with psychological effect in human population and saturation effect within avian population. On the basis of the deterministic model established by Iwami et al. [3], Zhang et al. [22] established the corresponding stochastic model by introducing density disturbance. All the papers mentioned above only focused on the extinction and persistence of stochastic avian influenza models. However, to the best of our knowledge, there is no results related to the asymptotic behavior of stochastic avian influenza model around the equilibria of the corresponding deterministic model.
Motivated by the above discussions, in this paper, we investigate the asymptotic behavior of a stochastic delayed avian influenza model with saturated incidence rate. This work differs from existing results [7,12,17,20,21,22] in that (a) time delays and white noise are taken into account to describe the latency period of avian influenza virus in both poultry and human population and the environmental fluctuations; (b) asymptotic behavior of a stochastic delayed avian influenza model is studied. Overview of the rest of the article is as follows: In section 3, we show that there exists a unique global positive solution of system (2.3) with the given initial value (2.4). In section 4, we prove that the solution of system (2.3) is going around E0 under certain conditions. Further, we derive that the solution of system (2.3) is going around E∗ under certain conditions in section 5. In section 6, some numerical examples are introduced to illustrate the effectiveness of theoretic results. Finally, some conclusions are given in section 7.
Although the avian influenza virus spreads between wild birds and poultry, and between poultry and humans, we will only consider the transmission dynamics of avian influenza between poultry and humans because poultry is the main source of infection. Moreover, we assume that the virus is not spread between humans and mutate. We denote the total population of poultry and humans at time t by Na(t) and Nh(t), respectively. When the susceptible poultry contact with the infected poultry closely, there is usually no quick way to detect whether they are infected or the detection cost is too high, which makes it impossible to distinguish whether the close contacts of poultry are infected with the avian influenza virus. Therefore, the poultry population is divided into three sub-populations depending on the state of the disease: susceptible poultry Sa(t), exposed poultry Ea(t) and infected poultry Ia(t). The total poultry population at time t is denoted by Na(t)=Sa(t)+Ea(t)+Ia(t). The human population is divided into three sub-populations, which are susceptible human Sh(t), infected human with avian influenza Ia(t) and recovered human from avian influenza Rh(t). The total population of human at time t is given by Nh(t)=Sh(t)+Ih(t)+Rh(t).
The reason why we do not consider the exposed class for human population is that the close contacts of human beings are usually isolated and tested to determine whether they are infected with the avian influenza virus. The poultry in Ea either shows symptoms after incubation period and move to Ia, or always stays in Ea until natural death. The number of susceptible poultry (human) is increased by new recruitment, but decreases by natural death and infection (moving to class Ia (Ih)). The number of infected poultry (human) is increased by the infection of susceptible poultry (human) and reduced through natural and disease-related death. In addition, the number of infected humans is also reduced by recovery from the disease (moving to class Rh). Based on the above discussions, we obtain the schematic diagram of our model (see Figure 1).
The corresponding avian influenza model can be represented by the following equations:
{dSa(t)dt=Λa−μaSa(t)−βaSa(t)Ia(t)1+α1Ia(t),dEa(t)dt=βae−μaτaSa(t−τa)Ia(t−τa)1+α1Ia(t−τa)−(μa+γa)Ea(t),dIa(t)dt=γaEa(t)−(μa+δa)Ia(t),dSh(t)dt=Λh−μhSh(t)−βhSh(t)Ia(t)1+α2Ia(t),dIh(t)dt=βhe−μhτhSh(t−τh)Ia(t−τh)1+α2Ia(t−τh)−(μh+δh+θh)Ih(t),dRh(t)dt=θhIh(t)−μhRh(t). | (2.1) |
All parameters in model (2.1) are assumed non-negative and described in Table 1.
Parameter | Description |
Λa | new recruitment of the poultry populations |
Λh | new recruitment of the human population |
βa | the transmission rate from infective poultry to susceptible poultry |
βh | the transmission rate from infective poultry to susceptible human |
μa | the natural death rate of poultry populations |
μh | the natural death rate of human populations |
δa | the disease-related death rate of poultry populations |
δh | the disease-related death rate of humans populations |
γa | the transfer rate of exposed poultry to infected poultry |
θh | the recovery rate of the infective human |
αi(i=1,2) | parameters that measure the inhibitory effect |
Because the removed human populations Rh(t) has no effect on the dynamics of the first five equations, system (2.1) can be decoupled to the following system:
{dSa(t)dt=Λa−μaSa(t)−βaSa(t)Ia(t)1+α1Ia(t),dEa(t)dt=βae−μaτaSa(t−τa)Ia(t−τa)1+α1Ia(t−τa)−(μa+γa)Ea(t),dIa(t)dt=γaEa(t)−(μa+δa)Ia(t),dSh(t)dt=Λh−μhSh(t)−βhSh(t)Ia(t)1+α2Ia(t),dIh(t)dt=βhe−μhτhSh(t−τh)Ia(t−τh)1+α2Ia(t−τh)−(μh+δh+θh)Ih(t). | (2.2) |
A realistic avian influenza system would not remain in this stable state due to environmental fluctuations. In this paper, we will reveal how the environmental white noise affects the spread of avian influenza through investigating the dynamics of a stochastic delayed avian influenza model with saturated incidence rate. Taking the same approach as the literatures [23,24], we assume that the environmental white noise is directly proportional to the variables Sa(t), Ea(t), Ia(t), Sh(t) and Ih(t), respectively. Then, corresponding to system (2.2), the stochastic avian influenza model with time delay is of the following form
{dSa(t)=(Λa−μaSa(t)−βaSa(t)Ia(t)1+α1Ia(t))dt+σ1Sa(t)dB1(t),dEa(t)=(βae−μaτaSa(t−τa)Ia(t−τa)1+α1Ia(t−τa)−(μa+γa)Ea(t))dt+σ2Ea(t)dB2(t),dIa(t)=(γaEa(t)−(μa+δa)Ia(t))dt+σ3Ia(t)dB3(t),dSh(t)=(Λh−μhSh(t)−βhSh(t)Ia(t)1+α2Ia(t))dt+σ4Sh(t)dB4(t),dIh(t)=(βhe−μhτhSh(t−τh)Ia(t−τh)1+α2Ia(t−τh)−(μh+δh+θh)Ih(t))dt+σ5Ih(t)dB5(t), | (2.3) |
in which Bi(t)(i=1,2,⋯,5) are mutually independent standard Brownian motions defined on a complete probability space (Ω,F,P) with a filtration{Ft}t≥0 satisfying the usual conditions (i.e., it is increasing and right continuous while F0 contains all P-null sets), σi(i=1,2,⋯,5) denote the intensities of the white noises. The initial value of system (2.3) are
{Sa(θ)=φ1(θ),Ea(θ)=φ2(θ),Ia(θ)=φ3(θ),Sh(θ)=φ4(θ),Ih(θ)=φ5(θ),φi(θ)∈C([−τ,0],R5+),i=1,2,3,4,5,τ=max{τa,τh}, | (2.4) |
where C is the Banach space C([−τ,0];R5+) of continuous functions mapping the interval [−τ,0] into R5+, and R5+={x=(x1,x2,x3,x4,x5):xi>0,i=1,2,3,4,5}. By a biological meaning, we assume that φi(0)>0(i=1,2,3,4,5).
In this section, we prove that the solution of system (2.3) is global and positive for any initial value (2.4).
Theorem 1. For any initial value (2.4), system (2.3) has a unique positive solution (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t)) on t≥0 and the solution will remain in R5+ with probability one, in other words, (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t))∈R5+ for all t≥0 almost surely.
Proof. Since the coefficients of system (2.3) satisfy the local Lipschitz conditions, then for any initial value (2.4), there exists a unique local solution (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t)) on t∈[−τ,τe), where τe is the explosive time. To show this solution is global, we only need to show that τe=∞ a.s. To this end, let k0≥1 be sufficiently large such that (Sa(θ),Ea(θ),Ia(θ),Sh(θ),Ih(θ))(θ∈[−τ,0]) all lie within the interval [1k0,k0]. For each integer k≥k0, define the stopping time as
τk=inf{t∈[0,τe):Sa(t)∉(1k,k) or Ea(t)∉(1k,k) or Ia(t)∉(1k,k) or Sh(t)∉(1k,k) or Ih(t)∉(1k,k)}. |
We set inf∅=∞. Obviously, τk increasing when k→∞. Let τ∞=limk→∞τk, where τ∞≤τe a.s. If we can verify τ∞=∞ a.s., then τe=∞ and (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t))∈R5+ a.s. for all t≥0. That is to say, to complete the proof we only need to show that τ∞=∞ a.s. If this assertion is not true, then there is a pair of constants T>0 and ε∈(0,1) such that
P{τ∞≤T}>ε. |
There exists an integer k1≥k0 such that
P{τk≤T}≥ε for all k≥k1. | (3.1) |
Define a C2-function V: R5+→R+ by
V(Sa,Ea,Ia,Sh,Ih)=e−μaτa(Sa−a−alnSaa)+(Ea−1−lnEa)+(Ia−1−lnIa)+βae−μaτa∫tt−τaSa(s)Ia(s)1+α1Ia(s)ds+e−μhτh(Sh−b−blnShb)+(Ih−1−lnIh)+βhe−μhτh∫tt−τhSh(s)Ia(s)1+α2Ia(s)ds, |
in which a and b are positive constants to be determined later. The nonnegativity of this function can be derived from x−1−lnx≥0 for any x>0. Applying the Itô's formula to V, we get
dV=e−μaτa(1−aSa)dSa+e−μaτaa2S2a(dSa)2+(1−1Ea)dEa+a2E2a(dEa)2+(1−1Ia)dIa+a2I2a(dIa)2+βae−μaτaSaIa1+α1Ia−βae−μaτaSa(t−τa)Ia(t−τa)1+α1Ia(t−τa)+e−μhτh(1−bSh)dSh+e−μhτhb2S2h(dSh)2+(1−1Ih)dIh+12I2h(dIh)2+βhe−μhτhShIh1+α2Ia−βhe−μhτhSh(t−τh)Ia(t−τh)1+α2Ia(t−τh)=LVdt+e−μaτaσ1(Sa−a)dB1(t)+σ2(Ea−1)dB2(t)+σ3(Ia−1)dB3(t)+e−μhτhσ4(Sh−b)dB4(t)+σ5(Ih−1)dB5(t), | (3.2) |
where
LV=e−μaτa(1−aSa)(Λa−μaSa)−(1−1Ea)(μa+γa)Ea+(1−1Ia)(γaEa−(μa+δa)Ia)+e−μhτh(1−bSh)(Λh−μhSh)−(1−1Ih)(μh+δh+θh)Ih+e−μaτaaσ212+σ222+σ232+e−μhτhbσ242+σ252≤e−μaτaΛa+aμae−μaτa+aσ212e−μaτa+2μa+δa+γa+12σ22+12σ23+e−μhτhΛh+bμhe−μhτh+μh+δh+θh+bσ242e−μhτh+12σ25+(aβae−μaτa+bβhe−μhτh−(μa+δa))Ia. |
Choose a=μaeμaτaβa and b=δaeμhτhβh or a=δaeμaτaβa and b=μaeμhτhβh such that
aβae−μaτa+bβhe−μhτh−(μa+δa)=0. |
Then, we can get
LV(Sa,Ea,Ia,Sh,Ih)≤e−μaτaΛa+aμae−μaτa+e−μhτhΛh+bμhe−μhτh+2μa+γa+δa+μh+δh+θh+aσ212e−μaτa+bσ242e−μhτh+12(σ22+σ23+σ25)=:K, |
where K is a positive constant. It thus follows from (3.2) that
dV(Sa,Ea,Ia,Sh,Ih)≤Kdt+e−μaτaσ1(Sa−a)dB1(t)+σ2(Ea−1)dB2(t)+σ3(Ia−1)dB3(t)+e−μhτhσ4(Sh−b)dB4(t)+σ5(Ih−1)dB5(t). | (3.3) |
Integrating both sides of (3.3) from 0 to τk∧T=min{τk,T} and then taking the expectation results in
EV(Sa(τk∧T),Ea(τk∧T),Ia(τk∧T),Sh(τk∧T),Ih(τk∧T))≤V(Sa(0),Ea(0),Ia(0),Sh(0),Ih(0))+KE(τk∧T)≤V(Sa(0),Ea(0),Ia(0),Sh(0),Ih(0))+KT. | (3.4) |
Set Ωk={τk≤T} for k≥k1, and according to (3.1), we have P(Ωk)≥ε. For every ω∈Ωk, there exists Sa(τk,ω) or Ea(τk,ω) or Ia(τk,ω) or Sh(τk,ω) or Ih(τk,ω) equals either k or 1k. Therefore, V(Sa(τk,ω),Ea(τk,ω),Ia(τk,ω),Sh(τk,ω),Ih(τk,ω)) is no less either k−1−lnk or 1k−1−ln1k or k−a−alnka or 1k−a+alnak or k−b−blnkb or 1k−b+blnbk.
Therefore, we have
V(Sa(τk,ω),Ea(τk,ω),Ia(τk,ω),Sh(τk,ω),Ih(τk,ω))≥(k−1−lnk)∧(1k−1+lnk)∧(k−a−alnka)∧(1k−a+alnak)∧(k−b−blnkb)∧(1k−b+blnbk). |
It follows from (3.4) that
V(Sa(0),Ea(0),Ia(0),Sh(0),Ih(0))+KT≥E[1ΩkV(Sa(τk,ω),Ea(τk,ω),Ia(τk,ω),Sh(τk,ω),Ih(τk,ω))]≥ε[(k−1−lnk)∧(1k−1+lnk)∧(k−a−alnka)∧(1k−a+alnak)∧(k−b−blnkb)∧(1k−b+blnbk)], |
where 1Ωk denotes the indicator function of Ωk. Letting k→∞, then
∞>V(Sa(0),Ea(0),Ia(0),Sh(0),Ih(0))+KT=∞, |
which leads to the contradiction. This completes the proof.
In this section, we will investigate the solution of system (2.3) around disease-free equilibrium E0 under certain conditions. It is worthwhile to mention that, if R0=βaγaΛae−μaτaμa(μa+δa)(μa+γa)<1, the deterministic system (2.2) is globally asymptotically stable around the unique disease-free equilibrium E0=(S0a,0,0,S0h,0)=(Λaμa,0,0,Λhμh,0), but E0 is not the equilibrium of the stochastic system (2.3). Thus, the result concerning the solution of stochastic system (2.3) around E0 is presented by the following theorem.
Theorem 2. Let (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t)) be the solution of system (2.3) with the initial value (2.4). If R0<1 and the following conditions hold
σ21<μa,σ22<μa+γa,σ23<μa+δa,σ24<μh,σ25<μh+δh+θh, |
then,
lim supt→∞1tE∫t0(Sa−Λaμa)2ds≤σ21Λ2aμ2a(μa−σ21),lim supt→∞1tE∫t0(E2a+I2a)ds≤P1M1,lim supt→∞1tE∫t0(Sh−Λhμh)2ds≤Λ2hμ2h(μh−σ24)(σ24+βhα2),lim supt→∞1tE∫t0I2hds≤P2, |
where
M1=min{μa+γa−σ224,(μa+γa−σ22)(μa+δa−σ23)(μa+δa)4γ2a},P1=e−2μaτaσ21Λ2aμ2a[1μa−σ21(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)+1],P2=2e−2μhτhΛhμ2h(μh+δh+θh−σ25)[α2σ24+βhα2(μh−σ24)(2μ2h+2μhδh+2μhθh+(δh+θh)22(μh+δh+θh)+σ24)+σ24]. |
Proof. Since (S0a,0,0,S0h,0) is the disease-free equilibrium of system (2.2), then
Λa=μaS0a,Λh=μhS0h. |
According to system (2.3), we can obtain that
dSa(t)=[−μa(Sa−Λaμa)−βaSaIa1+α1Ia]dt+σ1SadB1(t)=[−μa(Sa−Λaμa)−βa(Sa−Λaμa)Ia1+α1Ia−βaΛaμaIa1+α1Ia]dt+σ1SadB1(t), | (4.1) |
and
d[Ea(t+τa)+μa+γaγaIa(t+τa)]=dEa(t+τa)+μa+γaγadIa(t+τa)≤[βae−μaτa(Sa−Λaμa)Ia1+α1Ia−(μa+γa)(μa+δa)γaIa(t+τa)+βae−μaτaΛaμaIa]dt+σ2Ea(t+τa)dB2(t)+σ3(μa+γa)γaIa(t+τa)dB3(t)≤[βae−μaτa(Sa−Λaμa)Ia1+α1Ia+(μa+γa)(μa+δa)γa(Ia(t)−Ia(t+τa))]dt+σ2Ea(t+τa)dB2(t)+σ3(μa+γa)γaIa(t+τa)dB3(t). | (4.2) |
Let V1=12(Sa−Λaμa)2, then applying the Itô's formula to V1, together with (4.1), we have
dV1=[(Sa−Λaμa)(−μa(Sa−Λaμa)−βa(Sa−Λaμa)Ia1+α1Ia−βaΛaμaIa1+α1Ia)+12σ21S2a]dt+σ1Sa(Sa−Λaμa)dB1(t)=[−μa(Sa−Λaμa)2−βa(Sa−Λaμa)2Ia1+α1Ia−βaΛaμa(Sa−Λaμa)Ia1+α1Ia+12σ21S2a]dt+σ1Sa(Sa−Λaμa)dB1(t)=:LV1dt+σ1Sa(Sa−Λaμa)dB1(t), |
where
LV1≤−μa(Sa−Λaμa)2−βaΛaμa(Sa−Λaμa)Ia1+α1Ia+σ21(Sa−Λaμa)2+σ21Λ2aμ2a=−(μa−σ21)(Sa−Λaμa)2−βaΛaμa(Sa−Λaμa)Ia1+α1Ia+σ21Λ2aμ2a. | (4.3) |
Similarly, let V2=Ea(t+τa)+μa+γaγaIa(t+τa)+(μa+γa)(μa+δa)γa∫t+τatIa(s)ds, it follows from (4.2) that
dV2≤βae−μaτa(Sa−Λaμa)Ia1+α1Ia+σ2Ea(t+τa)dB2(t)+σ3(μa+γa)γaIa(t+τa)dB3(t). |
Define ˉV=e−μaτaV1+ΛaμaV2, then
dˉV≤[−e−μaτa(μa−σ21)(Sa−Λaμa)2+e−μaτaσ21Λ2aμ2a]dt+σ1Sa(Sa−Λaμa)dB1(t)+σ2Ea(t+τa)dB2(t)+σ3(μa+γa)γaIa(t+τa)dB3(t). | (4.4) |
Integrating both sides of (4.4) from 0 to t and taking expectation, we get
EˉV(t)−EˉV(0)≤−e−μaτa(μa−σ21)E∫t0(Sa−Λaμa)2ds+e−μaτaσ21Λ2aμ2at. |
Therefore, we can obtain
lim supt→∞1tE∫t0(Sa−Λaμa)2ds≤σ21Λ2aμ2a(μa−σ21). |
Similarly, we define
V3=12[e−μaτa(Sa−Λaμa)+Ea(t+τa)]2, |
then,
LV3=−e−2μaτaμa(Sa−Λaμa)2−e−μaτa(2μa+γa)(Sa−Λaμa)Ea(t+τa)−(μa+γa)E2a(t+τa)+12e−2μaτaσ21S2a+12σ22E2a(t+τa)≤−e−2μaτaμa(Sa−Λaμa)2+μa+γa2E2a(t+τa)+(2μa+γa)2e−2μaτa2(μa+γa)(Sa−Λaμa)2−(μa+γa)E2a(t+τa)+e−2μaτaσ21(Sa−Λaμa)2+e−2μaτaσ21Λ2aμ2a+12σ22E2a(t+τa)=e−2μaτa(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)(Sa−Λaμa)2−12(μa+γa−σ22)E2a(t+τa)+e−2μaτaσ21Λ2aμ2a. |
Let V4=V3+12(μa+γa−σ22)∫t+τatE2a(s)ds, we get
LV4≤e−2μaτa(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)(Sa−Λaμa)2−12(μa+γa−σ22)E2a+e−2μaτaσ21Λ2aμ2a. |
Let V5=12I2a, the derivative of V5 can be calculated as
LV5=γaEaIa−(μa+δa)I2a+12σ23I2a≤μa+δa2I2a+γ2a2(μa+δa)E2a−(μa+δa)I2a+12σ23I2a=γ2a2(μa+δa)E2a−12(μa+δa−σ23)I2a. |
The Young's inequality is used above. Let
˜V=V4+e−μaτaμa−σ21(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)ˉV+(μa+γa−σ22)(μa+δa)2γ2aV5, |
which implies that
L˜V≤−12(μa+γa−σ22)E2a+e−2μaτaσ21Λ2aμ2a+e−2μaτaσ21Λ2aμ2a(μa−σ21)(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)+14(μa+γa−σ22)E2a−(μa+γa−σ22)(μa+δa−σ23)(μa+δa)4γ2aI2a=−14(μa+γa−σ22)E2a−(μa+γa−σ22)(μa+δa−σ23)(μa+δa)4γ2aI2a+e−2μaτaσ21Λ2aμ2a[1μa−σ21(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)+1]. | (4.5) |
Integrating both sides of (4.5) from 0 to t and then taking expectation yields
E˜V(t)−E˜V(0)≤−14(μa+γa−σ22)E∫t0E2a(s)ds−(μa+γa−σ22)(μa+δa−σ23)(μa+δa)4γ2aE∫t0I2a(s)ds+e−2μaτaσ21Λ2aμ2a[1μa−σ21(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)+1]t. |
Consequently, we can obtain
lim supt→∞1tE∫t0(E2a(s)+I2a(s))ds≤P1M1, |
where M1 and P1 are defined in Theorem 2. Further, according to system (2.3), we have
dSh(t)=[−μh(Sh−Λhμh)−βhShIa1+α2Ia]dt+σ4ShdB4(t)=[−μh(Sh−Λhμh)−(Sh−Λhμh)βhIa1+α2Ia−βhΛhIaμh(1+α2Ia)]dt+σ4ShdB4(t), | (4.6) |
and
dIh(t+τh)=[βhe−μhτhShIa1+α2Ia−(μh+δh+θh)Ih(t+τh)]dt+σ5Ih(t+τh)dB5(t)≤[βhe−μhτhIa1+α2Ia(Sh−Λhμh)+βhΛhe−μhτhα2μh−(μh+δh+θh)Ih(t+τh)]dt+σ5Ih(t+τh)dB5(t). | (4.7) |
Let V6=12(Sh−Λhμh)2. Noting (4.6), we have
LV6=−μh(Sh−Λhμh)2−βh(Sh−Λhμh)2Ia1+α2Ia−βhΛhμh(Sh−Λhμh)Ia1+α2Ia+12σ24S2h≤−μh(Sh−Λhμh)2−βhΛhμh(Sh−Λhμh)Ia1+α2Ia+σ24(Sh−Λhμh)2+σ24Λ2hμ2h=−(μh−σ24)(Sh−Λhμh)2−βhΛhμh(Sh−Λhμh)Ia1+α2Ia+σ24Λ2hμ2h. |
Let V7=e−μhτhV6+ΛhμhIh(t+τh), it follows from (4.7) that
LV7≤−e−μhτh(μh−σ24)(Sh−Λhμh)2+e−μhτhσ24Λ2hμ2h+βhΛ2he−μhτhα2μ2h−Λhμh(μh+δh+θh)Ih(t+τh)≤−e−μhτh(μh−σ24)(Sh−Λhμh)2+e−μhτhΛ2hμ2h(σ24+βhα2). | (4.8) |
Integrating both sides of (4.8) from 0 to t and then taking the expectation yields
EV7(t)−EV7(0)≤−e−μhτh(μh−σ24)E∫t0(Sh−Λhμh)2ds+e−μhτhΛ2hμ2h(σ24+βhα2)t, |
therefore, we can get
lim supt→∞1tE∫t0(Sh−Λhμh)2ds≤Λ2hμ2h(μh−σ24)(σ24+βhα2). |
Let V8=12[e−μhτh(Sh−Λhμh)+Ih(t+τh)]2, then
LV8=(e−μhτh(Sh−Λhμh)+Ih(t+τh))[e−μhτh(Λh−μhSh)−(μh+δh+θh)Ih(t+τh)]+12e−2μhτhσ24S2h+12σ25I2h(t+τh)≤−e−2μhτhμh(Sh−Λhμh)2+(2μh+δh+θh)2e−2μhτh2(μh+δh+θh)(Sh−Λhμh)2+μh+δh+θh2I2h(t+τh)−(μh+δh+θh)I2h(t+τh)+e−2μhτhσ24(Sh−Λhμh)2+e−2μhτhσ24Λ2hμ2h+12σ25I2h(t+τh)=e−2μhτh(2μ2h+2μhδh+2μhθh+(δh+θh)22(μh+δh+θh)+σ24)(Sh−Λhμh)2−12(μh+δh+θh−σ25)I2h(t+τh)+e−2μhτhσ24Λ2hμ2h. |
Defining
V9=V8+e−μhτhμh−σ24(2μ2h+2μhδh+2μhθh+(δh+θh)22(μh+δh+θh)+σ24)V7+12(μh+δh+θh−σ25)∫t+τhtI2h(s)ds, |
we get
LV9≤−12(μh+δh+θh−σ25)I2h+e−2μhτhΛ2hμ2h[1μh−σ24(σ24+βhα2)(2μ2h+2μhδh+2μhθh+(δh+θh)22(μh+δh+θh)+σ24)+σ24]. | (4.9) |
Integrating both sides of (4.9) from 0 to t and taking expectation, we obtain
EV9(t)−EV9(0)≤−12(μh+δh+θh−σ25)E∫t0I2h(s)ds+e−2μhτhΛ2hμ2h[1μh−σ24(σ24+βhα2)(2μ2h+2μhδh+2μhθh+(δh+θh)22(μh+δh+θh)+σ24)+σ24]t. |
Consequently, we can obtain
lim supt→∞E∫t0I2h(s)ds≤P2, |
where P2 is defined in Theorem 2. This completes the proof.
If R0>1, there exists an endemic equilibrium E∗=(S∗a,E∗a,I∗a,S∗h,I∗h) of system (2.2), but it is not the equilibrium of system (2.3), where S∗a=Λa(1+α1I∗a)μa(1+α1I∗a)+βaI∗a, E∗a=βaΛae−μaτaI∗a(μa+γa)[μa(1+α1I∗a)+βaI∗a], I∗a=μa(R0−1)α1μa+βa, S∗h=Λh(1+α2I∗a)μh(1+α2I∗a)+βhI∗a, E∗a=βhe−μhτhS∗hI∗a(μh+δh+θh)(1+α2I∗a). In this section, we show that the solution of system (2.3) is going around E∗ under certain conditions.
Theorem 3. Let (Sa(t),Ea(t),Ia(t),Sh(t),Ih(t)) be the solution of system (2.3) with initial value (2.4). If R0>1 and the following conditions hold
(i) σ21<μa,σ22<12(μa+γa),σ23<12(μa+δa),σ24<μh,σ25<μh+δh+θh;
(ii) max(√P3,√P4,√P5,√P6)<d(E∗,E0),
then
lim supt→∞1tE∫t0(Sa−S∗a)2ds≤P3,lim supt→∞E∫t0[(Ea(s)−E∗a)2+(Ia(s)−I∗a)2]ds≤L1L2=:P4,lim supt→∞E∫t0(Sh−S∗h)2ds≤P5,lim supt→∞E∫t0(Ih−I∗h)2ds≤P6, |
where
d(E∗,E0)=√(S∗a−Λaμa)2+(E∗a)2+(I∗a)2+(S∗h−Λhμh)2+(I∗h)2P3=1μa−σ21[σ21(S∗a)2+σ21S∗aL32μa+(eμaτaS∗a+L3μae−μaτa)(12σ22E∗a+μa+γa2γaσ23I∗a)],P4=L1L2,P5=σ24(S∗h)2μh−σ24,P6=σ24L24(μh−σ24)(μh+δh+θh−σ25)2+2σ25(I∗h)2μh+δh+θh−σ25,L1=e−μaτaμa−σ21(2μ2a+2μaγa+γ2a2(μa+γa)+σ21)[σ21(S∗a)2+σ21S∗aL32μa+(eμaτaS∗a+L3μae−μaτa)(12σ22E∗a+12μa+γaγaσ23I∗a)]+e−2μaτaσ21(S∗a)2+σ22(E∗a)2+σ23(μa+δa)(μa+γa−2σ22)2γ2a(I∗a)2,L2=min{14(μa+γa−2σ22),(μa+δa)(μa+γa−2σ22)(μa+δa−2σ23)4γ2a},L3=βaS∗aI∗a1+α1I∗a,L4=β∗hS∗hI∗a1+α2I∗a. |
Proof. Since (S∗a,E∗a,I∗a,S∗h,I∗h) is the interior equilibrium of system (2.2), then
Λa=μaS∗a+βaS∗aI∗a1+α1I∗a,(μa+γa)E∗a=βae−μaτaS∗aI∗a1+α1I∗a,I∗aE∗a=γaμa+δa,Λh=μhS∗h+βhS∗hI∗a1+α2I∗a,(μh+δh+θh)I∗h=βhe−μhτhS∗hI∗a1+α2I∗a. | (5.1) |
Define the Lyapunov function W1 as W1=Sa−S∗a−S∗alnSaS∗a, from which we have
dW1=(Λa−μaSa−βaSaIa1+α1Ia−ΛaS∗aSa+μaS∗a+βaS∗aIa1+α1Ia+12S∗aσ21)dt+σ1(Sa−S∗a)dB1(t)=[(μaS∗a+βaS∗aI∗a1+α1I∗a)(2−S∗aSa−SaS∗a)+βaS∗aI∗a1+α1I∗a(−SaIa(1+α1I∗a)S∗aI∗a(1+α1Ia)+SaS∗a+Ia(1+α1I∗a)I∗a(1+α1Ia)−1)+12S∗aσ21]dt+σ1(Sa−S∗a)dB1(t)=LW1dt+σ1(Sa−S∗a)dB1(t), |
where
LW1=−(μa+βaI∗a1+α1I∗a)(Sa−S∗a)2Sa−βa(Sa−S∗a)(Ia1+α1Ia−I∗a1+α1I∗a)+12S∗aσ21. | (5.2) |
Similarly, we can define W2 as
W2=Ea(t+τa)−E∗a−E∗alnEa(t+τa)E∗a+μa+γaγa(Ia(t+τa)−I∗a−I∗alnIa(t+τa)I∗a). |
By using the Itô's formula, the derivative of W2 is calculated as follows
LW2=(1−E∗aEa(t+τa))(βae−μaτaSaIa1+α1Ia−(μa+γa)Ea(t+τa))+μa+γaγa(1−I∗aIa(t+τa))(γaEa(t+τa)−(μa+δa)Ia(t+τa))+12σ22E∗a+μa+γa2γaσ23I∗a=βae−μaτaI∗a1+α1I∗a(Sa−S∗a)(1+α1I∗aI∗aIa1+α1Ia−1)+βae−μaτaS∗aI∗a1+α1I∗a(SaS∗a−1+α1I∗aS∗aI∗aSaIa1+α1IaE∗aEa(t+τa)+1+α1I∗aI∗aIa1+α1Ia−Ia(t+τa)I∗a−Ea(t+τa)E∗aI∗aIa(t+τa))+12σ22E∗a+12μa+γaγaσ23I∗a. | (5.3) |
Since x−1−lnx≥0 for x>0, the following estimate can be obtained
1+α1I∗aS∗aI∗aSaIa1+α1IaE∗aEa(t+τa)≥1+ln(1+α1I∗aS∗aI∗aSaIa1+α1IaE∗aEa(t+τa))=1+lnSaS∗a−lnIa(t+τa)I∗a+lnIa(1+α1I∗a)I∗a(1+α1Ia)−lnEa(t+τa)E∗aI∗aIa(t+τa). | (5.4) |
Substituting (5.4) into (5.3), we can get
LW2≤βae−μaτaI∗a1+α1I∗a(Sa−S∗a)(1+α1I∗aI∗aIa1+α1Ia−1)+βae−μaτaS∗aI∗a1+α1I∗a(SaS∗a−1−lnSaS∗a+lnIa(t+τa)I∗a−lnIa(1+α1I∗a)I∗a(1+α1Ia)+lnEa(t+τa)E∗aI∗aIa(t+τa)+1+α1I∗aI∗aIa1+α1Ia−Ia(t+τa)I∗a−Ea(t+τa)E∗aI∗aIa(t+τa))+12σ22E∗a+12μa+γaγaσ23I∗a=βae−μaτaI∗a1+α1I∗a(Sa−S∗a)(1+α1I∗aI∗aIa1+α1Ia−1)+βae−μaτaS∗aI∗a1+α1I∗a[(SaS∗a−lnSaS∗a)−(Ia(t+τa)I∗a−lnIa(t+τa)I∗a)+(Ia(1+α1I∗a)I∗a(1+α1Ia)−lnIa(1+α1I∗a)I∗a(1+α1Ia))−(Ea(t+τa)E∗aI∗aIa(t+τa)−lnEa(t+τa)E∗aI∗aIa(t+τa))−1]+12σ22E∗a+12μa+γaγaσ23I∗a. | (5.5) |
Choose W3=W2+βae−μaτaS∗aI∗a1+α1I∗a∫t+τat(Ia(s)I∗a−lnIa(s)I∗a−1)ds. Therefore, LW3 can be obtained as follows by using (5.5):
LW3≤βae−μaτaI∗a1+α1I∗a(Sa−S∗a)(Ia(1+α1I∗a)I∗a(1+α1Ia)−1)+βae−μaτaS∗aI∗a1+α1I∗a[(SaS∗a−lnSaS∗a)−(Ia(t+τa)I∗a−lnIa(t+τa)I∗a)+(Ia(1+α1I∗a)I∗a(1+α1Ia)−lnIa(1+α1I∗a)I∗a(1+α1Ia))−(Ea(t+τa)E∗aI∗aIa(t+τa)−lnEa(t+τa)E∗aI∗aIa(t+τa))−1]+12σ22E∗a+12μa+γaγaσ23I∗a+βae−μaτaS∗aI∗a1+α1I∗a(Ia(t+τa)I∗a−lnIa(t+τa)I∗a−1)−βae−μaτaS∗aI∗a1+α1I∗a(IaI∗a−lnIaI∗a−1)≤βae−μaτaI∗a1+α1I∗a(Sa−S∗a)(Ia(1+α1I∗a)I∗a(1+α1Ia)−1)+βae−μaτaS∗aI∗a1+α1I∗a[SaS∗a+S∗aSa−1−IaI∗a+Ia(1+α1I∗a)I∗a(1+α1Ia)+lnI∗a(1+α1Ia)Ia(1+α1I∗a)IaI∗a−1]+12σ22E∗a+12μa+γaγaσ23I∗a. | (5.6) |
Noting that x−1−lnx≥0 holds for x>0, we also have
−IaI∗a+Ia(1+α1I∗a)I∗a(1+α1Ia)+lnI∗a(1+α1Ia)Ia(1+α1I∗a)IaI∗a≤−IaI∗a+Ia(1+α1I∗a)I∗a(1+α1Ia)+I∗a(1+α1Ia)Ia(1+α1I∗a)IaI∗a−1≤I∗a(1+α1Ia)Ia(1+α1I∗a)IaI∗a(Ia(1+α1I∗a)I∗a(1+α1Ia)I∗aIa−1)(Ia(1+α1I∗a)I∗a(1+α1Ia)−1)=(1+α1Ia)(1+α1I∗a)I∗a(11+α1Ia−11+α1I∗a)(Ia1+α1Ia−I∗a1+α1I∗a)<0, | (5.7) |
substituting (5.7) into (5.6) and using SaS∗a+S∗aSa−2=(Sa−S∗a)2SaS∗a, we know that
\begin{equation} LW_3 \leq \frac{\beta_a e^{-\mu_a \tau_a} I_a^*}{1+\alpha_1 I_a^*} (S_a-S_a^*) \left( \frac{I_a(1+\alpha_1 I_a^*)}{I_a^*(1+\alpha_1 I_a)} -1 \right) + \frac{\beta_a e^{-\mu_a \tau_a} I_a^*}{1+\alpha_1 I_a^*} \frac{(S_a -S_a^*)^2}{S_a} + \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^*. \end{equation} | (5.8) |
Let W_4 = W_1 + \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left(\mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right)W_3 . Applying the Itô's formula, together with (5.2) and (5.8), derives that
\begin{equation} \begin{split} LW_4 = & LW_1 + \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right)LW_3\\ \leq & -\left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right)\frac{(S_a -S_a^*)^2}{S_a} - \beta_a (S_a -S_a^*) \left( \frac{I_a}{1+\alpha_1 I_a} - \frac{I_a^*}{1+\alpha_1 I_a^*} \right)+ \frac{1}{2} \sigma_1 ^2S_a^* \\ & + \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \Bigg[ \frac{\beta_a e^{-\mu_a \tau_a} I_a^*}{1+\alpha_1 I_a^*} (S_a-S_a^*) \left( \frac{I_a(1+\alpha_1 I_a^*)}{I_a^*(1+\alpha_1 I_a)} -1 \right) \\ &+ \frac{\beta_a e^{-\mu_a \tau_a} I_a^*}{1+\alpha_1 I_a^*} \frac{(S_a -S_a^*)^2}{S_a} + \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \Bigg]\\ = & \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) (S_a-S_a^*) \left( \frac{I_a(1+\alpha_1 I_a^*)}{I_a^*(1+\alpha_1 I_a)} -1 \right) - \beta_a (S_a-S_a^*) \bigg(\frac{ I_a}{1+\alpha_1 I_a} \\ &-\frac{I_a^*}{1+\alpha_1 I_a^*} \bigg) + \frac{1}{2} \sigma_1 ^2S_a^* + \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right) \\ = & (S_a-S_a^*) \left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) \left[ \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \frac{1+\alpha_1 I_a^*}{I_a^*} - \beta_a\right] + \frac{1}{2} \sigma_1 ^2S_a^*\\ & + \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right) \\ = & \frac{\mu_a(1+\alpha_1 I_a^*)}{I_a^*} (S_a-S_a^*) \left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) + \frac{1}{2} \sigma_1 ^2S_a^*\\ &+ \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right). \\ \end{split} \end{equation} | (5.9) |
Choose Lyapunov function W_5 as W_5 = \frac{(S_a-S_a^*)^2}{2} , then its derivative is
\begin{align*} LW_5 = & (S_a-S_a^*)\left[\Lambda_a -\mu_a S_a - \frac{\beta_a S_a I_a}{1+\alpha_1 I_a} \right] + \frac{1}{2} \sigma_1 ^2S_a^2\\ = &(S_a-S_a^*)\left[\mu_aS_a^*-\mu_a S_a + \frac{\beta_a S_a^* I_a^*}{1+\alpha_1 I_a^*}- \frac{\beta_a S_a I_a}{1+\alpha_1 I_a} \right] + \frac{1}{2} \sigma_1 ^2S_a^2\\ = & -\mu_a (S_a-S_a^*)^2 - \beta_a S_a^* (S_a-S_a^*)\left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) - \beta_a (S_a-S_a^*)^2 \frac{ I_a}{1+\alpha_1 I_a} + \frac{1}{2} \sigma_1 ^2S_a^2\\ \leq & -\mu_a (S_a-S_a^*)^2 - \beta_a S_a^* (S_a-S_a^*)\left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) + \sigma_1 ^2 (S_a-S_a^*)^2 + \sigma_1 ^2 (S_a^*)^2\\ = &-(\mu_a - \sigma_1 ^2) (S_a-S_a^*)^2 - \beta_a S_a^* (S_a-S_a^*)\left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) + \sigma_1 ^2 (S_a^*)^2. \end{align*} |
Let \bar{W} = W_5 + \frac{\beta_a S_a^* I_a^*}{\mu_a(1+\alpha_1 I_a^*)}W_4 , one can derive that
\begin{equation} \begin{split} L\bar{W} \leq&-(\mu_a - \sigma_1 ^2) (S_a-S_a^*)^2 - \beta_a S_a^* (S_a-S_a^*)\left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) + \sigma_1 ^2 (S_a^*)^2\\ & + \frac{\beta_a S_a^* I_a^*}{\mu_a(1+\alpha_1 I_a^*)}\bigg[ \frac{\mu_a(1+\alpha_1 I_a^*)}{I_a^*} (S_a-S_a^*) \left(\frac{ I_a}{1+\alpha_1 I_a} -\frac{I_a^*}{1+\alpha_1 I_a^*} \right) + \frac{1}{2} \sigma_1 ^2S_a^*\\ &+ \frac{1+\alpha_1 I_a^*}{\beta_a e^{-\mu_a \tau_a} I_a^*} \left( \mu_a + \frac{\beta_a I_a^*}{1+\alpha_1 I_a^*} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right)\bigg]\\ = & -(\mu_a - \sigma_1 ^2) (S_a-S_a^*)^2 + \sigma_1 ^2 (S_a^*)^2 + \frac{\beta_a S_a^* I_a^*}{2\mu_a(1+\alpha_1 I_a^*)}\sigma_1 ^2 S_a^* \\ & + \left(e^{\mu_a \tau_a}S_a^* +\frac{\beta_a S_a^* I_a^*}{\mu_ae^{-\mu_a \tau_a}(1+\alpha_1 I_a^*)} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right). \end{split} \end{equation} | (5.10) |
Integrating both sides of (5.10) from 0 to t and then taking expectation yields
\begin{equation*} \begin{split} E\bar{W}(t) -E\bar{W}(0) \leq & -(\mu_a - \sigma_1 ^2) E\int_0^t (S_a(s)-S_a^*)^2 ds + \bigg[\sigma_1 ^2 (S_a^*)^2 +\frac{\beta_a S_a^* I_a^*}{2\mu_a(1+\alpha_1 I_a^*)}\sigma_1 ^2 S_a^* \\ & + \left(e^{\mu_a \tau_a}S_a^* +\frac{\beta_a S_a^* I_a^*}{\mu_ae^{-\mu_a \tau_a}(1+\alpha_1 I_a^*)} \right) \left( \frac{1}{2} \sigma_2 ^2E_a^* + \frac{1}{2} \frac{\mu_a+\gamma_a}{\gamma_a}\sigma_3 ^2I_a^* \right)\bigg] t. \end{split} \end{equation*} |
Then, we can get
\begin{equation*} \limsup\limits_{t\rightarrow\infty} \frac{1}{t}E\int_0^t (S_a(s)-S_a^*)^2 ds \leq P_{3}, \end{equation*} |
where P_3 is defined in Theorem 3. Defining W_6 = \frac{1}{2} \left[ e^{-\mu_a \tau_a}(S_a-S_a^*) + E_a(t+\tau_a) -E_a^* \right]^2 , the use of Itô's formula yields that
\begin{align*} LW_6 = & -\mu_a e^{-2\mu_a \tau_a} (S_a - S_a^*)^2 - (\mu_a+\gamma_a) (E_a(t+\tau_a) -E_a^*)^2 - (2\mu_a+\gamma_a) e^{-\mu_a \tau_a}(S_a - S_a^*)\\ &(E_a(t+\tau_a) -E_a^*) + \frac{1}{2} e^{-2\mu_a \tau_a}\sigma_1 ^2S_a^2 + \frac{1}{2} \sigma_2 ^2E_a^2(t+\tau_a)\\ \leq & -\mu_a e^{-2\mu_a \tau_a} (S_a - S_a^*)^2 - (\mu_a+\gamma_a) (E_a(t+\tau_a) -E_a^*)^2 + \frac{\mu_a+\gamma_a}{2} (E_a(t+\tau_a) -E_a^*)^2 \\ &+ \frac{(2\mu_a+\gamma_a)^2 e^{-2\mu_a \tau_a}}{2(\mu_a+\gamma_a)} (S_a - S_a^*)^2 + e^{-2\mu_a \tau_a} \sigma_1 ^2 (S_a - S_a^*)^2 + e^{-2\mu_a \tau_a} \sigma_1 ^2 (S_a^*)^2 \\ &+ \sigma_2 ^2 (E_a(t+\tau_a) -E_a^*)^2 + \sigma_2 ^2 (E_a^*)^2\\ = & e^{-2\mu_a \tau_a} \left( \frac{2\mu_a^2 +2\mu_a\gamma_a +\gamma_a^2}{2(\mu_a +\gamma_a)} +\sigma_1 ^2 \right)(S_a - S_a^*)^2 -\left( \frac{\mu_a+\gamma_a}{2} -\sigma_2 ^2 \right) (E_a(t+\tau_a) -E_a^*)^2\\ &+ e^{-2\mu_a \tau_a} \sigma_1 ^2 (S_a^*)^2 + \sigma_2 ^2 (E_a^*)^2. \end{align*} |
Let W_7 = W_6 + \left(\frac{\mu_a+\gamma_a}{2} -\sigma_2 ^2 \right) \int_t^{t+\tau_a}(E_a(s) -E_a^*)^2ds and W_8 = \frac{1}{2} (I_a -I_a^*)^2 . We have
\begin{equation} \begin{split} LW_7 \leq & e^{-2\mu_a \tau_a} \left( \frac{2\mu_a^2 +2\mu_a\gamma_a +\gamma_a^2}{2(\mu_a +\gamma_a)} +\sigma_1 ^2 \right)(S_a - S_a^*)^2 -\left( \frac{\mu_a+\gamma_a}{2} -\sigma_2 ^2 \right) (E_a -E_a^*)^2\\ &+ e^{-2\mu_a \tau_a} \sigma_1 ^2 (S_a^*)^2 + \sigma_2 ^2 (E_a^*)^2, \end{split} \end{equation} | (5.11) |
and
\begin{equation} \begin{split} LW_8 = & (I_a -I_a^*)(\gamma_aE_a -(\mu_a+\delta_a)I_a) + \frac{1}{2} \sigma_3 ^2 I_a^2\\ = & \gamma_a(E_a -E_a^*)(I_a -I_a^*) - (\mu_a+\delta_a)(I_a-I_a^*)^2 + \frac{1}{2} \sigma_3 ^2 I_a^2\\ \leq & \frac{\mu_a+\delta_a}{2} (I_a-I_a^*)^2 + \frac{\gamma_a^2}{2(\mu_a+\delta_a)} (E_a -E_a^*)^2 \\ &- (\mu_a+\delta_a)(I_a-I_a^*)^2 + \sigma_3 ^2 (I_a-I_a^*)^2 + \sigma_3 ^2(I_a^*)^2\\ = & \frac{\gamma_a^2}{2(\mu_a+\delta_a)} (E_a -E_a^*)^2 -\left( \frac{\mu_a+\delta_a}{2} -\sigma_3 ^2 \right) (I_a-I_a^*)^2 + \sigma_3 ^2(I_a^*)^2.\\ \end{split} \end{equation} | (5.12) |
Let \tilde{W} = W_7+ \frac{e^{-\mu_a \tau_a}}{\mu_a - \sigma_1 ^2} \left(\frac{2\mu_a^2 +2\mu_a\gamma_a +\gamma_a^2}{2(\mu_a +\gamma_a)} +\sigma_1 ^2 \right)\bar{W} + \frac{ (\mu_a+\delta_a) (\mu_a+\gamma_a-2\sigma_2 ^2)}{2\gamma_a^2} W_8 . Making use of (5.10), (5.11) and (5.12) yields that
\begin{equation} \begin{split} L\tilde{W} = & LW_7+ \frac{e^{-\mu_a \tau_a}}{\mu_a - \sigma_1 ^2} \left( \frac{2\mu_a^2 +2\mu_a\gamma_a +\gamma_a^2}{2(\mu_a +\gamma_a)} +\sigma_1 ^2 \right)L\bar{W} + \frac{ (\mu_a+\delta_a) (\mu_a+\gamma_a-2\sigma_2 ^2)}{2\gamma_a^2} LW_8\\ \leq & -\frac{1}{4} (\mu_a+\gamma_a-2\sigma_2 ^2)(E_a -E_a^*)^2 - \frac{(\mu_a+\delta_a)(\mu_a+\gamma_a-2\sigma_2 ^2)(\mu_a+\delta_a-2\sigma_3 ^2)}{4\gamma_a^2} (I_a-I_a^*)^2 + L_1. \end{split} \end{equation} | (5.13) |
Integrating both sides of (5.13) from 0 to t and then taking expectation yields
\begin{equation*} \begin{split} E\tilde{W}(t) - E\tilde{W}(0) \leq & -\frac{1}{4} (\mu_a+\gamma_a-2\sigma_2 ^2) E\int_0^t (E_a(s)-E_a^*)^2 ds \\ & - \frac{(\mu_a+\delta_a)(\mu_a+\gamma_a-2\sigma_2 ^2)(\mu_a+\delta_a-2\sigma_3 ^2)}{4\gamma_a^2} E\int_0^t (I_a(s)-I_a^*)^2 ds + L_1 t. \end{split} \end{equation*} |
Therefore, we can obtain
\begin{equation*} \limsup\limits_{t\rightarrow\infty} E\int_0^t [(E_a(s)-E_a^*)^2 + (I_a(s)-I_a^*)^2] ds \leq \frac{L_1}{L_2} = : P_{4}, \end{equation*} |
where L_{1}, L_{2} have been defined in Theorem 3. Taking U_1 = \frac{1}{2}(S_h - S_h^*)^{2} , we have
\begin{equation} \begin{split} LU_1 = & (S_h - S_h^*)(\Lambda_h -\mu_h S_h - \frac{\beta_h S_h I_a^*}{1+\alpha_2I_a^*})+ \frac{1}{2} \sigma_4 ^2S_h^*\\ = & (S_h - S_h^*)\left[\mu_h S_h^*-\mu_h S_h +\frac{\beta_h S_h^* I_a^*}{1+\alpha_2I_a^*}-\frac{\beta_h S_h I_a^*}{1+\alpha_2I_a^*}\right]+ \frac{1}{2} \sigma_4 ^2 S_h^*\\ = & -(\mu_h+ \frac{\beta_hI_a^*}{1+\alpha_2I_a^*})(S_h - S_h^*)^{2}+\sigma_4 ^2(S_h - S_h^*)^{2}+\sigma_4 ^2(S_h^*)^{2}\\ \leq& -(\mu_h-\sigma_4 ^2)(S_h - S_h^*)^{2}+\sigma_4 ^2(S_h^*)^{2}. \end{split} \end{equation} | (5.14) |
Integrating both sides of (5.14) from 0 to t and then taking expectation, we get
\begin{equation*} \begin{split} EU_{1}(t) - EU_{1}(0) \leq & -(\mu_h-\sigma_4 ^2)E\int_0^t (S_h - S_h^*)^{2} ds+\sigma_4 ^2(S_h^*)^{2}t. \end{split} \end{equation*} |
Therefore, we can obtain
\begin{equation*} \limsup\limits_{t\rightarrow\infty} E\int_0^t (S_h - S_h^*)^{2} ds \leq \frac{\sigma_4 ^2(S_h^*)^{2}}{\mu_h-\sigma_4 ^2}. \end{equation*} |
Let U_2 = \frac{1}{2}[I_h(t+\tau_{h}) - I_h^*]^{2} , we have
\begin{equation*} \begin{split} LU_2 = & (I_h(t+\tau_{h}) - I_h^*)\left[\frac{\beta_h S_h I_a^*}{1+\alpha_2I_a^*}-(\mu_h+\delta_{h}+\theta_{h})I_h(t+\tau_{h})\right]+ \frac{1}{2} \sigma_5 ^2I_h^{2}(t+\tau_{h})\\ = & \frac{\beta_hI_a^*}{1+\alpha_2I_a^*}(I_h(t+\tau_{h}) - I_h^*)(S_h - S_h^*)-(\mu_h+\delta_{h}+\theta_{h})(I_h(t+\tau_{h})-I_h^*)^{2}+ \frac{1}{2} \sigma_5 ^2I_h^{2}(t+\tau_{h})\\ \leq& \frac{\beta_h^{2}(I_a^*)^{2}}{2(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}(S_h - S_h^*)^{2}-\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}(I_h(t+\tau_{h})-I_h^*)^{2}\\ &-(\mu_h+\delta_{h}+\theta_{h})(I_h(t+\tau_{h})-I_h^*)^{2}+ \sigma_5 ^2(I_h(t+\tau_{h})-I_h^*)^{2}+\sigma_5 ^2(I_h^*)^{2}\\ = & \frac{\beta_h^{2}(I_a^*)^{2}(S_h - S_h^*)^{2}}{2(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}-\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}(I_h(t+\tau_{h})-I_h^*)^{2}+\sigma_5 ^2(I_h^*)^{2}. \end{split} \end{equation*} |
Let \overline{U} = \frac{\beta_h^{2}(I_a^*)^{2}}{2(\mu_h-\sigma_4 ^2)(1+\alpha_2I_a^*)^{2} (\mu_h+\delta_{h} +\theta_{h} -\sigma_5 ^2)}U_{1}+U_{2} , then
\begin{equation*} L\overline{U} = -\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}(I_h(t+\tau_{h})-I_h^*)^{2}+\frac{\beta_h^{2}(I_a^*)^{2}\sigma_4 ^2(S_h^*)^{2}}{2(\mu_h-\sigma_4 ^2)(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}+\sigma_5 ^2(I_h^*)^{2}. \end{equation*} |
Let U_{3} = \frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}\int_t^{t+\tau_{h}} (I_h(s) - I_h^*)^{2} ds , we obtain
\begin{equation*} LU_{3} = \frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}\left[(I_h(t+\tau_{h})-I_h^*)^{2}+(I_h-I_h^*)^{2}\right]. \end{equation*} |
Let \widetilde{U} = \overline{U}+U_{3} , then,
\begin{equation} \begin{split} L\widetilde{U} = & -\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}(I_h(t+\tau_{h})-I_h^*)^{2}+\frac{\beta_h^{2}(I_a^*)^{2}\sigma_4 ^2(S_h^*)^{2}}{2(\mu_h-\sigma_4 ^2)(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}\\ &+ \sigma_5 ^2(I_h^*)^{2}+\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}\left[(I_h(t+\tau_{h})-I_h^*)^{2}+(I_h-I_h^*)^{2}\right]\\ = & -\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}(I_h-I_h^*)^{2}+\frac{\beta_h^{2}(I_a^*)^{2}\sigma_4 ^2(S_h^*)^{2}}{2(\mu_h-\sigma_4 ^2)(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}+ \sigma_5 ^2(I_h^*)^{2}. \end{split} \end{equation} | (5.15) |
Integrating both sides of (5.15) from 0 to t and then taking expectation, we have
\begin{align*} E\widetilde{U}(t) - E\widetilde{U}(0) \leq& -\frac{\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2}{2}E\int_0^t (I_h - I_h^*)^{2} ds \\ &+\frac{\beta_h^{2}(I_a^*)^{2}\sigma_4 ^2(S_h^*)^{2}}{2(\mu_h-\sigma_4 ^2)(1+\alpha_2I_a^*)^{2}(\mu_h+\delta_{h}+\theta_{h}-\sigma_5 ^2)}t+ \sigma_5 ^2(I_h^*)^{2}t. \end{align*} |
Therefore, we can obtain
\begin{equation*} \limsup\limits_{t\rightarrow\infty} E\int_0^t (I_h - I_h^*)^{2} ds \leq P_{6}, \end{equation*} |
where P_{6} has been defined in Theorem 3. The proof is completed.
This section is devoted to illustrating the theoretical results by numerical examples. The parameters of system (2.3) are selected as in Table 2, \alpha_1 and \alpha_2 are varying parameters that is taken value from 0.001 to 0.1 , and \sigma_1 = 0.01, \, \sigma_2 = \sigma_3 = \sigma_5 = 0.04, \, \sigma_4 = 0.008 . The initial conditions of system (2.3) are S_a(\theta) = 3,000,000, E_a(\theta) = 1,000, I_a(\theta) = 10, S_h(\theta) = 1,000, I_h(\theta) = 5, \, \theta\in[-\tau, 0] . The Milstein method [25] is used to obtain the discrete form of system (2.3) as follows:
\left\{ \begin{align}{ S_a(k+1) = S_a(k) + \left(\Lambda_a-\mu_aS_a(k) -\frac{\beta_a S_a(k) I_a(k)}{1+\alpha_1 I_a(k)} \right)\Delta t +\sigma_1 S_a(k)\sqrt{\Delta t} \xi_1(k) \\ \qquad\qquad\quad + \frac12 \sigma_1^2 S_a(k) (\xi_1^2(k) -1) \Delta t,\\ E_a(k+1) = E_a(k) + \left(\frac{\beta_a e^{-\mu_a \tau_a}S_a(k - \frac{\tau_a}{\Delta t})I_a(k - \frac{\tau_a}{\Delta t})}{1+\alpha_1 I_a(k - \frac{\tau_a}{\Delta t})}-(\mu_a+\gamma_a)E_a(k)\right)\Delta t \\ \qquad\qquad\quad +\sigma_2 E_a(k)\sqrt{\Delta t} \xi_2(k) + \frac12 \sigma_2^2 E_a(k) (\xi_2^2(k) -1) \Delta t,\\ I_a(k+1) = I_a(k) + \left(\gamma_aE_a(k)-(\mu_a+\delta_a)I_a(k)\right) \Delta t +\sigma_3 I_a(k)\sqrt{\Delta t} \xi_3(k) + \frac12 \sigma_3^2 I_a(k) (\xi_3^2(k) -1) \Delta t,\\ S_h(k+1) = S_h(k) + \left(\Lambda_h-\mu_hS_h(k)-\frac{\beta_h S_h(k)I_a(k)}{1+\alpha_2I_a(k)}\right) \Delta t +\sigma_4 S_h(k)\sqrt{\Delta t} \xi_4(k) \\ \qquad\qquad\quad + \frac12 \sigma_4^2 S_h(k) (\xi_4^2(k) -1) \Delta t,\\ I_h(k+1) = I_h(k) + \left(\frac{\beta_h e^{-\mu_h \tau_h}S_h(k-\frac{\tau_h}{\Delta t})I_a(k-\frac{\tau_h}{\Delta t})}{1+\alpha_2 I_a(k-\frac{\tau_h}{\Delta t})} -(\mu_h+\delta_h+\theta_h)I_h(k)\right) \Delta t \\ \qquad\qquad\quad +\sigma_5 I_h(k)\sqrt{\Delta t} \xi_5(k) + \frac12 \sigma_5^2 I_h(k) (\xi_5^2(k) -1) \Delta t, } \end{align} \right. | (6.1) |
Parameter | Value | Source of data |
\Lambda_a | 30000 | Assumed |
\Lambda_{h} | \mu_h\times1000 | Assumed |
\beta_{a} | (0.5 --- 12.5)\times10^{-6} \text{day}^{-1} | [10] |
\beta_{h} | 3\times10^{-4} | [10] |
\mu_{a} | 1/100 \text{day}^{-1} | [10] |
\mu_{h} | 200/(70\times365)\text{day}^{-1} | Assumed |
\delta_{a} | 5 \text{day}^{-1} | [10] |
\delta_{h} | 0.03 \text{day}^{-1} | [10,11] |
\gamma_{a} | 0.3 \text{day}^{-1} | [11] |
\theta_{h} | 0.16 \text{day}^{-1} | [11] |
\tau_a | 7 day | Assumed |
\tau_h | 14 day | Assumed |
where \xi_i(k) \sim N(0, 1) (i = 1, \cdots, 5; k = 1, 2, \cdots) are independent Gaussian random variables. Initially, we study the effect of \mathscr{R}_0 , which, by Theorems 4.1 and 5.1, can govern the asymptotic behavior.
Example 1 Effect of basic reproduction number \mathscr{R}_0 .
Choose different \beta_a such that \mathscr{R}_0 take different values, which are shown in Table 3. Since \sigma_1^2 = 10^{-4} < \mu_a = 10^{-2}, \sigma_2^2 = 0.0016 < \frac{1}{2} (\mu_a+\gamma_a) = 0.155, \sigma_3^2 = 0.0016 < \frac{1}{2} (\mu_a+\delta_a) = 5.01, \sigma_4^2 = 0.000064 < \mu_h = 0.0078, \sigma_5^2 = 0.0016 < \mu_h+\delta_h+\theta_h = 0.1978 , the condition (ⅰ) of Theorem 3 is satisfied. From Table 3, we see that for each \mathscr{R}_0 , the inequality P_{m} < d_E holds, which means the condition (ⅱ) of Theorem 3 is also satisfied. Thus, all the conclusions of Theorem 3 hold. It follows from Table 3 that the change of \mathscr{R}_0 can result in different values of E^* , which also illustrate the value of E^* is related to \mathscr{R}_0 . By the discrete form of system (2.3), the numerical results under different \mathscr{R}_0 are presented by Figures 2 and 3 when \mathscr{R}_0 > 1 , which show that the solution of system (2.3) goes around the endemic equilibrium E^* . The effectiveness of Theorem 3 is also indicated by these two figures. In addition, we can see from Figures 2 and 3 and Table 3 that the number of infected poultry and humans will reduce with the decrease of \mathscr{R}_0 . On the other hand, in order to explore if the results of Theorem 3 hold, we enhance the intensity of perturbation as \sigma = (\sigma_1, \cdots, \sigma_5) = (0.02, 0.08, 0.08, 0.016, 0.08) (Case Ⅰ: condition (ⅰ) of Theorem 3 is satisfied but condition (ⅱ) is not satisfied), \sigma = (0.06, 0.24, 0.24, 0.048, 0.24) (Case Ⅱ: Both conditions (ⅰ) and (ⅱ) are not satisfied) and \sigma = (0.10, 0.40, 0.40, 0.080, 0.40) (Case Ⅲ: Both conditions (ⅰ) and (ⅱ) are not satisfied). The simulation results are presented in Figure 4, which are obtained by computing the average of 800 simulations. The equilibrium of corresponding deterministic model is E^* = (2.3931\times10^{6}, 1.8254\times10^{4}, 0.7812\times10^{3}, 227.3975, 11.5855) . From Figure 4, we see that the curves will move away from the equilibrium point E^* with the increasing of intensity of perturbation, which violate the conclusions of Theorem 3.
\mathscr{R}_0 | S_a^*(\times10^{6}) | E_a^*(\times10^{4}) | I_a^*(\times10^{3}) | S_h^* | I_h^* | P_{m} \, (\times 10^5)^{\sharp} | d_E \, (\times 10^5)^{\sharp} |
4.8269 | 0.7977 | 6.6239 | 2.8348 | 212.6795 | 11.8062 | 2.5226 | 22.033 |
3.2823 | 1.1336 | 5.6137 | 2.4024 | 213.7049 | 11.7908 | 2.7378 | 18.673 |
1.9308 | 1.7948 | 3.6249 | 1.5513 | 217.3720 | 11.7359 | 3.2294 | 12.057 |
1.3515 | 2.3931 | 1.8254 | 0.7812 | 227.3975 | 11.5855 | 3.7266 | 6.072 |
^{\sharp}\, P_{m}=\max(\sqrt{P_{3}}, \sqrt{P_{4}}, \sqrt{P_{5}}, \sqrt{P_{6}}), d_E = d(E^*,E^0). |
According to the values of \sigma_1, \cdots, \sigma_5 and the parameters values in Table 2, we easily verify the conditions of Theorem 2 are satisfied. Therefore, from Theorem 2 we know that the solution of system (2.3) will go around the disease-free equilibrium E^0 when \mathscr{R}_0 < 1 . The numerical simulation results of \mathscr{R}_0 are presented in Figures 5 and 6. These figures show E_a, I_a and I_h all go to zero when \mathscr{R}_0 < 1 , which illustrate the effectiveness of the theoretical results in Theorem 2. Meanwhile, Figures 5 and 6 also show that the rate of E_a, I_a and I_h converges to zero is increasing with the decrease of \mathscr{R}_0 . The conditions of Theorem 2 are only a sufficient ones, so we want to know whether the conclusions of Theorem 2 hold when the intensity of perturbation increase such that these conditions are not satisfied. Thus, we choose \sigma = (\sigma_1, \cdots, \sigma_5) = (0.02, 0.08, 0.08, 0.016, 0.08) (Case I), \sigma = (0.06, 0.24, 0.24, 0.048, 0.24) (Case II) and \sigma = (0.12, 0.48, 0.48, 0.096, 0.48) (Case III), and the simulation results are presented in Figure 7. Figure 7 shows E_a, I_a and I_h converge to zero for each cases, so the results of Theorem 2 also hold.
Example 2 Effect of time delays \tau_a and \tau_h .
In order to study the effect of time delays, we consider the average peak values of E_a , I_a and I_h , and the time of reaching average peak values by 300 simulation runs. The simulation results are shown in Figures 8 and 9. It follows from Figure 8 that the increase of time delay \tau_a or \tau_h can reduce the peak value of both infected poultry and human population. Meanwhile, from Figure 9, we know that the large time delay also lead to the delay of reaching peak value. Thus, we may conclude that time delays have significate influence for the spread of avian influenza. According to the practical meaning of \tau_a and \tau_h , related department can adopt some measures to increase the spread delay to suppress the outbreak of influenza, such as isolation. In addition, the adopting of those control measures will win time for taking drug control.
Example 3 Effect of saturation constants \alpha_1 and \alpha_2 .
According to the analysis of Introduction, the saturation constants \alpha_1 and \alpha_2 are important parameters for avian influenza. We thus explore the effects of \alpha_1 and \alpha_2 in this example. In order to explore the effect of \alpha_1 under fixed \alpha_2 , we run 1000 simulations and take their average values. The results are shown in Figure 10. It follows from Figure 10 that \alpha_1 can influence the rate of convergence to the equilibria of the poultry population, while it can not significantly influence the rate of convergence to the equilibria of the human population. In addition, we study the influence of \alpha_2 under fixed \alpha_1 . The simulation results are presented in Figure 11, which implies that \alpha_2 can not change the rate of convergence to the equilibria of the poultry population. Figure 11 also means that \alpha_2 can not increase the rate of convergence to the equilibria of the human population, but it can evidently reduce the peak value of I_h(t) . In summary, \alpha_1 and \alpha_2 have evidently influence to the spreading of avian influenza among both avian and human population.
In this paper, we establish a stochastic delayed avian influenza model with saturated incidence rate. To begin with, we investigate the existence and uniqueness of the global positive solution to the system (2.3) with any positive initial value (2.4). Since there is no equilibrium point in the system (2.3) at this time, thus, the asymptotic behaviors of the disease-free equilibrium and the endemic equilibrium are given by constructing some suitable Lyapunov functions and applying the Young's inequality and Hölder's inequality. Theorem 2 shows that if \mathscr{R}_0 < 1 , then the solution of system (2.3) is going around E^0 while from Theorem 3, we obtain that if \mathscr{R}_0 > 1 , then the solution of system (2.3) is going around E^* . Finally, some numerical examples are given to illustrate the accuracy of the theoretical results.
There are some interesting issues deserve further investigations. On the one hand, we can formulate some more realistic but complex avian influenza models, such as considering the effects of Lévy jumps or impulsive perturbations on system (2.3). On the other hand, the coefficients in our model studied in this paper are all constants. If the coefficients are with Markov switching, how will the properties change? We leave these investigations as our future work.
The research was supported by the National Natural Science Foundation of China (11661064), Ningxia Natural Science Foundation Project (2019AAC03069) and the Funds for Improving the International Education Capacity of Ningxia University (030900001921).
The authors declare that they have no conflict of interest.
[1] | Centers for Disease Control and Prevention (CDC), Avian influenza, 2017. Available from: https://www.cdc.gov/flu/avianflu/influenza-a-virus-subtypes.htm. |
[2] | S. Liu, S. Ruan, X. Zhang, Nonlinear dynamics of avian influenza epidemic models, Math. Biosci., 283 (2017), 118-135. |
[3] | S. Iwami, Y. Takeuchi, X. Liu, Avian-human influenza epidemic model, Math. Biosci., 207 (2007), 1-25. |
[4] | A. B. Gumel, Global dynamics of a two-strain avian influenza model, Int. J. Comput. Math., 86 (2009), 85-108. |
[5] | Q. Tang, J. Ge, Z. Lin, An SEI-SI avian-human influenza model with diffusion and nonlocal delay, Appl. Math. Comput., 247 (2014), 753-761. |
[6] | Y. Chen, Y. Wen, Global dynamic analysis of a H7N9 avian-human influenza model in an outbreak region, J. Theor. Biol., 367 (2015), 180-188. |
[7] | S. Liu, S. Ruan, X. Zhang, On avian influenza epidemic models with time delay, Theor. Biosci., 134 (2015), 75-82. |
[8] | N. S. Chong, J. M. Tchuenche, J. S. Robert, A mathematical model of avian influenza with halfsaturated incidence, Theor. Biosci., 133 (2013), 23-38. |
[9] | Z. Liu, C. T. Fang, A modeling study of human infections with avian influenza A H7N9 virus in mainland China, Int. J. Infect. Dis., 41 (2015), 73-78. |
[10] | C. Modnak, J. Wang, An avian influenza model with latency and vaccination, Dynam. Syst., 34 (2018), 195-217. |
[11] | A. K. Muhammad, F. Muhammad, I. Saeed, B. Ebenezer, Modeling the transmission dynamics of avian influenza with saturation and psychological effect, Discrete Contin. Dyn. Syst. Ser. S, 12 (2019), 455-474. |
[12] | T. Kang, Q. Zhang, L. Rong, A delayed avian influenza model with avian slaughter: Stability analysis and optimal control, Physica A, 529 (2019), 121544. |
[13] | Y. Kuang, Delay Differential Equations with Applications in Population Dynamics, Academic Press, New York, 1993. |
[14] | E. Beretta, Y. Takeuchi, Global stability of an sir epidemic model with time delays, J. Math. Biol., 33 (1995), 250-260. |
[15] | M. Liu, C. Bai, K. Wang, Asymptotic stability of a two-group stochastic SEIR model with infinite delays, Commun. Nonlinear Sci. Numer. Simul., 19 (2014), 3444-3453. |
[16] | Q. Liu, D. Jiang, N. Shi, T. Hayat, A. Alsaedi, Asymptotic behavior of a stochastic delayed seir epidemic model with nonlinear incidence, Physica A, 462 (2016), 870-882. |
[17] | G. P. Samanta, Permanence and extinction for a nonautonomous avian-human influenza epidemic model with distributed time delay, Math. Comput. Model., 52 (2010), 1794-1811. |
[18] | J. Semenza, B. Menne, Climate change and infectious diseases in europe, Lancet. Infect. Dis., 9 (2009), 365-375. |
[19] | A. Lowen, J. Steel, Roles of humidity and temperature in shaping influenza seasonality, J. Virol., 88 (2014), 7692-7705. |
[20] | X. Zhang, Global dynamics of a stochastic avian-human influenza epidemic model with logistic growth for avian population, Nonlinear Dynam., 90 (2017), 2331-2343. |
[21] | F. Zhang, X. Zhang, The threshold of a stochastic avian-human influenza epidemic model with psychological effect, Physica A, 492 (2018), 485-495. |
[22] | X. Zhang, Z. Shi, Y. Wang, Dynamics of a stochastic avian-human influenza epidemic model with mutation, Physica A, 534 (2019), 12940. |
[23] | D. Jiang, J. Yu, C. Ji, N. Shi, Asymptotic behavior of global positive solution to a stochastic sir model, Math. Comput. Model., 54 (2011), 221-232. |
[24] | Q. Liu, D. Jiang, T. Hayat, A. Alsaedi, Threshold behavior in a stochastic delayed sis epidemic model with vaccination and double diseases, J. Franklin I., 356 (2019), 7466-7485. |
[25] | J. H. Desmond, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525-546. |
1. | Baoquan Zhou, Daqing Jiang, Yucong Dai, Tasawar Hayat, Threshold Dynamics and Probability Density Function of a Stochastic Avian Influenza Epidemic Model with Nonlinear Incidence Rate and Psychological Effect, 2023, 33, 0938-8974, 10.1007/s00332-022-09885-8 | |
2. | Ting Kang, Qimin Zhang, Dynamics of a stochastic delayed avian influenza model with mutation and temporary immunity, 2021, 14, 1793-5245, 2150029, 10.1142/S1793524521500297 | |
3. | Xubin Jiao, Xiuxiang Liu, Rich dynamics of a delayed Filippov avian-only influenza model with two-thresholds policy, 2024, 182, 09600779, 114710, 10.1016/j.chaos.2024.114710 | |
4. | Xuezhong Bao, Xiaoling Han, Dynamics and simulations of stochastic time delay avian influenza model using spectral method, 2025, 44, 2238-3603, 10.1007/s40314-025-03167-1 |
Parameter | Description |
\Lambda_a | new recruitment of the poultry populations |
\Lambda_{h} | new recruitment of the human population |
\beta_{a} | the transmission rate from infective poultry to susceptible poultry |
\beta_{h} | the transmission rate from infective poultry to susceptible human |
\mu_{a} | the natural death rate of poultry populations |
\mu_{h} | the natural death rate of human populations |
\delta_{a} | the disease-related death rate of poultry populations |
\delta_{h} | the disease-related death rate of humans populations |
\gamma_{a} | the transfer rate of exposed poultry to infected poultry |
\theta_{h} | the recovery rate of the infective human |
\alpha_{i}(i=1, 2) | parameters that measure the inhibitory effect |
Parameter | Value | Source of data |
\Lambda_a | 30000 | Assumed |
\Lambda_{h} | \mu_h\times1000 | Assumed |
\beta_{a} | (0.5 --- 12.5)\times10^{-6} \text{day}^{-1} | [10] |
\beta_{h} | 3\times10^{-4} | [10] |
\mu_{a} | 1/100 \text{day}^{-1} | [10] |
\mu_{h} | 200/(70\times365)\text{day}^{-1} | Assumed |
\delta_{a} | 5 \text{day}^{-1} | [10] |
\delta_{h} | 0.03 \text{day}^{-1} | [10,11] |
\gamma_{a} | 0.3 \text{day}^{-1} | [11] |
\theta_{h} | 0.16 \text{day}^{-1} | [11] |
\tau_a | 7 day | Assumed |
\tau_h | 14 day | Assumed |
\mathscr{R}_0 | S_a^*(\times10^{6}) | E_a^*(\times10^{4}) | I_a^*(\times10^{3}) | S_h^* | I_h^* | P_{m} \, (\times 10^5)^{\sharp} | d_E \, (\times 10^5)^{\sharp} |
4.8269 | 0.7977 | 6.6239 | 2.8348 | 212.6795 | 11.8062 | 2.5226 | 22.033 |
3.2823 | 1.1336 | 5.6137 | 2.4024 | 213.7049 | 11.7908 | 2.7378 | 18.673 |
1.9308 | 1.7948 | 3.6249 | 1.5513 | 217.3720 | 11.7359 | 3.2294 | 12.057 |
1.3515 | 2.3931 | 1.8254 | 0.7812 | 227.3975 | 11.5855 | 3.7266 | 6.072 |
^{\sharp}\, P_{m}=\max(\sqrt{P_{3}}, \sqrt{P_{4}}, \sqrt{P_{5}}, \sqrt{P_{6}}), d_E = d(E^*,E^0). |
Parameter | Description |
\Lambda_a | new recruitment of the poultry populations |
\Lambda_{h} | new recruitment of the human population |
\beta_{a} | the transmission rate from infective poultry to susceptible poultry |
\beta_{h} | the transmission rate from infective poultry to susceptible human |
\mu_{a} | the natural death rate of poultry populations |
\mu_{h} | the natural death rate of human populations |
\delta_{a} | the disease-related death rate of poultry populations |
\delta_{h} | the disease-related death rate of humans populations |
\gamma_{a} | the transfer rate of exposed poultry to infected poultry |
\theta_{h} | the recovery rate of the infective human |
\alpha_{i}(i=1, 2) | parameters that measure the inhibitory effect |
Parameter | Value | Source of data |
\Lambda_a | 30000 | Assumed |
\Lambda_{h} | \mu_h\times1000 | Assumed |
\beta_{a} | (0.5 --- 12.5)\times10^{-6} \text{day}^{-1} | [10] |
\beta_{h} | 3\times10^{-4} | [10] |
\mu_{a} | 1/100 \text{day}^{-1} | [10] |
\mu_{h} | 200/(70\times365)\text{day}^{-1} | Assumed |
\delta_{a} | 5 \text{day}^{-1} | [10] |
\delta_{h} | 0.03 \text{day}^{-1} | [10,11] |
\gamma_{a} | 0.3 \text{day}^{-1} | [11] |
\theta_{h} | 0.16 \text{day}^{-1} | [11] |
\tau_a | 7 day | Assumed |
\tau_h | 14 day | Assumed |
\mathscr{R}_0 | S_a^*(\times10^{6}) | E_a^*(\times10^{4}) | I_a^*(\times10^{3}) | S_h^* | I_h^* | P_{m} \, (\times 10^5)^{\sharp} | d_E \, (\times 10^5)^{\sharp} |
4.8269 | 0.7977 | 6.6239 | 2.8348 | 212.6795 | 11.8062 | 2.5226 | 22.033 |
3.2823 | 1.1336 | 5.6137 | 2.4024 | 213.7049 | 11.7908 | 2.7378 | 18.673 |
1.9308 | 1.7948 | 3.6249 | 1.5513 | 217.3720 | 11.7359 | 3.2294 | 12.057 |
1.3515 | 2.3931 | 1.8254 | 0.7812 | 227.3975 | 11.5855 | 3.7266 | 6.072 |
^{\sharp}\, P_{m}=\max(\sqrt{P_{3}}, \sqrt{P_{4}}, \sqrt{P_{5}}, \sqrt{P_{6}}), d_E = d(E^*,E^0). |