1,655
Views
7
CrossRef citations to date
0
Altmetric
Articles

Asymptotic analysis of a vector-borne disease model with the age of infection

, , &
Pages 332-367 | Received 22 Feb 2019, Accepted 09 Mar 2020, Published online: 23 Apr 2020

Abstract

Vector-borne infectious diseases may involve both horizontal transmission between hosts and transmission from infected vectors to susceptible hosts. In this paper, we incorporate these two transmission modes into a vector-borne disease model that includes general nonlinear incidence rates and the age of infection for both hosts and vectors. We show the existence, uniqueness, nonnegativity, and boundedness of solutions for the model. We study the existence and local stability of steady states, which is shown to be determined by the basic reproduction number. By showing the existence of a global compact attractor and uniform persistence of the system, we establish the threshold dynamics using the Fluctuation Lemma and the approach of Lyapunov functionals. When the basic reproduction number is less than one, the disease-free steady state is globally asymptotically stable and otherwise the disease will be established when there is initial infection force for the hosts. We also study a model with the standard incidence rate and discuss the effect of different incidence rates on the disease dynamics.

1. Introduction

Vectors are living organisms that can transmit infectious diseases between humans (horizontal transmission) or from animals to humans. Many of these vectors are bloodsucking insects, which ingest disease-producing microorganisms during a blood meal from an infected host (human or animal) and later inject these microorganisms into a new host during their subsequent blood meal [Citation59]. Mosquitoes are the best known disease vector. Other vectors include ticks, flies, sandflies, fleas, and some freshwater aquatic snails. Vector-borne diseases are infections caused by pathogens and parasites in human populations which are transmitted by the bite of infected vectors. They have been a substantial threat to human health [Citation9]. According to the WHO [Citation59], vector-borne diseases account for a significant proportion of infectious diseases that affect humans. Every year there are more than one billion cases and over one million deaths from vector-borne diseases such as malaria, West Nile virus, leishmaniasis, dengue (the arboviral human disease), human African trypanosomiasis, Chagas disease, and so on. Therefore, it is imperative to understand the transmission dynamics of vector-borne diseases and, based on these results, design appropriate control and prevention measures.

Mathematical modeling is an efficient way to investigate infectious diseases dynamics. Over the past few decades, mathematical models have made significant contribution to the understanding of the transmission dynamics of vector-borne diseases. Many models are described by ordinary differential equations (see [Citation2,Citation14,Citation16,Citation18,Citation23,Citation31,Citation39,Citation43,Citation46,Citation48,Citation49,Citation52,Citation53] and references therein). A fundamental assumption in these models is that individuals in each epidemiological class are homogeneous. For example, infectious individuals have the same infectivity during their infectious period. Although it may provide a reasonable approximation to the biological process that is modeled, this is not always the case. To describe such phenomena as variable infectivity in the infectious period, continuous age structures have been introduced into mathematical models described by ordinary differential equations.

There is a rich literature on SIR (susceptible-infectious-recovered) epidemic models and their extensions with continuous age structures. Many diseases can simply spread because of the existence of biting vectors, which is important in the transmission dynamics of vector-borne diseases. The age at which a vector becomes infectious can affect the number of secondary infectious individuals. When infection occurs at the beginning of the vector's life, it makes a higher number of bites. From the point of biological modeling, Novoseltsev et al. [Citation44] provided an extension and refinement of the vectorial capacity paradigm by introducing an age-structured extension to the model. They assumed that insect vectors may have age-dependent mortality in the model, which was shown to greatly influence the pathogen transmission dynamics. Bellan [Citation3] developed a simple model to assess how relaxing the classical assumption of constant mortality rate affects the predicted effectiveness of anti-vectorial interventions. Comparing with a more realistic age-dependent model, he concluded that models with constant mortality overestimated the sensitivity of disease transmission to interventions that reduce mosquito survival. Pigeault et al. [Citation45] carried out experiments by using the avian malaria parasite (Plasmodium relictum) and its natural vector in the field, the mosquito (Culex pipiens). Their results indicated that older insects are often more resistant to infections than younger ones. These results suggested that structural and functional alterations in mosquito physiology with age may be more important than immunity in determining the probability of a Plasmodium infection in old mosquitoes. In view of mathematical modeling, age structure in vector-borne transmission models is not in itself new, although vector-borne age-structured models are predominantly focused on the age in the host population rather than in the vector population [Citation13,Citation26,Citation36,Citation54,Citation55,Citation57]. In [Citation13], Dang et al. proposed a vector-borne disease model that includes both incubation age of exposed hosts and infection age of infectious hosts. By using the method of Lyapunov functionals, threshold dynamics is established. Inaba and Sekine [Citation26] developed a mathematical model for Chagas disease with infection-age-dependent infectivity. Under the assumption that the population size of vectors is constant, the local and global stability of steady states were investigated. In particular, it was shown that bifurcation of endemic steady states can occur even when the basic reproduction number is less than one. In [Citation55], Vogt-Geisse et al. first obtained a partial differential equation model, which is reduced to an ordinary differential equation model with multiple age groups coupled with age since infection. The theoretical results may help generate insights into effective control measures by targeting age groups in an optimal way.

In addition to the transmission from vectors to hosts, vector-borne diseases may also involve horizontal transmission between hosts. Horizontal transmission is the spread of viruses from one individual to another, usually through contact with bodily excretions or fluids, such as sputum or blood, that contain pathogens. The horizontal transmission risk may not be high in general, but it cannot be neglected [Citation26]. For example, for Chagas disease, the prevalence rate among blood donors in Bolivia reached the level beyond 10% in 1993-1994, but in other countries of Central and South America, prevalence rates were at most several percent [Citation50]. Some models have been developed for vector-borne diseases with horizontal transmission [Citation1,Citation15,Citation20,Citation31–33]. They are described by ordinary differential equations. Wei et al. [Citation58] also used a delay differential equation system to study vector-borne diseases.

It is well-known that an important factor affecting the transmission dynamics is the incidence rate, which depends on the population behavior and the infectivity of the disease. Mathematically, incidence rate is the number of new infections per unit time. In the above mentioned references, the incidence rates are either bilinear or standard. Experiments and field work have indicated that nonlinear incidences may be more reasonable. Liu et al. [Citation35] showed that epidemiological models with nonlinear incidence rates exhibit much richer dynamics than those with bilinear incidence rates. By describing how the biology of host-parasite association could determine the functional form of the transmission rate, Georgescu and Hsieh [Citation21] and Korobeinikov et al. [Citation29] formulated a variety of models with incidence rates having the form f(S(t))g(I(t)), where f(S(t)) is the contact rate function dependent on susceptible class S and g(I(t)) represents the force of infection by infectious class I. Thereafter, Korobeinikov et al. [Citation27,Citation28] used a very general form f(S,I) for the incidence rate. For more other nonlinear incidence rates, we refer readers to [Citation5,Citation10,Citation11,Citation17,Citation34,Citation44,Citation56].

Motivated by the above studies, in this paper we will further the study on vector-borne diseases by incorporating age structures into both infectious hosts and infectious vectors and employing general nonlinear incidence rates. This paper is different from a previous study [Citation57] in that we include two routes of transmission of the infectious disease. We evaluate the influence of the horizontal transmission and exposed classes on disease dynamics. We also compare the results under different incidence rates. The model is formulated and some preliminary results are presented in Section 2. In Section 3, we study the existence and local stability of steady states. Using the Fluctuation Lemma and Lyapunov functional approaches, we study the asymptotic dynamics of the model in Section 4. We show that if the basic reproduction number is less than one, then the disease-free steady state is globally asymptotically stable and hence the infection can be eradicated; if the basic reproduction number is larger than one, the infectious steady state is globally asymptotically stable. If the total population of hosts or vectors is not constant, then the standard incidence rate does not satisfy the condition for the stability of the infectious steady state. As a result, in Section 5, for variable populations of both hosts and vectors, we study a model with the standard incidence rates. We extend some results to this model but the stability of the infectious steady state is challenging due to the complexity of the transcendental characteristic equation. The paper concludes with a brief discussion.

2. The model and preliminary results

The host population Nh in our model includes four classes, namely the susceptible Sh, exposed Eh, infectious Ih, and recovered Rh. We assume that the recovered hosts have acquired permanent immunity and hence in the sequel there is no need to include its dynamics in the model. The vector population Nv has three classes: susceptible Sv, exposed Ev, and infectious Iv. It is assumed that there are no recovered vectors. Letting a be the age of infection, we have ih(t,a) and iv(t,a) as infectious hosts and vectors with age a at time t, respectively. The total numbers of infectious hosts and infectious vectors at time t are Ih(t)=0ih(t,a)da and Iv(t)=0iv(t,a)da, respectively.

Let the natural death rate of the host be μh and the recruitment rate of hosts be λh. We assume that all the recruited are susceptible. Infectious hosts can infect susceptible hosts at the rate ϕ(Sh(t),0βh(a)ih(t,a)da) and infectious vectors can also infect them at the rate ψ(Sh(t),0βv(a)iv(t,a)da), where βh(a) denotes the age-dependent horizontal transmission rate of the disease from infectious hosts to susceptible hosts and βv(a) is the transmission rate from infectious vectors to susceptible hosts. Note that βh(a) or βv(a) lumps all the transmission-related parameters such as the contact or biting rate and the probability of successful transmission. The transition rate from exposed hosts to infectious hosts is αh. The parameter γh(a) is the per capita recovery rate of infectious hosts and ρh(a) is the disease-induced death rate of infectious hosts.

Unlike many of the existing models, we do not assume the population size of vectors to be constant. We let λv be the birth rate of vectors and μv be the natural death rate. For susceptible vectors, the force of infection is f(Sv(t),0kh(a)ih(t,a)da), where kh(a) is the transmission rate from infectious hosts to susceptible vectors due to biting. The transition rate from exposed vectors to infectious vectors is αv and the infection-related death rate of infectious vectors is ρv(a).

Based on the above assumptions, we develop the following model (1) dSh(t)dt=λhμhSh(t)ϕ(Sh(t),0βh(a)ih(t,a)da)ψ(Sh(t),0βv(a)iv(t,a)da),dEh(t)dt=ϕ(Sh(t),0βh(a)ih(t,a)da)+ψ(Sh(t),0βv(a)iv(t,a)da)(αh+μh)Eh(t),ih(t,a)t+ih(t,a)a=(μh+γh(a)+ρh(a))ih(t,a),ih(t,0)=αhEh(t),dSv(t)dt=λvf(Sv(t),0kh(a)ih(t,a)da)μvSv(t),dEv(t)dt=f(Sv(t),0kh(a)ih(t,a)da)(αv+μv)Ev(t),iv(t,a)t+iv(t,a)a=(μv+ρv(a))iv(t,a),iv(t,0)=αvEv(t),Sh(0)=Sh0R+,Eh(0)=Eh0R+,ih(0,)=ih0L+1(0,),Sv(0)=Sv0R+,Ev(0)=Ev0R+,iv(0,)=iv0L+1(0,).(1) Here R+=[0,) and L+1 is the positive cone of L1(0,).

Denote Jh(t)=0βh(a)ih(t,a)da,Jv(t)=0βv(a)iv(t,a)da,J(t)=0kh(a)ih(t,a)da,δh(a)=μh+γh(a)+ρh(a),δv(a)=μv+ρv(a). Clearly, δh(a)μh and δv(a)μv for all aR+. Using the new notations, model (Equation1) can be rewritten as (2) dSh(t)dt=λhμhSh(t)ϕ(Sh(t),Jh(t))ψ(Sh(t),Jv(t)),dEh(t)dt=ϕ(Sh(t),Jh(t))+ψ(Sh(t),Jv(t))(αh+μh)Eh(t),ih(t,a)t+ih(t,a)a=δh(a)ih(t,a),ih(t,0)=αhEh(t),dSv(t)dt=λvf(Sv(t),J(t))μvSv(t),dEv(t)dt=f(Sv(t),J(t))(αv+μv)Ev(t),iv(t,a)t+iv(t,a)a=δv(a)iv(t,a),iv(t,0)=αvEv(t),Sh(0)=Sh0R+,Eh(0)=Eh0R+,ih(0,)=ih0L+1(0,),Sv(0)=Sv0R+,Ev(0)=Ev0R+,iv(0,)=iv0L+1(0,).(2) For model (Equation2), we assume ih(0,0)=ih0(0) and iv(0,0)=iv0(0). Thus (3) ih0(0)=αhEh0andiv0(0)=αvEv0.(3)

We also assume the following for the parameters and the nonlinear incidence functions in (Equation2).

(C1)

βh, βv, γh, ρh, ρv, khL+(0,), where L+(0,) is the positive cone of the Banach space L(0,). The functions βh, βv, and kh are uniformly continuous and bounded. Moreover, at least one of βh, βv and kh cannot be the zero function. Otherwise, there is no transmission of the disease.

(C2)

ϕ(u,v), ψ(u,v) and f(u,v) are differentiable with ϕ(u,v)u>0, ψ(u,v)u>0, f(u,v)u>0 for u>0 and v>0, and ϕ(u,v)v>0, ψ(u,v)v>0, f(u,v)v>0 for u>0 and v0. Furthermore, ϕ(u,0)=ϕ(0,v)=ψ(u,0)=ψ(0,v)=f(u,0)=f(0,v)=0 for all u, vR+.

(C3)

ϕ(u,v)v, ψ(u,v)v and f(u,v)v are continuous at (Sh0,0), (Sh0,0) and (Sv0,0), respectively, where Sh0=λhμh and Sv0=λvμv. Moreover, 2ϕ(u,v)v20, 2ψ(u,v)v20, and 2f(u,v)v20 for all u, v>0.

(C4)

ϕ(u,v), ψ(u,v) and f(u,v) are all locally Lipschitz continuous, that is, given C>0, there exist positive values Lu, Lv, Ku, Kv, Mu and Mv such that |ϕ(u,v)ϕ(u~,v)|Lu|uu~|,|ϕ(u,v)ϕ(u,v~)|Lv|vv~|,|ψ(u,v)ψ(u~,v)|Ku|uu~|,|ψ(u,v)ψ(u,v~)|Kv|vv~|,|f(u,v)f(u~,v)|Mu|uu~|,|f(u,v)f(u,v~)|Mv|vv~|for any u, u~, v, v~ in the interval [0,C].

For positive u and v, we know that ϕ(u,v)>0, ψ(u,v)>0 and f(u,v)>0. From assumption (C2), these functions are also strictly increasing in terms of one independent variable when the other is fixed and positive. (C3) implies that the incidence rates are concave down for the first fixed variable.

The state space of (Equation2) is X+=R+×R+×L+1(0,)×R+×R+×L+1(0,), the positive cone of the Banach space R×R×L1(0,)×R×R×L1(0,) equipped with the norm (Sh,Eh,ih,Sv,Ev,iv)=|Sh|+|Eh|+ih1+|Sv|+|Ev|+iv1 for (Sh,Eh,ih,Sv,Ev,iv)R×R×L1(0,)×R×R×L1(0,).

It follows from [Citation6,Citation25,Citation37] that system (Equation2) has a unique continuous solution in X+ if the initial value (Sh0,Eh0,ih0,Sv0,Ev0,iv0)X+ satisfies (Equation3). This leads to a solution semiflow Φ:R+×X+X+ defined by Φ(t,(Sh0,Eh0,ih0,Sv0,Ev0,iv0))=(Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,)), where tR+ and (Sh(t),Eh(t),ih(t,a),Sv(t),Ev(t),iv(t,a)) is the unique solution of (Equation2) through the initial value (Sh0,Eh0,ih0,Sv0,Ev0,iv0)X+.

Let Gh(t)=Sh(t)+Eh(t)+0ih(t,a)da,Nv(t)=Sv(t)+Ev(t)+0iv(t,a)da. It follows from (Equation2) that (4) dGh(t)dtλhμhSh(t)μhEh(t)0δh(a)ih(t,a)daλhμhGh(t)(4) and (5) dNv(t)dtλvμvNv(t),(5) which respectively imply that lim suptGh(t)λhμhandlim suptNv(t)λvμv. Furthermore, one can easily see from (Equation4) and (Equation5) that the set Ω=(Sh,Eh,ih,Sv,Ev,iv)X+Sh+Eh+ih1λhμh,Sv+Ev+iv1λvμv is a positively invariant and attracting subset for system (Equation2).

3. The existence and local stability of steady states

Obviously, system (Equation2) always has the disease-free steady state E0=(Sh0,0,0,Sv0,0,0). Recall that Sh0=λhμh and Sv0=λvμv.

For the convenience of notation, we denote ξh=0βh(a)σh(a)da,ξv=0βv(a)σv(a)da,η=0kh(a)σh(a)da,σh(a)=exp0aδh(s)dsfor aR+,σv(a)=exp0aδv(s)dsfor aR+. It turns out that the structure of steady states of (Equation2) is determined by the basic reproduction number R0, which is given by R0=ϕ(Sh0,0)Jhξhαhαh+μh+ψ(Sh0,0)Jvf(Sv0,0)Jαhηαh+μhξvαvαv+μv. The expression of R0 is derived in the coming discussion of the existence of steady states and the local stability of the disease-free steady state. R0 is the number of infectious hosts generated by the introduction of an infectious host into a population consisting of only susceptible hosts, which agrees with the biological interpretation of the basic reproduction number. The first term in R0 is the number generated by the horizontal transmission while the second term is the number generated through vectors.

Let E=(Sh,Eh,ih,Sv,Ev,iv)X+ (in fact in Ω) be a steady state of (Equation2), that is, (6a) 0=λhμhShϕ(Sh,Jh)ψ(Sh,Jv),(6a) (6b) 0=ϕ(Sh,Jh)+ψ(Sh,Jv)(αh+μh)Eh,(6b) (6c) dih(a)da=δh(a)ih(a),(6c) (6d) 0=λvf(Sv,J)μvSv,(6d) (6e) f(Sv,J)=(αv+μv)Ev,(6e) (6f) div(a)da=δv(a)iv(a),(6f) (6g) Jh=0βh(a)ih(a)da=ξhih(0)=ξhαhEh,(6g) (6h) Jv=0βv(a)iv(a)da=ξviv(0)=ξvαvEv,(6h) (6i) J=0kh(a)ih(a)da=ηih(0)=ηαhEh.(6i) From (Equation6a) and (Equation6b), we have (7) Sh=λh(αh+μh)Ehμh.(7) By applying the implicit function theorem, we see from (Equation6d) and (Equation6i) that there exists a function p such that Sv=p(Eh), where p(0)=Sv0 and p(Eh)=ηαhf(p(Eh),ηαhEh)Jμv+f(p(Eh),ηαhEh)Sv. This, combined with (Equation6e) and (Equation6i), gives (8) Ev=f(Sv,J)αv+μv=f(p(Eh),ηαhEh)αv+μv.(8) It follows from (Equation6b), (Equation6g), (Equation6h), (Equation7), and (Equation8) that Eh is a nonnegative zero of H(x), where H(x)=ϕλh(αh+μh)xμh,ξhαhx+ψλh(αh+μh)xμh,ξvαvf(p(x),ηαhx)αv+μv(αh+μh)x. Obviously, H(0)=0, which gives the disease-free steady state. Moreover, from the above relations, one can easily see that if Eh0 then none of ih, Ev, and iv is zero. This means that a steady state rather than the disease-free one is an infectious steady state. In the following, we discuss their existence.

First, we assume R01. Then for any x>0, it follows from (C2) and (C3) that H(x)ϕ(Sh0,0)Jhξhαh+ψ(Sh0,0)Jvξvαvαv+μvf(Sv0,0)Jηαh(αh+μh)x=(αh+μh)(R01)x0. This implies that there is no infectious steady state when R01.

Next, we suppose R0>1. Note that H(0)=0, H(λhαh+μh)=λh<0, and H(0)=ϕ(Sh0,0)Jhξhαh+ψ(Sh0,0)Jvξvαvαv+μvf(Sv0,0)Jηαh(αh+μh)=(αh+μh)( R01)>0. It follows that H(x)>0 if x>0 is sufficiently small. By the Intermediate Value Theorem, H(x) has at least one positive zero and hence there is at least one infectious steady state. We claim that (Equation2) has a unique infectious steady state. Otherwise, suppose that there exist at least two infectious steady states, say E=(Sh,Eh,ih,Sv,Ev,iv) and E¯=(S¯h,E¯h,i¯h,S¯v,E¯v,i¯v). Without loss of generality, we can assume that E¯h>Eh. This implies S¯v=p(E¯h)<p(Eh)=Sv since p is decreasing. Let b=E¯hEh, which is larger than 1. We have Ev=f(Sv,J)αv+μv>f(S¯v,J)αv+μv=f(S¯v,1bJ¯)αv+μv>1bf(S¯v,J¯)αv+μv=1bE¯v. The first inequality comes from the monotonicity of f in the first independent variable while the second inequality comes from the concavity of f with respect to the second independent variable. Similarly, we have Eh=ϕ(Sh,Jh)+ψ(Sh,Jv)αh+μh>ϕ(S¯h,αhξhEh)+ψ(S¯h,αvξvEv)αh+μh>ϕ(S¯h,1bαhξhE¯h)+ψ(S¯h,1bαvξvE¯v)αh+μh1bϕ(S¯h,ξhαhE¯h)+ψ(S¯h,ξvαvE¯v)αh+μh=1bE¯h, which is a contradiction to Eh=1bE¯h. This proves the uniqueness of infectious steady states.

The following theorem summarizes the conditions in terms of the basic reproduction number on the existence of steady states.

Theorem 3.1

  1. If R01, then system (Equation2) only has the disease-free steady state E0.

  2. If R0>1, then besides E0, system (Equation2) also has a unique infectious steady state, denoted by E.

Now, we investigate the local stability of steady states. Let E¯=(S¯h,E¯h,i¯h, S¯v,E¯v,i¯v) be a steady state of (Equation2). Linearizing (Equation2) at the steady state leads to (9) dSh(t)dt=(μh+Λ)Sh(t)ϕ(S¯h,J¯h)Jh0βh(a)ih(t,a)daψ(S¯h,J¯v)Jv0βv(a)iv(t,a)da,dEh(t)dt=ΛSh(t)(αh+μh)Eh(t)+ϕ(S¯h,J¯h)Jh0βh(a)ih(t,a)da+ψ(S¯h,J¯v)Jv0βv(a)iv(t,a)da,ih(t,a)t+ih(t,a)a=δh(a)ih(t,a),ih(t,0)=αhEh(t),dSv(t)dt=f(S¯v,J¯)J0kh(a)ih(t,a)da(μv+f(S¯v,J¯)Sv)Sv(t),dEv(t)dt=f(S¯v,J¯)J0kh(a)ih(t,a)da+f(S¯v,J¯)SvSv(t)(αv+μv)Ev(t),iv(t,a)t+iv(t,a)a=δv(a)iv(t,a),iv(t,0)=αvEv(t),(9) where Λ=ϕ(S¯h,J¯h)Sh+ψ(S¯h,J¯v)Sh. Letting Sh(t)=x0eτt, Eh(t)=y0eτt, ih(t,a)=z0(a)eτt, and Sv(t)=u0eτt, Ev(t)=v0eτt, iv(t,a)=w0(a)eτt, and substituting them into (Equation9), from the third equation and the seventh equation, we can get z0(a)=z0(0)σh(a)eτa and w0(a)=w0(0)σv(a)eτa. Then system (Equation9) becomes (τ+μh+Λ)x0+ϕ(S¯h,J¯h)Jhξhˆ(τ)z0(0)+ψ(S¯h,J¯v)Jvξvˆ(τ)w0(0)=0,Λx0+(τ+αh+μh)y0ϕ(S¯h,J¯h)Jhξhˆ(τ)z0(0)ψ(S¯h,J¯v)Jvξvˆ(τ)w0(0)=0,αhy0z0(0)=0,f(S¯v,J¯)Jηˆ(τ)z0(0)+(τ+μv+f(S¯v,J¯)Sv)u0=0,f(S¯v,J¯)Jηˆ(τ)z0(0)f(S¯v,J¯)Svu0+(τ+αv+μv)v0=0,αvv0w0(0)=0, where ξˆh(τ)=0βh(a)σh(a)eτada,ξˆv(τ)=0βv(a)σv(a)eτada,ηˆ(τ)=0kh(a)σh(a)eτada. This leads to the characteristic equation

Δ(E¯)=τ+μh+Λ0ϕ(S¯h,J¯h)Jhξˆh(τ)00ψ(S¯h,J¯v)Jvξˆv(τ)Λτ+αh+μhϕ(S¯h,J¯h)Jhξˆh(τ)00ψ(S¯h,J¯v)Jvξˆv(τ)0αh100000f(S¯v,J¯)Jηˆ(τ)τ+μv+f(S¯v,J¯)Sv0000f(S¯v,J¯)Jηˆ(τ)f(S¯v,J¯)Svτ+αv+μv00000αv1=0.The steady state E¯ is locally asymptotically stable if all eigenvalues of the characteristic equation have negative real parts and it is unstable if at least one eigenvalue has a positive real part. For more detail, we refer the readers to Martcheva and Thieme [Citation40].

Theorem 3.2

  1. When R0<1, the infection-free steady state E0 of (Equation2) is locally asymptotically stable. When R0>1, it is unstable.

  2. The infectious steady state E of (Equation2) is locally asymptotically stable when R0>1.

Proof.

(i) Note that ϕ(Sh0,0)Sh=ψ(Sh0,0)Sh=f(Sv0,0)Sv=0 by (C2). The characteristic equation at E0 is Δ(E0)=(τ+μh)(τ+μv)(τ+αh+μh)(τ+αv+μv)g1(τ)=0, where g1(τ)=1ϕ(Sh0,0)Jhξˆh(τ)αhτ+αh+μhαhηˆ(τ)τ+αh+μhαvξˆv(τ)τ+αv+μvψ(Sh0,0)Jvf(Sv0,0)J. The roots of g1(τ)=0 determines the stability of E0. If R0>1, then g1(0)=1R0<0. We also have limτg1(τ)=1>0. Therefore, the equation g1(τ)=0 admits a positive root, which shows that E0 is unstable. Next, we claim that the root of g1(τ)=0 has negative real parts when R0<1. Prove by contradiction. If τ0 is a root with non-negative real part, then from the definition of g1 we have 1=αhξˆh(τ0)τ0+αh+μhϕ(Sh0,0)Jh+αhηˆ(τ0)τ0+αh+μhαvξv(τ0)τ0+αv+μvψ(Sh0,0)Jvf(Sv0,0)Jαhξhαh+μhϕ(Sh0,0)Jh+αhηαh+μhαvξvαv+μvψ(Sh0,0)Jvf(Sv0,0)J=R0, which leads to a contradiction. This proves the claim and hence E0 is locally asymptotically stable when R0<1.

(ii) When R0>1, the characteristic equation at the infectious steady state E is (10) Δ(E)=(τ+μh+Λ)(τ+αh+μh)τ+μv+f(Sv,J)Sv(τ+αv+μv)αh(τ+μh)ϕ(Sh,Jh)Jhξˆh(τ)τ+μv+f(Sv,J)Sv(τ+αv+μv)αh(τ+μh)αvξˆv(τ)ψ(Sh,Jv)Jv(τ+μv)f(Sv,J)Jηˆ(τ)=0,(10) where Λ=ϕ(Sh,Jh)Sh+ψ(Sh,Jv)Sh. We now prove that all roots of (Equation10) have negative real parts with contradictive arguments. If τ0 is a solution with nonnegative real part, then we have (11) 1=αh(τ0+μh)ϕ(Sh,Jh)Jhξˆh(τ0)(τ0+μv+f(Sv,J)Sv)(τ0+αv+μv)(τ0+μh+Λ)(τ0+αh+μh)(τ0+μv+f(Sv,J)Sv)(τ0+αv+μv)+αh(τ0+μh)αvξˆv(τ0)ψ(Sh,Jv)Jv(τ0+μv)f(Sv,J)Jηˆ(τ0)(τ0+μh+Λ)(τ0+αh+μh)(τ0+μv+f(Sv,J)Sv)(τ0+αv+μv)<αhξhαh+μhϕ(Sh,Jh)Jh+αhηαh+μhαvξvαv+μvψ(Sh,Jv)Jvf(Sv,J)J.(11) However, since iv(0)=αvEv=αvαv+μvfSv,0kh(a)ih(a)da=αvαv+μvf(p(Eh),ηαhEh) and ih(0)=αhEh=αhαh+μhϕSh,0βh(a)ih(a)da+ψSh,0βv(a)iv(a)da, we have 1=1αh+μhϕ(Sh,0βh(a)ih(a)da)Eh+ψ(Sh,0βv(a)iv(a)da)Eh=1αh+μhϕ(Sh,ξhαhEh)Eh+1αh+μhψ(Sh,ξvαvEv)Eh=1(αh+μh)Ehϕ(Sh,ξhαhEh)+1(αh+μh)EhψSh,ξvαvf(p(Eh),ηαhEh)αv+μvξhαhαh+μhϕ(Sh,Jh)Jh+ηαhαh+μhψ(Sh,Jv)Jvξvαvαv+μvf(Sv,J)J, which is a contradiction to (Equation11). This completes the proof.

4. Threshold dynamics

We first establish the global stability of the disease-free steady state E0 of (Equation2) using the Fluctuation Lemma. Let u=lim inftu(t) and u=lim suptu(t).

Lemma 4.1

Fluctuation Lemma [Citation24]

Let u:R+R be bounded and continuously differentiable. Then there exist two sequences {sn} and {tn} such that sn, tn, u(sn)u, u(sn)0, u(tn)u, and u(tn)0 as n.

Lemma 4.2

[Citation25]

Suppose kL+1(0,) and B is a bounded function from R+ to R. Then lim supt0tk(θ)B(tθ)dθBk1.

Theorem 4.3

The disease-free steady state E0 of (Equation2) is globally asymptotically stable when R0<1.

Proof.

On the basis of the local stability, it suffices to prove that E0 is globally attractive.

We first get the expressions of ih(t,a) and iv(t,a). Consider ih(t,a)t+ih(t,a)a=δh(a)ih(t,a)ih(t,0)=αhEh(t)ih(0,0)=ih0 and iv(t,a)t+iv(t,a)a=δv(a)iv(t,a)iv(t,0)=αvEv(t)iv(0,0)=iv0 Solving them by the method of characteristic lines gives the solutions (12) ih(t,a)=σh(a)αhEh(ta),0atσh(a)σh(at)ih0(at),a>t>0(12) and iv(t,a)=σv(a)αvEv(ta),0at,σv(a)σv(at)iv0(at),a>t>0.

Next, we prove that Eh=Ev=0. From the second and sixth equations of (Equation2), we see that (13) Eh(ϕ(Sh,Jh))+(ψ(Sh,Jv))αh+μhandEv(f(Sv,J))αv+μv.(13) By (C3) and (Equation12), we have ϕ(Sh(t),Jh(t))ϕ(Sh0,0)JhJh(t)=ϕ(Sh0,0)Jh0βh(a)ih(t,a)da=ϕ(Sh0,0)Jh0tβh(a)αhEh(ta)σh(a)da+tβh(a)ih0(at)σh(a)σh(at)daϕ(Sh0,0)Jh0tβh(a)αhEh(ta)σh(a)da+eμhtβhih01. It follows from Lemma 4.2 that (14) (ϕ(Sh,Jh))ϕ(Sh0,0)JhξhαhEh.(14) Similarly, (15) (ψ(Sh,Jv))ψ(Sh0,0)JvξvαvEvand(f(Sv,J))f(Sv0,0)JηαhEh.(15) Therefore, it follows from (Equation13)–(Equation15) that EhαhEhαh+μhξhϕ(Sh0.0)Jh+ξvαvαv+μvψ(Sh0,0)Jvηf(Sv0,0)J=R0Eh, which implies that Eh=0 because R0<1. Then Ev=0 follows immediately from the second inequalities in both (Equation13) and (Equation15).

Similarly as in the proof of (Equation14), we can get (ih(t,)1)=(iv(t,)1)=0 and hence limtih(t,)1=limtiv(t,)1=0.

Finally, we show limtSh(t)=Sh0 and limtSv(t)=Sv0. We only give the proof of limtSh(t)=Sh0 as that for the other is similar. Recall that ShSh0. It suffices to show (Sh)Sh0. By Lemma 4.1, there exists a sequence {tn} such that tn, Sh(tn)(Sh), and dSh(tn)dt0 as n. Note that limt0βh(a)ih(t,a)da=0andlimt0βv(a)iv(t,a)da=0 because limtih(t,)1=limtiv(t,)1=0. It follows from dSh(tn)dt=λhμhSh(tn)ϕSh(tn),0βh(a)ih(tn,a)daψSh(tn),0βv(a)iv(tn,a)da by letting n that 0=λhμh(Sh). Hence (Sh)=λhμh=Sh0.

In summary, we have shown that limt(Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,))=E0. This completes the proof.

In the remaining of this section, we establish the attractivity of the infectious steady state E by constructing suitable Lyapunov functionals. For this purpose, we need the persistence of (Equation2).

We start with the existence of a compact attractor A that attracts all points in X+. The existence is established by applying the following result.

Lemma 4.4

[Citation22]

If Φ(t):X+X+, tR+ is asymptotically smooth, point dissipative, and orbits of bounded sets are bounded, then there exists a global attractor.

From Smith [Citation51], if each forward invariant bounded closed set is attracted by a non-empty compact set, then the semiflow is asymptotically smooth.

Lemma 4.5

Φ(t)=Φ1(t)+Φ2(t) is asymptotically smooth if the following are satisfied: (1) Φ2(t) is completely continuous;(2) There exists a continuous function u~ such that u~(t,r~)0 as t and u~(t,r~)Φ1(t)x when x<r~.

Theorem 4.6

If R0>1, then there exists a global compact attractor A for the solution semiflow Φ of (Equation2) in X+.

Proof.

It follows from (Equation4) and (Equation5) that Φ is point dissipative and orbits of bounded sets are bounded. In order to apply Lemma 4.4, we only need to prove that Φ is asymptotically smooth. For any tR+ and x=(Sh0,Eh0,ih0,Sv0,Ev0,iv0)X+, let Φ1(t,x)=(0,0,iˆh(t,),0,0,iˆv(t,)) and Φ2(t,x)=(Sh(t),Eh(t),i~h(t,),Sv(t),Ev(t),i~v(t,)), where (16) iˆh(t,a)=0,0at,σh(a)σh(at)ih0(at),0<t<a,(16) (17) iˆv(t,a)=0,0at,σv(a)σv(at)iv0(at),0<t<a,(17) and i~h(t,a)=ih(t,a)iˆh(t,a)=σh(a)αhEh(ta),0at,0,0<t<a,i~v(t,a)=iv(t,a)iˆv(t,a)=σv(a)αvEv(ta),0at,0,0<t<a. Then Φ=Φ1+Φ2. It is easy to see that iˆh, iˆv, i~h, and i~v are nonnegative. Using (Equation16) and (Equation17), we get Φ1(t,x)=iˆh(t,)1+iˆv(t,)1=tih0(at)σh(a)σh(at)da+tiv0(at)σv(a)σv(at)da=0ih0(a)σh(a+t)σh(a)da+0iv0(a)σv(a+t)σv(a)daeμhtih01+eμvtiv01eμtx, where μ=min{μh,μv}. Thus, the second condition in Lemma 4.5 regarding Φ1 is satisfied.

Now, we show Φ2 is completely continuous, i.e., for a positive t and a bounded BX+, the set {Φ2(t,x)|x=(Sh0,Eh0,ih0,Sv0,Ev0,iv0)B}, denoted by Bt, is precompact. As in Chen et al. [Citation12], we only need to prove that Bt(ih,iv)={(i~h(t,),i~v(t,))|(Sh(t),Eh(t),i~h(t,),Sv,Ev,i~v(t,))Bt} is precompact by employing the Fréchet-Kolmogrov Theorem. We only verify the second condition (as the others are trivial to verify) that Bt(ih,iv) is uniformly continuous or (18) limk0+i~h(t,)i~h(t,+k)1=0=limk0+i~v(t,)i~v(t,+k)1(18) uniformly in Bt(ih,iv).

Note that, by (Equation4) and (Equation5) again, there exists C>0 such that Gh(t)C and Nv(t)C for all t0 and xB.

We only show the first part of (Equation18) holds as the proof for the second one is similar. Clearly, it holds when t = 0 since i~h(0,)=0. Without losing generality, let t be positive. When k0+, we let k be in the interval (0,t). It follows that (19) i~h(t,)i~h(t,+k)1=0|i~h(t,a)i~h(t,a+k)|da=0tk|αhEh(tak)σh(a+k)αhEh(ta)σh(a)|da+tktαhEh(ta)σh(a)da0tkαhEh(tak)|σh(a+k)σh(a)|da+0tkαh|Eh(tak)Eh(ta)|σh(a)da+Ckαh.(19) Recall that σh(a)=e0a(μh+ρh(s)+γh(s))ds is a monotone decreasing function of a. It is clear that σh(a)σh(a+k)=σh(a)1σh(a+k)σh(a)eμha1eaa+kδh(s)dseμhaaa+kδh(s)dseμhakδh. Moreover, dEh(t)dtϕ(Sh(t),Jh(t))+ψ(Sh(t),Jv(t))+(αh+μh)Cϕ(C,Jh(t))+ψ(C,Jv(t))+(αh+μh)Cϕ(C,0)Jh0βh(a)ih(t,a)da+ψ(C,0)Jv0βv(a)iv(t,a)da+(αh+μh)Cϕ(C,0)Jhβh+ψ(C,0)Jvβv+αh+μhCLE. Therefore, it follows from (Equation19) that i~h(t,)i~h(t,+k)1αhCkδhμh+αhLEkt+Ckαh. This completes the proof.

Clearly, the global attractor AΩ. Since it is invariant, it can only contain points with a total trajectory through it. From the reference [Citation51], a total trajectory x:RX+ of Φ requires Φ(s,x(t))=x(s+t) for tR and sR+. Thus ih(t,a)=ih(ta,0)σh(a) and iv(t,a)=iv(ta,0)σv(a) for a>0. The α-limit of a total trajectory x(t) that passes through the initial value x(0)=x0 is α(x0)=t0st{x(s)}¯. We will apply Theorem 5.2 in [Citation51] to establish the uniform persistence. For this purpose, we define ρ:X+R+ by ρ(Sh,Eh,ih,Sv,Ev,iv)=0βh(a)ih(a)da+0βv(a)iv(a)da for (Sh,Eh,ih,Sv,Ev,iv)X+. Clearly ρ is continuous and not identically zero. A biological interpretation of ρ is the initial infection force for the host. Let X0={x0=(Sh0,Eh0,ih0,Sv0,Ev0,iv0)X+|ρ(Φ(t,x0))=0 for all t0}. We say that Φ is uniformly weakly ρ-persistent if there exists γ>0 such that lim suptρ(Φ(t,x0))>γwhenever ρ(x0)>0 and is uniformly ρ-persistent if we can replace limsup by liminf above (see [Citation51, pp. 126]). Note that ρ(Φ(t,x0))=Jh(t)+Jv(t) for all t0 and x0X+.

To prove the uniform weak ρ-persistence, we need the following result. Though the result has been used in the literature, to the best of our knowledge, it has been explicitly stated and proved recently for the first time by Bentout et al. [Citation4].

Lemma 4.7

Let x0X+ such that ρ(x0)>0. Then there exists a T0 such that ρ(Φ(t,x0))>0 for tT.

Proof.

It is easy to see that Sh(t)>0 and Sv(t)>0 for all t>0. Also by (4) and (Equation5), we know that {Φ(t,x0)|t0} is bounded. Let C>0 such that Φ(t,x0)C for t0. Then for t0, using the Lipschitz continuity on ϕ and ψ, we have dSh(t)dtλh(μh+Lu+Ku)Sh(t), which implies that lim inftSh(t)λhμh+Lu+Ku. Similarly, lim inftSv(t)λvμv+Mu. Therefore, without loss of generality (we can make a translation), we can assume that ρ(x0)>0, and there exist ζ>0 and γ>0 such that Sh(t)ζ, Sv(t)ζ, Jh(t)γ, Jv(t)γ, and J(t)γ for t0.

Now, using (Equation12) and the monotonicity and concavity of ϕ and ψ, we get Jh(t)=tβh(a)ih(t,a)da+0tβh(a)ih(t,a)da=tβh(a)σh(a)σh(at)ih0(at)da+0tβh(a)αhσh(a)Eh(ta)da=0βh(t+a)σh(a+t)σh(a)ih0(a)da+0tαhβh(a)σh(a)[e(αh+μh)(ta)Eh(0)+0tae(αh+μh)(tas)[ϕ(Sh(s),Jh(s))+ψ(Sh(s),Jv(s))]dsda+tβh(a)ih(t,a)da0t0taαhβh(a)σh(a)e(αh+μh)(tas)×ϕ(ζ,γ)JhJh(s)+ψ(ζ,γ)JvJv(s)dsda+Rh(t), where Rh(t)=0βh(t+a)σh(a+t)σh(a)ih0(a)da. Changing the order of integration gives us Jh(t)αhminϕ(ζ,γ)Jh,ψ(ζ,γ)Jv0t[Jh(s)+Jv(s)]κh(ts)ds+Rh(t), where κh(t)=0tβh(a)σh(a)e(αh+μh)(ta)da. With a similar but tedious computation, we can get Jv(t)αhαvf(ζ,γ)Jminϕ(ζ,γ)Jh,ψ(ζ,γ)Jv0t[Jh(s)+Jv(s)]κv(ts)ds+Rv(t), where κv(t)=0t0taτtaβv(a)σv(a)kh(τ)σh(τ)e(αv+μv)(taη)e(αh+μh)(ητ)dηdτda and Rv(t)=0βv(t+a)σv(t+a)σv(a)iv0(a)da. In summary, we have obtained, ρ(Φ(t,x0))0tρ(Φ(s,x0))κ(ts)ds+R(t), where κ(t)=αhminϕ(ζ,γ)Jh,ψ(ζ,γ)Jvκh(t)+αvf(ζ,γ)Jκv(t) and R(t)=Rh(t)+Rv(t). Note that R(0)=ρ(x0)>0 and R(t) is locally integrable. Moreover, by (C1), we know that κ is not zero a.e. By Corollary B.6 [Citation51], there exists T0 (dependent only on κ and not on R) such that ρ(Φ(t,x0))>0 for tT.

Theorem 4.8

Suppose R0>1. Then model (Equation2) is uniformly weakly ρ-persistent.

Proof.

By contradiction, we suppose that (Equation2) is not uniformly weakly ρ-persistent. Then for any ε>0, there is an xε=(Sh0ε,Eh0ε,ih0ε,Sv0ε,Ev0ε,iv0ε)X+ such that ρ(xε)>0 and lim suptρ(Φ(t,xε))ε. Since R0>1, we can choose a positive ε0 such that ε1=λh(Lv+Kv)ε0μhε0>0,ε2=αhη(Lv+Kv)ε0αh+μh+ε0,ε3=λvMvε2μvε0>0, and (20) ϕ(ε1,ε0)Jhαhβhσhˆ(ε0)αh+μh+ε0+ψ(ε1,ε0)Jvf(ε3,ε2)Jαhkhσhˆα+h+μh+ε0αvβvσvˆ(ε0)αv+μv+ε0>1.(20) Here and in the remaining of the proof, the notation ˆ represents the Laplace transform. Given this ε0, we can find an xε02X+ with ρ(xε02)>0 such that lim suptρ(Φ(t,xε02))ε02. For simplicity of notation, we write x for xε02. In the following, we will first establish some estimates on Sh, Sv, and J, and then use Laplace transform to obtain a contradiction.

From lim suptρ(Φ(t,x))ε02, we can find t0>0 such that when tt0, ρ(Φ(t,x))<ε0. We further assume that t0 is 0 since we can replace x with Φ(t0,x). By the first equation of (Equation2), (C2), and (C4), we have dSh(t)dtλh(Lv+Kv)ε0μhSh(t)for tR+, which implies that lim inftSh(t)λh(Lv+Kv)ε0μh. Without losing generality, we let Sh(t)ε1. Similarly, from the second equation of (Equation2), we have lim suptEh(t)(Lv+Kv)ε0αh+μh. This, combined with J(t)=0tkh(a)σh(a)αhEh(ta)da+tkh(a)σh(a)σh(at)ih0(at)da0tkh(a)σh(a)αhEh(ta)da+eμhttkh(a)ih0(at)da0tkh(a)σh(a)αhEh(ta)da+eμhtkhih01 and Lemma 4.2, gives lim suptJ(t)αhη(Lv+Kv)ε0αh+μh. Again, we let J(t)ε2. Using the similar argument as those for obtaining Sh(t)ε1, we can get Sv(t)ε3 for tR+.

Now, from the second equation of (Equation2), (C2), and (C3), we get dEh(t)dtϕ(ε1,ε0)JhJh(t)+ψ(ε1,ε0)JvJv(t)(αh+μh)Eh(t)for tR+. We take Laplace transforms to get λEˆh(λ)Eh(0)ϕ(ε1,ε0)JhJˆh(λ)+ψ(ε1,ε0)JvJˆv(λ)(αh+μh)Eˆh(λ)for λR+, or (21) Eˆh(λ)ϕ(ε1,ε0)JhJˆh(λ)αh+μh+λ+ψ(ε1,ε0)JvJˆv(λ)αh+μh+λfor λR+(21) since Eh(0)0. By Lemma 4.7, Jˆh(λ)+Jˆv(λ)>0 for λR+, which combined with (Equation21) implies that Eˆh(λ)0 for any λR+. Similarly, (22) Eˆv(λ)f(ε3,ε2)Jαv+μv+λJˆ(λ)for λR+.(22) Moreover, from Jh(t)=0βh(a)ih(t,a)da0tβh(a)σh(a)αhEh(ta)dafor tR+, we obtain (23) Jˆh(λ)αhβhσhˆ(λ)Eˆh(λ)for {λR+}.(23) Similarly, for λR+, (24) Jˆv(λ)αvβvσvˆ(λ)Eˆv(λ),(24) (25) Jˆ(λ)αhkhσhˆ(λ)Eˆh(λ).(25) It follows from (Equation21)–(Equation25) that Eˆh(λ)ϕ(ε1,ε0)Jhαhβhσhˆ(λ)αh+μh+λ+ψ(ε1,ε0)Jvf(ε3,ε2)Jαhkhσhˆ(λ)αh+μh+λαvβvσvˆ(λ)αv+μv+λEˆh(λ), which is impossible at λ=ε0 because of (Equation20) and Eˆh(ε0)>0. This completes the proof.

In order to apply Theorem 5.2 [Citation51], it leaves to check hypothesis (H1) [Citation51, page 125]. This is done by the following two lemmas.

Lemma 4.9

Let x(t)=(Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,)) be a total trajectory of (Equation2) in A. If ρ(x(t))=0 for all t0 then ρ(x(t))=0 for all t>0.

Proof.

Clearly, for t0, dEh(t)dt=(α+μh)Eh(t). Since x(t)AΩ, we must have Eh(0)=0 (and hence Eh(t)=0 for t0) otherwise limtEh(t)=. This implies that, for t0, ih(t,a)=ih(ta,0)σh(a)=αhEh(ta)σh(a)=0 and hence J(t)=0. Using similar arguments as above we can get Ev(t)=0 and iv(t,a)=0 for t0. Then, for t0, we have dSh(t)dt=λhμhSh(t)anddSv(t)dt=λvμvSv(t). Using x(t)A again, we see that Sh(t)=Sh0 and Sv(t)=Sv0 for t0. By the uniqueness of solutions, we get x(t)E0. In particular, we have that Jh(t)+Jv(t)=0 for t>0.

Lemma 4.10

For a total trajectory x(t)=(Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,)) in A, both Sh(t) and Sv(t) are positive and either ρ(x(t))0 or ρ(x(t)) is positive on R.

Proof.

For the first part, suppose by way of contradiction, there exists some t0R such that Sh(t0)=0. Then dSh(t0)dt=λh>0. This implies that Sh(t)<0 for t<t0 sufficiently small, a contradiction to x(t)AΩ. Thus Sh(t) is positive. Similarly, we can show that Sv(t) is positive.

Now, suppose ρ(x(t))0. We claim that ρ(x(t))>0 for all tR. Otherwise, with the assistance of Lemma 4.9, there exists a sequence {tn} such that tn as n and ρ(x(t))>0 for all n. For every n, define xn:RA by xn(t)=x(t+tn). Then using similar arguments as those in the proof of Lemma 4.7, there exists a T0 (independent of n) such that ρ(xn(t))>0 for tT and all n. This is a contradiction and hence the proof is completed.

By now we have verified all the assumptions of Theorem 5.2 [Citation51]. Therefore, we have the following result.

Theorem 4.11

Suppose R0>1. Then there exists a global compact attractor A1 in X+X0 and the semi-flow Φ is uniformly ρ-persistent.

Corollary 4.12

Suppose R0>1. Then there is an ε0 such that Sh(t), Eh(t), ih(t,0), Sv(t),Ev(t), iv(t,0)ε0 for all tR, where (Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,)) is any total trajectory in A1.

Proof.

It follows from the first equation of (Equation2) and (C4) that dSh(t)dtλhμhSh(t)(Lu+Ku)Sh(t), which implies lim inftSh(t)λhμh+Lu+Kuεh1. Similarly, lim inftSv(t)λvμv+Muεv1. By invariance, Sh(t)εh1 and Sv(t)εv1.

From Theorem 4.11, we know there is positive ε~ such that 0βh(a)ih(t,a)da+0βv(a)iv(t,a)daε~ By the concavity in (C3), we have dEh(t)dtϕ(εh1,Jh(t))+ψ(εh1,Jv(t))(αh+μh)Eh(t)ϕ(εh1,βhλhμh)JhJh(t)+ψ(εh1,βvλvμv)JvJv(t)(αh+μh)Eh(t)Aε~(αh+μh)Eh(t), which implies lim inftEh(t)Aε~αh+μhεh2, where A=minϕ(εh1,βhλhμh)Jh,ψ(εh1,βvλvμv)Jv.

By invariance again, Eh(t)εh2 for tR. It follows that ih(t,0)=αhEh(t)αhεh2εh3for {tR}. Thirdly, since J(t)=0kh(a)ih(t,a)da=0kh(a)ih(ta,0)σh(a)daηεh3 for tR, it follows from dEv(t)dtf(εv1,ηεh3)(αv+μv)Ev(t) for tR that lim inftEv(t)f(εv1,ηεh3)αv+μvεv2. Again by invariance, Ev(t)εv2 for tR and hence iv(t,0)=αvEv(t)αvεv2εv3for {tR}. Letting ε0=min{εh1,εh2,εh3,εv1,εv2,εv3}, we immediately complete the proof.

Now, we are ready to establish the global stability of the infectious steady state E under the following additional assumption on the nonlinear incidences.

(C5)

For Sh>0 and Sv>0, JhJhShϕ(Sh,Jh)Shϕ(Sh,Jh)1,0<JhJh,1Shϕ(Sh,Jh)Shϕ(Sh,Jh)JhJh,0<JhJh,JvJvShψ(Sh,Jv)Shψ(Sh,Jv)1,0<JvJv,1Shψ(Sh,Jv)Shψ(Sh,Jv)JvJv,0<JvJv,JJSvf(Sv,J)Svf(Sv,J)1,0<JJ,1Svf(Sv,J)Svf(Sv,J)JJ,0<JJ.

Theorem 4.13

Under the conditions R0>1 and (C1)(C5), the infectious steady state E of (Equation2) is globally asymptotically stable in X+X0.

Proof.

Since E is locally asymptotically stable, we only need to prove that A1={E}. The approach is the technique of Lyapunov functionals. Let x(t)=(Sh(t),Eh(t),ih(t,), Sv(t),Ev(t),iv(t,)) be a total trajectory in A1. From Corollary 4.12, we can find an ε~0 ensuring g(x)[0,ε~0], where x can be any of Sh(t)Sh, Eh(t)Eh, ih(t,a)ih(a), Sv(t)Sv, Ev(t)Ev and iv(t,a)iv(a). Here g:(0,)xx1lnxR. Clearly, g is nonnegative and g(x)=0 if and only if x = 1.

Define a Lyapunov functional by L(t)=L(Sh(t),Eh(t),ih(t,),Sv(t),Ev(t),iv(t,))=L1(t)+L2(t)+L3(t)+L4(t)+L5(t), where L1(t)=ShgSh(t)Sh+EhgEh(t)Eh,L2(t)=ψ(Sh,Jv)f(Sv,J)SvgSv(t)Sv+EvgEv(t)Ev,L3(t)=ψ(Sh,Jv)ηαhEh0η(a)ih(a)gih(t,a)ih(a)da,L4(t)=ϕ(Sh,Jh)ξhαhEh0ξh(a)ih(a)gih(t,a)ih(a)da,L5(t)=ψ(Sh,Jv)ξvαvEv0ξv(a)iv(a)giv(t,a)iv(a)da, with ξh(a)=aβh(θ)eaθδh(s)dsdθ,ξv(a)=aβv(θ)eaθδv(s)dsdθ,η(a)=akh(θ)eaθδh(s)dsdθ. Then L(t) is bounded on the solution x(t). In the following, we calculate the time derivatives of Li (i=1,2,,5).

With the assistance of λh=ϕ(Sh,Jh)+ψ(Sh,Jv)+μhSh and αh+μh=ϕ(Sh,Jh)+ψ(Sh,Jv)Eh, we have dL1(t)dt=1ShSh(t)(λhϕ(Sh(t),Jh(t))ψ(Sh(t),Jv(t))μhSh(t))+1EhEh(t)(ϕ(Sh(t),Jh(t))+ψ(Sh(t),Jv(t))(αh+μh)Eh(t))=1ShSh(t)(μhShμhSh(t))+1ShSh(t)(ϕ(Sh,Jh)+ψ(Sh,Jv))+ShSh(t)(ϕ(Sh(t),Jh(t))+ψ(Sh(t),Jv(t)))Eh(t)Eh(ϕ(Sh,Jh)+ψ(Sh,Jv))EhEh(t)(ϕ(Sh(t),Jh(t))+ψ(Sh(t),Jv(t)))+ϕ(Sh,Jh)+ψ(Sh,Jv)=μhSh2ShSh(t)Sh(t)Sh(ϕ(Sh,Jh)+ψ(Sh,Jv))ShSh(t)1lnShSh(t)+ϕ(Sh,Jh)Shϕ(Sh(t),Jh(t))Sh(t)ϕ(Sh,Jh)1lnShϕ(Sh(t),Jh(t))Sh(t)ϕ(Sh,Jh)+ψ(Sh,Jv)Shψ(Sh(t),Jv(t))Sh(t)ψ(Sh,Jv)1lnShψ(Sh(t),Jv(t))Sh(t)ψ(Sh,Jv)(ϕ(Sh,Jh)+ψ(Sh,Jv))Eh(t)Eh1lnEh(t)Ehϕ(Sh,Jh)Ehϕ(Sh(t),Jh(t))Eh(t)ϕ(Sh,Jh)1lnEhϕ(Sh(t),Jh(t))Eh(t)ϕ(Sh,Jh)ψ(Sh,Jv)Ehψ(Sh(t),Jv(t))Eh(t)ψ(Sh,Jv)1lnEhψ(Sh(t),Jv(t))Eh(t)ψ(Sh,Jv)=μhSh2ShSh(t)Sh(t)Sh(ϕ(Sh,Jh)+ψ(Sh,Jv))gShSh(t)+ϕ(Sh,Jh)gShϕ(Sh(t),Jh(t))Sh(t)ϕ(Sh,Jh)+ψ(Sh,Jv)gShψ(Sh(t),Jv(t))Sh(t)ψ(Sh,Jv)(ϕ(Sh,Jh)+ψ(Sh,Jv))gEh(t)Ehϕ(Sh,Jh)gEhϕ(Sh(t),Jh(t))Eh(t)ϕ(Sh,Jh)ψ(Sh,Jv)gEhψ(Sh(t),Jv(t))Eh(t)ψ(Sh,Jv). Similarly, dL2(t)dt=ψ(Sh,Jv)f(Sv,J)μvSv2SvSv(t)Sv(t)Svψ(Sh,Jv)gSvSv(t)+ψ(Sh,Jv)gSvf(Sv(t),J(t))Sv(t)f(Sv,J)ψ(Sh,Jv)gEv(t)Evψ(Sh,Jv)gEvf(Sv(t),J(t))Ev(t)f(Sv,J).

Next, dL3(t)dt=ψ(Sh,Jv)ηαhEh0η(a)1ih(a)ih(t,a)ih(t,a)tda=ψ(Sh,Jv)ηαhEh0η(a)1ih(a)ih(t,a)ih(t,a)aδh(a)ih(t,a)da. Note that ih(a)agih(t,a)ih(a)=ih(a)agih(ta,0)ih(0)=ih(a)tgih(ta,0)ih(0)=ih(a)1ih(0)ih(ta,0)tih(ta,0)ih(0)=ih(a)1ih(a)ih(t,a)tih(t,a)ih(a)=1ih(a)ih(t,a)ih(t,a)a+δh(a)ih(t,a). Using η(0)=η, ih(0)=αhEh, and ih(t,0)=αhEh(t), we get dL3(t)dt=ψ(Sh,Jv)ηαhEh0η(a)ih(a)agih(t,a)ih(a)da=ψ(Sh,Jv)ηαhEhη(a)ih(a)gih(t,a)ih(a)|a=0a=+ψ(Sh,Jv)ηαhEh0gih(t,a)ih(a)(η(a)δh(a)η(a))ih(a)da=ψ(Sh,Jv)gEh(t)Ehψ(Sh,Jv)ηαhEh0kh(a)ih(a)gih(t,a)ih(a)da. Similarly, we obtain dL4(t)dt=ϕ(Sh,Jh)gEh(t)Ehϕ(Sh,Jh)ξhαhEh0βh(a)ih(a)gih(t,a)ih(a)da, and dL5(t)dt=ψ(Sh,Jv)gEv(t)Evψ(Sh,Jv)ξvαvEv0βv(a)iv(a)giv(t,a)iv(a)da. To sum up, we have achieved that dL(t)dt=μhSh2ShSh(t)Sh(t)Sh(ϕ(Sh,Jh)+ψ(Sh,Jv))gShSh(t)+ϕ(Sh,Jh)gShϕ(Sh(t),Jh(t))Sh(t)ϕ(Sh,Jh)+ψ(Sh,Jv)gShψ(Sh(t),Jv(t))Sh(t)ψ(Sh,Jv)ϕ(Sh,Jh)gEhϕ(Sh(t),Jh(t))Eh(t)ϕ(Sh,Jh)ψ(Sh,Jv)gEhψ(Sh(t),Jv(t))Eh(t)ψ(Sh,Jv)ψ(Sh,Jv)ηαhEh0kh(a)ih(a)gih(t,a)ih(a)daϕ(Sh,Jh)ξhαhEh0βh(a)ih(a)gih(t,a)ih(a)daψ(Sh,Jv)ξvαvEv0βv(a)iv(a)giv(t,a)iv(a)da+ψ(Sh,Jv)f(Sv,J)μvSv2SvSv(t)Sv(t)Svψ(Sh,Jv)gSvSv(t)+ψ(Sh,Jv)gSvf(Sv(t),J(t))Sv(t)f(Sv,J)ψ(Sh,Jv)gEvf(Sv(t),J(t))Ev(t)f(Sv,J). It follows from the monotonicity of g, (C5), and the Jensen Inequality that ϕ(Sh,Jh)gShϕ(Sh(t),Jh(t))Sh(t)ϕ(Sh,Jh)ϕ(Sh,Jh)gJh(t)Jh=ϕ(Sh,Jh)g0βh(a)ih(t,a)da0βh(a)ih(a)da=ϕ(Sh,Jh)0βh(a)ih(a)da0βh(a)ih(a)g0βh(a)ih(a)ih(t,a)ih(a)da0βh(a)ih(a)dadaϕ(Sh,Jh)ξhαhEh0βh(a)ih(a)gih(t,a)ih(a)da, and similarly ψ(Sh,Jv)gShψ(Sh(t),Jv(t))Sh(t)ψ(Sh,Jv)ψ(Sh,Jv)ξvαvEv0βv(a)iv(a)giv(t,a)iv(a)da,ψ(Sh,Jv)gSvf(Sv(t),J(t))Sv(t)f(Sv,J)ψ(Sh,Jv)ηαhEh0kh(a)ih(a)gih(t,a)ih(a)da. Therefore, dL(t)dt0. Because the Lyapunov functional L is bounded on x(t), we conclude that the alpha limit of x() is in the largest invariant subset M in {dL(t)dt=0}. However, M is exactly {E}. Therefore, L(x(t))L(E) and A1={E}. This completes the proof.

5. A model with standard incidence rates

If the total populations of hosts Nh(t) and vectors Nv(t) are not constants, then the standard incidence rate does not satisfy the condition (C5). In this section, we study the following system with standard incidence rates (26) dSh(t)dt=λhμhSh(t)Sh(t)Jh(t)Nh(t)Sh(t)Jv(t)Nh(t),dEh(t)dt=Sh(t)Jh(t)Nh(t)+Sh(t)Jv(t)Nh(t)(αh+μh)Eh(t),ih(t,a)t+ih(t,a)a=δh(a)ih(t,a),ih(t,0)=αhEh(t),dRh(t)dt=0γh(a)ih(t,a)daμhRh(t),dSv(t)dt=λvSv(t)J(t)Nh(t)μvSv(t),dEv(t)dt=Sv(t)J(t)Nh(t)(αv+μv)Ev(t),iv(t,a)t+iv(t,a)a=δv(a)iv(t,a),iv(t,0)=αvEv(t),Sh(0)=Sh0R+,Eh(0)=Eh0R+,ih(0,)=ih0L+1(0,),Rh(0)=Rh0R+,Sv(0)=Sv0R+,Ev(0)=Ev0R+,iv(0,)=iv0L+1(0,).(26) The variables and parameters in the above model are the same as those in system (Equation2). The total host population is Nh(t)=Sh(t)+Eh(t)+0ih(t,a)da+Rh(t).

System (Equation26) always has the disease-free steady state E0~=(Sh0~,0,0,0,Sv0~,0,0), where Sh0~=Sh0=λhμh and Sv0~=Sv0=λvμv. The basic reproduction number R~0 is give by R~0=ξhαhαh+μh+αhηαh+μhξvαvαv+μvSv0Nh0, where Nh0=Sh0.

If E~=(Sh~,Eh~,ih~,Rh~,Sv~,Ev~,iv~) is an infectious steady state, then Sh~=λh(αh+μh)Eh~μh,Rh~=γαhEh~μh,γ=0γh(a)σh(a)da,σ=0σh(a)da,ih~(a)=σh(a)αhEh~,Sv~=λvμv+ηαhEh~Nh~,Nh~=λhμh+γ1+σμhμhαhEh~,Ev~=Sv~J~Nh~(αv+μv)=λvμv+ηαhEh~Nh~ηαhEh~(αv+μv)Nh~,iv~(a)=σv(a)αvEv~, and Eh~ is a positive root of the following quadratic equation (27) H(Eh~)=mEh~2+nEh~+q=0,(27) where m=(αh+μh)μhμv(m12+ξhαhμh+ηαhμvm1+ξhαhμhηαhμv),m1=(γ1+σμhμh)αh,n=λhξhαh2η(αh+μh)λhμhμvξhαh+ξvαvαv+μvηαhSv0Nh0+ηαhμhμv+(λhξhαhμv2(αh+μh)μhμvλhμh)(γ1+σμhμh)αh,q=(αh+μh)λh2μhμv(R~01).

Using the relationship between roots and coefficients of the quadratic equation H(Eh~)=0, given in  (Equation27), we list all the possibilities for the number of positive roots of H(Eh~)=0 in Table . From all the cases enumerated in Table , we have the following results: The system (Equation26) has a unique infectious steady state Eh~ if either (i) R~0>1, m<0 or (ii) R~0<1, m>0 is satisfied. The system (Equation26) could have two infectious steady states if either (iii) R~0>1, m>0, n<0, n24mq>0 or (iv) R~0<1, m<0, n>0, n24mq>0 is satisfied. Therefore, it is possible for the system with standard incidence rates to have backward bifurcation.

Table 1. Number of positive zeros of H given in  (Equation27).

Theorem 5.1

When R~0<1, the disease-free steady state E0~ of (Equation26) is locally asymptotically stable. When R~0>1, it is unstable.

Proof.

Similar to the proof of Theorem 3.2, the characteristic equation at E0~ is Δ(E0~)=(τ+μh)2(τ+μv)(τ+αh+μh)(τ+αv+μv)g2(τ)=0, where g2(τ)=1αhξhˆ(τ)τ+αh+μhαhξvˆ(τ)αvηˆ(τ)Sv0(τ+αh+μh)(τ+αv+μv)Nh0. The roots of g2(τ)=0 determine the stability of E0~. If R~0>1, then g2(0)=1R~0<0, limτg2(τ)=1. Therefore, g2(τ)=0 admits a positive root, which means that E0~ is unstable. If R~0<1, we show that all roots of g2(τ)=0 have negative real parts. By contradiction, we suppose that τ0 is a root with non-negative real part. Then 1=αhξhˆ(τ0)τ0+αh+μh+αhξvˆ(τ0)αvηˆ(τ0)Sv0~(τ0+αh+μh)(τ0+αv+μv)Nh0αhξhαh+μh+αhξvαvηSv0(αh+μh)(αv+μv)Nh0=R~0, which yields a contradiction. This completes the proof.

When an infectious steady state E~ exists, the characteristic equation evaluated at E~ is Δ(E)=(τ+μh)(τ4+a1τ3+a2τ2+a3τ+a4)=0, where a1=αv+2μv+JNh+2μh+αh+A+Λ1αhB,a2=(αv+μv)μv+JNh+(2μh+αh+A)αv+2μv+JNh+μh(μh+αh+A)+2μv+JNh+αv+αh+μhΛ1αhB2μv+JNh+αv+μh+Aαhγˆ(τ)αhαvCD2,a3=(2μh+αh+A)(αv+μv)μv+JNh+μh(μh+αh+A)αv+2μv+JNh+Λ1μv+JNh(αv+μv+αh+μh)+(αh+μh)(αv+μv)Λ1αvCD(αh+μh+μv)+CDαv(μv+μh)αhαvCD2(μv+μh)αhB(αv+μv+μh)(μv+JNh)+μh(αv+μv)+Aαhγˆ(τ)αv+2μv+JNh+αhαvCDγˆ(τ),a4=μh(μh+αh+A)(αv+μv)μv+JNh+CDαvμhμv+Λ1(αh+μh)(αv+μv)μv+JNhαvCD(αh+μh)μvαhBμh(αv+μv)μv+JNh+Aαhγˆ(τ)μv+JNh(αv+μv)+αhαvCDγˆ(τ)μvαhαvCD2μhμv, and A=Sh(Jh+Jv)Nh2,Λ1=(NhSh)(Jh+Jv)Nh2,B=Shξˆh(τ)NhShJhσˆ(τ)Nh2,C=Shξˆv(τ)Nh,D=SvJNh2,D2=Shηˆ(τ)NhDσˆ(τ),γˆ(τ)=0γh(a)σh(a)eτada,σˆ(τ)=0σh(a)eτada.ξˆh(τ), ξˆv(τ), and ηˆ(τ) are the same as previous definitions. The analysis of the distribution of the zeros of this complicated transcendental equation is challenging and remains to be further investigated.

6. Conclusion and discussion

Infection age is an important factor in studying the transmission of vector-borne diseases such as malaria. We included the age of infection for both infectious hosts and vectors in a mathematical model that studies both horizontal host transmission and direct vector transmission. Because of the extra disease-induced death, the total population of hosts or vectors does not converge to a constant. We assumed that incidences of horizontal transmission, transmission from vectors to hosts, and transmission from hosts to vectors are all in general nonlinear forms. Because the model is an infinite dimensional system consisting of coupled ordinary differential equations and partial differential equations, the analysis is not trivial. Using the Fluctuation Lemma, the approach of Lyapunov functionals, and some techniques in [Citation38,Citation42], we have established the threshold dynamics, which are completely determined by the basic reproduction number R0. Specifically, the disease-free steady state is globally asymptotically stable if R0<1 while the infectious steady state is globally asymptotically stable if R0>1. In other words, if R0>1 and there is initial infection force for the hosts then the solution of the system converges to the infectious steady state.

From the expression of R0, R0=ϕ(Sh0,0)Jhξhαhαh+μh+ψ(Sh0,0)Jvf(Sv0,0)Jαhηαh+μhξvαvαv+μv, we see that R0 is the number of infectious hosts generated by the introduction of an infectious host into a population consisting of only susceptible hosts, which agrees with the biological interpretation of the basic reproduction number. The first term in R0 is the number generated by horizontal transmission while the second term is the number generated through vectors. Ignoring the horizontal transmission would underestimate the value of R0, which may result in suboptimal and unsuccessful treatment.

Transmission is a key process in the host-vector interaction. In many models, the transmission is described by the mass action law. Given the number of susceptible hosts Sh(t) and number of infected hosts Ih(t) (or vectors Iv(t)), the number of new infected hosts per unit of time caused by horizontal transmission is βhSh(t)Ih(t) (or by vector transmission is βvSh(t)Iv(t)), where βh (or βv) is the transmission rate. Transmission depends on the number of susceptible individuals with which an infectious individual might interact. This implies that the number of susceptible individuals in a ‘neighborhood’ of an infectious individual is important, rather than the total number of susceptible individuals in the population. McCallum et al. [Citation41] suggested the use of mass action might not be accurate in some scenarios and that other functions could be used to describe the transmission. In our paper, using ϕ(Sh,Jh) where Jh=0βh(a)ih(t,a)da as an example, the transmission can be used to describe the Holling Type II functional response βShJh1+αJh [Citation47] and kShln(1+βJhk) [Citation41]. These functions, as well as the bilinear mass action term, are special cases of our nonlinear incidence rates satisfying (C1)–(C5).

Another popular form used to describe infection is the standard incidence rate. When the total population is not constant, the standard incidence rate does not belong to our previous class of incidence rates. Therefore, in Section 5, we studied the corresponding model with standard incidences. In this case, the horizontal transmission between hosts is Sh(t)0βh(a)ih(t,a)daNh(t), the transmission from vectors to hosts is Sh(t)0βv(a)iv(t,a)daNh(t), and the transmission to susceptible vectors because of biting infectious hosts is given by Sv(t)0kh(a)ih(t,a)daNh(t), where Nh(t)=Sh(t)+Eh(t)+0ih(t,a)da+Rh(t). For this model, we listed the conditions for the existence of one or more infectious steady states in Table . The disease-free steady sate always exists and when the basic reproduction number R~0 is less than 1 it is locally asymptotically stable. In addition to the disease-free steady state, one or two infectious steady states may also exist and be stable when R~0<1. Thus, backward bifurcation may take place. However, at this moment we are not able to carry out the stability analysis of the infectious steady state due to the complexity of the transcendental characteristic equation.

For models described by ordinary differential equations, lots of studies have indicated that standard incidence rate coupled with other factors can generate bifurcations. For example, Garba et al. [Citation19] investigated a deterministic model with a standard incidence rate for the transmission dynamics of dengue disease. They showed that the use of standard incidence can generate the backward bifurcation phenomenon of dengue disease. Roop-O et al. [Citation47] studied the effect of the choice of the incidence function for the occurrence of backward bifurcation in a malaria model. They found that three methods may eliminate the backward bifurcation: (i) reducing the reproductive number to below a sub-threshold, (ii) the rate of malaria-induced mortality in humans is zero, (iii) the incidence function takes a nonlinear incidence function rather than the standard incidence rate. In a recent paper by Cai et al. [Citation8], the authors also found that the backward bifurcation can be generated in an ordinary differential equation malaria model by a standard incidence rate coupled with a nonzero disease-induced mortality rate or superinfection (i.e., asymptomatic individuals can be reinfected and move to the symptomatic class). Cai et al. [Citation7] proposed delayed vector-host disease models with two time delays, one describing the incubation period in the vector population and the other representing the incubation period in the host population. They also assumed that the total populations of both host Nh and vector Nv are constants. Then the standard incidence rate can be reduced to the bilinear incidence rate, which satisfies our conditions (C1)(C5). Their analytic results revealed that the global dynamics of such vector-host disease models with time delays are completely determined by the basic reproduction number by constructing suitable Liapunov functionals. Kribs-Zaleta and Martcheva [Citation30] considered partial differential equation models for a disease with acute and chronic infective stages, in which the standard incidence rate was also used. Their models for SIRS and SIS disease cycles without vectors exhibit backward bifurcations under certain conditions.

Besides a more detailed description of the dynamics of model (Equation26), future work may include models with the physiological heterogeneity in susceptibility. Highly susceptible individuals tend to acquire infection first, while resistant individuals acquire infection later and at a slower rate. This may yield a nonlinear relation between time and number of new infections acquired [Citation41]. Therefore, different transmission modes should be evaluated in studying vector-borne disease dynamics by mathematical models.

Acknowledgments

The authors would like to thank the editors and the anonymous reviewers for their very valuable comments and suggestions on improving the presentation of the paper.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

The research of Wang was supported by NSFC (No. 11771374), the Nanhu Scholar Program for Young Scholars of Xinyang Normal University. Chen was supported by NSERC of Canada. Martcheva and Rong were supported by the NSF grant DMS-1515661 and DMS-1758290, respectively.

References

  • M.Z. Ahmed, S.J. Li, X. Xue, X.J. Yin, S.X. Ren, F.M. Jiggins, J.M. Greeff, and B.L. Qiu, The intracellular bacterium Wolbachia uses parasitoid wasps as phoretic vectors for efficient horizontal transmission, PLoS Pathog. 11(2) (2015), pp. e1004672.
  • E. Avila-Vales and B. Buonomo, Analysis of a mosquito-borne disease transmission model with vector stages and nonlinear forces of infection, Ric. Mat. 64(2) (2015), pp. 377–390.
  • S.E. Bellan, The importance of age dependent mortality and the extrinsic incubation period in models of mosquito-borne disease transmission and control, PLoS One 5(4) (2010), pp. e10165.
  • S. Bentout, Y. Chen, and S. Djilali, Global dynamics of an SEIR model with two age structures and a nonlinear incidence, preprint.
  • C.J. Briggs and H.C.J. Godfray, The dynamics of insect-pathogen interactions in stage-structured populations, Am. Nat. 145 (1995), pp. 855–887.
  • C.J. Browne and S.S. Pilyugin, Global analysis of age-structured within-host virus model, Discrete Cont. Dyn. Syst. Ser. B 18(8) (2013), pp. 1999–2017.
  • L.M. Cai, X.Z. Li, B. Fang, and S. Ruan, Global properties of vector-host disease models with time delays, J. Math. Biol. 74(6) (2017), pp. 1397–1423.
  • L. Cai, X. Li, N. Tuncer, M. Martcheva, and A. Ali Lashari, Optimal control of a malaria model with asymptomatic class and superinfection, Math. Biosci. 288 (2017), pp. 94–108.
  • C. Caminade, M.K. McIntyre, and A.E. Jones, Climate change and vector-borne diseases: where are we next heading? J. Infect. Dis. 214(9) (2016), pp. 1300–1301.
  • V. Capasso and G. Serio, A generalization of the Kermack-McKendrick deterministic epidemic model, Math. Biosci. 42 (1978), pp. 41–61.
  • V. Capasso, E. Grosso, and G. Serio, I modelli matematici nella indagine epidemiologica. Applicazione all'epidemia di colera verificatasi in Bari nel 1973 (italian), Annali Sciavo 19 (1977), pp. 193–208.
  • Y. Chen, S. Zou, and J. Yang, Global analysis of an SIR epidemic model with infection age and saturated incidence, Nonlinear Anal. Real World Appl. 30 (2016), pp. 16–31.
  • Y.X. Dang, Z.P. Qiu, X.Z. Li, and M. Martcheva, Global dynamics of a vector-host epidemic model with age of infection, Math. Biosci. Eng. 14(5 & 6) (2017), pp. 1159–1186.
  • K. Dietz, L. Molineaux, and A. Thomas, A malaria model tested in the African Savannah, Bull. World Health Organ. 50(3-4) (1974), pp. 347.
  • S.H. Faeth, K.P. Hadeler, and H.R. Thieme, An apparent paradox of horizontal and vertical disease transmission, J. Biol. Dyn. 1 (2007), pp. 45–62.
  • Z. Feng and J.X. Velasco-Hernändez, Competitive exclusion in a vector-host model for the dengue fever, J. Math. Biol. 35(5) (1997), pp. 523–544.
  • X. Feng, S. Ruan, Z. Teng, and K. Wang, Stability and backward bifurcation in a malaria transmission model with applications to the control of malaria in China, Math. Biosci. 266 (2015), pp. 52–64.
  • F. Forouzannia and A.B. Gumel, Mathematical analysis of an age-structured model for malaria transmission dynamics, Math. Biosci. 247(2) (2014), pp. 80–94.
  • S.M. Garba, A.B. Gumel, and M.R.A. Bakar, Backward bifurcations in dengue transmission dynamics, Math. Biosci. 215(1) (2008), pp. 11–25.
  • I.N. Gavrilovskaya, N.S. Apekina and A.D. Bernshtein, V.T. Demina, N.M. Okulova, Y.A. Myasnikov, and M.P. Chumakov, Pathogenesis of hemorrhagic fever with renal syndrome virus infection and mode of horizontal transmission of hantavirus in bank voles, in Hemorrhagic Fever with Renal Syndrome, Tick-and Mosquito-Borne Viruses, Springer, Vienna, 1990, pp. 57–62.
  • P. Georgescu and Y.H. Hsieh, Global stability for a virus dynamics model with nonlinear incidence of infection and removal, SIAM J. Appl. Math. 67 (2006), pp. 337–353.
  • J.K. Hale, Asymptotic Behavior of Dissipative Systems, American Mathematical Society, Providence, RI, 1988.
  • H.W. Hethcote, The mathematics of infectious diseases, SIAM Rev. 42(4) (2000), pp. 599–653.
  • W.M. Hirsch, H. Hanisch, and J.-P. Gabriel, Differential equation models of some parasitic infections: Methods for the study of asymptotic behavior, Commun. Pure Appl. Math. 38(6) (1985), pp. 733–753.
  • M. Iannelli, Applied Mathematics Monographs 7. Mathematical Theory of Age-Structured Population Dynamics, Comitato Nazionale per le Scienze Matematiche, Consiglio Nazionale delle Ricerche (CNR), Giardini, Pisa, 1995.
  • H. Inaba and H. Sekine, A mathematical model for chagas disease with infection-age-dependent infectivity, Math. Biosci. 190(1) (2004), pp. 39–69.
  • A. Korobeinikov, Global properties of infectious disease models with nonlinear incidence, Bull. Math. Biol. 69 (2007), pp. 1871–1886.
  • A. Korobeinikov, Global asymptotic properties of virus dynamics models with dose-dependent parasite reproduction and virulence and nonlinear incidence rate, Math. Med. Biol. 26 (2009), pp. 225–239.
  • A. Korobeinikov and P.K. Maini, Non-linear incidence and stability of infectious disease models, Math. Med. Biol. J. IMA 22(2) (2005), pp. 113–128.
  • C.M. Kribs-Zaleta and M. Martcheva, Vaccination strategies and backward bifurcation in an age-since-infection structured model, Math. Biosci. 177 (2002), pp. 317–332.
  • A.A. Lashari and G. Zaman, Global dynamics of vector-borne diseases with horizontal transmission in host population, Comput. Math. Appl. 61(4) (2011), pp. 745–754.
  • A.A. Lashari and G. Zaman, Optimal control of a vector borne disease with horizontal transmission, Nonlinear Anal. Real World Appl. 13(1) (2012), pp. 203–212.
  • K.S. Lee and A.A. Lashari, Stability analysis and optimal control of pine wilt disease with horizontal transmission in vector population, Appl. Math. Comput. 226 (2014), pp. 793–804.
  • W. Liu, S.A. Levin, and Y. Iwasa, Influence of nonlinear incidence rates upon the behavior of SIRS epidemiological models, J. Math. Biol. 23(2) (1986), pp. 187–204.
  • W. Liu, H.W. Hethcote, and S.A. Levin, Dynamical behavior of epidemiological models with nonlinear incidence rates, J. Math. Biol. 25(4) (1987), pp. 359–380.
  • H. Liu, J. Yu, and G. Zhu, Analysis of a vector-host malaria model with impulsive effect and infection-age, Adv. Complex Syst. 9(3) (2006), pp. 237–248.
  • P. Magal, Compact attractors for time periodic age-structured population models, Electron. J. Differ. Eq. 65 (2001), pp. 1–35.
  • P. Magal, C.C. McCluskey, and G.F. Webb, Lyapunov functional and global asymptotic stability for an infection-age model, Appl. Anal. 89(7) (2010), pp. 1109–1140.
  • S. Mandal, R.R. Sarkar, and S. Sinha, Mathematical models of malaria-a review, Malaria J. 10(1) (2011), pp. 202.
  • M. Martcheva and H.R. Thieme, Progression age enhanced backward bifurcation in an epidemic model with super-infection, J. Math. Biol. 46(5) (2003), pp. 385–424.
  • H. McCallum, N. Barlow, and J. Hone, How should pathogen transmission be modelled? Trends Ecol. Evol. 16(6) (2001), pp. 295-–300.
  • A.V. Melnik and A. Korobeinikov, Lyapunov functions and global stability for SIR and SEIR models with age-dependent susceptibility, Math. Biosci. Eng. 10(2) (2013), pp. 369–378.
  • G.A. Ngwa and W.S. Shu, A mathematical model for endemic malaria with variable human and mosquito populations, Math. Comput. Model. 32 (2000), pp. 747–763.
  • V.N. Novoseltsev, A.I. Michalski, J.A. Novoseltseva, A.I. Yashin, J.R. Carey, and A.M. Ellis, An age-structured extension to the vectorial capacity model, PloS One 7(6) (2012), pp. e39479.
  • R. Pigeault, A. Nicot, S. Gandon, and A. Rivero, Mosquito age and avian malaria infection, Malaria J. 14(1) (2015), pp. 383, 11pp.
  • Z. Qiu, Dynamical behavior of a vector-host epidemic model with demographic structure, Comput. Math. Appl. 56(12) (2008), pp. 3118–3129.
  • O.P. Roop, W. Chinviriyasit, and S. Chinviriyasit, The effect of incidence function in backward bifurcation for malaria model with temporary immunity, Math. Biosci. 265 (2015), pp. 47–64.
  • S. Ruan, D. Xiao, and J.C. Beier, On the delayed Ross-Macdonald model for malaria transmission, Bull. Math. Biol. 70(4) (2008), pp. 1098–1114.
  • A. Saul, Transmission dynamics of Plasmodium falciparum, Parasitol. Today 12 (1996), pp. 74–79.
  • G.A. Schmunis, F. Zicker, F. Pinheiro, and D. Brandling-Bennett, Risk for transfusion-transmitted infectious diseases in Central and South America, Emerg. Inefect. Dis. 4 (1998), pp. 5–11.
  • H.L. Smith and H.R. Thieme, Dynamical Systems and Population Persistence, American Mathematical Society, Providence, RI, 2011.
  • J. Tumwiine, J.Y.T. Mugisha, and L.S. Luboobi, A mathematical model for the dynamics of malaria in a human host and mosquito vector with temporary immunity, Appl. Math. Comput. 189(2) (2007), pp. 1953–1965.
  • C. Vargas-De-León, Global analysis of a delayed vector-bias model for malaria transmission with incubation period in mosquitoes, Math. Biosci. Eng. 9(1) (2012), pp. 165–174.
  • C. Vargas-De-León, L. Esteva, and A. Korobeinikov, Age-dependency in host-vector models: The global analysis, Appl. Math. Comput. 243 (2014), pp. 969–981.
  • K. Vogt-Geisse, C. Lorenzo, and Z. Feng, Impact of age-dependent relapse and immunity on Malaria dynamics, J. Biol. Syst. 21(4) (2013), pp. 1340001, 49pp.
  • X. Wang, Y. Lou, and X. Song, Age-structured within-host HIV dynamics with multiple target cells, Stud. Appl. Math. 138(1) (2017), pp. 43–76.
  • X. Wang, Y. Chen, and S. Liu, Global dynamics of a vector-borne disease model with infection ages and general incidence rates, Comput. Appl. Math. 37(4) (2018), pp. 4055–4080.
  • H.M. Wei, X.Z. Li, and M. Martcheva, An epidemic model of a vector-borne disease with direct transmission and time delay, J. Math. Anal. Appl. 342(2) (2008), pp. 895–908.
  • World Health Organization, Fact sheets about Vector-borne diseases. February 2016. Available at http://www.who.int/mediacentre/factsheets/fs387/en/.