
In this paper, the dynamical behavior of a delayed SIQR stochastic epidemic model with Lévy noise is presented and studied. First, we prove the existence and uniqueness of positive solution. Then, we establish the threshold Rl0 as a sufficient condition for the extinction and persistence in mean of the disease. Finally, some numerical simulations are presented to support our theoretical results and we infer that the white and Lévy noises affect the transmission dynamics of the system.
Citation: Yubo Liu, Daipeng Kuang, Jianli Li. Threshold behaviour of a triple-delay SIQR stochastic epidemic model with Lévy noise perturbation[J]. AIMS Mathematics, 2022, 7(9): 16498-16518. doi: 10.3934/math.2022903
[1] | Yuanfu Shao . Dynamics and optimal harvesting of a stochastic predator-prey system with regime switching, S-type distributed time delays and Lévy jumps. AIMS Mathematics, 2022, 7(3): 4068-4093. doi: 10.3934/math.2022225 |
[2] | Xiaodong Wang, Kai Wang, Zhidong Teng . Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence. AIMS Mathematics, 2023, 8(2): 2829-2855. doi: 10.3934/math.2023148 |
[3] | Shuo Ma, Jiangman Li, Qiang Li, Ruonan Liu . Adaptive exponential synchronization of impulsive coupled neutral stochastic neural networks with Lévy noise and probabilistic delays under non-Lipschitz conditions. AIMS Mathematics, 2024, 9(9): 24912-24933. doi: 10.3934/math.20241214 |
[4] | Zhengwen Yin, Yuanshun Tan . Threshold dynamics of stochastic SIRSW infectious disease model with multiparameter perturbation. AIMS Mathematics, 2024, 9(12): 33467-33492. doi: 10.3934/math.20241597 |
[5] | Tian Xu, Ailong Wu . Stabilization of nonlinear hybrid stochastic time-delay neural networks with Lévy noise using discrete-time feedback control. AIMS Mathematics, 2024, 9(10): 27080-27101. doi: 10.3934/math.20241317 |
[6] | Yassine Sabbar, Aeshah A. Raezah . Modeling mosquito-borne disease dynamics via stochastic differential equations and generalized tempered stable distribution. AIMS Mathematics, 2024, 9(8): 22454-22485. doi: 10.3934/math.20241092 |
[7] | Chuanfu Chai, Yuanfu Shao, Yaping Wang . Analysis of a Holling-type IV stochastic prey-predator system with anti-predatory behavior and Lévy noise. AIMS Mathematics, 2023, 8(9): 21033-21054. doi: 10.3934/math.20231071 |
[8] | Yassine Sabbar, Aeshah A. Raezah, Mohammed Moumni . Enhancing epidemic modeling: exploring heavy-tailed dynamics with the generalized tempered stable distribution. AIMS Mathematics, 2024, 9(10): 29496-29528. doi: 10.3934/math.20241429 |
[9] | Hong Qiu, Yanzhang Huo, Tianhui Ma . Dynamical analysis of a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps. AIMS Mathematics, 2022, 7(8): 14492-14512. doi: 10.3934/math.2022799 |
[10] | Yassine Sabbar, Aeshah A. Raezah . Threshold analysis of an algae-zooplankton model incorporating general interaction rates and nonlinear independent stochastic components. AIMS Mathematics, 2024, 9(7): 18211-18235. doi: 10.3934/math.2024889 |
In this paper, the dynamical behavior of a delayed SIQR stochastic epidemic model with Lévy noise is presented and studied. First, we prove the existence and uniqueness of positive solution. Then, we establish the threshold Rl0 as a sufficient condition for the extinction and persistence in mean of the disease. Finally, some numerical simulations are presented to support our theoretical results and we infer that the white and Lévy noises affect the transmission dynamics of the system.
Governments have been prioritizing public health policies and taking decisions, plans and actions to save human lives from deadly infectious diseases. For this issue, computational biologists study the dynamics of epidemics in order to prevent and control the spread of infections in the population [1,2]. Recently, an epidemic named Corona Virus Disease 2019 (COVID-19) by the World Health Organization (WHO) has been spreading worldwide, especially in the United States, Brazil, India and South Africa, and the spread of the epidemic has caused a huge impact on industrial production and social life [2,3,4,5,6,7,8]. The virus has spread widely from person to person, although its origin remains unclear [9,10,11,12]. According to the data released by WHO Coronavirus (COVID-19) Dashboard. As of 14 February 2022, there have been 416,614,051 confirmed cases of COVID-19, including 5,844,097 deaths. COVID-19 has generated many mutant strains so far, and some of them have higher transmission and lethality rates, posing new challenges to the prevention and control of the epidemic.
In the 14th century, the authorities of the city of Venice adopted quarantine measures for access to the port, where each crew member of each ship was examined and could be cleared from land once the entire population was free of symptoms. This idea was adopted as the main measure to prevent the spread of infectious diseases such as Ebola and malaria. Recently, quarantine measures have proven to be effective in the extinction of COVID-19 disease in China [13], which has led many countries to adopt this strategy in the absence of a vaccine or cure for Neocoronavirus. To understand the effect of quarantine on epidemic behavior [14,15,16,17,18]. Liu et al. [19] proposed a model with quarantine to describe isolated individuals in a segregation model.
When studying the spread of epidemics, researchers now consider the impact of environmental noise, such as high temperature, freezing, drought, humidity, hurricanes, and so on. And they show that the existence of random factors such that the development of infectious diseases can be interfered [20]. The stochastic model can make up for the shortcomings of the deterministic model. Gard points out that the population dynamics is often disturbed by random perturbations [21], Cai et al. revealed that disease outbreaks can be suppressed by white noise [22]. Du et al. [23,24] propose the following model
{ dS=[Λ−F(S,I)−ρS]dt+σ1SdB1(t), dI=[F(S,I)−(ρ+γ)I]dt+σ2IdB2(t), dR=[γI−ρR]dt+σ3RdB3(t), |
where F(S,I)=βSI1+α1S+α2I, α1,α2 are positive constants measuring the suppression effect. On the other hand, a novel delayed stochastic model is proposed to describe the role of time delays in reality [25], which leads to a more complex behavior of dynamical system stability. This concept was described as temporary immunity in [26] and as a vaccine effect in [27]. However, temporary immunity can also affect isolated individuals. To better reflect reality, motivated by the study of [23,24], we propose the following triple-delay SIQR epidemic model with vaccination and isolation strategies
{ dS=[Λ−f(S,I)I−(ρ+q)S+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3]dt+σ1SdB1(t), dI=[f(S,I)I−(ρ+ω+γ+δ)I]dt+σ2IdB2(t), dQ=[δI−(ρ+μ+ε)Q]dt+σ3QdB3(t), dR=[γI+qS+εQ−ρR−qS(t−τ1)e−μτ1−γI(t−τ2)e−μτ2−εQ(t−τ3)e−μτ3]dt+σ4RdB4(t), | (1.1) |
where S(t) stands for the susceptible individuals, I(t) for infected individuals, R(t) for recovered compartment and Q(t) for isolated or quarantined compartment. The parameter δ, ϵ, Λ, β and ρ denotes the rate of infectious individuals who were isolated, the recovered people coming from isolation, the population recuritment rate, the transmission coefficient from susceptible to infected individuals, the natural death rate respectively. γ,ω,μ and q represents the recovery rate of the infective individual, the death rate for infected, quarantined individuals due to infection complications and the proportional coefficient of vaccinated for the susceptible respectively. The time τ1>0, τ2>0 and τ3>0 represents the delay for the efficiency of vaccine, the length of the immunity period, the delay for isolated individuals to get back their immunity respectively. The term S(t−τ1)e−μτ1 reflects the fact that some individuals remains susceptible even after the vaccine for a specific time. The term I(t−τ2)e−μτ2 represents the individuals who became susceptible because of the lose of immunity for a specific time. The term Q(t−τ3)e−μτ3 represents the individuals coming out from isolation with immunity impairment. The Bi(t)(i=1,2,3,4) are independent standard Brownian motions defined on a complete probability space (Ω,F,P) with the filtration (Ft)t≥0, satisfying the usual conditions, and σi≥0 represent the intensities of Bi(t). The incidence of model (1.1) is of the form
f(S,I)=βS1+α1S+α2I+α3SI+α4I2, |
where α1,α2,α3 and α4 are constants measuring the suppression effect.
As we know, population systems may suffer severe environmental perturbations, such as tsunami, volcanoes, avian influenza, hurricanes, earthquakes, toxic pollutants, etc. These phenomena cannot be described by stochastic continuous models. And so it is feasible to introduce a jump process into the underlying population systems (see e.g., [28,29,30]). Our goal in this work is to extend the model presented in [23] to a model with Lévy noise perturbation and also take in consideration a special incidence f(S,I) and this model can be practically applied to describe hepatitis B epidemic [31].
{ dS=[Λ−f(S,I)I−(ρ+q)S+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3]dt+σ1SdB1(t)+∫Yη1(y)S(t−)˜N(dt,dy), dI=[f(S,I)I−(ρ+ω+γ+δ)I]dt+σ2IdB2(t)+∫Yη2(y)S(t−)˜N(dt,dy), dQ=[δI−(ρ+μ+ε)Q]dt+σ3QdB3(t)+∫Yη3(y)Q(t−)˜N(dt,dy), dR=[γI+qS+εQ−ρR−qS(t−τ1)e−μτ1−γI(t−τ2)e−μτ2−εQ(t−τ3)e−μτ3]dt+σ4RdB4(t)+∫Yη4(y)R(t−)˜N(dt,dy), | (1.2) |
where S(t−),I(t−),Q(t−) and R(t−) is the left limit of S(t),I(t),Q(t) and R(t). ˜N=N(dt,dy) is a poisson counting measure with the stationary compensator v(dy)dt. v defined on a measurable subset Y of [0,∞) with v(Y)<∞ and ηi>−1,i=1,2,3,4.
Noticing the first three stochastic differential equations in system (1.2) do not depend on the function R(t), and so we can exclude the fourth one without loss of generality. Hence, we will only discuss the following system
{ dS=[Λ−f(S,I)I−(ρ+q)S+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3]dt+σ1SdB1(t)+∫Yη1(y)S(t−)˜N(dt,dy), dI=[f(S,I)I−(ρ+ω+γ+δ)I]dt+σ2IdB2(t)+∫Yη2(y)S(t−)˜N(dt,dy), dQ=[δI−(ρ+μ+ε)Q]dt+σ3QdB3(t)+∫Yη3(y)Q(t−)˜N(dt,dy), | (1.3) |
R3+={(x1,x2,x3)∈R3,x1>0,x2>0,x3>0}. Let C([−τ,0],R3+) be the Banach space of continuous function mappings [−τ,0]intoR3+ with norm ‖ϕ‖=sup−τ≤θ≤0|ϕ(θ)|, where τ=max{τ1,τ2,τ3}. We assume that
S(θ)=ϕ1(θ),I(θ)=ϕ2(θ),Q(θ)=ϕ3(θ), |
ϕi(θ)>0,∀θ∈[−τ,0], i=1,2,3, |
ϕi∈C([−τ,0],R+)fori∈{1,2,3}. |
The innovation of this paper as follow:
● We consider the delay and Lévy noise based on the model in [23], a threshold Rl0 of model (1.3) is obtained. If we disregard Lévy jumps, then Rs0=Rl0, here Rs0 is the threshold of the random model (1.1).
● A complex incidence function f(S,I) is considered, and the function can contain the following form:
(1) Holling Type II incidence f(S,I)I=βSI1+α2S [32], Saturation incidence f(S,I)I=βSI1+α1I [33], Bilinear functional response f(S,I)I=βSI [34].
(2) if α4=α3=0, we obtain Beddington-DeAngelis rate f(S,I)I=βSI1+α1I+α2S [35].
(3) if α4=0,α3=α1α2, Crowly-Martin functional response f(S,I)I=βSI1+α1I+α2S+α1α2SI [36].
(4) if α1=α2=α3=0, Non-monotonous incidence f(S,I)I=βSI1+α4I2 [37].
(5) if α1=α3=0, Holling Type IV incidence f(S,I)I=βSI1+α2I+α4I2 [38].
● The numerical simulations compare the effects of Lévy noise and white noise on infectious diseases, and further conclude that Lévy noise can make diseases extinct.
Throughout this paper, we define the operator L associated with the following n-dimensional stochastic differential equation(SDE)
dX(t)=f(t,X(t))dt+g(t,X(t))dB(t)+∫YH(X(t−),y)˜N(dt,dy), |
X=(x1,x2,⋯,xn). If L acts on a function G∈C1,2(Rn;R+), then
LG(X(t))=Gx(X(t−))f(t,X(t−))+12trace(gT(t,X(t−))Gxx(X(t))g(t,X(t−)))+∫Y[G(X(t−)+H(X(t−),y))−G(X(t−))−Gx(X(t−))H(X(t−),y)]v(dy), |
and
Gx=(∂G(X(t))∂x1,...,∂G(X(t))∂xn),Gxx=(∂2G(X(t))∂xi∂xj)n×n. |
then by Itô's formula, we obtain
dG(X(t))=LG(X(t−))dt+Gx(X(t−))g(t,X)dB(t)+∫Y[G(X(t−)+H(X(t−),y))−G(X(t−))]˜N(dt,dy). |
The rest of this article is organized as follows. In Section 2, the existence and uniqueness of the global positive solution of a stochastic system with Lévy noise is proven. In Section 3, the result of the analysis is the extinction can be determined when Rl0<1. In Section 4, we show that disease will persistence in the mean when Rl0>1. In Section 5, some numerical simulations to summarize related results, and provides direction for future research.
Throughout this section, we will establish the existence of a global positive solution for our delayed stochastic epidemic model with jumps. For the sake of convenience, we shall impose a standard assumption (H1), which is essential to prove the existence and uniqueness of a global positive solution of (1.3).
(H1) 1+ηi(y)>0,y∈Y,i=1,2,3,|ln(1+ηi(y))|≤C, where C is a positive constant.
Theorem 2.1. For any initial condition (S(0),I(0),Q(0))∈L1([−τ,0];R3+). There is a unique solution (S(t),I(t),Q(t)) of the stochastic system (1.3) for t≥−τ and the solution will remain in R3+ with probability one.
Proof. For any initial condition (S(0),I(0),Q(0))∈R3+, the local Lipschitz condition can make the system (1.3) exist solution (S(t),I(t),Q(t)) for any t∈[−τ,τe) almost surely (τe is the explosion time [39]). So, in order to prove the existence of a global positive solution, we need to prove that τe=∞. In other words, (S(t),I(t),Q(t)) will not explode to infinity in a finite time. We choose a positive constant m0>0. In order to keep S(0), I(0), and Q(0) all lie within the interval [1m0,m0], we let m0 be sufficiently large. Next, we construct a set {τm,m≥m0} related to this positive number m0>0:
τm=inf{t∈[−τ,τe):min{S(t),I(t),R(t)}≤1mormax{S(t),I(t),R(t)}≥m}. |
Clearly, we can find that τm is a monotonically increasing function of the independent variable m. According to the definition above, set τ∞=limm→∞τm, we know that τ∞≤τe holds. In order to prove τe=∞ we just need to ensure that τ∞=∞ for (S(0),I(0),R(0))∈R3+. We write ϕ as the empty set and define infϕ=∞ in this paper. We assume that τ∞<∞ holds, then there exist a pair of constant T>0 and ϵ∈(0,1) such that P(τ∞≤T)≥ϵ. From the above discussion, we know that there is an integer m1≥m0 such that
P(τm≤T)≥ϵ, | (2.1) |
for all m≥m1.
Define C2−functionV(S,I,Q):R3+⟶R+ by
V(S,I,Q)=(S−a−alnSa)+(I−1−lnI)+(Q−1−lnQ)+qe−μτ1∫tt−τ1Sds+γe−μτ2∫tt−τ2Ids+εe−μτ3∫tt−τ3Qds, |
where a is a positive constant determined later, the non-negativity of this function occurs from u−1−lnu≥0 for ∀u>0. With Itô's formula, then
LV=Λ−f(S,I)I−(ρ+q)S+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3+f(S,I)−(ρ+ω+γ+δ)I+δI−(ρ+μ+ε)Q−1I[f(S,I)−(ρ+ω+γ+δ)I]−aS[Λ−f(S,I)I−(ρ+q)S+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3]+aσ212+σ222+σ232−1Q[δI−(ρ+μ+ε)Q]+∫Y[aη1(y)−aln(1+η1(y))]v(dy)+∫Y[η2(y)−ln(1+η2(y))]v(dy)+∫Y[η3(y)−ln(1+η3(y))]v(dy)≤Λ−ρS−(ρ+ω)I−(ρ+μ)Q−(1−e−μτ1)qS−(1−e−μτ2)γI−(1−e−μτ3)εQ−ΛaS+af(S,I)IS+a(ρ+q)+(ρ+ω+γ+δ)+(ρ+μ+ε)+aσ212+σ222+σ232+∫Y[aη1(y)−aln(1+η1(y))]v(dy)+∫Y[η2(y)−ln(1+η2(y))]v(dy)+∫Y[η3(y)−ln(1+η3(y))]v(dy). |
From f(S,I)IS≤βI and a=ρ+ωβ, we get
LV≤Λ+a(ρ+q)+(ρ+ω+γ+δ)+(ρ+μ+ε)+aσ212+σ222+σ232+∫Y[aη1(y)−aln(1+η1(y))]v(dy)+∫Y[η2(y)−ln(1+η2(y))]v(dy)+∫Y[η3(y)−ln(1+η3(y))]v(dy):=W. |
Then,
dV≤Wdt+σ1(S−a)dB1(t)+σ2(I−1)dB2(t)+σ3(Q−1)dB3(t)+∫Y[aη1(y)−aln(1+η1(y))]˜N(dt,dy)+∫Y[η2(y)−ln(1+η2(y))]˜N(dt,dy)+∫Y[η3(y)−ln(1+η3(y))]˜N(dt,dy). | (2.2) |
Integrating both side of (2.2) from 0 to τm∧T=min{τm,T}, then
∫τm∧T0dV(S,I,Q)≤∫τm∧T0Wdt+∫τm∧T0σ1(S−a)dB1(t)+∫τm∧T0σ2(I−1)dB2(t)+∫τm∧T0σ3(R−1)dB3(t)+∫τm∧T0∫Y[aη1(y)−aln(1+η1(y))]˜N(dt,dy)+∫τm∧T0∫Y[η2(y)−ln(1+η2(y))]˜N(dt,dy)+∫τm∧T0∫Y[η3(y)−ln(1+η3(y))]˜N(dt,dy). | (2.3) |
Take the expectations to (2.3)
EV(S(τm∧T),I(τm∧T),Q(τm∧T))≤V(S(0),I(0),Q(0))+WE(τm∧T), |
so, we have
EV(S(τm∧T),I(τm∧T),Q(τm∧T))≤V(S(0),I(0),Q(0))+WT. |
Let Ωm={τm∧T} for m≥m1, and by (2.1), we derive that P(Ωm)≥ϵ. For ∀ω∈Ωm, there is at least one of S(τm∧T),I(τm∧T) and Q(τm∧T) that equals either 1m or m. It follows that V(S(τm∧T),I(τm∧T),Q(τm∧T)) is no less than m−1−lnm or 1m−1−ln1m or am−1−lnam or am−1−lnam. From this we obtain
V(S(0),I(0),Q(0))+WT≥E(1Ωm(ω)V(S(τm∧T),I(τm∧T),Q(τm∧T)))=E(1Ωm(ω)V(S(τm,ω),I(τm,ω),Q(τm,ω)))≥ϵmin{(m−1−lnm),(1m−1+lnm),(am−1−lnam),(am−1−lnam)}, |
where 1Ωm(ω) is a indicator function of Ωm(ω). Let m⟶∞, then we attain
∞>V(S(0),I(0),Q(0))+WT≥∞. |
From this we conclude that the above equation is a contradiction, then S(t),I(t) and Q(t) will not explode in a finite time.
We consider the region
Γ={(S,I,Q)∈R3+:S+I+Q≤Λρ+q(1−e−μτ1)}. |
Theorem 2.2. The region Γ is almost surely (a.s.) positive invariant for the stochastic model (1.3).
Proof. Suppose (S(θ),I(θ),Q(θ))∈Γ,θ∈[−τ,0] and n0≥0 be sufficiently large such that S(θ)∈(1n0,¯M], I(θ)∈(1n0,¯M] and Q(θ)∈(1n0,¯M]. For each integer n≥n0, the stopping times are defined as follows
τn=inf{t>0|(S(t),I(t),Q(t))=X(t)∈Γ,(S(t),I(t),Q(t))∉(1n,¯M]3},τ=inf{t>0|(S(t),I(t),Q(t))∉Γ}. |
We need to show that P(τ<t)=0 for all t>0.
Notice that P(τ<t)≤P(τn<t), then we have to prove limsupn→+∞P(τn<t)=0. Define the function
W(S,I,Q)=1S+1I+1Q, |
then
dW=LWdt−σ1SdB1(t)−σ2IdB2(t)−σ3QdB3(t)−∫Y(η1(y)S(t−)+η1(y)S(t−)+η2(y)I(t−)+η2(y)I(t−)+η3(y)Q(t−)+η3(y)Q(t−))˜N(dt,dy), |
here,
LW=−ΛS2+ρ+qS+βf(S,I)S2−qS(t−τ1)e−μτ1S2−γI(t−τ2)e−μτ2S2−εQ(t−τ3)e−μτ3S2+σ21S+σ22I+σ23Q−βf(S,I)I2+ρ+ω+γ+δI−δIQ2+ρ+μ+εQ+∫Yη21(y)S(1+η1(y))v(dy)+∫Yη22(y)I(1+η2(y))v(dy)+∫Yη23(y)Q(1+η3(y))v(dy)≤[ρ+q+βf(S,I)S+σ21+∫Yη21(y)S(1+η1(y))v(dy)]1Sdt+[ρ+ω+γ+δ+σ22+∫Yη22(y)I(1+η2(y))v(dy)]1Idt+[ρ+μ+ε+σ23+∫Yη23(y)Q(1+η3(y))v(dy)]1Qdt. |
Then
dW≤ηLWdt−σ1SdB1(t)−σ2IdB2(t)−σ3QdB3(t)−∫Y(η1(y)S(t−)+η1(y)S(t−)+η2(y)I(t−)+η2(y)I(t−)+η3(y)Q(t−)+η3(y)Q(t−))˜N(dt,dy), | (2.4) |
where
η=max{ρ+q+βΛρ+q(1−e−μτ1)+σ21+∫Yη21(y)S(1+η1(y))v(dy);+ρ+ω+γ+δ+σ22+∫Yη22(y)I(1+η2(y))v(dy);ρ+μ+ε+σ23+∫Yη23(y)Q(1+η3(y))v(dy)}. |
Taking integral and expectation on both sides of (2.4) and by virtue of Fubini Theorem, then we derive
E(W(X(s)))≤W(X0)+η∫s0E(W(X(ξ)))dξ. |
Applying Gronwall Lemma, we obtain that
E(W(X(s)))≤W(X0)eηs. |
for all s∈[0,t∧τn]. Thus,
E(W(X(t∧τn)))≤W(X0)eη(t∧τn)≤W(X0)eηt,t≥0. |
In consideration of W(X(t∧τn))>0 and some component of X(τn) being less than or equal to 1n, we achieve
E(W(X(t∧τn)))≥E(W(X(τn))1{τn<t})≥nP(τn<t). | (2.5) |
By (2.5), we obtain that
P(τn<t)≤W(X0)eηtn, |
for all t≥0. Therefore,
lim supn→+∞P(τn<t)=0. |
The proof is completed.
In the study of infectious disease models, the search for thresholds is an important aspect, and in this section we investigate a threshold condition that can determine the extinction and persistence of the disease. Define a parameter
R0=βΛ(ρ+q(1−e−μτ1)+α1Λ)(ρ+ω+γ+δ). |
Let Rs0=R0−1(ρ+ω+γ+δ)(σ222) be the threshold of our stochastic model (1.3) and Rl0 be the threshold of our model (1.3) defined as follows
Rl0=R0−1(ρ+ω+γ+δ)[σ222+∫Y[η2(y)−ln(1+η2(y))]v(dy)]=Rs0−1(ρ+ω+γ+δ)[∫Y[η2(y)−ln(1+η2(y))]v(dy)]. |
To simplify, we consider the following notation ⟨x(t)⟩=1t∫t0x(s)ds.
Lemma 3.1. [40] M={Mt}t≥0 be a real-valued continuous local martingle vanishing at t=0 then
limt→∞⟨M,M⟩=∞a.s⇒limt→∞Mt⟨M,M⟩=0,a.s. |
and also
lim supt→∞⟨M,M⟩t<∞a.s.⇒limt→∞Mtt=0,a.s. |
Theorem 3.1. For any initial value (S(θ),I(θ),Q(θ))∈Γ,θ∈[−τ,0], let (S(t),I(t),Q(t)) be the solution of stochastic system (1.3). If Rl0<1, then
lim supt→∞lnI(t)t≤(ρ+ω+γ+δ)(Rl0−1)<0. |
The disease will be extinct exponentially. Moreover
limt→∞⟨S(t)⟩=βΛρ+q(1−e−μτ1),limt→∞⟨Q(t)⟩=0. |
Proof. We consider the following function
V=S+I+εe−μτ3Qρ+μ+ε+qe−μτ1∫tt−τ1Sds+γe−μτ2∫tt−τ2Ids+εe−μτ3∫tt−τ3Qds. |
Using the Itô's formula, we get
LV=[Λ−(ρ+q)S−f(S,I)I+qS(t−τ1)e−μτ1+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3]+[f(S,I)I−(ρ+ω+γ+δ)I]+εe−μτ3ρ+μ+ε[δI−(ρ+μ+ε)Q]+qe−μτ1+γe−μτ2−(qS(t−τ1)e−μτ1+εe−μτ3+γI(t−τ2)e−μτ2+εQ(t−τ3)e−μτ3), |
and
dV=LVdt+σ1SdB1(t)+σ2IdB2(t)+εe−μτ3ρ+μ+εσ3QdB3(t)+∫Y[η1(y)S(t−)+η2(y)I(t−)+εe−μτ3ρ+μ+εη3(y)Q(t−)]˜N(dt,dy). |
Then,
dV=[Λ−(ρ+q(1−e−μτ1))S−(ρ+γ(1−e−μτ1)+ω+δ−δεe−μτ3ρ+μ+ε)I]dt+∫Y[η1(y)S(t−)+η2(y)I(t−)+εe−μτ3ρ+μ+εη3(y)Q(t−)]˜N(dt,dy)+σ1SdB1(t)+σ2IdB2(t)+εe−μτ3ρ+μ+εσ3QdB3(t). | (3.1) |
Therefore, intergrating both sides of (3.1), we obtain
⟨S(t)⟩=Λρ+q(1−e−μτ1)[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]⟨I(t)⟩−ϕ(t), | (3.2) |
where
ϕ(t)=S(t)+I(t)+εe−μτ3ρ+μ+εQ(t)+qe−μτ1∫tt−τ1Sds+γe−μτ2∫tt−τ2Ids+εe−μτ3∫tt−τ3Qds[ρ+q(1−e−μτ1)]t−S(0)+I(0)+εe−μτ3ρ+μ+εQ(0)+qe−μτ1∫0−τ1Sds+γe−μτ2∫0−τ2Ids+εe−μτ3∫0−τ3Qds[ρ+q(1−e−μτ1)]t+∫t0∫Y[η1(y)S(t−)+η2(y)I(t−)+εe−μτ3ρ+μ+εη3(y)Q(t−)]˜N(dt,dy)[ρ+q(1−e−μτ1)]t+∫t0σ1SdB1(t)+∫t0σ2IdB2(t)+εe−μτ3ρ+μ+ε∫t0σ3QdB3(t)[ρ+q(1−e−μτ1)]t. |
Since (S,I,Q)∈Γ, and taking expectation of ϕ(t), we obtain limt→∞ϕ(t)=0. Now, applying Itô's formula to the function lnI(t) we get
dlnI(t)=[f(S,I)−(ρ+ω+γ+δ+12σ22)−∫Yη2(y)−ln(1+η2(y))v(dy)]dt+σ2dB2(t)+∫Yln(1+η2(y))˜N(dt,dy). | (3.3) |
Noticing that the function f(S,I) can be written as
f(S,I)=−(Λρ+q(1−e−μτ1)−S)[β(ρ+q(1−e−μτ1))(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)]−βΛα2I(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)+βΛρ+q(1−e−μτ1)+α1Λ−βΛα3SI(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)−βΛα4I2(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2). |
Then
f(S,I)≤βΛρ+q(1−e−μτ1)+α1Λ. |
Hence, integrating both sides of (3.3) and by dividing by t we obtain
lnI(t)t≤lnI(0)t+βΛρ+q(1−e−μτ1)+α1Λ−(ρ+ω+δ+γ+σ222)+1t∫t0∫Yln(1+η2(y))˜N(dt,dy)+1t∫t0σ2dB2(t). |
By the strong of large number law for local martingales [41], and for Rl0<1, we get
lim supt→∞lnI(t)t≤βΛρ+q(1−e−μτ1)+α1Λ−(ρ+ω+δ+γ+σ222)+lim supt→∞1t∫t0∫Yln(1+η2(y))˜N(dt,dy)+lim supt→∞1t∫t0σ2dB2(t)=(ρ+ω+γ+δ)(Rl0−1), |
which leads to limt→∞I(t)=0.
From (3.2) we obtain
limt→∞⟨S(t)⟩=−[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]limt→∞⟨I(t)⟩−limt→∞ϕ(t)+Λρ+q(1−e−μτ1)=Λρ+q(1−e−μτ1). |
Integrating third equation of system (1.3),
Q(t)−Q(0)t=δ1t∫t0I(s)ds−(ρ+μ+ε)1t∫t0Q(s)ds+1t∫t0σ3QdB3(t)+1t∫t0∫Yη3(y)Q(t−)˜N(dt,dy). |
Hence, from limt→∞I(t)=0 and by the strong law of large numbers for local martingales
limt→∞1t∫t0Q(s)ds=0a.s. |
So our proof is complete.
In this section, we investigate the persistence of the disease. First, we recall the following definition.
Definition 4.1. [42] The solution of the stochastic model (1.3) is said to persistence in the mean, if
lim inft→∞⟨x(t)⟩>0almostsure(a.s.). |
Lemma 4.1. [42] Let f∈C([0,∞),(0,+∞)) and F∈C([0,+∞),R) such thatif there exist positive constants m1,m2 and T, such that
lnf(t)≥m1t−m2∫t0f(x)dx+F(t)a.s.forallt≥T, |
and limt→∞F(t)t=0 a.s., then
lim inft→∞⟨f(t)⟩≥m1m2a.s. |
Let
λ=[β(ρ+q(1−e−μτ1))(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))(ρ+q(1−e−μτ1)+α1Λ)+βΛα2ρ+q(1−e−μτ1)+α1Λ(1+βΛα3ρ+q(1−e−μτ1)+βΛα4ρ+q(1−e−μτ1))]. |
Theorem 4.1. Let (S(t),I(t),Q(t)) be the solution of system (1.3) with initial value (S(θ),I(θ),Q(θ))∈Γ,θ∈[−τ,0]. If Rl0>1, then
lim inft→∞⟨I(t)⟩≥(Rl0−1)(ρ+ω+δ+γ)λ=I∗>0, |
Λ(ρ+q+βΛρ+q(1−e−μτ1))≤lim inft⟶∞⟨S(t)⟩≤lim supt→∞⟨S(t)⟩≤Λρ+q(1−e−μτ1)−[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]I∗, |
lim inft→∞⟨Q(t)⟩≥δρ+μ+εI∗>0, |
where I∗=(Rl0−1)(ρ+ω+δ+γ)λ.
Proof. From Theorem 2.2, we can get
f(S,I)=−(Λρ+q(1−e−μτ1)−S)[β(ρ+q(1−e−μτ1))(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)]−βΛα2I(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)+βΛρ+q(1−e−μτ1)+α1Λ−βΛα3SI(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)−βΛα4I2(ρ+q(1−e−μτ1)+α1Λ)(1+α1S+α2I+α3SI+α4I2)≥−(Λρ+q(1−e−μτ1)−S)[β(ρ+q(1−e−μτ1))(ρ+q(1−e−μτ1)+α1Λ)]+βΛρ+q(1−e−μτ1)+α1Λ−βΛα2Iρ+q(1−e−μτ1)+α1Λ−βΛα3SIρ+q(1−e−μτ1)+α1Λ−βΛα4I2ρ+q(1−e−μτ1)+α1Λ≥β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1ΛS−[βα2Λρ+q(1−e−μτ1)+α1Λ−βΛ2α3(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))−βΛ2α4(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))]I. |
Applying the Itô's formula to the second equation of model (1.3) yields
dlnI(t)=LVdt+σ2dB2(t)+∫Yln(1+η2(y))˜N(dt,dy), |
where
LV=f(S,I)−(ρ+ω+γ+δ+σ222)−∫Y(η2(y)−ln(1+η2(y))v(dy). |
Then
dlnI(t)≥[β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1ΛS−(ρ+ω+γ+δ+σ222+∫Y(η2(y)−ln(1+η2(y))v(dy))]dt−[βα2Λρ+q(1−e−μτ1)+α1Λ−βΛ2α3(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))−βΛ2α4(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))]Idt+σ2dB2(t)+∫Yln(1+η2(y))˜N(dt,dy). | (4.1) |
From the result (3.2) and integrating (4.1) between 0 and t we have
lnI(t)≥β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1ΛΛρ+q(1−e−μτ1)−β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1Λ[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]⟨I(t)⟩t−[βα2Λρ+q(1−e−μτ1)+α1Λ−βΛ2α3(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))−βΛ2α4(ρ+q(1−e−μτ1)+α1Λ)(ρ+q(1−e−μτ1))]⟨I(t)⟩t−(ρ+ω+δ+γ+σ222+∫Y(η2(y)−ln(1+η2(Y)))v(dy))t+φ(t), |
where
φ(t)=−β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1Λϕ(t)t+∫t0σ2dB2(t)+∫t0∫Yln(1+η2(y))˜N(dt,dy)+lnI(0). |
By the strong law of large numbers for local martingales that
lim inft→∞φ(t)t=0, a.s. |
Hence, by Lemma 4.1 we get
[β(ρ+q(1−e−μτ1))ρ+q(1−e−μτ1)+α1Λ((ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1)))+βΛρ+q(1−e−μτ1)+α1Λ(α2+βΛα3ρ+q(1−e−μτ1)+βΛα4ρ+q(1−e−μτ1))]⟨I(t)⟩≥(Rl0−1)(ρ+ω+δ+γ). |
Then, we get
lim inft→∞⟨I(t)⟩≥(Rl0−1)(ρ+ω+δ+γ)λ=I∗>0. | (4.2) |
From (3.2) and (4.2), we can check that
lim supt→∞⟨S(t)⟩≤Λρ+q(1−e−μτ1)−lim inft→∞[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]⟨I(t)⟩≤Λρ+q(1−e−μτ1)−[(ρ+γ(1−e−μτ1)+ω+δ)(ρ+μ)+ε(ρ+γ(1−e−μτ2)+ω)(ρ+μ+ε)(ρ+q(1−e−μτ1))]I∗. |
Also, from Theorem 2.2 and the first equation of (1.3) gives
dS(t)≥[Λ−(ρ+q+βΛρ+q(1−e−μτ1))S(t)]dt+σ1S(t)dB1(t)+∫Yη1(y)S(t−)˜N(dt,dy). |
Then
(ρ+q+βΛρ+q(1−e−μτ1))⟨S(t)⟩≥Λ−S(t)−S(0)t+1t∫t0σ1S(t)dB1(t)+1t∫t0∫Yη1(y)S(t−)˜N(dt,dy). |
By the law of large numbers for martingales and S(t)∈Γ, we obtain
lim inft⟶∞⟨S(t)⟩≥Λ(ρ+q+βΛρ+q(1−e−μτ1)). |
From the third equation of the system (1.3), we have
Q(t)−Q(0)t=δ1t∫t0I(s)ds−(ρ+μ+ε)1t∫t0Q(s)ds+1t∫t0σ3QdB3(t)+1t∫t0∫Yη3(y)Q(t−)˜N(dt,dy). |
Hence, from the strong law of large numbers for local martingales we get
lim inft→∞⟨Q(t)⟩=δρ+μ+εlim inft→∞⟨I(t)⟩≥δρ+μ+εI∗>0. |
So our proof is complete.
In this section, we shall use Euler-Maruyama numerical approximation [43] to illustrate the rigor of our analytical results. The two examples are given below concern the results obtained in Theorems 3.1 and 4.1. Moreover, we numerically simulate the solution of a corresponding system (1.2) for the comparison.
Example 5.1. According to the parameters in the paper [31], Choose Λ=0.5, β=0.2, ρ=0.1, q=0.1, δ=0.15, γ=0.11,ω=0.12,μ=0.2,ϵ=0.3,τ1=0.1,τ2=0.5,τ3=0.5,α1=0.22,α2=α3=α4=0.2, η1=0.01,η2=0.1,η3=0.03 and Y=(0,∞),v(Y)=1, σ1=σ2=σ3=0.1. Then R0=1.04>1,Rs0=1.024>1, however Rl0=0.9921<1.
The computer simulation illustrated by Figure 1, support the result of Theorem 3.1. That is to say, the disease in system (1.3) (with jump) dies out exponentially with probability one, although the disease in system (1.1) (without jumps) persists. If we decrease β to 0.19, we get R0=0.995<1,Rs0=0.97<1, and Rl0=0.9421<1. By Theorem 3.1, the disease will tend to zero exponentially with probability one.
Example 5.2. Choose β=0.19 and other parameters be the same as Example 5.1. Then R0=1.15>1, Rs0=1.12>1, Rl0=1.09121>1. By Theorem 4.1, We can get limt→∞⟨S(t)⟩=1.0921>0, limt→∞⟨I(t)⟩=0.593>0, limt→∞⟨Q(t)⟩=0.1201>0. This means that the disease persists almost surely. The come simulations showed in Figure 2 support the result 4.1 clearly.
Since the role of isolation has been shown to be meaningful for the prevention and control of infectious diseases such as for the recent influenza disease COVID-19. Therefore the dynamical behavior of a delayed SIQR stochastic epidemic model with Lévy noise is studied. In comparison with the studies of [23], we explore a new response function f(S,I) and consider the Lévy noise. Where the reaction function can contain forms such as Holling Type II incidence f(S,I)I=βSI1+α2S, Saturation rate f(S,I)I=βSI1+α1I, Bilinear functional response f(S,I)I=βSI, Beddington-DeAngelis rate f(S,I)I=βSI1+α1I+α2S, Crowly-Martin functional response f(S,I)I=βSI1+α1I+α2S+α1α2SI, Non-monotonous incidence f(S,I)I=βSI1+α4I2, Holling Type IV incidence f(S,I)I=βSI1+α2I+α4I2. A threshold value Rl0 is derived
● If Rl0>1, the disease will persistence in mean.
● If Rl0<1, the disease will tend to extinction exponetially.
We can also compare the expressions for Rl0 and the parameter R0. Obviously, when we ignore the environmental noise and Lévy noise, we show that Rl0=Rs0=R0, this implies that the stochastic model is an extension of corresponding deterministic model.
The following topics deserve further discussion. Since white noise is a continuous stochastic perturbation, some discontinuous perturbations such as the color noises can be further investigated and the effect of the impulsive can also be considered. At the same time, we can also try to find the probability density function by solving the Fokker-Planck equation of stochastic model (1.3). We left the above topics for future work.
We will say thanks to reviewers for the valuable ideas and comments on improving the article. The said article is supported by Funding for National Nature Science Foundation of China (11571088) and National Nature Science Foundation of China (12071105).
The authors declare that there are no conflicts of interest.
[1] |
Q. Liu, D. Jiang, Stationary distribution of a stochastic cholera model with imperfect vaccination, Physica A, 550 (2020), 124031. https://doi.org/10.1016/j.physa.2019.124031 doi: 10.1016/j.physa.2019.124031
![]() |
[2] |
D. Okuonghae, A. Omame, Analysis of a mathematical model for COVID-19 population dynamics in Lagos, Nigeria, Chaos Soliton. Fract., 139 (2020), 110032. https://doi.org/10.1016/j.chaos.2020.110032 doi: 10.1016/j.chaos.2020.110032
![]() |
[3] |
J. Asamoahab, Z. Jin, G. Sun, B. Seidu, E. Yankson, A. Abidemi, et al., Sensitivity assessment and optimal economic evaluation of a new COVID-19 compartmental epidemic model with control interventions, Chaos Soliton. Fract., 146 (2021), 110885. https://doi.org/10.1016/j.chaos.2021.110885 doi: 10.1016/j.chaos.2021.110885
![]() |
[4] |
D. Kuang, Q. Yin, J. Li, Dynamics of stochastic HTLV-I infection model with nonlinear CTL immune response, Math. Method. Appl. Sci., 44 (2021), 14059–14078. https://doi.org/10.1002/mma.7674 doi: 10.1002/mma.7674
![]() |
[5] |
J. Asamoah, E. Okyere, A. Abidemi, S. Moore, G. Sun, Z. Jin, et al., Optimal control and comprehensive cost-effectiveness analysis for COVID-19, Results Phys., 33 (2022), 105177. https://doi.org/10.1016/j.rinp.2022.105177 doi: 10.1016/j.rinp.2022.105177
![]() |
[6] |
L. Chang, C. Liu, G. Sun, Z. Wang, Z. Jin, Delay-induced patterns in a predator-prey model on complex networks with diffusion, New. J. Phys., 21 (2019), 073035. https://doi.org/10.1088/1367-2630/ab3078 doi: 10.1088/1367-2630/ab3078
![]() |
[7] |
J. Asamoah, M. Owusu, Z. Jin, F. Oduro, A. Abidemi, E. Gyasi, Global stability and cost-effectiveness analysis of COVID-19 considering the impact of the environment: using data from Ghana, Chaos Soliton. Fract., 140 (2020), 110103. https://doi.org/10.1016/j.chaos.2020.110103 doi: 10.1016/j.chaos.2020.110103
![]() |
[8] |
G. Sun, H. Zhang, J. Wang, J. Li, Y. Wang, L. Li, et al., Mathematical modeling and mechanisms of pattern formation in ecological systems: a review, Nonlinear Dyn., 104 (2021), 1677–1696. https://doi.org/10.1007/s11071-021-06314-5 doi: 10.1007/s11071-021-06314-5
![]() |
[9] |
S. Nadim, I. Ghosh, J. Chattopadhyay, Short-term predictions and prevention strategies for COVID-19: a model-based study, Appl. Math. Comput., 404 (2021), 126251. https://doi.org/10.1016/j.amc.2021.126251 doi: 10.1016/j.amc.2021.126251
![]() |
[10] |
M. De la Sen, S. Alonso-Quesada, A. Ibeas, On the stability of an SEIR epidemic model with distributed time-delay and a general class of feedback vaccination rules, Appl. Math. Comput., 270 (2015), 953–976. https://doi.org/10.1016/j.amc.2015.08.099 doi: 10.1016/j.amc.2015.08.099
![]() |
[11] |
J. Mateus, C. Silva, A non-autonomous SEIRS model with general incidence rate, Appl. Math. Comput., 247 (2014), 169–189. https://doi.org/10.1016/j.amc.2014.08.078 doi: 10.1016/j.amc.2014.08.078
![]() |
[12] |
A. Lahrouz, L. Omari, D. Kiouach, Global analysis of a deterministic and stochastic nonlinear SIRS epidemic model, Nonlinear Anal.-Model., 16 (2011), 59–76. https://doi.org/10.15388/NA.16.1.14115 doi: 10.15388/NA.16.1.14115
![]() |
[13] |
X. B. Zhang, X. H. Zhang, The threshold of a deterministic and a stochastic SIQS epidemic model with varying total population size, Appl. Math. Model., 91 (2021), 749–767. https://doi.org/10.1016/j.apm.2020.09.050 doi: 10.1016/j.apm.2020.09.050
![]() |
[14] |
Y. Zhang, X. Ma, A. Din, Stationary distribution and extinction of a stochastic SEIQ epidemic model with a general incidence function and temporary immunity, AIMS Mathematics, 6 (2021), 12359–12378. https://doi.org/10.3934/math.2021715 doi: 10.3934/math.2021715
![]() |
[15] |
J. Dimi, T. Mbaya, Dynamics analysis of stochastic tuberculosis model transmission with immune response, AIMS Mathematics, 3 (2018), 391–408. https://doi.org/10.3934/Math.2018.3.391 doi: 10.3934/Math.2018.3.391
![]() |
[16] |
C. Qin, J. Du, Y. Hui, Dynamical behavior of a stochastic predator-prey model with Holling-type III functional response and infectious predator, AIMS Mathematics, 7 (2022), 7403–7418. https://doi.org/10.3934/math.2022413 doi: 10.3934/math.2022413
![]() |
[17] |
A. Din, A. Khan, D. Baleanu, Stationary distribution and extinction of stochastic coronavirus (COVID-19) epidemic model, Chaos Soliton. Fract., 139 (2020), 110036. https://doi.org/10.1016/j.chaos.2020.110036 doi: 10.1016/j.chaos.2020.110036
![]() |
[18] |
C. Chen, D. Gao, P. Guo, The stationary distribution of a stochastic rumor spreading model, AIMS Mathematics, 6 (2021), 1234–1248. https://doi.org/10.3934/math.2021076 doi: 10.3934/math.2021076
![]() |
[19] |
Q. Liu, D. Jiang, T. Hayat, A. Alsaedi, Dynamics of a stochastic multigroup SIQR epidemic model with standard incidence rates, J. Franklin. I., 356 (2019), 2960–2993. https://doi.org/10.1016/j.jfranklin.2019.01.038 doi: 10.1016/j.jfranklin.2019.01.038
![]() |
[20] |
Q. Liu, D. Jiang, T. Hayat, A. Alsaedi, Dynamical behavior of a stochastic epidemic model for cholera, J. Franklin. I., 356 (2019), 7486–7514. https://doi.org/10.1016/j.jfranklin.2018.11.056 doi: 10.1016/j.jfranklin.2018.11.056
![]() |
[21] |
T. Gard, Persistence in stochastic food web models, Bltn. Mathcal. Biology, 46 (1984), 357–370. https://doi.org/10.1007/BF02462011 doi: 10.1007/BF02462011
![]() |
[22] |
Y. Cai, Y. Kang, M. Banerjee, W. Wang, A stochastic SIRS epidemic model with infectious force under intervention strategies, J. Differ. Equations, 259 (2015), 7463–7502. https://doi.org/10.1016/j.jde.2015.08.024 doi: 10.1016/j.jde.2015.08.024
![]() |
[23] |
N. Du, N. Nhu, Permanence and extinction for the stochastic SIR epidemic model, J. Differ. Equations, 269 (2020), 9619–9652. https://doi.org/10.1016/j.jde.2020.06.049 doi: 10.1016/j.jde.2020.06.049
![]() |
[24] |
N. Du, N. Dieu, N. Nhu, Conditions for permanence and ergodicity of certain SIR epidemic models, Acta Appl. Math., 160 (2019), 81–99. https://doi.org/10.1007/s10440-018-0196-8 doi: 10.1007/s10440-018-0196-8
![]() |
[25] |
S. Ruschel, T. Pereira, S. Yanchuk, L. Young, An SIQ delay differential equations model for disease control via isolation, J. Math. Biol., 79 (2019), 249–279. https://doi.org/10.1007/s00285-019-01356-1 doi: 10.1007/s00285-019-01356-1
![]() |
[26] |
K. Fan, Y. Zhang, S. Gao, X. Wei, A class of stochastic delayed SIR epidemic models with generalized nonlinear incidence rate and temporary immunity, Physica A, 481 (2017), 198–208. https://doi.org/10.1016/j.physa.2017.04.055 doi: 10.1016/j.physa.2017.04.055
![]() |
[27] |
K. Fan, Y. Zhang, S. Guo, S. Chen, A delayed vaccinated epidemic model with nonlinear incidence rate and Lévy jumps, Physica A, 544 (2020), 123379. https://doi.org/10.1016/j.physa.2019.123379 doi: 10.1016/j.physa.2019.123379
![]() |
[28] |
Z. Cao, Y. Shi, X. Wen, L. Liu, J. Hu, Analysis of a hybrid switching SVIR epidemic model with vaccination and Lévy noise, Physica A, 537 (2020), 122749. https://doi.org/10.1016/j.physa.2019.122749 doi: 10.1016/j.physa.2019.122749
![]() |
[29] |
H. Yang, Z. Jin, Stochastic SIS epidemic model on network with Lévy noise, Stoch. Anal. Appl., 40 (2022), 520–538. https://doi.org/10.1080/07362994.2021.1930051 doi: 10.1080/07362994.2021.1930051
![]() |
[30] |
M. Fatini, I. Sekkak, A. Laaribi, A threshold of a delayed stochastic epidemic model with Crowly-Martin functional response and vaccination, Physica A, 520 (2019), 151–160. https://doi.org/10.1016/j.physa.2019.01.014 doi: 10.1016/j.physa.2019.01.014
![]() |
[31] |
T. Khan, A. Khan, G. Zaman, The extinction and persistence of the stochastic hepatitis B epidemic model, Chaos Soliton. Fract., 108 (2018), 123–128. https://doi.org/10.1016/j.chaos.2018.01.036 doi: 10.1016/j.chaos.2018.01.036
![]() |
[32] |
L. Huo, J. Jiang, S. Gong, B. He, Dynamical behavior of a rumor transmission model with Holling-type II functional response in emergency event, Physica A, 450 (2016), 228–240. https://doi.org/10.1016/j.physa.2015.12.143 doi: 10.1016/j.physa.2015.12.143
![]() |
[33] |
S. Ruan, W. Wang, Dynamical behavior of an epidemic model with a nonlinear incidence rate, J. Differ. Equations, 188 (2003), 135–163. https://doi.org/10.1016/S0022-0396(02)00089-X doi: 10.1016/S0022-0396(02)00089-X
![]() |
[34] |
N. Dieu, D. Nguyen, N. Du, G. Yin, Classification of asymptotic behavior in a stochastic SIR model, SIAM J. Appl. Dyn. Syst., 15 (2016), 1062–1084. https://doi.org/10.1137/15M1043315 doi: 10.1137/15M1043315
![]() |
[35] |
N. Du, N. Nhu, Permanence and extinction of certain stochastic SIR models perturbed by a complex type of noises, Appl. Math. Lett., 64 (2017), 223–230. https://doi.org/10.1016/j.aml.2016.09.012 doi: 10.1016/j.aml.2016.09.012
![]() |
[36] |
P. Naik, J. Zu, M. Ghoreishi, Stability analysis and approximate solution of SIR epidemic model with Crowley-Martin type functional response and Holling type-II treatment rate by using homotopy analysis method, J. Appl. Anal. Comput., 10 (2020), 1482–1515. https://doi.org/10.11948/20190239 doi: 10.11948/20190239
![]() |
[37] |
X. Zhang, H. Huo, H. Xiang, X. Meng, An SIRS epidemic model with pulse vaccination and non-monotonic incidence rate, Nonlinear Anal.-Hybri., 8 (2013), 13–21. https://doi.org/10.1016/j.nahs.2012.08.001 doi: 10.1016/j.nahs.2012.08.001
![]() |
[38] |
D. Xu, M. Liu, X. Xu, Analysis of a stochastic predator-prey system with modified Leslie-Gower and Holling-type IV schemes, Physica A, 537 (2020), 122761. https://doi.org/10.1016/j.physa.2019.122761 doi: 10.1016/j.physa.2019.122761
![]() |
[39] |
Q. Liu, D. Jiang, Dynamical behavior of a higher order stochastically perturbed HIV/AIDS model with differential infectivity and amelioration, Chaos Soliton. Fract., 141 (2020), 110333. https://doi.org/10.1016/j.chaos.2020.110333 doi: 10.1016/j.chaos.2020.110333
![]() |
[40] | X. Mao, Stochastic differential equations and applications, Cambridge: Woodhead Publishing, 2007. |
[41] |
R. Liptser, A strong law of large numbers for local martingales, Stochastics, 3 (1980), 217–228. https://doi.org/10.1080/17442508008833146 doi: 10.1080/17442508008833146
![]() |
[42] |
C. Ji, D. Jiang, Threshold behaviour of a stochastic SIR model, Appl. Math. Model., 38 (2014), 5067–5079. https://doi.org/10.1016/j.apm.2014.03.037 doi: 10.1016/j.apm.2014.03.037
![]() |
[43] |
P. Protter, D. Talay, The Euler scheme for Lévy driven stochastic differential equations, Ann. Probab., 25 (1997), 393–423. https://doi.org/10.1214/aop/1024404293 doi: 10.1214/aop/1024404293
![]() |
1. | Xiaodong Wang, Kai Wang, Zhidong Teng, Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence, 2023, 8, 2473-6988, 2829, 10.3934/math.2023148 | |
2. | Mohamed Mehdaoui, Abdesslem Lamrani Alaoui, Mouhcine Tilioua, Analysis of a stochastic SVIR model with time‐delayed stages of vaccination and Lévy jumps, 2023, 0170-4214, 10.1002/mma.9198 | |
3. | S. El Attouga, D. Bouggar, M. El Fatini, A. Hilbert, R. Pettersson, Lévy noise with infinite activity and the impact on the dynamic of an SIRS epidemic model, 2023, 03784371, 128701, 10.1016/j.physa.2023.128701 | |
4. | Kai Wang, Hongjie Fan, Yanling Zhu, Dynamics and application of a generalized SIQR epidemic model with vaccination and treatment, 2023, 0307904X, 10.1016/j.apm.2023.03.036 | |
5. | Qigui Yang, Jiabing Huang, A STOCHASTIC MULTI-SCALE COVID-19 MODEL WITH INTERVAL PARAMETERS, 2024, 14, 2156-907X, 515, 10.11948/20230298 |