Abstract

In this paper, a virus infection model with time delay and absorption is studied. By analyzing the corresponding characteristic equations, the local stability of each of feasible equilibria of the model is established. By using comparison arguments, it is shown that the infection free equilibrium is globally asymptotically stable when the basic reproduction ratio is less than unity. When the basic reproduction ratio is greater than unity, sufficient conditions are derived for the global stability of the virus-infected equilibrium. Numerical simulations are carried out to illustrate the theoretical results.

1. Introduction

Mathematical modelling has been proven to be valuable in studying the dynamics of virus infection, such as HIV/AIDS, HBV, and HCV. In recent years, great attention has been paid by many researchers to the pathogen infectious agent, or germ, which can cause disease or illness to its host (see, e.g., [13]). Based on the clinical experiment of chronic HBV carriers treated with various doses of lamivudine, Nowak and Bangham [4] and Bonhoeffer et al. [5] proposed a classical mathematical model describing the interaction between the susceptible host cells (𝑥), infected host cells (𝑦), and free virus particles (𝑣), which is formulated by the following differential equations: ̇̇𝑥(𝑡)=𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣(𝑡),̇𝑦(𝑡)=𝛽𝑥(𝑡)𝑣(𝑡)𝑎𝑦(𝑡),𝑣(𝑡)=𝑘𝑦(𝑡)𝑢𝑣(𝑡),(1.1) where hepatocytes are produced at a rate 𝜆, die at rate 𝑑𝑥, and become infected at rate 𝛽𝑥𝑣; infected hepatocytes are produced at rate 𝛽𝑥𝑣 and die at rate 𝑎𝑦; free viruses are produced from infected cells at rate 𝑘𝑦 and are removed at rate 𝑢𝑣. It is assumed that parameters 𝑎, 𝑑, 𝑘, 𝑢, 𝜆, and 𝛽 are positive constants. The basic reproductive ratio 0 was found in [4]. Furthermore, in [6], by constructing suitable Lyapunov function, Korobeinikov showed that the infection-free equilibrium is globally asymptotically stable if 0<1, and the virus-infected equilibrium is globally asymptotically stable if 0>1.

It is assumed in model (1.1) that the infection process follows the principle of mass action [7], namely, for each uninfected cell and free virus particle is assumed to be constant 𝛽 between the rate of infection. However, studies of parasitic infections have shown that the relationship between the dose and rate of infection is clearly nonlinear. In [8], a more general saturated infection rate 𝛽𝑥𝑣𝑝/(1+𝛼𝑣𝑞) was suggested, where 𝑝, 𝑞, and 𝛼 are positive constants.

The virus life cycle plays a crucial role in disease progression. The binding of a viral particle to a receptor on a target cell initiates a cascade of events that can ultimately lead to the target cell becoming productively infected, that is, producing new virus. In model (1.1), it is assumed that as soon as virus contacts a target cell, the cell begins producing virus. This is not biologically sensible. In reality, there is a time delay between initial viral entry into a cell and subsequent viral production [9]. There have been some works on virus infection model in the literature (see, e.g., [914]). In [10], Herz et al. examined the effect of including a constant delay in the source term for productively infected T-cells. In [11], Li and Ma considered the following more general HIV-1 infection model with saturation incidence rate and time delay: ̇𝑥(𝑡)=𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣(𝑡),1+𝑣(𝑡)̇𝑦(𝑡)=𝛽𝑥(𝑡𝜏)𝑣(𝑡𝜏)̇𝑣1+𝑣(𝑡𝜏)𝑎𝑦(𝑡),(𝑡)=𝑘𝑦(𝑡)𝑢𝑣(𝑡).(1.2) By analyzing the transcendental characteristic equations, sufficient conditions for the local asymptotic stability of the equilibria were studied, and by using Lyapunov-LaSalle invariance principal, the global asymptotic stability of the viral-free equilibrium was given.

We note that in (1.1) and (1.2), the loss of pathogens due to the absorption into uninfected cells are ignored. In reality, when a pathogen enters an uninfected cell, the number of pathogens in the blood decreases by one. This is called the absorption effect [7]. In [15], considering the erythrocytic cycle in the absence of an immunological response by the host, Anderson et al. presented the following model for malaria infection: ̇̇𝑥(𝑡)=𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣(𝑡),̇𝑦(𝑡)=𝛽𝑥(𝑡)𝑣(𝑡)𝑎𝑦(𝑡),𝑣(𝑡)=𝑘𝑦(𝑡)𝛽𝑥(𝑡)𝑣(𝑡)𝑢𝑣(𝑡),(1.3) where the term 𝛽𝑥𝑣 in the third equation of (1.3) represents the absorption effect.

Motivated by the works of Anderson et al. [15], Li and Ma [11], and Song and Neumann [8], in this paper, we study the following virus infection model with a time delay and absorption: ̇𝑥(𝑡)=𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣(𝑡),1+𝛼𝑣(𝑡)̇𝑦(𝑡)=𝛽𝑥(𝑡𝜏)𝑣(𝑡𝜏)̇1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)=𝑘𝑦(𝑡)𝛽𝑥(𝑡)𝑣(𝑡)1+𝛼𝑣(𝑡)𝑢𝑣(𝑡).(1.4) The initial conditions for system (1.4) take the form 𝑥(𝜃)=𝜙1(𝜃),𝑦(𝜃)=𝜙2(𝜃),𝑣(𝜃)=𝜙3𝜙(𝜃),1(𝜃)0,𝜙2(𝜃)0,𝜙3([],𝜙𝜃)0,𝜃𝜏,01(0)>0,𝜙2(0)>0,𝜙3(0)>0,(1.5) where (𝜙1(𝜃),𝜙2(𝜃),𝜙3(𝜃))𝐶([𝜏,0],3+0), here 3+0={(𝑥1,𝑥2,𝑥3)𝑥𝑖0,𝑖=1,2,3}.

It is easy to show that all solutions of system (1.4) with initial condition (1.5) is defined on [0,+) and remain positive for all 𝑡0.

The organization of this paper is as follows. In the next section, we introduce some notations and state several lemmas which will be essential to our proofs. In Section 3, by analyzing the corresponding characteristic equations, the local stability of each of nonnegative equilibria of system (1.4) is discussed. In Section 4, by using an iteration technique, we study the global stability of the uninfected equilibrium of system (1.4). By comparison arguments, we discuss the global stability of the virus-infected equilibrium of system (1.4). Numerical simulations are carried out in Section 5 to illustrate the main theoretical results. The paper ends with conclusions in Section 6.

2. Preliminaries

In this section, based on the work developed by Xu and Ma [16], we introduce some notations and state several results which will be useful in next section.

Let 𝑛+ be the cone of nonnegative vectors in 𝑛. If 𝑥,𝑦𝑛, we write 𝑥𝑦(𝑥<𝑦) if 𝑥𝑖𝑦𝑖(𝑥𝑖<𝑦𝑖) for 1𝑖𝑛. Let {𝑒1,𝑒2,,𝑒𝑛} denote the standard basis in 𝑛. Suppose, 𝑟0 and let 𝐶=𝐶([𝑟,0],𝑛) be the Banach space of continuous functions mapping the interval [𝑟,0] into 𝑛 with supremum norm. If 𝜙,𝜓𝐶, we write 𝜙𝜓(𝜙<𝜓) when the indicated inequality holds at each point of [𝑟,0]. Let 𝐶+={𝜙𝐶𝜙0}, and let denote the inclusion 𝑛𝐶([𝑟,0],𝑛) by 𝑥̂𝑥, ̂𝑥(𝜃)=𝑥, 𝜃[𝑟,0]. Denote the space of functions of bounded variation on [𝑟,0] by BV[𝑟,0]. If 𝑡0, 𝐴0 and 𝑥𝐶([𝑡0𝑟,𝑡0+𝐴],𝑛), then for any 𝑡[𝑡0,𝑡0+𝐴], we let 𝑥𝑡𝐶 be defined by 𝑥𝑡(𝜃)=𝑥(𝑡+𝜃), 𝑟𝜃0.

We now consider ̇𝑥(𝑡)=𝑓𝑡,𝑥𝑡.(2.1)

We assume throughout this section that 𝑓×𝐶𝑛 is continuous, 𝑓(𝑡,𝜙) is continuously differentiable in 𝜙,𝑓(𝑡+𝑇,𝜙)=𝑓(𝑡,𝜙) for all (𝑡,𝜙)×𝐶+, and some 𝑇>0. Then, by [17], there exists a unique solution of (2.1) through (𝑡0,𝜙) for 𝑡0, 𝜙𝐶+. This solution will be denoted by 𝑥(𝑡,𝑡0,𝜙) if we consider the solution in 𝑛, or by 𝑥𝑡(𝑡0,𝜙) if we work in the space 𝐶. Again, by [17], 𝑥(𝑡,𝑡0,𝜙)(𝑥𝑡(𝑡0,𝜙)) is continuously differentiable in 𝜙. In the following, the notation 𝑥𝑡0=𝜙 will be used as the condition of the initial data of (2.1), by which we mean that we consider the solution 𝑥(𝑡) of (2.1) which satisfies 𝑥(𝑡0+𝜃)=𝜙(𝜃), 𝜃[𝑟,0].

To proceed further, we need the following results. Let 𝑟=(𝑟1,𝑟2,,𝑟𝑛)𝑛+, |𝑟|=max𝑖{𝑟𝑖}, and define 𝐶𝑟=𝑛𝑖=1𝐶𝑟𝑖.,0,(2.2)

We write 𝜙=(𝜙1,𝜙2,,𝜙𝑛) for a generic point of 𝐶𝑟. Let 𝐶+𝑟={𝜙𝐶𝑟𝜙0}. Due to the ecological applications, we choose 𝐶+𝑟 as the state space of (2.1) in the following discussions.

Fix 𝜙0𝐶+𝑟 arbitrarily. Then, we set L(𝑡,)=𝐷𝜙0𝑓(𝑡,𝜙0),𝐷𝜙0𝑓(𝑡,𝜙0) denotes the Frechet derivation of 𝑓 with respect to 𝜙0. It is convenient to have the standard representation of 𝐿=(𝐿1,𝐿2,,𝐿𝑛) as 𝐿𝑖(𝑡,𝜙)=𝑛𝑗=10𝑟𝑗𝜙𝑗(𝜃)𝑑𝜃𝜂𝑖𝑗(𝜃,𝑡),1𝑖𝑛,(2.3) in which 𝜂𝑖𝑗× satisfies𝜂𝑖𝑗(𝜃,𝑡)=𝜂𝑖𝑗𝜂(0,𝑡),𝜃0,𝑖𝑗(𝜃,𝑡)=0,𝜃𝑟𝑗,𝜂𝑖𝑗(,𝑡)BV𝑟𝑗,,0(2.4) where 𝜂𝑖𝑗(,𝑡) is continuous from the left in (𝑟𝑗,0).

We make the following assumptions for (2.1).(h0)If 𝜙,𝜓𝐶+, 𝜙𝜓, and 𝜙𝑖(0)=𝜓𝑖(0) for some 𝑖, then 𝑓𝑖(𝑡,𝜙)𝑓𝑖(𝑡,𝜓). (h1)For all 𝜙𝐶+𝑟 with 𝜙𝑖(0)=0, 𝐿𝑖(𝑡,𝜙)0 for 𝑡. (h2)The matrix 𝐴(𝑡) defined by 𝐴𝐿(𝑡)=col𝑡,̂𝑒1,𝐿𝑡,̂𝑒2,,𝐿𝑡,̂𝑒𝑛=𝜂𝑖𝑗(0,𝑡)(2.5) is irreducible for each 𝑡. (h3)For each 𝑗, for which 𝑟𝑗>0, there exists 𝑖 such that for all 𝑡 and for positive constant 𝜀 sufficiently small, 𝜂𝑖𝑗(𝑟𝑗+𝜀,𝑡)>0.(h4)If 𝜙=0, then 𝑥(𝑡,𝑡0,𝜙)0 for all 𝑡𝑡0.

The following result was established by Wang et al. [18].

Lemma 2.1. Let (h1)(h4) hold. Then, the hypothesis (h0) is valid (i)If 𝜙 and 𝜓 are distinct elements of 𝐶+𝑟 with 𝜙𝜓 and [𝑡0,𝑡0+𝜎) with 𝑛|𝑟|<𝜎 is the intersection of the maximal intervals of existence of 𝑥(𝑡,𝑡0,𝜙) and 𝑥(𝑡,𝑡0,𝜓), then 𝑥0𝑡,𝑡0,𝜙𝑥𝑡,𝑡0,𝜓for𝑡0𝑡<𝑡0𝑥+𝜎,𝑡,𝑡0,𝜙<𝑥𝑡,𝑡0,𝜓for𝑡0+𝑛|𝑟|𝑡<𝑡0+𝜎.(2.6)(ii)If 𝜙𝐶+𝑟, 𝜙0, 𝑡0 and 𝑥(𝑡,𝑡0,𝜙) is defined on [𝑡0,𝑡0+𝜎) with 𝜎>𝑛|𝑟|, then 0<𝑥𝑡,𝑡0,𝜙for𝑡0+𝑛|𝑟|𝑡<𝑡0+𝜎.(2.7)

This lemma shows that if (h1)(h4) hold, then the positivity of solutions of (2.1) follows.

The following definition and results are useful in proving our main result in this section.

Definition 2.2. Let 𝐴=(𝑎𝑖𝑗)𝑛×𝑛 be an 𝑛×𝑛 matrix, and let 𝑃1,,𝑃𝑛 be distinct points of the complex plane. For each nonzero element 𝑎𝑖𝑗 of 𝐴, connect 𝑃𝑖 to 𝑃𝑗 with a directed line 𝑃𝑖𝑃𝑗. The resulting figure in the complex plane is a directed graph for 𝐴. We say that a directed graph is strongly connected if, for each pair of nodes 𝑃𝑖,𝑃𝑗 with 𝑖𝑗, there is a directed path 𝑃𝑖𝑃𝑘1,𝑃𝑘1𝑃𝑘2𝑃,,𝑘𝑟1𝑃𝑗,(2.8) connecting 𝑃𝑖 and 𝑃𝑗. Here, the path consists of 𝑟 directed lines.

Lemma 2.3 (see [19]). A square matrix is irreducible if and only if its directed graph is strongly connected.

Lemma 2.4 (see [20]). If (2.1) is cooperative and irreducible in 𝐷, where 𝐷 is an open subset of 𝐶, and the solutions with positive initial data is bounded, then the trajectory of (2.1) tends to some single equilibrium.

We now consider the following delay differential system: ̇𝑢1𝑎(𝑡)=1𝑢2(𝑡𝜏)1+𝛼𝑢2(𝑡𝜏)𝑎𝑢1(𝑡),̇𝑢2(𝑡)=𝑘𝑢1𝑎(𝑡)2𝑢2(𝑡)1+𝛼𝑢2(𝑡)𝑢𝑢2(𝑡)(2.9) with initial conditions 𝑢𝑖(𝑠)=𝜙𝑖[(𝑠)0,𝑠𝜏,0),𝜙𝑖(0)>0,𝜙𝑖[𝐶𝜏,0),+(𝑖=1,2).(2.10)

System (2.9) always has a trivial equilibrium 𝐴0(0,0). If 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then system (2.9) has a unique positive equilibrium 𝐴(𝑢1,𝑢2), where 𝑢1=𝑎1𝑘𝑎1𝑎𝑎2𝑎𝑢𝛼𝑎𝑘𝑎1𝑎𝑎2,𝑢2=𝑘𝑎1𝑎𝑎2𝑎𝑢𝛼𝑎𝑢.(2.11)

The characteristic equation of system (2.9) at the positive equilibrium 𝐴0 takes the form 𝜆2+𝑝1𝜆+𝑝0+𝑞0𝑒𝜆𝜏=0,(2.12) where 𝑝0𝑎=𝑎2+𝑢,𝑝1=𝑎+𝑎2+𝑢,𝑞0=𝑘𝑎1.(2.13) Noting that 𝑝1>0,𝑝0+𝑞0=𝑎𝑢𝑘𝑎1𝑎𝑎2,(2.14) if 𝑘𝑎1𝑎𝑎2<𝑎𝑢, then the equilibrium 𝐴0 is locally stable when 𝜏=0. If 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then 𝐴0 is unstable when 𝜏=0.

It is easy to show that 𝑝212𝑝0=𝑎2+(𝑎2+𝑢)2>0. If 𝑘𝑎1𝑎𝑎2<𝑎𝑢, then 𝑝20𝑞20>0. By Kuang [21], we see that the equilibrium 𝐴0 is locally asymptotically stable for all 𝜏>0. If 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then 𝐴0 is unstable for all 𝜏>0.

The characteristic equation of system (2.9) at the positive equilibrium 𝐴 is of the form 𝜆2+𝑔1𝜆+𝑔0+0𝑒𝜆𝜏=0,(2.15) where 𝑔0𝑎=𝑎𝑢+21+𝛼𝑢22,𝑔1𝑎=𝑎+𝑢+21+𝛼𝑢22,0=𝑘𝑎11+𝛼𝑢22.(2.16) Note that 𝑔1>0,𝑔0+0=𝑎𝑢𝑎𝑢𝑘𝑎1𝑎𝑎2𝑘𝑎1𝑎𝑎2.(2.17) Hence, if 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then the positive equilibrium 𝐴 is locally stable when 𝜏=0. If 𝑘𝑎1𝑎𝑎2<𝑎𝑢, then 𝐴 is unstable when 𝜏=0.

It is easy to see that 𝑔212𝑔0=𝑎2+𝑎𝑢+21+𝛼𝑢222>0.(2.18) If 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then 𝑔2020>0. By Kuang [21], we see that the positive equilibrium 𝐴 is locally asymptotically stable for all 𝜏>0. If 𝑘𝑎1𝑎𝑎2<𝑎𝑢, then 𝐴 is unstable for all 𝜏>0.

Lemma 2.5. For system (2.9), one hase the following. (i)If 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then the positive equilibrium 𝐴(𝑢1,𝑢2) is globally stable. (ii)If 𝑘𝑎1𝑎𝑎2<𝑎𝑢, then the equilibrium 𝐴0(0,0) is globally stable.

Proof. We represent the right-hand side of (2.9) by 𝑓(𝑡,𝑥𝑡)=(𝑓1(𝑡,𝑥𝑡),𝑓2(𝑡,𝑥𝑡)), and set 𝐿(𝑡,)=𝐷𝜙𝑓(𝑡,𝜙).(2.19) By direct calculation, we have 𝐿1𝑎(𝑡,)=11+𝛼𝜙2(𝜏)22(𝜏)𝑎1𝐿(0),2(𝑡,)=𝑘1𝑎(0)21+𝛼𝜙2(0)22(0)𝑢2(0).(2.20) We now claim that the hypotheses (h1)(h4) hold for system (2.9). It is easily seen that (h1) and (h4) hold for system (2.9). We need only to verify that (h2) and (h3) hold.
The matrix 𝐴(𝑡) takes the form𝑎𝑎11+𝛼𝜙2(𝜏)2𝑎𝑘21+𝛼𝜙2(0)2𝑢.(2.21) Clearly, the matrix 𝐴(𝑡) is irreducible for each 𝑡.
From the definition of 𝐴(𝑡) and 𝜂𝑖𝑗, it is readily seen that 𝜂12(𝜃,𝑡)=𝜂12(0,𝑡)=𝑎1/(1+𝛼𝜙2(𝜏))2, 𝜂21(𝜃,𝑡)=𝜂21(0,𝑡)=𝑘 for 𝜃0, 𝜂𝑖𝑗(𝜃,𝑡)=0, 𝑖𝑗 for 𝜃𝜏, and 𝜂𝑖𝑗(,𝑡)BV[𝜏,0], where 𝜂𝑖𝑗 is a positive Borel measure on [𝜏,0]. Therefore, 𝜂𝑖𝑗(,𝑡)>0. Thus, for each 𝑗, there is 𝑖𝑗 such that 𝜂𝑖𝑗(𝑟𝑗+𝜀,𝑡)=𝜂𝑖𝑗(𝜏+𝜀,𝑡)>0 for all 𝑡 and for 𝜀>0 sufficiently small, 𝑖=1,2. Hence, (h3) holds.
Thus, the conditions of Lemma 2.1 are satisfied. Therefore, the positivity of solutions of system (2.9) follows. It is easy to see that system (2.9) is cooperative. By Lemma 2.3, we see that any solution starting from 𝐷=𝐶+𝜏 converges to some single equilibrium. However, system (2.9) has only two equilibria: 𝐴0 and 𝐴. Note that if 𝑘𝑎1𝑎𝑎2>𝑎𝑢, then the positive equilibrium 𝐴 is locally stable, and the equilibrium 𝐴0 is unstable. Hence, any solution starting from 𝐷 converges to 𝐴(𝑢1,𝑢2) if 𝑘𝑎1𝑎𝑎2>𝑎𝑢. Using a similar argument one can show the global stability of the equilibrium 𝐴0 when 𝑘𝑎1𝑎𝑎2<𝑎𝑢. This completes the proof.

3. Local Stability

In this section, we discuss the local stability of equilibria of system (1.4) by analyzing the corresponding characteristic equations.

System (1.4) always has an infection free equilibrium 𝐸0(𝜆/𝑑,0,0).

Let 0=𝜆𝛽(𝑘𝑎).𝑎𝑢𝑑(3.1)0 is called the basic reproduction ratio of system (1.4). It is easy to show that if 0>1, system (1.4) has a virus infected equilibrium 𝐸(𝑥,𝑦,𝑣), where 𝑥=𝑎𝑢1+𝛼𝑣𝛽(𝑘𝑎),𝑦=𝑢𝑣𝑘𝑎,𝑣=𝜆𝛽(𝑘𝑎)𝑎𝑢𝑑.𝑎𝑢(𝛼𝑑+𝛽)(3.2)

The characteristic equation of system (1.4) at the infection free equilibrium 𝐸0 is of the form 𝑠(𝑠+𝑑)2+𝑝1𝑠+𝑝0+𝑞0𝑒𝑠𝜏=0,(3.3) where 𝑝0=𝑎𝑢+𝜆𝛽𝑑,𝑝1=𝑎+𝑢+𝜆𝛽𝑑,𝑞0=𝑘𝜆𝛽𝑑.(3.4) Obviously, (3.3) always has a negative real root 𝑠=𝑑. All other roots of (3.3) are determined by 𝑠2+𝑝1𝑠+𝑝0+𝑞0𝑒𝑠𝜏=0.(3.5) It is easy to show that 𝑝1>0, 𝑝0+𝑞0>0, then the infection free equilibrium 𝐸0 of system (1.4) is locally asymptotically stable when 𝜏=0.

If i𝜔(𝜔>0) is a solution of (3.3), by calculating, we have the following: 𝜔4+𝑝212𝑝0𝜔2+𝑝20𝑞20=0.(3.6) Note that 𝑝212𝑝0=𝑎2+𝑢+𝜆𝛽𝑑𝑝>0,20𝑞20=1𝑑2[].𝑎𝑢𝑑+𝜆𝛽(𝑘+𝑎)][𝑎𝑢𝑑+𝜆𝛽(𝑎𝑘)(3.7) If 0<1, then 𝑝20𝑞20>0. Therefore, (3.6) has no positive roots. Accordingly, if 0<1, the infection free equilibrium 𝐸0 of system (1.4) is locally asymptotically stable for all 𝜏>0; if 0>1, (3.6) has at least a positive real root. Accordingly, 𝐸0 is unstable.

The characteristic equation of system (1.4) at the virus infected equilibrium 𝐸(𝑥,𝑦,𝑣) takes the form 𝑠3+𝑔2𝑠2+𝑔1𝑠+𝑔0+1𝑠+0𝑒𝑠𝜏=0,(3.8) where 𝑔0=𝑎𝑑𝑢+𝑑𝛽𝑥(1+𝛼𝑣)2+𝑢𝛽𝑣1+𝛼𝑣,𝑔1=𝑎𝑢+𝑑𝑢+𝑎𝑑+(𝑎+𝑑)𝛽𝑥(1+𝛼𝑣)2+(𝑎+𝑢)𝛽𝑣1+𝛼𝑣,𝑔2=𝑎+𝑢+𝑑+𝛽𝑥(1+𝛼𝑣)2+𝛽𝑣1+𝛼𝑣,0=𝑘𝑑𝛽𝑥(1+𝛼𝑣)2,1=𝑘𝛽𝑥(1+𝛼𝑣)2.(3.9) When 𝜏=0, (3.8) becomes 𝑠3+𝑔2𝑠2+𝑔1+1𝑠+𝑔0+0=0.(3.10) By direct calculation, we have 𝑔0+0𝑣=𝑎𝑢(𝛼𝑑+𝛽)1+𝛼𝑣𝑔>0,2𝑔1+1𝑔0+0=𝑎2𝑢𝛼𝑣1+𝛼𝑣+𝑎2𝜆𝑥+𝑘𝑑𝛽𝑥(1+𝛼𝑣)2+𝜆𝑢+𝑥+𝛽𝑥(1+𝛼𝑣)2×𝑎𝑢𝛼𝑣1+𝛼𝑣+(𝑢+𝑎)𝜆𝑥+𝑑𝛽𝑥(1+𝛼𝑣)2>0.(3.11) Clearly, all roots of (3.10) have only negative real parts.

If i𝜔(𝜔>0) is a solution of (3.8), separating real and imaginary parts, it follows that 𝜔3𝑔1𝜔=1𝜔cos𝜔𝜏𝑔0𝑔sin𝜔𝜏,2𝜔2𝑔0=1𝜔sin𝜔𝜏+0cos𝜔𝜏.(3.12) Squaring and adding the two equations of (3.12), we derive that 𝜔6+𝐶1𝜔4+𝐶2𝜔2+𝐶3=0,(3.13) where 𝐶1=𝑔222𝑔1,𝐶2=𝑔212𝑔0𝑔221,𝐶3=𝑔2020.(3.14) Clearly, 𝐶3=𝑔2020>0. It is easy to show that 𝐶1=𝑎2+𝑢2+𝑑2+𝛽𝑥1+𝛼𝑣2𝑢𝑑+1+𝛼𝑣+11+1+𝛼𝑣2𝐶>0,2=𝑎2+𝑢2𝑑+𝛽𝑣1+𝛼𝑣2+𝑎𝑢𝛼𝑣1+𝛼𝑣(𝑎𝑢+𝑘+𝑎)𝛽𝑥(1+𝛼𝑣)2𝑎+22𝛽+𝑑𝑢2𝑥𝑣(1+𝛼𝑣)3+𝑑2𝛽𝑥(1+𝛼𝑣)22𝑢+𝛽𝑥(1+𝛼𝑣)2>0.(3.15) Hence, (3.13) has no positive roots. Accordingly, if 0>1, the virus-infected equilibrium 𝐸 of system (1.4) exists and is locally asymptotically stable for all 𝜏>0.

Based on the discussions above, we have the following result.

Theorem 3.1. For system (1.4), one has the following. (i)If 0<1, the infection free equilibrium 𝐸0(𝜆/𝑑,0,0) is locally asymptotically stable. If 0>1, then 𝐸0(𝜆/𝑑,0,0) is unstable. (ii)If 0>1, the virus infected equilibrium 𝐸(𝑥,𝑦,𝑣) is locally asymptotically stable.

4. Global Stability

In this section, we discuss the global stability of the uninfected equilibrium and the virus infected equilibrium of system (1.4), respectively. The technique of proofs is to use a comparison argument and an iteration scheme [22].

Theorem 4.1. Let 0>1. The virus infected equilibrium 𝐸(𝑥,𝑦,𝑣) of system (1.4) is globally asymptotically stable provided that (H1)𝛼𝑑(𝑘𝑎)>𝑎𝛽, (H2)0<(𝑘+𝑎)[𝜆𝛼(𝑘+𝑎)𝑎𝑢]<4𝑘𝜆𝑎𝛼.

Proof. Let (𝑥(𝑡),𝑦(𝑡),𝑣(𝑡)) be any positive solution of system (1.4) with initial condition (1.5). Let 𝑈1=limsup𝑡+𝑥(𝑡),𝑉1=liminf𝑡+𝑈𝑥(𝑡),2=limsup𝑡+𝑦(𝑡),𝑉2=liminf𝑡+𝑈𝑦(𝑡),3=limsup𝑡+𝑣(𝑡),𝑉3=liminf𝑡+𝑣(𝑡).(4.1) Now, we claim that 𝑈1=𝑉1=𝑥, 𝑈2=𝑉2=𝑦, 𝑈3=𝑉3=𝑣.
It follows from the first equation of system (1.4) thaṫ𝑥(𝑡)𝜆𝑑𝑥(𝑡).(4.2) By comparison, we derive that 𝑈1=limsup𝑡+𝜆𝑥(𝑡)𝑑=𝑀𝑥1.(4.3) Hence, for 𝜀>0 sufficiently small, there exists a 𝑇1>0 such that if 𝑡>𝑇1, 𝑥(𝑡)𝑀𝑥1+𝜀. We therefore, derive from the second and the third equations of system (1.4) that for 𝑡>𝑇1+𝜏, 𝛽𝑀̇𝑦(𝑡)𝑥1𝑣+𝜀(𝑡𝜏)̇1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)𝑘𝑦(𝑡)𝑢𝑣(𝑡).(4.4) Consider the following auxiliary equations: ̇𝑢1𝛽𝑀(𝑡)=𝑥1𝑢+𝜀2(𝑡𝜏)1+𝛼𝑢2(𝑡𝜏)𝑎𝑢1(𝑡),̇𝑢2(𝑡)=𝑘𝑢1(𝑡)𝑢𝑢2(𝑡).(4.5) Since 0>1, by Lemma 2.5, it follows from (4.5) that lim𝑡+𝑢1𝑀(𝑡)=𝑘𝛽𝑥1+𝜀𝑎𝑢,𝑘𝑎𝛼lim𝑡+𝑢2𝑀(𝑡)=𝑘𝛽𝑥1+𝜀𝑎𝑢.𝑎𝑢𝛼(4.6) By comparison, we obtain that 𝑈2=limsup𝑡+𝑀𝑦(𝑡)𝑘𝛽𝑥1+𝜀𝑎𝑢,𝑈𝑘𝑎𝛼3=limsup𝑡+𝑀𝑣(𝑡)𝑘𝛽𝑥1+𝜀𝑎𝑢.𝑎𝑢𝛼(4.7) Since these inequalities are true for arbitrary 𝜀>0, it follows that 𝑈2𝑀𝑦1, 𝑈3𝑀𝑣1, where 𝑀𝑦1=𝑘𝛽𝑀𝑥1𝑎𝑢𝑘𝑎𝛼,𝑀𝑣1=𝑘𝛽𝑀𝑥1𝑎𝑢.𝑎𝑢𝛼(4.8) Hence, for 𝜀>0 sufficiently small, there is a 𝑇2𝑇1+𝜏 such that if 𝑡>𝑇2, 𝑦(𝑡)𝑀𝑦1+𝜀,𝑣(𝑡)𝑀𝑣1+𝜀.
For 𝜀>0 sufficiently small, we derive from the first equation of system (1.4) that for 𝑡>𝑇2,𝑀̇𝑥(𝑡)𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣1+𝜀𝑀1+𝛼𝑣1+𝜀.(4.9) A comparison argument shows that 𝑉1=liminf𝑡+𝜆𝑀𝑥(𝑡)1+𝛼𝑣1+𝜀𝑀𝑑+(𝛼𝑑+𝛽)𝑣1+𝜀.(4.10) Since these inequalities are true for arbitrary 𝜀>0, it follows that 𝑉1𝑁𝑥1, where 𝑁𝑥1=𝜆1+𝛼𝑀𝑣1𝑑+(𝛼𝑑+𝛽)𝑀𝑣1.(4.11) Hence, for 𝜀>0 sufficiently small, there is a 𝑇3𝑇2 such that if 𝑡>𝑇3, 𝑥(𝑡)𝑁𝑥1𝜀.
For 𝜀>0 sufficiently small, it follows from the second and the third equations of system (1.4) that for 𝑡>𝑇3+𝜏,𝛽𝑁̇𝑦(𝑡)𝑥1𝑣𝜀(𝑡𝜏)̇𝛽𝑀1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)𝑘𝑦(𝑡)𝑥1+𝜀𝑣(𝑡)1+𝛼𝑣(𝑡)𝑢𝑣(𝑡).(4.12) Consider the following auxiliary equations: ̇𝑢1𝛽𝑁(𝑡)=𝑥1𝑢𝜀2(𝑡𝜏)1+𝛼𝑢2(𝑡𝜏)𝑎𝑢1(𝑡),̇𝑢2(𝑡)=𝑘𝑢1𝛽𝑀(𝑡)𝑥1𝑢+𝜀2(𝑡)1+𝛼𝑢2(𝑡)𝑢𝑢2(𝑡).(4.13) Since (H1) holds, by Lemma 2.5, it follows from (4.13) that lim𝑡+𝑢1𝛽𝑁(𝑡)=𝑥1𝑁𝜀𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1+𝜀𝑎𝑢𝑁𝑎𝛼𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1,+𝜀lim𝑡+𝑢2𝑁(𝑡)=𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1+𝜀𝑎𝑢.𝑎𝑢𝛼(4.14) By comparison, we derive that 𝑉2=liminf𝑡+𝛽𝑁𝑦(𝑡)𝑥1𝑁𝜀𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1+𝜀𝑎𝑢𝑁𝑎𝛼𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1,𝑉+𝜀3=liminf𝑡+𝑁𝑣(𝑡)𝑘𝛽𝑥1𝑀𝜀𝑎𝛽𝑥1+𝜀𝑎𝑢.𝑎𝑢𝛼(4.15) Since these two inequalities hold for arbitrary 𝜀>0 sufficiently small, we conclude that 𝑉2𝑁𝑦1, 𝑉3𝑁𝑣1, where 𝑁𝑦1=𝛽𝑁𝑥1𝑘𝛽𝑁𝑥1𝑎𝛽𝑀𝑥1𝑎𝑢𝑎𝛼𝑘𝛽𝑁𝑥1𝑎𝛽𝑀𝑥1,𝑁𝑣1=𝑘𝛽𝑁𝑥1𝑎𝛽𝑀𝑥1𝑎𝑢𝑎𝑢𝛼.(4.16) Therefore, for 𝜀>0 sufficiently small, there exists a 𝑇4𝑇3+𝜏 such that if 𝑡>𝑇4, 𝑦(𝑡)𝑁𝑦1𝜀,𝑣(𝑡)𝑁𝑣1𝜀.
For 𝜀>0 sufficiently small, it follows from the first equation of system (1.4) that for 𝑡>𝑇4,𝑁̇𝑥(𝑡)𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣1𝜀𝑁1+𝛼𝑣1𝜀.(4.17) By comparison, we derive that 𝑈1=limsup𝑡+𝜆𝑁𝑥(𝑡)1+𝛼𝑣1𝜀𝑁𝑑+(𝛼𝑑+𝛽)𝑣1𝜀.(4.18) Since this is true for arbitrary 𝜀>0, it follows that 𝑈1𝑀𝑥2, where 𝑀𝑥2=𝜆1+𝛼𝑁𝑣1𝑑+(𝛼𝑑+𝛽)𝑁𝑣1.(4.19) Hence, for 𝜀>0 sufficiently small, there is a 𝑇5𝑇4 such that if 𝑡>𝑇5, 𝑥(𝑡)𝑀𝑥2+𝜀. It therefore, follows from the second and the third equations of system (1.4) that for 𝑡>𝑇5+𝜏, 𝛽𝑀̇𝑦(𝑡)𝑥2𝑣+𝜀(𝑡𝜏)̇𝛽𝑁1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)𝑘𝑦(𝑡)𝑥2𝜀𝑣(𝑡𝜏)1+𝛼𝑣(𝑡𝜏)𝑢𝑣(𝑡).(4.20) By Lemma 2.5 and a comparison argument, we derive from (4.20) that 𝑈2=limsup𝑡+𝛽𝑀𝑦(𝑡)𝑥2𝑀+𝜀𝑘𝛽𝑥2𝑁+𝜀𝑎𝛽𝑥1𝜀𝑎𝑢𝑀𝑎𝛼𝑘𝛽𝑥2𝑁+𝜀𝑎𝛽𝑥1,𝑈𝜀3=limsup𝑡+𝑀𝑣(𝑡)𝑘𝛽𝑥2𝑁+𝜀𝑎𝛽𝑥1𝜀𝑎𝑢.𝑎𝑢𝛼(4.21) Since these inequalities are true for arbitrary 𝜀>0, it follows that 𝑈2𝑀𝑦2, 𝑈3𝑀𝑣2, where 𝑀𝑦2=𝛽𝑀𝑥2𝑘𝛽𝑀𝑥2𝑎𝛽𝑁𝑥1𝑎𝑢𝑎𝛼𝑘𝛽𝑀𝑥2𝑎𝛽𝑁𝑥1,𝑀𝑣2=𝑘𝛽𝑀𝑥2𝑎𝛽𝑁𝑥1𝑎𝑢𝑎𝑢𝛼.(4.22) Hence, for 𝜀>0 sufficiently small, there is a 𝑇6𝑇5+𝜏 such that if 𝑡>𝑇6, 𝑦(𝑡)𝑀𝑦2+𝜀, 𝑣(𝑡)𝑀𝑣2+𝜀.
Again, for 𝜀>0 sufficiently small, we derive from the first equation of system (1.4) that for 𝑡>𝑇6,𝑀̇𝑥(𝑡)𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝑣2+𝜀𝑀1+𝛼𝑣2+𝜀.(4.23) A comparison argument shows that 𝑉1=liminf𝑡+𝜆𝑀𝑥(𝑡)1+𝛼𝑣2+𝜀𝑀𝑑+(𝛼𝑑+𝛽)𝑣2+𝜀.(4.24) Since this is true for arbitrary 𝜀>0, it follows that 𝑉1𝑁𝑥2, where 𝑁𝑥2=𝜆1+𝛼𝑀𝑣2𝑑+(𝛼𝑑+𝛽)𝑀𝑣2.(4.25) Hence, for 𝜀>0 sufficiently small, there is a 𝑇7𝑇6 such that if 𝑡>𝑇7, 𝑥(𝑡)𝑁𝑥2𝜀.
For 𝜀>0 sufficiently small, it follows from the second and the third equations of system (1.4) that for 𝑡>𝑇7+𝜏,𝛽𝑁̇𝑦(𝑡)𝑥2𝑣𝜀(𝑡𝜏)̇𝛽𝑀1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)𝑘𝑦(𝑡)𝑥2+𝜀𝑣(𝑡)1+𝛼𝑣(𝑡)𝑢𝑣(𝑡).(4.26) Since (H1) holds, by Lemma 2.5 and a comparison argument, it follows from (4.26) that 𝑉2=liminf𝑡+𝛽𝑁𝑦(𝑡)𝑥2𝑁𝜀𝑘𝛽𝑥2𝑀𝜀𝑎𝛽𝑥2+𝜀𝑎𝑢𝑁𝑎𝛼𝑘𝛽𝑥2𝑀𝜀𝑎𝛽𝑥2,𝑉+𝜀3=liminf𝑡+𝑁𝑣(𝑡)𝑘𝛽𝑥2𝑀𝜀𝑎𝛽𝑥2+𝜀𝑎𝑢.𝑎𝑢𝛼(4.27) Since these two inequalities hold for arbitrary 𝜀>0 sufficiently small, we conclude that 𝑉2𝑁𝑦2, 𝑉3𝑁𝑣2, where 𝑁𝑦2=𝛽𝑁𝑥2𝑘𝛽𝑁𝑥2𝑎𝛽𝑀𝑥2𝑎𝑢𝑎𝛼𝑘𝛽𝑁𝑥2𝑎𝛽𝑀𝑥2,𝑁𝑣2=𝑘𝛽𝑁𝑥2𝑎𝛽𝑀𝑥2𝑎𝑢𝑎𝑢𝛼.(4.28) Therefore, for 𝜀>0 sufficiently small, there exists a 𝑇8𝑇7+𝜏 such that if 𝑡>𝑇8, 𝑦(𝑡)𝑁𝑦2𝜀, 𝑣(𝑡)𝑁𝑣2𝜀.
Continuing this process, we derive six sequences 𝑀𝑥𝑛, 𝑀𝑦𝑛, 𝑀𝑣𝑛, 𝑁𝑥𝑛, 𝑁𝑦𝑛, 𝑁𝑣𝑛  (𝑛=1,2,) such that for 𝑛2,𝑀𝑥𝑛=𝜆1+𝛼𝑁𝑣𝑛1𝑑+(𝛼𝑑+𝛽)𝑁𝑣𝑛1,𝑀𝑦𝑛=𝛽𝑀𝑥𝑛𝑘𝛽𝑀𝑥𝑛𝑎𝛽𝑁𝑥𝑛1𝑎𝑢𝑎𝛼𝑘𝛽𝑀𝑥𝑛𝑎𝛽𝑁𝑥𝑛1,𝑀𝑣𝑛=𝑘𝛽𝑀𝑥𝑛𝑎𝛽𝑁𝑥𝑛1𝑎𝑢,𝑁𝑎𝑢𝛼𝑥𝑛=𝜆1+𝛼𝑀𝑣𝑛𝑑+(𝛼𝑑+𝛽)𝑀𝑣𝑛,𝑁𝑦𝑛=𝛽𝑁𝑥𝑛𝑘𝛽𝑁𝑥𝑛𝑎𝛽𝑀𝑥𝑛𝑎𝑢𝑎𝛼𝑘𝛽𝑁𝑥𝑛𝑎𝛽𝑀𝑥𝑛,𝑁𝑣𝑛=𝑘𝛽𝑁𝑥𝑛𝑎𝛽𝑀𝑥𝑛𝑎𝑢.𝑎𝑢𝛼(4.29) Clearly, we have 𝑁𝑥𝑛𝑉1𝑈1𝑀𝑥𝑛,𝑁𝑦𝑛𝑉2𝑈2𝑀𝑦𝑛,𝑁𝑣𝑛𝑉3𝑈3𝑀𝑣𝑛.(4.30) It is easy to show that the sequences 𝑀𝑥𝑛, 𝑀𝑦𝑛, and 𝑀𝑣𝑛 are nonincreasing, and the sequences 𝑁𝑥𝑛, 𝑁𝑦𝑛, 𝑁𝑣𝑛 are nondecreasing. Hence, the limit of each sequence in 𝑀𝑥𝑛, 𝑀𝑦𝑛, 𝑀𝑣𝑛, 𝑁𝑥𝑛, 𝑁𝑦𝑛, and 𝑁𝑣𝑛 exists. Denote 𝑥=lim𝑛+𝑀𝑥𝑛,𝑥=lim𝑛+𝑁𝑥𝑛,𝑦=lim𝑛+𝑀𝑦𝑛,𝑦=lim𝑛+𝑁𝑦𝑛,𝑣=lim𝑛+𝑀𝑣𝑛,𝑣=lim𝑛+𝑁𝑣𝑛.(4.31) We, therefore, obtain from (4.29) and (4.31) that 𝑎(𝛼𝑑+𝛽)𝑥2𝑘(𝛼𝑑+𝛽)𝑥𝑥=𝑎(𝜆𝛼𝑢)𝑥𝑘𝜆𝛼𝑥,(4.32)𝑎(𝛼𝑑+𝛽)𝑥2𝑘(𝛼𝑑+𝛽)𝑥𝑥=𝑎(𝜆𝛼𝑢)𝑥𝑘𝜆𝛼𝑥.(4.33) (4.32) minus (4.33), 𝑎(𝛼𝑑+𝛽)𝑥2𝑥2=(𝜆𝑎𝛼+𝜆𝑘𝛼𝑎𝑢)𝑥𝑥.(4.34) Assume that 𝑥𝑥. Then, we derive from (4.34) that 𝑥+𝑥=𝜆𝑎𝛼+𝜆𝑘𝛼𝑎𝑢𝑎.(𝛼𝑑+𝛽)(4.35) (4.32) plus (4.33), 𝑎(𝛼𝑑+𝛽)𝑥+𝑥22(𝑘+𝑎)(𝛼𝑑+𝛽)𝑥𝑥=(𝜆𝑎𝛼𝜆𝑘𝛼𝑎𝑢)𝑥+𝑥.(4.36) On substituting (4.35) into (4.36), it follows that 𝑥𝑥=𝜆𝑘𝛼𝑎(𝑘+𝑎)(𝛼𝑑+𝛽)2(𝜆𝑎𝛼+𝜆𝑘𝛼𝑎𝑢).(4.37) Note that 𝑥>0, 𝑥>0. Let (H2) hold. It follows from (4.35) and (4.37) that 𝑥+𝑥24𝑥𝑥=(𝜆𝑎𝛼+𝜆𝑘𝛼𝑎𝑢)𝑎2(𝛼𝑑+𝛽)2(𝜆𝑎𝛼+𝜆𝑘𝛼𝑎𝑢)4𝑘𝜆𝑎𝛼.𝑘+𝑎(4.38) Hence, we have (𝑥+𝑥)24𝑥𝑥<0. This is a contradiction. Accordingly, we have 𝑥=𝑥. We, therefore, derive from (4.31) that 𝑦=𝑦, 𝑣=𝑣. Note that if (H1) and (H2) hold, by Theorem 3.1, the virus-infected equilibrium 𝐸 is locally stable, we conclude that 𝐸 is globally stable. The proof is complete.

Theorem 4.2. If 0<1 holds, the infection-free equilibrium 𝐸0(𝜆/𝑑,0,0) of system (1.4) is globally asymptotically stable.

Proof. Let (𝑥(𝑡),𝑦(𝑡),𝑣(𝑡)) be any positive solution of system (1.4) with initial condition (1.5).
If 0<1, choose 𝜀>0 sufficiently small satisfying𝜆𝑘𝛽𝑑+𝜀<𝑎𝑢.(4.39) It follows from the first equation of system (1.4) that ̇𝑥(𝑡)𝜆𝑑𝑥(𝑡).(4.40) By comparison, we derive that limsup𝑡+𝜆𝑥(𝑡)𝑑.(4.41) Hence, for 𝜀>0 sufficiently small satisfying (4.39), there exists a 𝑇1>0 such that if 𝑡>𝑇1, 𝑥(𝑡)(𝜆/𝑑)+𝜀. We, therefore, derive from the second and the third equations of system (1.4) that for 𝑡>𝑇1+𝜏, ̇𝑦(𝑡)𝛽((𝜆/𝑑)+𝜀)𝑣(𝑡𝜏)̇1+𝛼𝑣(𝑡𝜏)𝑎𝑦(𝑡),𝑣(𝑡)𝑘𝑦(𝑡)𝑢𝑣(𝑡).(4.42) Consider the following auxiliary equation: ̇𝑢1(𝑡)=𝛽((𝜆/𝑑)+𝜀)𝑢2(𝑡𝜏)1+𝛼𝑢2(𝑡𝜏)𝑎𝑢1(𝑡),̇𝑢2(𝑡)=𝑘𝑢1(𝑡)𝑢𝑢2(𝑡).(4.43) If 0<1, then by Lemma 2.5, it follows from (4.41) and (4.43) that lim𝑡+𝑢1(𝑡)=0,lim𝑡+𝑢1(𝑡)=0.(4.44) By comparison, we obtain that lim𝑡+𝑦(𝑡)=0,lim𝑡+𝑣(𝑡)=0.(4.45) Therefore, for 𝜀>0 sufficiently small, there is a 𝑇2>𝑇1+𝜏 such that if 𝑡>𝑇2, 𝑦(𝑡)<𝜀, 𝑣(𝑡)<𝜀.
It follows from the first equation of system (1.4) that for 𝑡>𝑇2,̇𝑥(𝑡)𝜆𝑑𝑥(𝑡)𝛽𝑥(𝑡)𝜀.1+𝛼𝜀(4.46) By comparison, we derive that liminf𝑡+𝑥(𝑡)𝜆(1+𝛼𝜀).𝑑+(𝛼𝑑+𝛽)(1+𝛼𝜀)(4.47) Letting 𝜀0, it follows that liminf𝑡+𝜆𝑥(𝑡)𝑑.(4.48) Noting that (4.41) holds, we conclude that lim𝑡+𝜆𝑥(𝑡)=𝑑.(4.49) This completes the proof.

5. Numerical Examples

In this section, we give two examples to illustrate the main theoretical results above.

Example 5.1. In system (1.4), let 𝛼=2, 𝛽=2, 𝜆=4, 𝑎=1, 𝑑=1.5, 𝑘=2, 𝑢=3, and 𝜏=1. By calculation, we have 0=16/9>1, and system (1.4) has a virus-infected equilibrium 𝐸(11/5,7/10,7/30). Clearly, (H1) and (H2) hold. By Theorem 4.1, we see that 𝐸 is globally asymptotically stable. Numerical simulation illustrates the above result (see Figure 1).

Example 5.2. In system (1.4), let 𝛼=1, 𝛽=1, 𝜆=2.5, 𝑎=2, 𝑑=1, 𝑘=1, 𝑢=2, and 𝜏=3. Noting that 0=5/8<1, system (1.4) always has an infection-free equilibrium 𝐸0(5/2,0,0). By Theorem 4.2, we see that 𝐸0 is globally asymptotically stable. Numerical simulation illustrates this fact (see Figure 2).

6. Conclusions

In this paper, we have discussed a virus infection model with time delay, absorption, and saturation incidence. The basic reproduction ratio 0 was found. We investigated the global asymptotic stability of the infection-free equilibrium and the virus-infected equilibrium of system (1.4), respectively. When the basic reproduction ratio is greater than unity, by using the iteration scheme, we have established sufficient conditions for the global stability of the virus-infected equilibrium of system (1.4). By Theorem 4.1, we see that when 0>1 and (H1) and (H2) hold, the virus-infected equilibrium is globally stable. Biologically, these indicate that when the death rate of infected cells and the production rate of free viruses from infected cells are sufficiently large, then the solutions of system (1.4) tend to the virus infected equilibrium which means that the virus persists in the host. On the other hand, by Theorem 4.2, we see that if the basic reproduction ratio is less than unity, the infection free equilibrium is globally asymptotically stable. Biologically, if the rate at which new uninfected cells are generated and the average number of effective contacts of one infective individual per unit time are small enough and the death rates of uninfected cells, infected cells and pathogens are large enough such that 0<1, then the virus is cleared. We would like to point out here that Theorem 4.1 has room for improvement, we leave this for future work.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (no. 11071254) and the Scientific Research Foundation for the Returned Overseas Chinese Scholars, State Education Ministry, and the Science Research Foundation of JCB (no. JCB 1005).