Research article Special Issues

A theta-scheme approximation of basic reproduction number for an age-structured epidemic system in a finite horizon

  • This paper focuses on numerical approximation of the basic reproduction number R0, which is the threshold defined by the spectral radius of the next-generation operator in epidemiology. Generally speaking, R0 cannot be explicitly calculated for most age-structured epidemic systems. In this paper, for a deterministic age-structured epidemic system and its stochastic version, we discretize a linear operator produced by the infective population with a theta scheme in a finite horizon, which transforms the abstract problem into the problem of solving the positive dominant eigenvalue of the next-generation matrix. This leads to a corresponding threshold R0,n. Using the spectral approximation theory, we obtain that R0,nR0 as n+. Some numerical simulations are provided to certify the theoretical results.

    Citation: Wenjuan Guo, Ming Ye, Xining Li, Anke Meyer-Baese, Qimin Zhang. A theta-scheme approximation of basic reproduction number for an age-structured epidemic system in a finite horizon[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 4107-4121. doi: 10.3934/mbe.2019204

    Related Papers:

    [1] Shuixian Yan, Sanling Yuan . Critical value in a SIR network model with heterogeneous infectiousness and susceptibility. Mathematical Biosciences and Engineering, 2020, 17(5): 5802-5811. doi: 10.3934/mbe.2020310
    [2] Baoxiang Zhang, Yongli Cai, Bingxian Wang, Weiming Wang . Dynamics and asymptotic profiles of steady states of an SIRS epidemic model in spatially heterogenous environment. Mathematical Biosciences and Engineering, 2020, 17(1): 893-909. doi: 10.3934/mbe.2020047
    [3] Mostafa Adimy, Abdennasser Chekroun, Claudia Pio Ferreira . Global dynamics of a differential-difference system: a case of Kermack-McKendrick SIR model with age-structured protection phase. Mathematical Biosciences and Engineering, 2020, 17(2): 1329-1354. doi: 10.3934/mbe.2020067
    [4] F. Berezovskaya, G. Karev, Baojun Song, Carlos Castillo-Chavez . A Simple Epidemic Model with Surprising Dynamics. Mathematical Biosciences and Engineering, 2005, 2(1): 133-152. doi: 10.3934/mbe.2005.2.133
    [5] Abdelheq Mezouaghi, Salih Djillali, Anwar Zeb, Kottakkaran Sooppy Nisar . Global proprieties of a delayed epidemic model with partial susceptible protection. Mathematical Biosciences and Engineering, 2022, 19(1): 209-224. doi: 10.3934/mbe.2022011
    [6] J. Amador, D. Armesto, A. Gómez-Corral . Extreme values in SIR epidemic models with two strains and cross-immunity. Mathematical Biosciences and Engineering, 2019, 16(4): 1992-2022. doi: 10.3934/mbe.2019098
    [7] Kento Okuwa, Hisashi Inaba, Toshikazu Kuniya . Mathematical analysis for an age-structured SIRS epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 6071-6102. doi: 10.3934/mbe.2019304
    [8] Yanmei Wang, Guirong Liu . Dynamics analysis of a stochastic SIRS epidemic model with nonlinear incidence rate and transfer from infectious to susceptible. Mathematical Biosciences and Engineering, 2019, 16(5): 6047-6070. doi: 10.3934/mbe.2019303
    [9] Xia Wang, Shengqiang Liu . Global properties of a delayed SIR epidemic model with multiple parallel infectious stages. Mathematical Biosciences and Engineering, 2012, 9(3): 685-695. doi: 10.3934/mbe.2012.9.685
    [10] Yan-Xia Dang, Zhi-Peng Qiu, Xue-Zhi Li, Maia Martcheva . Global dynamics of a vector-host epidemic model with age of infection. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1159-1186. doi: 10.3934/mbe.2017060
  • This paper focuses on numerical approximation of the basic reproduction number R0, which is the threshold defined by the spectral radius of the next-generation operator in epidemiology. Generally speaking, R0 cannot be explicitly calculated for most age-structured epidemic systems. In this paper, for a deterministic age-structured epidemic system and its stochastic version, we discretize a linear operator produced by the infective population with a theta scheme in a finite horizon, which transforms the abstract problem into the problem of solving the positive dominant eigenvalue of the next-generation matrix. This leads to a corresponding threshold R0,n. Using the spectral approximation theory, we obtain that R0,nR0 as n+. Some numerical simulations are provided to certify the theoretical results.


    Infectious diseases are known to have caused huge devastation and loss of human life since ancient times, and they have been posing a great threat, especially to developing countries. Over the last two decades, human beings have witnessed numbers of major outbreaks of infectious diseases such as SARS (2003), H7N9 (2013) in China, and the Ebola virus (2014) in West Africa [1,2]. These outbreaks have huge negative impacts on people's health and social stability. Hence, how to prevent the spread of diseases effectively becomes an important and hot issue. Meanwhile, more and more researchers have been focusing on studying the transmission mechanism of epidemic diseases in recent years; see [3,4,5,6,7].

    It has been widely agreed that the mathematical modeling of infectious diseases is an effective tool and plays a crucial role in investigating the transmission mechanism and dynamics behaviors of various communicable diseases [8]. One of the pioneer works in this field was done by Kermack and McKendrick [9,10,11], where the authors established two classic compartmental epidemic models called susceptible-infected-recovered (SIR) and susceptible-infected-susceptible (SIS) models. From then on, many mathematical models about the prevalence of diseases which descend from the two fundamental epidemic models have been widely used in investigating the spreading mechanics of epidemic diseases [3,5,6,12]. In SIR models, one usually supposes that recovered individuals could not be infected again because infected individuals become permanently immune once recovered. However, for some diseases (influenza, malaria), the recovered individuals may lose immunity after a while and they will be susceptible again. By taking this factor into account, SIRS models are proposed and investigated [3,5]. Moreover, in some bacterial agent diseases such as plague and venereal diseases, the recovered individuals cannot keep immunity for a long time. Therefore, it is more reasonable and appropriate to incorporate the case that a part of infected individuals may go back directly to the susceptible class after some treatments into epidemiology modeling [5,6,12]. In [6], the authors established an SIRS model with nonlinear incidence rate as well as transfer from infectious to susceptible and investigated the global dynamics of the model. In these compartmental epidemic models, there is always an underlying assumption that the population is sufficiently large and all individuals are mixed uniformly. It means that every individual has the same possibility to contact others, whereas it cannot reflect the realistic feature of the epidemic transmission mechanics completely since different individuals may have different numbers of acquaintances, and the contact behaviors of individuals exhibit heterogeneity [13,14]. Thus, it is essential to take the effect of contact heterogeneity into consideration in investigating the transmission mechanics of epidemic diseases.

    It is well known that complex networks provide a powerful framework for describing and quantifying the topological structures of various social, economic and biological systems [15]. In particular, the heterogeneity of contact patterns can be well characterized by degree fluctuation of heterogeneous networks [16]. In view of this, the structures of complex network are embedded into the traditional epidemic models, and the epidemic dynamics in complex networks have attracted increasing attention and been studied extensively in recent years [17,18,19,20,21,22,23,24,25,26,27]. In 2001, Pastor-Satorras and Vespignani [20] proposed and studied an SIS epidemic model in highly heterogeneous networks (i.e., scale-free networks) and showed that a highly heterogeneous contact pattern can result in the absence of any epidemic threshold for the first time, i, e., the threshold approaches zero in the limit of a large number of edges and nodes, and even quite a small infectious rate can lead to a major epidemic outbreaks. This is a new epidemic spreading phenomenon which is completely different from previous works. And this result was proved rigorously by Wang and Dai in [21]. The dynamics of SIR model in heterogeneous networks was also investigated in [23], and the authors found that the epidemic threshold will increase with network size on finite networks. In [27], Huo et al. proposed a fractional SIR model with birth and death rates on heterogeneous complex networks and obtained that the threshold value R0 determines the dynamics of the model. In order to investigate the heterogeneities of disease transmission about SIRS models, Li et al. [25] proposed a new SIRS model with nonlinear infectivity as well as birth and death rates on complex networks and obtained that the disease-free equilibrium is globally asymptotically stable when the basic reproduction number R0<1; when R0>1, then there exists a unique endemic equilibrium which is globally asymptotically stable. In [24] and [18], the authors proposed and investigated an SIRS epidemic model with infective individuals entering into the removed or susceptible on complex network respectively. But both of them only focused on the threshold of the epidemic models and the dynamics of equilibria were not discussed.

    To the best of our knowledge, the factor of population turnover is not considered in most of aforementioned models, and there are few studies [25,27,28,29,30] about incorporating demographics into network disease models. Inspired by the works of [18,24,25] and [28], in this paper, we shall establish an improved SIRS model on complex heterogeneous networks by taking both demographics and transfer from infectious to susceptible into consideration. Our main goal is to discuss the global epidemic dynamics of the model and provide some strategies to prevent the epidemic outbreaks. The remainder of this article is organized as follows. A network-based SIRS model with demographics and transfer from infectious to susceptible is proposed in Section 2. The basic reproduction number of the model and the stability of equilibria are investigated in Section 3 and Section 4. In Section 5, the effects of three major immunization strategies are investigated and compared. In Section 6, some numerical simulations are given to support the theoretical analysis. Finally, conclusions are drawn in Section 7.

    In this paper, we define the whole population as a finite size network. Every node of the network represents an individual and the edges are the interactions among individuals. The degree k of a node denotes the number of edges connected to the node. Let Sk(t), Ik(t) and Rk(t) be the number of the susceptible, infected and recovered nodes with degree k at time t, respectively, for k=1,2,...,n, where n is the maximal degree of the nodes among the network. The transmission sketch is shown in Figure 1.

    Figure 1.  Transition diagram for the SIRS epidemic model.

    Based on the mean-field theory, the dynamics of the network-based model can be described by the following differential equations

    {dSk(t)dt=bλ(k)Sk(t)Θ(t)dSk(t)+γIk(t)+δRk(t),dIk(t)dt=λ(k)Sk(t)Θ(t)(γ+d+α)Ik(t),dRk(t)dt=αIk(t)(d+δ)Rk(t), (2.1)

    where the parameter b denotes the number of newly born susceptible nodes with degree k per unit time; λ(k) (degree-dependent) is the infected rate of a susceptible individual; d represents the death rate; γ is the transfer rate from the infected class to the susceptible class; δ denotes the immunity loss rate; α represents the recovery rate from infected individuals. The parameters b,d,λ(k),α,γandδ are all nonnegative by their epidemiological meaning. Θ(t) represents the probability that a randomly chosen edge of a node degree k points to an infected node at time t and it satisfies the following equation.

    Θ(t)=ni=1P(ik)φ(i)iIi(t)Ni(t),

    where P(ik) represents the probability that a node with degree k is connected to a node with degree i. In this paper, we investigate epidemic transmission on uncorrelated networks, then the probability is considered independent of the connectivity of the node from which the link is emanating[20]. Hence P(ik) = iP(i)/k, where P(k) represents the probability of a randomly chosen node with degree k and nk=1P(k)=1. k=nk=1kP(k) is the average degree of the network. φ(i) represents the infectivity of a node with degree i, that is, the mean number of edges from which a node with degree i can transmit diseases [19]. Nk(t)=Sk(t)+Ik(t)+Rk(t) denotes the number of nodes with degree k at time t.

    Adding the three equations of model (2.1), we have dNk(t)dt=bdNk(t). Then Nk(t)=Nk(0)edt+bd(1edt). For the purpose of having a population of constant size, throughout this paper we always assume that

    (H1)Sk(0)+Ik(0)+Rk(0)=Nk(0)=bd.

    Hence, we have Sk(t)+Ik(t)+Rk(t)=Nk(t)=bd for all t0. Thus

    Θ(t)=1kdbnk=1φ(k)P(k)Ik(t).

    For the practical consideration, the initial conditions of system (2.1) take the form

    Sk(0)>0,Ik(0)0,Rk(0)0,Θ(0)>0,k=1,2,...,n. (2.2)

    In this paper, beside the hypothesis (H1), we also need some hypotheses shown as below:

    (H2) the infection rate is bounded, i.e., there exists two constants μ,ν>0, such that 0<μλ(k)ν;

    (H3) the network should be static, that is, the degree of each node is time invariant.

    Now, we reveal some properties of solutions of system (2.1).

    Lemma 2.1. Suppose (S1(t),I1(t),R1(t),,Sn(t),In(t),Rn(t)) is a solution of system (2.1) with the initial conditions (2.2). Then we have the following conclusion

    Sk(t)>0,Ik(t)0,Rk(t)0,foranyk=1,2,...,nandallt>0.

    Proof. Firstly, we prove Θ(t)>0 for all t>0. By the second equation of system (2.1), we have

    dΘ(t)dt=1kdbnk=1φ(k)P(k)[λ(k)Sk(t)Θ(t)(γ+d+α)Ik(t)]=1kdbnk=1λ(k)φ(k)P(k)Sk(t)Θ(t)1kdbnk=1(γ+d+α)φ(k)P(k)Ik(t)=Θ(t)[1kdbnk=1λ(k)φ(k)P(k)Sk(t)(γ+d+α)]. (2.3)

    Obviously, we have

    Θ(t)=Θ(0)exp[(γ+d+α)t+1kdbt0nk=1λ(k)φ(k)P(k)Sk(s)ds].

    Since Θ(0)>0, one can obtain that Θ(t)>0 for all t>0.

    Note that Sk(0)>0. Based on the fact of the first equation of system (2.1) and the continuity of Sk(t), we could find a sufficiently small t0>0 such that Sk(t)>0 for t(0,t0). We now prove that Sk(t)>0 for all t>0. Assume Sk(t)>0 is not always valid, then there exists i{1,2,...n} and the first time t1t0>0 such that Si(t1)=0 and Si(t)>0 for t(0,t1). From the second equation of system (2.1), we have dIi(t)dt>(γ+d+α)Ii(t) for t(0,t1). Thus, Ii(t)>Ii(0)e(γ+d+α)t0 for t(0,t1).

    Combining Ii(t)0 for t(0,t1) with the third equation of system (2.1), we have dRi(t)dt>(d+δ)Ri(t) for t(0,t1). Hence, one can obtain that Ri(t)>Ri(0)e(d+δ)t for t(0,t1). By the continuity of Ii(t) and Ri(t), we know Ii(t1)0 and Ri(t)0. Then the first equation of system (2.1) shows

    dSi(t)dtt=t1=b+γIi(t1)+δRi(t1)b>0.

    So, one can find that Si(t)<Si(t1) for t(0,t1). Due to Si(t1)=0, it contradicts to Sk(t)>0 obviously. Therefore, Sk(t)>0 for all t>0 and all k=1,2,...,n. By similar discussions, we also have Ik(t)0 and Rk(t)0 for all t>0. Therefore, this completes the proof.

    Hence, we only need to focus on the dynamics of solutions of system (2.1) in the following feasible bounded region

    Γ={(S1,I1,R1,,Sn,In,Rn):Sk>0,Ik0,Rk0,Sk+Ik+Rk=bd,k=1,2,,n}.

    Remark 2.1. It's remarkable that system (2.1) includes some special cases, if b=d=α=δ=0, system (2.1) turns into an SIS model [20]; if b=d=γ=0, model (2.1) is an SIR model [23]; if γ=0 and b=d, system (2.1) is an SIRS model [25].

    Remark 2.2. In recent years, many different types of λ(k) and φ(k) have been studied. The most common types for λ(k) is λ(k)=λk (see, [30,21]). For φ(k), one can see φ(k)=k is choosen in [30,18]. Furthermore, in [23], the authors adopt a general form of φ(k)=akα1+vkα, where 0α1,a>0 and v0. The types φ(k)=C (C is a constant) [22] and φ(k)=km (0<m1) [17] are discussed as well.

    In epidemiological investigations, the basic reproduction number R0 is widely used to measure the epidemic potential.The term basic reproduction number represents the mean number of secondary cases produced by a typical infectious individual during its entire infectious period in a completely susceptible population [31,32]. By the methods in [31], the transmission matrix F for the rate of appearance of new infections and the transition matrix V for the transfer rate of individuals among compartments around the disease-free equilibrium E0 are denoted as follows, respectively

    F=(λ(1)φ(1)P(1)kλ(1)φ(2)P(2)kλ(1)φ(n)P(n)kλ(2)φ(1)P(1)kλ(2)φ(2)P(2)kλ(2)φ(n)P(n)kλ(n)φ(1)P(1)kλ(n)φ(2)P(2)kλ(n)φ(n)P(n)k),

    and

    V=(γ+d+α000γ+d+α000γ+d+α).

    Hence, the basic reproduction number R0=ρ(FV1) is the spectral radius of the next generation matrix FV1. By direct calculations, we have

    R0=ρ(FV1)=λ(k)φ(k)(γ+d+α)k,

    where λ(k)φ(k)=nk=1λ(k)φ(k)P(k).

    Next, we will investigate the existence of equilibria of system (2.1).

    Theorem 3.1. Consider system (2.1), then we have following results:

    (i) There always exists a disease-free equilibrium E0=(bd,0,0,,bd,0,0).

    (ii) There is a unique endemic equilibrium E=(S1,I1,R1,,Sn,In,Rn) when R0>1, where

    Sk=(γ+d+α)λ(k)ΘIk,Rk=αd+δIk,
    Ik=b(d+δ)λ(k)Θ(d+δ)d+αdλ(k)Θ+d(d+δ)(γ+d+α)

    with Θ=1kdbnk=1φ(k)P(k)Ik,k=1,2,,n.

    Proof. Clearly, E0 is always a disease-free equilibrium of system (2.1).

    Now, let us consider the existence of E. Note that the equilibrium E satisfies the following equalities

    {bλ(k)SkΘdSk+γIk+δRk=0,λ(k)SkΘ(γ+d+α)Ik=0,αIk(d+δ)Rk=0. (3.1)

    From the second and the third equations of (3.1), we have

    Sk=(γ+d+α)λ(k)ΘIk, (3.2)
    Rk=αd+δIk. (3.3)

    Substituting (3.2) and (3.3) into the first equation of (3.1), one can obtain

    Ik=b(d+δ)λ(k)Θd(d+δ+α)λ(k)Θ+d(d+δ)(γ+d+α). (3.4)

    Substituting (3.4) into the expression of Θ, we obtain a self-consistency equation about Θ.

    Θ=1kdbnk=1φ(k)P(k)b(d+δ)λ(k)Θd(d+δ+α)λ(k)Θ+d(d+δ)(γ+d+α):=f(Θ). (3.5)

    Obviously, Θ=0 is a solution of (3.5). Note that

    {f(1)<1kdbnk=1φ(k)P(k)bd1,df(Θ)dΘ=1kdbnk=1φ(k)P(k)b(d+δ)2λ(k)d(γ+d+α)[d(d+δ+α)λ(k)Θ+d(d+δ)(γ+d+α)]2>0,df2(Θ)dΘ2=21kdbnk=1φ(k)P(k)b(d+δ)2λ2(k)d(γ+d+α)[d(d+δ+α)][d(d+δ+α)λ(k)Θ+d(d+δ)(γ+d+α)]3<0.

    If R0>1, then we have df(Θ)dΘ|Θ=0>1. Consequently, the equation (3.5) has a unique nontrivial solution Θ(Θ(0,1)). It means that there exists one and only one endemic equilibrium when R0>1. This finishes the proof.

    Remark 3.1. From Theorem 3.1, we know that the existence of the endemic equilibrium depends on R0. Obviously, R0 is in proportion to the heterogeneous parameter λ(k)φ(k)k which is related to the topology of the networks closely. We also find that R0 has nothing to do with the immunity loss rate δ. It seems that the transfer rate γ and the recovery rate α have the same effects on R0, i.e., increasing the transfer rate γ and the recovery rate α will decrease R0.

    In this section, we will consider the global stability of equilibria E0 and E. To begin with we will discuss the local asymptotical stability of the disease-free equilibrium E0.

    Theorem 4.1. The disease-free equilibrium E0 of system (2.1) is locally asymptotically stable when R0<1 and is unstable when R0>1.

    Proof. As shown in Section 2, for the purpose of having a population of constant size, we assume that the initial values always satisfy the hypothesis (H1). Then we can obtain Sk(t)+Ik(t)+Rk(t)=Nk(t)=bd. Hence, system (2.1) can be reduced in the following form:

    {dIk(t)dt=λ(k)[bdIk(t)Rk(t)]Θ(t)(γ+d+α)Ik(t),dRk(t)dt=αIk(t)(d+δ)Rk(t). (4.1)

    The Jacobian matrix of the disease-free equilibrium E0 of system (4.1) is given by

    J=(A11A12A13A1nA21A22A23A2nAn1An2An3Ann),

    where

    Aii=(λ(i)φ(i)P(i)k(γ+d+α)0α(d+δ)),Aij=(λ(i)φ(j)P(j)k000).

    By induction, the characteristic equation can be shown as follows

    (x+γ+d+α)n1(x+d+δ)n[(x+γ+d+α)1knk=1λ(k)φ(k)P(k)]=0. (4.2)

    It's obvious that J has n1 negative eigenvalues that equal to (γ+d+α) and n negative eigenvalues that equal to (d+δ). The 2n-th eigenvalue completely depends on

    (x+γ+d+α)1knk=1λ(k)φ(k)P(k)=0,i.e.,x=(γ+d+α)(R01).

    When R0<1, then x<0; when R0>1, then x>0. Consequently, E0 is locally asymptotically stable if R0<1 and is unstable if R0>1. The proof is complete.

    Next, we proceed to study the global attractivity of E0. For this purpose, we need the following results.

    Lemma 4.1. [33] If a>0,b>0, and dx(t)dtbax, when t0 and x(0)0, then lim suptx(t)ba.

    We are now in a position to establish the following results about the global asymptotical stability of the disease-free equilibrium E0 by employing above lemma.

    Theorem 4.2. The disease-free equilibrium E0 of system (2.1) is globally asymptotically stable when R0<1.

    Proof. Equation (2.3) implies that

    dΘ(t)dt=Θ(t)[1kdbnk=1λ(k)φ(k)P(k)(bdIk(t)Rk(t))(γ+d+α)]Θ(t)[λ(k)φ(k)k(γ+d+α)1kdbnk=1λ(k)φ(k)P(k)Ik(t)]Θ(t)[(γ+d+α)(R01)μΘ(t)]=Θ(t)(rμΘ(t)), (4.3)

    where r=(γ+d+α)(R01), it follows from R01 that r0.

    (i) if r=0, then we have dΘ(t)dtμΘ2(t). Using following comparison equation

    {dx(t)dt=μx2(t),x(0)=Θ(0).

    Obviously, we can see that Θ(t)x(t)=Θ(0)1+μΘ(0)t, and hence limtΘ(t)=0.

    (ii) if r<0, consider the comparison equation

    {dy(t)dt=y(t)(rμy(t)),y(0)=Θ(0).

    By using the comparison principle, one can obtain Θ(t)y(t)=rcert1+cμert, where c=Θ(0)rμΘ(0). Therefore, limtΘ(t)=0.

    From (i) and (ii), we have

    limt1kdbnk=1φ(k)P(k)Ik(t)=0.

    Since φ(k)>0 and P(k)>0 for all k, then we have limtIk(t)=0.

    Next, we will prove limtRk(t)=0. Owing to limtIk(t)=0, then for any given sufficiently small constant ε>0, there exists t1>0 such that 0Ik(t)ε for t>t1. According to the third equation of system (2.1), we have

    dRk(t)dtαε(d+δ)Rk(t).

    By Lemma 2, we obtain that lim supt+Rk(t)α(d+δ)ε. Setting ε0, then lim supt+Rk(t)0, and hence limtRk(t)=0.

    Finally, due to Sk(t)=bdIk(t)Rk(t) we have limtSk(t)=bd. This implies that the disease-free equilibrium E0 of system (2.1) is globally attractive when R01. Note that E0 is locally asymptotically stable if R0<1 by Theorem 4.1. Therefore, we obtain that E0 is globally asymptotically stable if R0<1. This completes the proof.

    In the following, we will prove the persistence of the disease by employing Theorem 4.6 in [34].

    Theorem 4.3. The disease is permanent when R0>1, i.e., there exists a ζ>0, such that

    lim inftnk=1P(k)Ik(t)ζ>0.

    Proof. Define

    ˜xt=(S1(t),I1(t),R1(t),,Sn(t),In(t),Rn(t)),
    Γ={xt:Sk(t),Ik(t),Rk(t)0,Sk(t)+Ik(t)+Rk(t)=bd,k=1,2,,n},
    X0={xtΓ,nk=1P(k)Ik(t)>0},
    X0=ΓX0.

    It is obvious that Γ is positively invariant with respect to system (2.1). Note that Θ(0)>0, one can deduce that Ik(0)>0 for some k and nk=1P(k)Ik(0)>0 for all k. Owing to

    dnk=1P(k)Ik(t)dt=nk=1P(k)dIk(t)dt(γ+d+α)nk=1P(k)Ik(t),

    then we have

    nk=1P(k)Ik(0)e(γ+d+α)t>0.

    Thus, X0 is positively invariant as well. Furthermore, there exists a compact set B in which all solutions of (2.1) initiated in Γ will enter and remain forever after. The compactness condition (C4.2) for the set B in [34] is verified easily. Let

    M={˜x0:x(t,˜x0)X0,t0},
    Ω={ω(˜x0):˜x0M},

    where ˜x0=(S1(0),I1(0),R1(0),,Sn(0),In(0),Rn(0)),ω(˜x0) is the omega limit set of the solutions of system (2.1) starting at ˜x0Γ. Restricting system (2.1) on M, we obtain that

    {dSk(t)dt=bdSk(t)+δRk(t),dIk(t)dt=(γ+d+α)Ik(t),dRk(t)dt=(d+δ)Rk(t). (4.4)

    It is clear that system (4.4) has a unique equilibrium E0 in Γ, and E0 is also the unique equilibrium of system (2.1) in M. Then, we can easily prove that E0 is globally asymptotically stable. Therefore Ω={E0}. E0 is a covering of Ω, which is isolated and acyclic (there exists no nontrivial solution in M which links E0 to itself). Finally, to complete the proof, we only need to show that E0 is a weak repeller for X0, i.e.

    lim suptd(Ψ(t;˜x0),E0)>0, (4.5)

    where Ψ(t;˜x0) is an arbitrarily solution starting in ˜x0X0. In order to prove (4.5), we need to verify Ws(E0)X0= by the method in [35], where Ws(E0) is the stable manifold of E0. Assume it is not true, then there exists a solution ˜xtX0, such that

    limtSk(t)=bd,limtIk(t)=0,limtRk(t)=0. (4.6)

    Hence, for any ξ>0, there exists a T>0 such that for all t>T

    bdξ2<Sk(t)<bd+ξ2,0Ik(t)<ξ2,0Rk(t)<ξ2.

    Since R0=λ(k)φ(k)(γ+d+α)k>1, namely λ(k)φ(k)k(γ+d+α)>0, then we can choose a sufficiently small ξ>0, such that dbλ(k)φ(k)k(bdξ)(γ+d+α)>0.

    Let L(t)=Θ(t), obviously, L(t) is bounded and the derivative of L along the solution is

    dL(t)dt=1kdbnk=1φ(k)P(k)[λ(k)Sk(t)Θ(t)(γ+d+α)Ik(t)]=Θ(t)[1kdbnk=1λ(k)φ(k)P(k)Sk(t)(γ+d+α)]=Θ(t)[1kdbnk=1λ(k)φ(k)P(k)(bdIk(t)Rk(t))(γ+d+α)]>Θ(t)[1kdbnk=1λ(k)φ(k)P(k)(bdξ)(γ+d+α)].

    Due to

    σ=dbλ(k)φ(k)k(bdξ)(γ+d+α)>0,

    then we have Θ(t)>Θ(0)eσt. Hence L(t)=Θ(t)+ as t+, which is inconsistent with the boundedness of L(t). This proof is complete.

    Finally, we will further show that the endemic equilibrium E is globally asymptotically stable by constructing a suitable Lyapunov function.

    Theorem 4.4. If R0>1 and δ>γ, then the endemic equilibrium E of system (2.1) is globally asymptotically stable.

    Proof. system (2.1) can be written as

    {dSk(t)dt=bΛλ(k)Sk(t)Θ(t)(d+γ)Sk(t)+(δγ)Rk(t),dIk(t)dt=λ(k)Sk(t)Θ(t)(γ+d+α)Ik(t),dRk(t)dt=αIk(t)(d+δ)Rk(t), (4.7)

    where Λ=(1+γd), and the components of endemic equilibrium E satisfy the following equations

    {bΛλ(k)SkΘ(d+γ)Sk(t)+(δγ)Rk(t)=0,λ(k)SkΘ(γ+d+α)Ik=0,αIk(d+δ)Rk=0. (4.8)

    For the second equation of (4.8), multiplying by φ(k)P(k) and summing both sides with respect to k, we have

    γ+d+α=1kdbnk=1λ(k)φ(k)P(k)Sk. (4.9)

    Submitting Ik(t)=bdSkRk into the third equation of (4.8), then we obtain

    αbd=αSk+(α+d+δ)Rk. (4.10)

    Now we investigate a non-negative solution (S1(t),I1(t),R1(t),,Sn(t),In(t),Rn(t)) of system (2.1). Motivated by Li et al. [25], we consider the following Lyapunov function

    V(t)=12nk=1{ρ1(k)(Sk(t)Sk)2+ρ2(k)(Rk(t)Rk)2}+(Θ(t)ΘΘlnΘ(t)Θ),

    where ρ1(k)=dφ(k)P(k)bkSk>0 and ρ2(k)=δγαρ1(k)>0. Clearly, V(t)>0 except at E, and taking the derivative of V(t) along the positive solution of (4.7), it follows that

    dV(t)dt=nk=1{ρ1(k)(Sk(t)Sk)dSk(t)dt+ρ2(k)(Rk(t)Rk)dRk(t)dt}+Θ(t)ΘΘ(t)dΘ(t)dt=nk=1{ρ1(k)(Sk(t)Sk)[bΛλ(k)Sk(t)Θ(t)(d+γ)Sk(t)+(δγ)Rk(t)]+ρ2(k)(Rk(t)Rk)[αbdαSk(t)(α+d+δ)Rk(t)]}+Θ(t)ΘΘ(t)(1kdbnk=1φ(k)P(k)[λ(k)Sk(t)Θ(t)(γ+d+α)Ik(t)]=nk=1[ρ1(k)(λ(k)Θ+d+γ)(Sk(t)Sk)2ρ1(k)λ(k)Sk(Θ(t)Θ)(Sk(t)Sk)+(δγ)ρ1(k)(Sk(t)Sk)(Rk(t)Rk)αρ2(k)(Sk(t)Sk)(Rk(t)Rk)ρ2(k)(α+d+δ)(Rk(t)Rk)2]+1kdbnk=1λ(k)φ(k)P(k)(SkSk)(Θ(t)Θ)=nk=1[ρ1(k)(λ(k)Θ(t)+d+γ)(Sk(t)Sk)2+ρ2(k)(α+d+δ)(Rk(t)Rk)2]0.

    Therefore, dV(t)dt=0 if and only if Sk=Sk,Ik=Ik,Rk=Rk for k=1,2,,n. It means that the largest compact invariant subset contained in the set {dV(t)dt=0} is the singleton {E}. From the Lasalle Invariance Principle [36], one can obtain that the endemic equilibrium E is globally asymptotically stable if R0>1 and δ>γ. This completes the proof.

    From Section 3 and Section 4, we know the basic reproduction number R0 is a key parameter which determines the global dynamics of the epidemic model, and if the basic reproduction number R0 can be reduced to a value less than one, the epidemic disease will die out. Moreover, immunization is one of effective ways in controlling preventable diseases [37,38,39,40] by reducing the basic reproduction number. However, it is always impossible to vaccinate all individuals in reality because of the high costs and effort. Therefore, in this section we will discuss the impacts of three immunization strategies (i.e., the random immunization, the targeted immunization and the acquaintance immunization) about system (2.1) on the heterogenous networks.

    Random immunization is the simplest immunization strategy which vaccinates a fraction of the population without selectivity. Let p (0<p<1) be the immunization rate, then system (2.1) can be written as

    {dSk(t)dt=bλ(k)(1p)Sk(t)Θ(t)dSk(t)+γIk(t)+δRk(t),dIk(t)dt=λ(k)(1p)Sk(t)Θ(t)(γ+d+α)Ik(t),dRk(t)dt=αIk(t)(d+δ)Rk(t). (5.1)

    Through similar derivations in Section 3, we obtain the basic reproduction number Rr0 for system (5.1)

    Rr0=(1p)λ(k)φ(k)(γ+d+α)k=(1p)R0.

    Then, similar results are obtained.

    Theorem 5.1. Assume R0>1. Define pc=11R0, then

    (i) If p>pc i.e. Rr0<1, then the disease free equilibrium of system (5.1) is globally asymptotically stable. That is to say, the disease will die out ultimately.

    (ii) If 0<p<pc (1<Rr0<R0), then there always exists a unique endemic equilibrium Er of system (5.1) which is globally asymptotically stable when δ>γ. That indicates that the random immunization is effective, but cannot eliminate the disease in the network.

    There is a fact that random immunization strategy ignores the heterogeneous connectivity of the network. In order to take this factor into account, the targeted immunization strategy was proposed [37,38]. This immunization strategy is to immunize the most highly connected nodes because they are vulnerable to selective attacks in the network. Assuming that all nodes with kkt will be immunized, where kt is the upper threshold. Then the immunization rate pk can be denoted as

    pk={1,0,kkt,k<kt.

    Then, system (2.1) becomes

    {dSk(t)dt=bλ(k)(1pk)Sk(t)Θ(t)dSk(t)+γIk(t)+δRk(t),dIk(t)dt=λ(k)(1pk)Sk(t)Θ(t)(γ+d+α)Ik(t),dRk(t)dt=αIk(t)(d+δ)Rk(t). (5.2)

    By similar derivations in Section 3, we obtain the basic reproduction number Rt0 for system (5.2)

    Rt0=1(γ+d+α)λ(k)φ(k)pkλ(k)φ(k)k=R0pkλ(k)φ(k)(γ+d+α)k. (5.3)

    Let ˉp=nk=1pkP(k) be the average immunization rate, then

    pkλ(k)φ(k)=pkλ(k)φ(k)+Cov(pk,λ(k)φ(k))=ˉpλ(k)φ(k)+(pkˉp)(λ(k)φ(k)λ(k)φ(k)),

    where Cov(,) represents the covariance of two variables. For appropriately small kt, pkˉp and λ(k)φ(k)λ(k)φ(k) have the same signs, except that k's where pk=ˉp or λ(k)φ(k)=λ(k)φ(k), then Cov(pk,λ(k)φ(k))>0. That is to say for some kt, pkλ(k)φ(k)>pkλ(k)φ(k). Hence, we have following conclusions.

    Theorem 5.2. Consider system (5.2), then we have following results:

    (i) If Rt0<1<R0, then the disease free equilibrium of system (5.2) is globally asymptotically stable, i.e. the disease can be controlled by the targeted immunization strategy.

    (ii) If 1<Rt0<R0, there will always exists a unique endemic equilibrium Et of system (5.2), which is globally asymptotically stable when δ>γ.

    (iii) For some kt, pkλ(k)φ(k)>pkλ(k)φ(k), then Rt0<Rr0. This shows that the targeted immunization is more effective than random immunization.

    Since the targeted immunization requires us to know the global connectivity information of each node in networks, and it is always much difficult to carry out in reality. In order to deal with this dilemma, R. Cohen, S. Havlin and D. Ben-Avraham [37] proposed the acquaintance immunization scheme for the immunization of random acquaintances of random nodes. Specifically, choose a fraction q of N nodes randomly, the probability that a particular node with degree k is selected for immunization is kP(k)/(Nk) [37,41]. Hence, substituting pk=pNkP(k)/(Nk)=kP(k)(p/k) to (5.3), then the basic reproduction number for acquaintance immunization is

    Ra0=1(γ+d+α)λ(k)φ(k)kP(k)(p/k)λ(k)φ(k)k=R0pkP(k)λ(k)φ(k)(γ+d+α)k2.

    This implies that acquaintance immunization scheme is effective as well.

    Theorem 5.3. For system (5.2), we have following results:

    (i) If Ra0<1<R0, then the disease can be controlled by this immunization scheme, i.e., the disease will be eradicated from the network ultimately.

    (ii) If 1<Ra0<R0, then the disease is uniformly persistent in the network and there always exists a unique endemic equilibrium Ea.

    In this section, we present several numerical simulations to demonstrate the correctness and effectiveness of the main results. These simulations are based on scale-free networks. We first start with a graph with 5 vertexes, a node with one edge is added into the graph in each time step, and then a BA scale-free network G with 100 nodes is obtained by using preferential attachment mechanism in [42]. There will give several numerical examples with different infectivities φ(k) to illustrate the theoretical analysis.

    In Figures 2, 3 and 4, we adopt the infected rate λ(k)=βk and give infectivity φ(k) three types such as linear, constant and nonlinear infectivity to verify the stability of the disease-free equilibrium.

    Figure 2.  The time evolution of the number of each states for Case ⅰ, where n=11,λ(k)=0.01k,φ(k)=k, b=2,γ=0.04,d=0.025,α=0.04 and δ=0.03.
    Figure 3.  The time evolution of the densities of each states for Case ⅱ, where the parameter n=13,φ(k)=4 and other parameters are the same as in Case ⅰ.
    Figure 4.  The time evolution of the number of each states for Case ⅲ, where the parameter n=13,φ(k)=k0.5 and other parameters are the same as in Case ⅰ.

    Case ⅰ. Let n=11,φ(k)=k, b=2,γ=0.04,β=0.01,d=0.025,α=0.04,δ=0.03 and the initial values: Sk(0)=40,Ik(0)=30,Rk(0)=10. Then we have R0=0.8884<1. From our results, the disease will die out if R0=0.8884<1, and the time evolution of the densities of each state for the initial value is drawn in Figure 2. It agrees with Theorem 4.2.

    Case ⅱ. Choose n=13,φ(k)=4 and other parameters are the same as in Case ⅰ. Then one can obtain R0=0.2857<1. It follows from Theorem 4.2, disease-free equilibrium E0 is globally asymptotically stable. The Figure 3 shows the time evolution of the number of each state for the initial value.

    Case ⅲ. We choose n=13, the nonlinear infectivity φ(k)=k0.5 and other parameters are the same as in Case ⅰ as well. Then R0=0.2369<1, and by Theorem 4.2 the disease-free equilibrium E0 is globally asymptotically stable. The time evolution of the densities of each state for the initial value is shown in Figure 3.

    Next, in the same network architecture in Subsection 6.1, we will verify the global asymptotical stability of the endemic equilibrium E. Similarly, we choose λ(k)=βk and give three types of the infectivity φ(k).

    Case ⅳ. We make n=13,φ(k)=k, b=2,γ=0.02,β=0.06,d=0.025,α=0.04,δ=0.03 and the initial value: Sk(0)=40,Ik(0)=30,Rk(0)=10. We obtain R0=5.9765>1, which implies that E is global asymptotical stable from Theorem 4.4. The time evolution of the densities of each state for the initial value is depicted in Figure 5.

    Figure 5.  The time evolution of the number of each states for Case ⅳ, where n=13,λ(k)=0.06k,φ(k)=k, b=2,γ=0.02,d=0.025,α=0.04 and δ=0.03.

    Case ⅴ. Setting n=15,φ(k)=4 and other parameters are the same as in Case ⅳ. Then one can verify R0=2.1176>1. Hence it follows from Theorem 4.4 that the endemic equilibrium E is globally asymptotically stable. The Figure 6 shows the time evolution of the densities of each state for the initial values.

    Figure 6.  The time evolution of the number of each states for Case ⅴ, where n=15,φ(k)=4 and other parameters are the same as in Case ⅳ.

    Case ⅵ. We choose n=13, the nonlinear infectivity φ(k)=k0.5 and other parameters are the same as in Case ⅳ as well. Then we obtain R0=1.7420>1, and the time evolution of the densities of each state for the initial value is presented in Figure 7. It agrees with Theorem 4.4.

    Figure 7.  The time evolution of the number of each states for Case ⅵ. The parameter n=13,φ(k)=k0.5 and other parameters are the same as in Case ⅳ.

    As shown in the Figures 5, 6 and 7, it follows from Theorem 4.4 that the endemic equilibrium E is globally asymptotically stable when R0>1 and δ>γ. In addition, if we choose n=14,λ(k)=βk, φ(k)=k, b=2,γ=0.04,β=0.06,d=0.025,α=0.04 and δ=0.03, then we can obtain that R0=4.1789>1 and δγ hold. The time evolution of the densities of each state is shown in Figure 8. It seems that the endemic equilibrium E is also globally attractive. However, we could not give a rigorous proof for the global attractivity of E when R0>1 and δγ, which we leave as our future work.

    Figure 8.  The time evolution of the number of each states for the case of R0>1 and δγ.

    By incorporating demographics and transfer from infectious to susceptible individuals into network disease models, we propose an improved SIRS model on complex networks in this paper. We derive the basic reproduction number R0 by using the next generation matrix method and obtain that the basic reproduction number R0 determines not only the existence of the endemic equilibrium E but also the global stability of equilibria E0 and E, i.e., the disease-free equilibrium E0 is globally asymptotically stable when R0<1; when R0>1, there exists a unique endemic equilibrium E and the disease is permanent on the network, which indicates that reducing the basic reproduction number R0 below one could eradicate the disease. Meanwhile, by employing the method of Lyapunov function, we prove that if R0>1 and δ>γ, the endemic equilibrium E is globally asymptotically stable. Compared with the model in literature [27], our model is more general by considering the degree-related infection rate λ(k) and the general form of infectivity φ(k). At the same time, the global dynamics of the disease-free equilibrium and the endemic equilibrium under different forms of infectivity are discussed by some numerical simulations. It also confirms the correctness of the theoretical analysis. In addition, from our numerical simulations, the endemic equilibrium E is globally asymptotically stable when R0>1 with δ<γ which is not proved in the present paper and will be investigated in further exploration.

    Some new insights in this paper may help us to understand the spreading of diseases and provide some effective intervening measures to prevent the epidemic outbreaks. As one can see, the basic reproduction number R0 is an important index in disease control, which is closely related to the topology of the networks and some model parameters. One can obtain that increasing the transfer rate γ and the recovery rate α will decrease R0 to restrict the spreading of the epidemic diseases. Moreover, it is worth noting that R0 is not dependent on the immunity loss rate δ, which indicates that strengthening the treatment is very important. According to the investigation on three immunization strategies about system (2.1) on the heterogenous networks, we can find that immunizations are also effective ways in controlling diseases transmission by decreasing R0. Furthermore, one can see that the targeted immunization is more effective than random immunization, and the acquaintance immunization is more economical and more easier to be carried out.

    The authors are grateful to the anonymous referee for the valuable comments. The work is partially supported by NNSF of China (Nos. 11771059, 11401051, 71471020), Hunan Provincial Natural Science Foundation (Nos. 2015JJ3013, 2016JJ1001) and the Scientific Research Fund of Hunan Provincial Education Department of China (No. 18B152).

    All authors declare no conflicts of interest in this paper.



    [1] X. Zhang, D. Jiang, A. Alsaedi, et al., Stationary distribution of stochastic SIS epidemic model with vaccination under regime switching, Appl. Math. Lett., 59 (2016), 87–93.
    [2] P. Driessche and J. Watmough, A simple SIS epidemic model with a backward bifurcation, J. Math. Biol., 40 (2000), 525–540.
    [3] W. Guo, Y. Cai, Q. Zhang, et al., Stochastic persistence and stationary distribution in an SIS epidemic model with media coverage, Physica A, 492 (2018), 2220–2236.
    [4] J. Pan, A. Gray, D. Greenhalgh, et al., Parameter estimation for the stochastic SIS epidemic model, J. Stat. Inference Stoch. Process, 17 (2014), 75–98.
    [5] Y. Cai, Y. Kang and W. Wang, A stochastic SIRS epidemic model with nonlinear incidence rate, Appl. Math. Comput., 305 (2017), 221–240.
    [6] S. Busenberg, M. Iannelli and H. Thieme, Global behavior of an age-structured epidemic model, Siam J. Math. Anal., 22 (1991), 1065–1080.
    [7] B. Cao, H. Huo and H. Xiang, Global stability of an age-structure epidemic model with imperfect vaccination and relapse, Physica A, 486 (2017), 638–655.
    [8] H. Inaba, Age-structured population dynamics in demography and epidemiology, Springer, Sin- gapore, 2017.
    [9] T.Kuniya, Globalstabilityanalysiswithadiscretizationapproachforanage-structuredmultigroup SIR epidemic model, Nonlinear Anal. Real World Appl., 12 (2011), 2640–2655.
    [10] K. Toshikazu, Numerical approximation of the basic reproduction number for a class of age- structured epidemic models, Appl. Math. Lett., 73 (2017), 106–112.
    [11] H. Inaba, Threshold and stability results for an age-structured epidemic model, J. Math. Biol., 28 (1990), 411–434.
    [12] O. Diekmann, J. Heesterbeek and J. Metz, On the definition and the computation of the basic reproduction ratio R 0 , in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28 (1990), 365–382.
    [13] T. Kuniya and R. Oizumi, Existence result for an age-structured SIS epidemic model with spatial diffusion, Nonlinear Anal. Real World Appl., 23 (2015), 196–208.
    [14] N. Bacaër, Approximation of the basic reproduction number R 0 for Vector-Borne diseases with a periodic vector population, B. Math. Biol., 69 (2007), 1067–1091.
    [15] Z. Xu, F. Wu and C. Huang, Theta schemes for SDDEs with non-globally Lipschitz continuous coefficients, J. Comput. Appl. Math., 278 (2015), 258–277.
    [16] X. Mao and L. Szpruch, Strong convergence and stability of implicit numerical methods for stochastic differential equations with non-globally Lipschitz continuous coefficients, J. Comput. Appl. Math., 238 (2013), 14–28.
    [17] F. Chatelin, The spectral approximation of linear operators with applications to the computation of eigenelements of differential and integral operators, Siam Rev., 23 (1981), 495–522.
    [18] A. Berman and R. Plemmons, Nonnegative matrices in the mathematical sciences, Academic press, New York, 1979.
    [19] K. Ito and F. Kappel, The Trotter-Kato theorem and approximation of PDEs, Math. Comput., 67 (1998), 21–44.
    [20] B. Pagter, Irreducible compact operators, Math. Z., 192 (1986), 149–153.
    [21] M. Krein, Linear operators leaving invariant acone in a Banach space, Amer. Math. Soc. Transl.,10 (1962), 3–95.
    [22] W. Guo, Q. Zhang, X. Li, et al., Dynamic behavior of a stochastic SIRS epidemic model with media coverage, Math. Method. Appl. Sci., 41 (2018), 5506-5525.
    [23] C. Mills, J. Robins and M. Lipsitch, Transmissibility of 1918 pandemic influenza, Nature, 432 (2004), 904–906.
  • This article has been cited by:

    1. Shuyun Jiao, Mingzhan Huang, An SIHR epidemic model of the COVID-19 with general population-size dependent contact rate, 2020, 5, 2473-6988, 6714, 10.3934/math.2020431
    2. Chaofan Qian, Yuhui Hu, Novel stability criteria on nonlinear density-dependent mortality Nicholson’s blowflies systems in asymptotically almost periodic environments, 2020, 2020, 1029-242X, 10.1186/s13660-019-2275-4
    3. Qian Cao, Xiaojin Guo, Anti-periodic dynamics on high-order inertial Hopfield neural networks involving time-varying delays, 2020, 5, 2473-6988, 5402, 10.3934/math.2020347
    4. Manickam Iswarya, Ramachandran Raja, Grienggrai Rajchakit, Jinde Cao, Jehad Alzabut, Chuangxia Huang, Existence, Uniqueness and Exponential Stability of Periodic Solution for Discrete-Time Delayed BAM Neural Networks Based on Coincidence Degree Theory and Graph Theoretic Method, 2019, 7, 2227-7390, 1055, 10.3390/math7111055
    5. Yadan Zhang, Minghui Jiang, Xue Fang, A New Fixed-Time Stability Criterion and Its Application to Synchronization Control of Memristor-Based Fuzzy Inertial Neural Networks with Proportional Delay, 2020, 52, 1370-4621, 1291, 10.1007/s11063-020-10305-9
    6. Qian Cao, Guoqiu Wang, Chaofan Qian, New results on global exponential stability for a periodic Nicholson’s blowflies model involving time-varying delays, 2020, 2020, 1687-1847, 10.1186/s13662-020-2495-4
    7. Hong Zhang, Qian Cao, Hedi Yang, Asymptotically almost periodic dynamics on delayed Nicholson-type system involving patch structure, 2020, 2020, 1029-242X, 10.1186/s13660-020-02366-0
    8. Anbalagan Pratap, Ramachandran Raja, Jehad Alzabut, Jinde Cao, Grienggrai Rajchakit, Chuangxia Huang, Mittag‐Leffler stability and adaptive impulsive synchronization of fractional order neural networks in quaternion field, 2020, 43, 0170-4214, 6223, 10.1002/mma.6367
    9. Chaofan Qian, New periodic stability for a class of Nicholson's blowflies models with multiple different delays, 2020, 0020-7179, 1, 10.1080/00207179.2020.1766118
    10. Umesh Kumar, Subir Das, Chuangxia Huang, Jinde Cao, Fixed-time synchronization of quaternion-valued neural networks with time-varying delay, 2020, 476, 1364-5021, 20200324, 10.1098/rspa.2020.0324
    11. Chuangxia Huang, Luanshan Yang, Jinde Cao, Asymptotic behavior for a class of population dynamics, 2020, 5, 2473-6988, 3378, 10.3934/math.2020218
    12. M. Syed Ali, G. Narayanan, Sumit Saroha, Bandana Priya, Ganesh Kumar Thakur, Finite-time stability analysis of fractional-order memristive fuzzy cellular neural networks with time delay and leakage term, 2021, 185, 03784754, 468, 10.1016/j.matcom.2020.12.035
    13. Sudesh Kumari, Renu Chugh, Jinde Cao, Chuangxia Huang, Multi Fractals of Generalized Multivalued Iterated Function Systems in b-Metric Spaces with Applications, 2019, 7, 2227-7390, 967, 10.3390/math7100967
    14. M. Iswarya, R. Raja, G. Rajchakit, J. Cao, J. Alzabut, C. Huang, A perspective on graph theory-based stability analysis of impulsive stochastic recurrent neural networks with time-varying delays, 2019, 2019, 1687-1847, 10.1186/s13662-019-2443-3
    15. Xin Long, Novel stability criteria on a patch structure Nicholson’s blowflies model with multiple pairs of time-varying delays, 2020, 5, 2473-6988, 7387, 10.3934/math.2020473
    16. Lihong Huang, Huili Ma, Jiafu Wang, Chuangxia Huang, GLOBAL DYNAMICS OF A FILIPPOV PLANT DISEASE MODEL WITH AN ECONOMIC THRESHOLD OF INFECTED-SUSCEPTIBLE RATIO, 2020, 10, 2156-907X, 2263, 10.11948/20190409
    17. Xin Yang, Shigang Wen, Xian Zhao, Chuangxia Huang, Systemic importance of financial institutions: A complex network perspective, 2020, 545, 03784371, 123448, 10.1016/j.physa.2019.123448
    18. Wentao Wang, Wei Chen, Persistence and extinction of Markov switched stochastic Nicholson’s blowflies delayed differential equation, 2020, 13, 1793-5245, 2050015, 10.1142/S1793524520500151
    19. Xin Yang, Shigang Wen, Zhifeng Liu, Cai Li, Chuangxia Huang, Dynamic Properties of Foreign Exchange Complex Network, 2019, 7, 2227-7390, 832, 10.3390/math7090832
    20. Qian Cao, Guoqiu Wang, Dynamic analysis on a delayed nonlinear density-dependent mortality Nicholson's blowflies model, 2020, 0020-7179, 1, 10.1080/00207179.2020.1725134
    21. Hong Zhang, Chaofan Qian, Convergence analysis on inertial proportional delayed neural networks, 2020, 2020, 1687-1847, 10.1186/s13662-020-02737-3
    22. Ajendra singh, Jitendra Nath Rai, Stability of Fractional Order Fuzzy Cellular Neural Networks with Distributed Delays via Hybrid Feedback Controllers, 2021, 1370-4621, 10.1007/s11063-021-10460-7
    23. Qian Cao, Xin Long, New convergence on inertial neural networks with time-varying delays and continuously distributed delays, 2020, 5, 2473-6988, 5955, 10.3934/math.2020381
    24. Chuangxia Huang, Xiaoguang Yang, Jinde Cao, Stability analysis of Nicholson’s blowflies equation with two different delays, 2020, 171, 03784754, 201, 10.1016/j.matcom.2019.09.023
    25. Jian Zhang, Chuangxia Huang, Dynamics analysis on a class of delayed neural networks involving inertial terms, 2020, 2020, 1687-1847, 10.1186/s13662-020-02566-4
    26. Qian Wang, Hui Wei, Zhiwen Long, A non-reduced order approach to stability analysis of delayed inertial genetic regulatory networks, 2021, 33, 0952-813X, 227, 10.1080/0952813X.2020.1735531
    27. Jiafu Wang, Su He, Lihong Huang, Limit Cycles Induced by Threshold Nonlinearity in Planar Piecewise Linear Systems of Node-Focus or Node-Center Type, 2020, 30, 0218-1274, 2050160, 10.1142/S0218127420501606
    28. Gang Yang, Qian Cao, Stability for patch structure Nicholson’s blowflies systems involving distinctive maturation and feedback delays, 2020, 0952-813X, 1, 10.1080/0952813X.2020.1836032
    29. Chuangxia Huang, Xin Long, Jinde Cao, Stability of antiperiodic recurrent neural networks with multiproportional delays, 2020, 43, 0170-4214, 6093, 10.1002/mma.6350
    30. Qian Cao, Guoqiu Wang, New findings on global exponential stability of inertial neural networks with both time-varying and distributed delays, 2021, 0952-813X, 1, 10.1080/0952813X.2021.1883744
    31. Ruihan Chen, Song Zhu, Yongqiang Qi, Yuxin Hou, Reachable set bounding for neural networks with mixed delays: Reciprocally convex approach, 2020, 125, 08936080, 165, 10.1016/j.neunet.2020.02.005
    32. Shigang Wen, Yu Tan, Mengge Li, Yunke Deng, Chuangxia Huang, Analysis of Global Remittance Based on Complex Networks, 2020, 8, 2296-424X, 10.3389/fphy.2020.00085
    33. Zhenhua Duan, Changjin Xu, Anti-periodic behavior for quaternion-valued delayed cellular neural networks, 2021, 2021, 1687-1847, 10.1186/s13662-021-03327-7
    34. Yi Wang, Jinde Cao, Gang Huang, Further dynamic analysis for a network sexually transmitted disease model with birth and death, 2019, 363, 00963003, 124635, 10.1016/j.amc.2019.124635
    35. Luogen Yao, Qian Cao, Anti-periodicity on high-order inertial Hopfield neural networks involving mixed delays, 2020, 2020, 1029-242X, 10.1186/s13660-020-02444-3
    36. Anbalagan Pratap, Ramachandran Raja, Jinde Cao, Chuangxia Huang, Michal Niezabitowski, Ovidiu Bagdasar, Stability of discrete‐time fractional‐order time‐delayed neural networks in complex field, 2021, 44, 0170-4214, 419, 10.1002/mma.6745
    37. Luogen Yao, Global exponential stability on anti-periodic solutions in proportional delayed HIHNNs, 2021, 33, 0952-813X, 47, 10.1080/0952813X.2020.1721571
    38. Yanli Xu, Qian Cao, Xiaojin Guo, Stability on a patch structure Nicholson’s blowflies system involving distinctive delays, 2020, 105, 08939659, 106340, 10.1016/j.aml.2020.106340
    39. Yanli Xu, Qian Cao, Global asymptotic stability for a nonlinear density-dependent mortality Nicholson’s blowflies system involving multiple pairs of time-varying delays, 2020, 2020, 1687-1847, 10.1186/s13662-020-02569-1
    40. Sudesh Kumari, Renu Chugh, Jinde Cao, Chuangxia Huang, On the construction, properties and Hausdorff dimension of random Cantor one pth set, 2020, 5, 2473-6988, 3138, 10.3934/math.2020202
    41. A. Pratap, R. Raja, Jinde Cao, J. Alzabut, Chuangxia Huang, Finite-time synchronization criterion of graph theory perspective fractional-order coupled discontinuous neural networks, 2020, 2020, 1687-1847, 10.1186/s13662-020-02551-x
    42. Qian Cao, Guoqiu Wang, Hong Zhang, Shuhua Gong, New results on global asymptotic stability for a nonlinear density-dependent mortality Nicholson’s blowflies model with multiple pairs of time-varying delays, 2020, 2020, 1029-242X, 10.1186/s13660-019-2277-2
    43. Rundong Zhao, Qiming Liu, Meici Sun, Dynamical behavior of a stochastic SIQS epidemic model on scale-free networks, 2021, 1598-5865, 10.1007/s12190-021-01550-9
    44. Xianhui Zhang, Convergence analysis of a patch structure Nicholson’s blowflies system involving an oscillating death rate, 2021, 0952-813X, 1, 10.1080/0952813X.2021.1908433
    45. Roberto Galizia, Petri T. Piiroinen, Regions of Reduced Dynamics in Dynamic Networks, 2021, 31, 0218-1274, 2150080, 10.1142/S0218127421500802
    46. Jian Zhang, Ancheng Chang, Gang Yang, Periodicity on Neutral-Type Inertial Neural Networks Incorporating Multiple Delays, 2021, 13, 2073-8994, 2231, 10.3390/sym13112231
    47. Jiaying Zhou, Yi Zhao, Yong Ye, Yixin Bao, Bifurcation Analysis of a Fractional-Order Simplicial SIRS System Induced by Double Delays, 2022, 32, 0218-1274, 10.1142/S0218127422500687
    48. Shuping Li, Xiaorong Zhao, Ruixia Zhang, Site-bond percolation model of epidemic spreading with vaccination in complex networks, 2022, 85, 0303-6812, 10.1007/s00285-022-01816-1
    49. Lian Duan, Lihong Huang, Chuangxia Huang, Spatial dynamics of a diffusive SIRI model with distinct dispersal rates and heterogeneous environment, 2021, 20, 1534-0392, 3539, 10.3934/cpaa.2021120
    50. Jie Li, Jiu Zhong, Yong-Mao Ji, Fang Yang, A new SEIAR model on small-world networks to assess the intervention measures in the COVID-19 pandemics, 2021, 25, 22113797, 104283, 10.1016/j.rinp.2021.104283
    51. Chaouki Aouiti, Mahjouba Ben Rezeg, Impulsive multidirectional associative memory neural networks: New results, 2021, 14, 1793-5245, 10.1142/S1793524521500601
    52. Hong Zhang, Qian Cao, Hedi Yang, Dynamics analysis of delayed Nicholson-type systems involving patch structure and asymptotically almost periodic environments, 2022, 34, 0952-813X, 725, 10.1080/0952813X.2021.1924869
    53. Reinhard Schlickeiser, Martin Kröger, Determination of a Key Pandemic Parameter of the SIR-Epidemic Model from Past COVID-19 Mutant Waves and Its Variation for the Validity of the Gaussian Evolution, 2023, 5, 2624-8174, 205, 10.3390/physics5010016
    54. Xiaojin Guo, Chuangxia Huang, Jinde Cao, Nonnegative periodicity on high-order proportional delayed cellular neural networks involving D operator, 2020, 6, 2473-6988, 2228, 10.3934/math.2021135
    55. Qian Cao, Attractivity analysis on a neoclassical growth system incorporating patch structure and multiple pairs of time-varying delays, 2021, 14173875, 1, 10.14232/ejqtde.2021.1.76
    56. Reinhard Schlickeiser, Martin Kröger, Key Epidemic Parameters of the SIRV Model Determined from Past COVID-19 Mutant Waves, 2023, 3, 2673-8112, 592, 10.3390/covid3040042
    57. Shixiang Han, Guanghui Yan, Huayan Pei, Wenwen Chang, Dynamical Analysis of an Improved Bidirectional Immunization SIR Model in Complex Network, 2024, 26, 1099-4300, 227, 10.3390/e26030227
    58. Guangyu Li, Haifeng Du, Jiarui Fan, Xiaochen He, Wenhua Wang, The Effect of Fangcang Shelter Hospitals under Resource Constraints on the Spread of Epidemics, 2023, 20, 1660-4601, 5802, 10.3390/ijerph20105802
    59. Chuangxia Huang, Jiafu Wang, Lihong Huang, Asymptotically almost periodicity of delayed Nicholson-type system involving patch structure, 2020, 2020, 1072-6691, 61, 10.58997/ejde.2020.61
    60. 德玉 郭, Construction and Dynamic Analysis of a Class of Hepatitis C Model with Population Heterogeneity, 2023, 12, 2324-7991, 4665, 10.12677/AAM.2023.1211458
    61. Guojin Wang, Wei Yao, An application of small-world network on predicting the behavior of infectious disease on campus, 2024, 9, 24680427, 177, 10.1016/j.idm.2023.12.007
    62. Bingjie Wu, Liang’an Huo, Studying the impact of individual emotional states on the co-evolution of information, behavior and disease in multiplex networks, 2025, 03784371, 130480, 10.1016/j.physa.2025.130480
    63. Samuel Lopez, Natalia L. Komarova, An optimal network that promotes the spread of an advantageous variant in an SIR epidemic, 2025, 00225193, 112095, 10.1016/j.jtbi.2025.112095
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(4620) PDF downloads(527) Cited by(7)

Figures and Tables

Figures(5)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog