1,374
Views
11
CrossRef citations to date
0
Altmetric
Articles

An age-structured within-host HIV-1 infection model with virus-to-cell and cell-to-cell transmissions

, &
Pages 89-117 | Received 28 Mar 2017, Accepted 07 Nov 2017, Published online: 24 Nov 2017

ABSTRACT

In this paper, a within-host HIV-1 infection model with virus-to-cell and direct cell-to-cell transmission and explicit age-since-infection structure for infected cells is investigated. It is shown that the model demonstrates a global threshold dynamics, fully described by the basic reproduction number. By analysing the corresponding characteristic equations, the local stability of an infection-free steady state and a chronic-infection steady state of the model is established. By using the persistence theory in infinite dimensional system, the uniform persistence of the system is established when the basic reproduction number is greater than unity. By means of suitable Lyapunov functionals and LaSalle's invariance principle, it is shown that if the basic reproduction number is less than unity, the infection-free steady state is globally asymptotically stable; if the basic reproduction number is greater than unity, the chronic-infection steady state is globally asymptotically stable. Numerical simulations are carried out to illustrate the feasibility of the theoretical results.

2000 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

In past decades, great attention has been paid to the within-host dynamics of HIV using mathematical modelling. Mathematical modelling combined with experimental measurements has yielded important insights into HIV-1 pathogenesis and has enhanced progress in the understanding of HIV-1 infection (see, e.g. [Citation1,Citation6,Citation16,Citation18–21]). These models mainly investigated the dynamics of the target cells and infected cells, viral production and clearance, and the effects of antiretroviral drugs treatment. For decades it was believed that the spreading of HIV-1 within a host was mainly through free circulation of the viral particles with a repeated process. Models used to study HIV-1 infection have involved the concentrations of uninfected target cells, T, infected cells that are producing virus, T, and virus, V. The following classic and basic mathematical model describing HIV-1 infection dynamics was proposed and studied in [Citation17,Citation21]: (1) T˙(t)=λdT(t)βT(t)V(t),T˙(t)=βT(t)V(t)aT(t),V˙(t)=kT(t)uV(t),(1) where uninfected, susceptible cells are produced at a rate, λ, and die at rate dT, and become infected at rate βTV, where β is the rate constant describing the infection process; infected cells are produced at rate βTV and die at rate aT; free virions are produced from infected cells at rate kT and are removed at rate uV.

However, recent studies have revealed that a large number of viral particles can also be transferred from infected cells to uninfected cells through the formation of virally induced structures termed virological synapses (see, [Citation2,Citation4,Citation8,Citation25,Citation26,Citation29]). Cell-to-cell spread of HIV-1 between CD4+T cells is an efficient means of viral dissemination [Citation25] and has been estimated to be several orders of magnitude more rapid than cell-free virus infection [Citation2]. Cell-to-cell spread not only facilitates the rapid viral dissemination but may also promote immune invasion and, thereby, influence the disease [Citation14]. It was shown in [Citation27] that cell-to-cell spread of HIV-1 does reduce the efficacy of antiretroviral therapy, because cell-to-cell infection can cause multiple infections of target cells, which can in turn reduce the sensitivity to the antiretroviral drugs. The relative contribution of the two transmission pathways to virus growth through multiple rounds of replication has been examined by Sourisseau et al. [Citation29], but it has not yet been quantified rigorously. In [Citation10], by fitting a mathematical model to data reported in [Citation29] as well as newly generated experimental data, Komarova et al. determined that free-virus and synaptic transmission make approximately equal contributions to virus growth in vitro.

In [Citation11], Lai and Zou considered the following dynamical system model to incorporate both cell-to-cell infection mechanism and virus-to-cell infection mode: (2) T˙(t)=λdTT(t)β1T(t)V(t)β2T(t)T(t),T˙(t)=β10f(s)eμsT(ts)V(ts)ds+β20f(s)eμsT(ts)T(ts)dsδT(t),V˙(t)=bT(t)cV(t),(2) where T(t),T(t) and V(t) are concentrations of uninfected T cells, infected T cells, and free viral particles at time t, respectively. In system (Equation2), uninfected, susceptible cells are produced at a rate λ, and die at rate dTT(t), and become infected at rate β1TV+β2TT, where β1 is the infection rate of free virus and β2 is the infection rate of productively infected cells; infected cells are produced at rate 0f(s)eμs[β1T(ts)V(ts)+β2T(ts)T(ts)]ds and die at rate δT. The time for infected cells to become productively infected may vary from individual to individual, and hence, a distribution function f(s) is introduced to account for such variance. The term eμs accounts for the survival rates of cells that are infected at time t and become productively infected s time units later. Free virions are produced from infected cells at rate bT and are removed at rate cV. In [Citation12], Lai and Zou further studied an HIV-1 infection model with diffusion-limited cell-free virus transmission and cell-to-cell transfer and logistic target cell growth.

Note that in systems (Equation1) and (Equation2), both the death rate and virus production rate of infected cells are assumed to be constant. It was reported in [Citation23] that virus production increases exponentially with the age of the infected cell in the case of simian immunodeficiency virus-infected CD4+ T cells in rhesus macaques. Recently, age-structured within-host HIV infection models have received increasing interest due to their greater flexibility in modelling variations in the death rate of productively infected T cells and the production rate of viral particles as a function of the length of time a T cell has been infected [Citation15]. In [Citation15], Nelson et al. developed the following age-structured within-host HIV-1 infection model: (3) T˙(t)=sdTT(t)βT(t)V(t),T(a,t)t+T(a,t)a=μ(a)T(a,t),V˙(t)=0p(a)T(a,t)dauV(t),(3) with boundary condition T(0,t)=βT(t)V(t). In system (Equation3), T(a,t) denotes the density of infected T cells of infection age a (i.e. the time that has elapsed since an HIV virion has penetrated the cell) at time t, μ(a) is the age-dependent per capita death rate of infected cells, p(a) is the viral production rate of an infected cell with age a. The age of cellular infection plays a key role in determining the rate of viral particle production per productively infected T cell and how long the productively infected T cell lives. In [Citation24], in order to assess the effect of different combination therapies on viral dynamics, Rong et al. incorporated treatment with three different classes of drugs into the age-structured model. In [Citation7], by using the direct Lyapunov method, Huang et al. established the global stability of feasible steady states of system (Equation3).

We note that the infection process in system (Equation3) is assumed to be governed by the mass-action principle, that is, the infection rate per host and per virus is a constant. However, experiments reported in [Citation3] strongly suggested that the infection rate of microparasitic infections is an increasing function of the parasite dose, and is usually sigmoidal in shape. In [Citation22], to place the model on more sound biological grounds, Regoes et al. replaced the mass-action infection rate with a dose-dependent infection rate.

Motivated by the works of Lai and Zou [Citation11], Nelson et al. [Citation15] and Regoes et al. [Citation22], in the present paper, we are concerned with the joint effects of age since infection, direct cell-to-cell transfer and virus-to-cell infection on the dynamics of HIV-1 infection. To this end, we consider the following within-host HIV-1 infection model: (4) x˙(t)=sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da,y(a,t)t+y(a,t)a=μ(a)y(a,t),v˙(t)=0k(a)y(a,t)dauv(t),(4) with boundary condition (5) y(0,t)=βx(t)v(t)1+αv(t)+x(t)0β1(a)y(a,t)da,t>0,(5) and initial condition (6) X0:=(x(0),y(,0),v(0))=(x0,y0(),v0)X,(6) where X=R+×L+1(0,)×R+, L+1(0,) is the set of all integrable functions from (0,) into R+=[0,).

In system (Equation4), x(t) represents the concentration of uninfected target T cells at time t, y(a,t) denotes the density of infected T cells of infection age a (i.e. the time that has elapsed since an HIV virion has penetrated cell) at time t, and v(t) denotes the concentration of infectious free virion at time t. The definitions of all parameters in system (Equation4) are listed in Table .

Table 1. The definitions of the parameters in system (Equation4).

We make the following assumptions on the parameters in system (Equation4).

(H1)

k,μ,β1L+1(0,), let k¯,μ¯,β¯1 be the essential supremums of k,μ,β1, respectively;

(H2)

β1(a) is Lipschitz continuous on R+ with Lipschitz coefficient Mβ1;

(H3)

There is a positive constant μ0min{d,u} such that μ(a)μ0 for all a0.

Using the theory of age-structured dynamical systems developed in [Citation9,Citation30], we can verify that system (Equation4) has a unique solution (x(t),y(,t),v(t)) satisfying the boundary condition (Equation5) and the initial condition (Equation6). Moreover, it is easy to show that all solutions of system (Equation4) with the boundary condition (Equation5) and the initial condition (Equation6) are defined on [0,+) and remain positive for all t0. Furthermore, X is positively invariant and system (Equation4) exhibits a continuous semi-flow Φ:R+×XX, namely, Φt(X0)=Φ(t,X0):=(x(t),y(,t),v(t)),t0,X0X. Given a point (x,ϕ,z)X, one has the norm (x,ϕ,z)X:=x+0ϕ(a)da+z.

In this paper, our primary goal is to carry out a complete mathematical analysis of system (Equation4) with the boundary condition (Equation5) and the initial condition (Equation6) and establish its global dynamics. The organization of this paper is as follows. In the next section, we are concerned with the asymptotic smoothness of the semi-flow generated by system (Equation4). In Section 3, we calculate the basic reproduction number and investigate the existence of feasible steady states of system (Equation4). In Section 4, by analysing corresponding characteristic equations, we study the local asymptotic stability of an infection-free steady state and a chronic-infection steady state of system (Equation4). In Section 5, using the persistence theory in infinite dimensional system developed by Hale and Waltman in [Citation5], the uniform persistence of the semi-flow generated by system (Equation4) is established when the basic reproduction number is greater than unity. In Section 6, we are concerned with the global stability of each of feasible steady states by constructing suitable Lyapunov functionals and using LaSalle's invariance principle. In Section 7, numerical examples are carried out to illustrate the feasibility of theoretical results. A brief discussion is given in Section 8 to conclude this work.

2. Asymptotic smoothness

In order to study the global dynamics of system (Equation4), in this section, we need to verify the asymptotic smoothness of the semi-flow {Φ(t)}t0 generated by system (Equation4).

2.1. Boundedness of solutions

Denote (7) π(a)=e0aμ(s)dsfor aR+.(7) It follows from (H1) and (H3) that 0<eμ¯aπ(a)eμ0a for all a0. Clearly, π(a) is a deceasing function.

Let Φt(X0)=Φ(t,X0):=(x(t),y(a,t),v(t)) be any non-negative solution of system (Equation4) with the boundary condition (Equation5) and the initial condition (Equation6). Integrating the second equation of system (Equation4) along the characteristic line ta=const. yields (8) y(a,t)={L(ta)π(a),0a<t,y0(at)π(a)π(at),0ta,(8) where L(t):=y(0,t)=βx(t)v(t)/(1+αv(t))+x(t)0β1(a)y(a,t)da.

Denote X0X=x0+0y0(a)da+v0 and N(t)=Φ(t,X0)X=x(t)+0y(a,t)da+v(t).

Proposition 2.1

For system (Equation4), the following statements hold.

  1. (d/dt)N(t)s+k¯max{s/μ0,X0X}μ0N(t) for all t0;

  2. N(t)max{s/μ0+(k¯/μ0)max{s/μ0,X0X},X0X} for all t0;

  3. lim supt+N(t)s(1+k¯/μ0)/μ0;

  4. Φt is point dissipative: there is a bounded set that attracts all points in X.

Proof.

It follows from Equations (Equation4)–(Equation6) that (9) ddt(x(t)+0y(a,t)da)=sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da0y(a,t)ada0μ(a)y(a,t)da=sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)day(a,t)|00μ(a)y(a,t)da.(9)

On substituting Equation (Equation5) into Equation (Equation9), one obtains that (10) ddt(x(t)+0y(a,t)da)sdx(t)0μ(a)y(a,t)dasμ0(x(t)+0y(a,t)da).(10) The variation of constants formula implies x(t)+0y(a,t)dasμ0eμ0t[sμ0(x0+0y0(a)da)]<sμ0eμ0t{sμ0X0X}, which yields (11) x(t)+0y(a,t)damax{sμ0,X0X}(11) for all t0.

We derive from Equation (Equation11) and the third equation of system (Equation4) that (12) dv(t)dt=0k(a)y(a,t)dauv(t)k¯max{sμ0,X0X}uv(t).(12) It follows from Equations (Equation10) and (Equation12) that (13) ddtN(t)sμ0(x(t)+0y(a,t)da)+k¯max{sμ0,X0X}uv(t)s+k¯max{sμ0,X0X}μ0N(t).(13) Again, using variation of constants formula we have from Equation (Equation13) that N(t)sμ0+k¯μ0max{sμ0,X0X}eμ0t{sμ0+k¯μ0max{sμ0,X0X}X0X} for all t0. This yields N(t)max{s/μ0+(k¯/μ0)max{s/μ0,X0X},X0X}. The proof is complete.

The following results are direct consequences of Proposition 2.1.

Proposition 2.2

If X0X and X0XK for some Ks(1+k¯/μ0)/μ0, then x(t)K,0y(a,t)daK,v(t)Kfor all t0.

Proposition 2.3

Let CX be bounded. Then

  1. Φt(C) is bounded for all t0;

  2. Φt is eventually bounded on C.

2.2. Asymptotic smoothness

In this section, we show the asymptotic smoothness of the semi-flow {Φ(t)}t0 generated by system (Equation4).

Denote A(t)=βx(t)v(t)/(1+αv(t)), B(t)=0β1(a)y(a,t)da.

Proposition 2.4

The function B(t) is Lipschitz continuous on R+.

Proof.

Let Kmax{s(1+k¯/u)/μ0,X0X}. By Proposition 2.1 we have ΦtXK for all t0. Fix t0 and h>0. Then (14) B(t+h)B(t)=0β1(a)y(a,t+h)da0β1(a)y(a,t)da=0hβ1(a)y(a,t+h)da+hβ1(a)y(a,t+h)da0β1(a)y(a,t)da.(14) On substituting Equation (Equation8) into Equation (Equation14), it follows that (15) B(t+h)B(t)=0hβ1(a)L(t+ha)π(a)da+hβ1(a)y(a,t+h)da0β1(a)y(a,t)da.(15) By Proposition 2.2, we have L(t)βK2/(1+αK)+β1¯K2. Noting that π(a)1, we obtain from Equation (Equation15) that (16) |B(t+h)B(t)|(β1¯βK21+αK+β1¯K2)h+|hβ1(a)y(a,t+h)da0β1(a)y(a,t)da|=(β1¯βK21+αK+β1¯K2)h+|0β1(σ+h)y(σ+h,t+h)dσ0β1(a)y(a,t)da|.(16) It follows from Equation (Equation8) that y(a+h,t+h)=y(a,t)π(a+h)π(a)=y(a,t)e0hμ(s)ds for all a0,t0,h0. Hence, Equation (Equation16) can be rewritten as (17) |B(t+h)B(t)|(β1¯βK21+αK+β1¯K2)h+|0β1(a+h)y(a,t)e0hμ(s)dsda0β1(a)y(a,t)da|(β1¯βK21+αK+β1¯K2)h+0β1(a+h)(1e0hμ(s)ds)y(a,t)da+0|β(a+h)β1(a)|y(a,t)da.(17) Noting that 1exx for x0, we obtain from Equation (Equation17) that (18) |B(t+h)B(t)|(β1¯βK21+αK+β1¯K2)h+β1¯μ¯Kh+Mβ1Kh,(18) where Mβ1 is defined in (H2). This completes the proof.

Proposition 2.5

The function A(t) is Lipschitz continuous on R+.

Proof.

Let Kmax{s(1+k¯/u)/μ0, X0X}. By Proposition 2.1 we have x(t)K, v(t)K for all t0. Fix t0 and h>0. Then (19) |A(t+h)A(t)|=|βx(t+h)v(t+h)1+αv(t+h)βx(t)v(t)1+αv(t)|=β|x(t+h)(v(t+h)v(t))(1+αv(t))(1+αv(t+h))+v(t)(x(t+h)x(t))1+αv(t)|βK|v(t+h)v(t)|+K|x(t+h)x(t)|.(19) Note that the Lipschitz continuity of x() and v() on R+ can be verified from Equation (Equation4) and Proposition 2.2. Hence, there are positive constants Mx and Mv such that (20) |x(t+h)x(t)|Mxh,|v(t+h)v(t)|Mvh.(20) It therefore follows from Equations (Equation19) and (Equation20) that (21) |A(t+h)A(t)|βK(Mv+Mx)h.(21) This completes the proof.

By Propositions 2.4 and 2.5, one can directly obtain the following result.

Proposition 2.6

The function L(t) is Lipschitz continuous on R+.

We now state two theorems introduced in [Citation28] which are useful in proving the asymptotic smoothness of the semi-flow {Φ(t)}t0.

Theorem 2.1

The semi-flow Φ:R+×X+X+ is asymptotically smooth if there are maps Θ,Ψ:R+×X+X+ such that Φ(t,x)=Θ(t,X)+Ψ(t,X) and the following hold for any bounded closed set CX+ that is forward invariant under Φ:

  1. limt+diamΘ(t,C)=0;

  2. there exists tC0 such that Ψ(t,C) has compact closure for each ttC.

Theorem 2.2

Let C be a subset of L1(R+). Then C has compact closure if and only if the following assumptions hold:

  1. supfC0|f(a)|da<;

  2. limrr|f(a)|da=0 uniformly in fC;

  3. limh0+0|f(a+h)f(a)|da=0 uniformly in fC;

  4. limh0+0h|f(a)|da=0 uniformly in fC.

We are now ready to state and prove the asymptotic smoothness of the semi-flow Φ.

Theorem 2.3

The semi-flow {Φ(t)}t0 generated by system (Equation4) is asymptotically smooth.

Proof.

To verify the two conditions in Theorem 2.1, we first decompose the semi-flow Φ into two parts: for t0, let Ψ(t,X0):=(x(t),y~(,t),v(t)),Θ(t,X0):=(0,φ~y(,t),0), where y~(a,t)={L(ta)π(a),0at,0,t<a,φ~y(a,t)={0,0at,y0(at)π(a)π(at),t<a. Clearly, for t0, we have Φ=Θ+Ψ.

Let C be a bounded subset of X and K>s(1+k¯/u)/μ0 the bound for C. Let Φ(t,X0)=(x(t),y(,t),v(t)), where X0=(x0,y0(a),v0)C. Then (22) Θ(t,X0)=φ~y(,t)L1=0|φ~y(a,t)|da=ty0(at)π(a)π(at)da.(22)

Letting at=σ, it follows from Equation (Equation22) that Θ(t,X0)=0y0(σ)π(σ+t)π(σ)dσ=0y0(σ)eσσ+tμ(s)dsdσeμ0t0y0(σ)dσKeμ0t, which yields limt+Θ(t,X0)=0. Hence, limt+diamΘ(t,C)=0. The assumption (1) in Theorem 2.1 holds.

In the following we show that Ψ(t,C) has compact closure for each ttC by verifying the assumptions (i)-(iv) of Theorem 2.2. From Proposition 2.2 we see that x(t) and v(t) remain in the compact set [0,K]. Next, we show that y~(a,t) remains in a pre-compact subset of L+1 independent of X0. It is easy to show that y~(a,t)L¯eμ0a, where L¯=βK2/(1+αK)+β¯1K. Hence, the assumptions (i),(ii) and (iv) of Theorem 2.2 follow directly. We need only to verify that (iii) of Theorem 2.2 holds. Since we are concerned with the limit as h0, we assume that h(0,t). In this case, we have (23) 0|y~(a+h,t)y~(a,t)|da=0th|L(tah)π(a+h)L(ta)π(a)|da+thtL(ta)π(a)da0thL(tah)|π(a+h)π(a)|da+0th|L(tah)L(ta)|π(a)da+thtL(ta)π(a)da.(23) It follows from Equations (Equation18) and (Equation21) that (24) |L(a+h)L(a)|MLh,(24) where ML=K(Mv+Mx)+[β¯1βK2/(1+αK)+β¯1K2]+β¯1μ¯Kh+Mβ1K.

We therefore obtain from Equations (Equation23)–(Equation24) that 0|y~(a+h,t)y~(a,t)|daL¯0thπ(a)(1eaa+hμ(s)ds)da+MLh0thπ(a)da+L¯thtπ(a)daL¯0thπ(a)aa+hμ(s)dsda+MLh+L¯h(μ¯L¯+ML+L¯)h. Hence, the condition (iii) of Theorem 2.2 holds. By Theorem 2.1, the asymptotic smoothness of the semi-flow {Φ(t)}t0 follows. This completes the proof.

The following result is immediate from Theorem 2.33 in [Citation28] and Theorem 2.3.

Theorem 2.4

There exists a global attractor A of bounded sets in X.

3. Steady states and basic reproduction number

In this section, we calculate the basic reproduction number and study the existence of feasible steady states of system (Equation4) with the boundary condition (Equation5).

Clearly, system (Equation4) always has an infection-free steady state E1(s/d,0,0). If system (Equation4) has a chronic-infection steady state (x,y(a),v), it must satisfy the following equations: (25) sdxβxv1+αvx0β1(a)y(a)da=0,y(a)=μ(a)y(a),0k(a)y(a)dauv=0,y(0)=βxv1+αv+x0β1(a)y(a)da.(25) It follows from the second and the third equations of (Equation25) that (26) y(a)=y(0)π(a),y(0)=uv0k(a)π(a)da,(26) where π(a) is defined in Equation (Equation7).

We obtain from the first and the fourth equations of (Equation25) and (Equation26) that (27) x=1d(suv0k(a)π(a)da).(27) On substituting Equation (Equation26) into the fourth equation of (Equation25), it follows that (28) x=u(1+αv)β0k(a)π(a)da+u(1+αv)0β1(a)π(a)da.(28) From Equations (Equation27)–(Equation28) we obtain that Av2+Bv+C=0, where (29) A=αu20β1(a)π(a)da,B=α0k(a)π(a)da(du(1R0)+sβ0k(a)π(a)da)+βu0k(a)π(a)da+u20β1(a)π(a)da,C=du0k(a)π(a)da(1R0),(29) where R0=sβ0k(a)π(a)dadu+s0β1(a)π(a)dad. R0 is called the basic reproduction number of system (Equation4) representing the number of newly infected cells produced by one infected cell during its lifespan.

Hence, if R0>1, in addition to the infection-free steady state E1, system (Equation4) has a unique chronic-infection steady state E(x,y(a),v), where v=(B+B24AC)/(2A), x and y(a) are defined in Equations (Equation28) and (Equation26), respectively, A,B and C are defined in Equation (Equation29). It is easy to see that if R0=1, system (Equation4) has only the infection-free steady state E1.

4. Local stability

In this section, we are concerned with the local stability of each of feasible steady states of system (Equation4).

We first consider the local stability of the steady state E1(x0,0,0), where x0=s/d. Letting x(t)=x1(t)+x0,y(a,t)=y1(a,t),v(t)=v1(t), and linearizing system (Equation4) at E1, we obtain that (30) x˙1(t)=dx1(t)βx0v1(t)x00β1(a)y1(a,t)da,y1(a,t)t+y1(a,t)a=μ(a)y1(a,t),v˙1(t)=0k(a)y1(a,t)dauv1(t),y1(0,t)=βx0v1(t)+x00β1(a)y1(a,t)da.(30) Looking for solutions of system (Equation30) of the form x1(t)=x11eλt,y1(a,t)=y11(a)eλt,v1(t)=v11eλt, where x11,y11(a) and v11 will be determined later, we obtain the following linear eigenvalue problem: (31) (λ+d)x11=βx0v11x00β1(a)y11(a)da,y11(a)=(λ+μ(a))y11(a),(λ+u)v11=0k(a)y11(a)da,y11(0)=βx0v11+x00β1(a)y11(a)da.(31) It follows from Equation (Equation31) that (32) y11(a)=y11(0)e0a(λ+μ(s))ds,v11=0k(a)y11(a)daλ+u.(32) On substituting Equation (Equation32) into the fourth equation of (Equation31), we obtain the characteristic equation of system (Equation4) at E1 of the form (33) f(λ)=1,(33) where f(λ)=βx0λ+u0k(a)e0a(λ+μ(s))dsda+x00β1(a)e0a(λ+μ(s))dsda. It is easy to show that f(0)=R0, and f(λ)<0,limλ+f(λ)=0. Hence, if R0>1, f(λ)=1 has a unique positive root. Accordingly, if R0>1, the steady state E1 is unstable.

If R0<1, we claim that all roots of Equation (Equation33) have negative real parts. Otherwise, Equation (Equation33) has at least one root λ0 satisfying Reλ00. In this case, we have that |f(λ0)|βx0|λ0+u||0k(a)e0a(λ0+μ(s))dsda|+x0|0β1(a)e0a(λ0+μ(s))dsda|βx0u0k(a)e0aμ(s)dsda+x00β1(a)e0aμ(s)dsda=R0, a contradiction. Hence, if R0<1, E1(x0,0,0) is locally asymptotically stable.

We now study the local stability of the steady state E(x,y(a),v) of system (Equation4). Letting x(t)=x2(t)+x,y(a,t)=y2(a,t)+y(a),v(t)=v2(t)+v, and linearizing system (Equation4) at E, one obtains that (34) x˙2(t)=(d+βv1+αv+0β1(a)y(a)da)x2(t)βx(1+αv)2v2(t)x0β1(a)y2(a,t)da,y2(a,t)t+y2(a,t)a=μ(a)y2(a,t),v˙2(t)=0k(a)y2(a,t)dauv2(t),y2(0,t)=(βv1+αv+0β1(a)y(a)da)x2(t)+βx(1+αv)2v2(t)+x0β1(a)y2(a,t)da.(34) Looking for solutions of system (Equation34) of the form x2(t)=x21eλt, y2(a,t)=y21(a)eλt, v1(t)=v21eλt, where x21,y21(a) and v21 will be determined later, we obtain the following linear eigenvalue problem: (35) λx21=(d+βv1+αv+0β1(a)y(a)da)x21βx(1+αv)2v21x0β1(a)y21(a)da,y21(a)=(λ+μ(a))y21(a),λv21=0k(a)y21(a)dauv21,y21(0)=(βv1+αv+0β1(a)y(a)da)x21+βx(1+αv)2v21+x0β1(a)y21(a)da.(35) It follows from system (Equation35) that (36) (λ+d)x21=y2(0),y21(a)=y21(0)e0a(λ+μ(s))ds,(36) and (37) v21=0k(a)y21(a)daλ+u.(37) On substituting Equations (Equation36)–(Equation37) into the fourth equation of (Equation35), we obtain the characteristic equation of system (Equation4) at E as follows: (38) λ+d+βv1+αv+0β1(a)y(a)daλ+d=βx(1+αv)20k(a)e0a(λ+μ(s))dsdaλ+u+x0β1(a)e0a(λ+μ(s))dsda.(38) We claim that if R0>1, all roots of Equation (Equation38) has negative real parts. Otherwise, Equation (Equation38) has at least one root λ1 satisfying Reλ10. In this case, it is readily seen that |λ1+d+βv1+αv+0β1(a)y(a)da|>|λ1+d|.

On the other hand, the modulus of the right-hand side of Equation (Equation38) satisfies |βx(1+αv)20k(a)e0a(λ1+μ(s))dsdaλ1+u+x0β1(a)e0a(λ1+μ(s))dsda|βx1+αv|0k(a)e0a(λ1+μ(s))dsda||λ1+u|+x|0β1(a)e0a(λ1+μ(s))dsda|βx1+αv0k(a)π(a)dau+x0β1(a)π(a)da=1, a contradiction. Hence, if R0>1, E is locally asymptotically stable.

In conclusion, we have the following result.

Theorem 4.1

For system (Equation4) with the boundary condition (Equation5), if R0<1, the infection-free steady state E1(s/d,0,0) is locally asymptotically stable; if R0>1, E1 is unstable and the chronic-infection steady state E(x,y(a),v) exists and is locally asymptotically stable.

5. Uniform persistence

In this section, we establish the uniform persistence of the semi-flow {Φ(t)}t0 generated by system (Equation4) when R0>1.

Define a¯1=inf{a:ak(u)du=0},a¯2=inf{a:aβ1(u)du=0}. Noting that k(),β1()L+1(0,), we have a¯1>0,a¯2>0.

Denote X=L+1(0,+)×R+,a¯=max{a¯1,a¯2},Y~={(y(,t),v(t))X:0a¯y(a,t)da>0 or v(t)>0}, and Y=R+×Y~,Y=XY,Y~=XY~. Following [Citation13], the following result is immediate.

Proposition 5.1

The subsets Y and Y are both positively invariant under the semi-flow {Φ(t)}t0, namely, Φ(t,Y)Y and Φ(t,Y)Y for t0.

The following result is useful in proving the uniform persistence of the semi-flow {Φ(t)}t0 generated by system (Equation4).

Theorem 5.1

The infection-free steady state E1(s/d,0,0) is globally asymptotically stable for the semi-flow {Φ(t)}t0 restricted to Y.

Proof.

Let (x0,y0(),v0)Y. Then (y0(),v0)Y~. We consider the following system y(a,t)t+y(a,t)a=μ(a)y(a,t),v˙(t)=0k(a)y(a,t)dauv(t),y(0,t)=βx(t)v(t)1+αv(t)+x(t)0β1(a)y(a,t)da,y(a,0)=y0(a),v(0)=0. Since lim supt+x(t)s/d, by comparison principle, we have y(a,t)yˆ(a,t), v(t)vˆ(t), where yˆ(a,t) and vˆ(t) satisfy (39) yˆ(a,t)t+yˆ(a,t)a=μ(a)yˆ(a,t),vˆ˙(t)=0k(a)yˆ(a,t)dauvˆ(t),yˆ(0,t)=βsdvˆ(t)+sd0β1(a)yˆ(a,t)da,yˆ(a,0)=y0(a),vˆ(0)=0.(39) Solving the first equation of system (Equation39), we obtain that (40) yˆ(a,t)={Lˆ(ta)π(a),0a<t,y0(at)π(a)π(at),0ta,(40) where Lˆ(t)=yˆ(0,t)=(βs/d)vˆ(t)+(s/d)0β1(a)yˆ(a,t)da.

On substituting Equation (Equation40) into the second equation of system (Equation39), it follows that (41) vˆ˙(t)=0tk(a)Lˆ(ta)π(a)dauvˆ(t)+G1(t),Lˆ(t)=βsdvˆ(t)+sd0tβ1(a)Lˆ(ta)π(a)da+G2(t),G1(t)=tk(a)y0(at)π(a)π(at)da,G2(t)=tβ1(a)y0(at)π(a)π(at)da,vˆ(0)=0.(41) Since (y0(),v0)Y, we have G1(t)0 and G2(t)0 for all t0. We obtain from Equation (Equation41) that (42) vˆ˙(t)=0tk(a)Lˆ(ta)π(a)dauvˆ(t),Lˆ(t)=βsdvˆ(t)+sd0tβ1(a)Lˆ(ta)π(a)da,vˆ(0)=0.(42) It is easy to show that system (Equation42) has a unique solution vˆ(t)=0,Lˆ(t)=0. It follows from Equation (Equation40) that yˆ(a,t)=0 for 0a<t. For at, we have yˆ(a,t)L1=y0(at)π(a)π(at)L1eμ0ty0L1, which yields limt+yˆ(a,t)=0. By comparison principle, it follows that limt+v(t)=0,limt+y(a,t)=0. From the first equation of system (Equation4) we have limt+x(t)=s/d. This completes the proof.

Theorem 5.2

If R0>1, then the semi-flow {Φ(t)}t0 generated by system (Equation4) is uniformly persistent with respect to the pair (Y,Y); that is, there exists an ε>0 such that limt+Φ(t,x)Xε for xY. Furthermore, there is a compact subset A0Y which is a global attractor for {Φ(t)}t0 in Y.

Proof.

Since the infection-free steady state E1(s/d,0,0) is globally asymptotically stable in Y, applying Theorem 4.2 in [Citation5], we need only to show that Ws(E1)Y=, where Ws(E1)={xY:limt+Φ(t,x)=E1}. Otherwise, there exists a solution uY such that Φ(t,u)E1 as t. In this case, one can find a sequence {un}Y such that Φ(t,un)u¯X<1/n, t0, where u¯=(x0,0,0), x0=s/d.

Denote Φ(t,un)=(xn(t),yn(,t),vn(t)) and un=(xn(0),yn(,0),vn(0)). Since R0>1, one can choose n sufficiently large satisfying x01/n>0 and (43) β(x01n)1+αn0k(a)π(a)da+(x01n)0β1(a)π(a)da>1.(43) For such an n>0, there exists a T>0 such that for t>T, x01n<xn(t)<x0+1n,0vn(t)1n. Consider the following auxiliary system (44) y~(a,t)t+y~(a,t)a=μ(a)y~(a,t),v~˙(t)=0k(a)y~(a,t)dauv~(t),y~(0,t)=β(x01n)v~(t)1+αn+(x01n)0β1(a)y~(a,t)da.(44) It is easy to show that if (Equation43) holds, then system (Equation44) has a unique steady state E0(0,0). Looking for solutions of system (Equation44) of the form y~(a,t)=y~1(a)eλt, v~(t)=v~1eλt, where y~1(a) and v~1 will be determined later, we obtain the following linear eigenvalue problem: y~1(a)=(λ+μ(a))y~1(a),0k(a)y~1(a)da=(λ+u)v~1,y~1(0)=β(x01n)1+αnv~1+(x01n)0β1(a)y~1(a)da. We therefore obtain the characteristic equation of system (Equation44) at E0 of the form (45) f1(λ)=1,(45) where f1(λ)=β(x01n)1+αn0k(a)e0a(λ+μ(s))dsdaλ+u+(x01n)0β1(a)e0a(λ+μ(s))dsda. Clearly, limλ+f1(λ)=0. From Equation (Equation43) we have f1(0)>1. Hence, if R0>1, Equation (Equation45) has at least one positive root. This implies that the solution (y~(a,t),v~(t)) of system (Equation44) is unbounded. By comparison principle, the solution Φ(t,yn) of system (Equation4) is unbounded, which contradicts Proposition 2.2. Therefore, the semi-flow {Φ(t)}t0 generated by system (Equation4) is uniformly persistent. Furthermore, there is a compact subset A0Y which is a global attractor for {Φ(t)}t0 in Y. This completes the proof.

6. Global stability

In this section, we study the global stability of each of feasible steady states of system (Equation4). The strategy of proofs is to use suitable Lyapunov functionals and LaSalle's invariance principle.

We first state and prove a result on the global stability of the infection-free steady state E1(s/d,0,0) of system (Equation4).

Theorem 6.1

The infection-free steady state E1(s/d,0,0) of system (Equation4) is globally asymptotically stable if R0<1.

Proof.

Let (x(t),y(a,t),v(t)) be any positive solution of system (Equation4) with the boundary condition (Equation5) and the initial condition (Equation6). Denote x0=s/d.

Define V1(t)=x(t)x0x0lnx(t)x0+0F1(a)y(a,t)da+k1v(t), where the non-negative kernel function F1(a) and the positive constant k1 will be determined later.

Calculating the derivative of V1(t) along positive solutions of system (Equation4), we obtain that (46) ddtV1(t)=(1x0x(t))[sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da]+0F1(a)y(a,t)tda+k1[0k(a)y(a,t)dauv(t)].(46) On substituting s=dx0 and y(a,t)/t=μ(a)y(a,t)y(a,t)/a into Equation (Equation46), it follows that (47) ddtV1(t)=(1x0x(t))[d(x(t)x0)]βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da+βx0v(t)1+αv(t)+x00β1(a)y(a,t)da0F1(a)[μ(a)+y(a,t)a]da+k10k(a)y(a,t)dak1uv(t).(47) Using integration by parts, we derive from Equation (Equation47) that (48) (dv100) ddtV1(t)=d(x(t)x0)2x(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da+βx0v(t)1+αv(t)+x00β1(a)y(a,t)da+F1(a)y(a,t)|0+0(F1(a)μ(a)F1(a))y(a,t)da+k10k(a)y(a,t)dak1uv(t).(48) (dv100) Choose F1(a)=x0aβ1(u)eauμ(s)dsdu+k1ak(u)eauμ(s)dsdu. By calculation, we have (49) F1(0)=x00β1(a)π(a)da+k10k(a)π(a)da,F1(a)=x0β1(a)k1k(a)+μ(a)F1(a),lima+F1(a)=0.(49) We obtain from Equations (Equation48)– (Equation49) that (50) ddtV1(t)=d(x(t)x0)2x(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da+βx0v(t)1+αv(t)+x00β1(a)y(a,t)da+(x00β1(a)π(a)da+k10k(a)π(a)da)y(0,t)0(x0β1(a)+k1k(a))y(a,t)da+k10k(a)y(a,t)dak1uv(t).(50) On substituting Equation (Equation5) into Equation (Equation50), one has (51) ddtV1(t)=d(x(t)x0)2x(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da+βx0v(t)1+αv(t)k1uv(t)+(x00β1(a)π(a)da+k10k(a)π(a)da)βx(t)v(t)1+αv(t)+x(t)0β1(a)y(a,t)da(x00β1(a)π(a)da+k10k(a)π(a)da).(51) Noting that if R0<1, one can choose k1>0 satisfying x00β1(a)π(a)da+k10k(a)π(a)da=1. It therefore follows from Equation (Equation51) that ddtV1(t)=d(x(t)x0)2x(t)+v(t)1+αv(t)(u0k(a)π(a)da(R01)k1uαv(t)). Clearly, if R0<1, V1(t)0 holds and V1(t)=0 implies that x(t)=x0,v(t)=0. It is readily seen that the largest invariant subset of {V1(t)=0} is the singleton E1(x0,0,0). By Theorem 4.1, we see that if R0<1, E1 is locally asymptotically stable. Hence, the global asymptotic stability of E1 follows from LaSalle's invariance principle. This completes the proof.

Remark

From the proof of Theorem 6.1, we see that if R0=1, the infection-free steady state E1(s/d,0,0) is globally attractive.

In the following, we establish the global asymptotic stability of the chronic-infection steady state E(x,y(a),v) of system (Equation4).

Theorem 6.2

If R0>1, the chronic-infection steady state E(x,y(a),v) of system (Equation4) is globally asymptotically stable.

Proof.

Let (x(t),y(a,t),v(t)) be any positive solution of system (Equation4) with the boundary condition (Equation5) and the initial condition (Equation6).

Define V2(t)=xG(x(t)x)+0F(a)y(a)G(y(a,t)y(a))da+k2vG(v(t)v), where the function G(x)=x1lnx for x>0, the non-negative kernel function F(a) and the positive constant k2 will be determined later.

Calculating the derivative of V2(t) along positive solutions of system (Equation4), we obtain that (52) ddtV2(t)=(1xx(t))[sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da]+0F(a)y(a)tG(y(a,t)y(a))da+k2(1vv(t))[0k(a)y(a,t)dauv(t)]=(1xx(t))[sdx(t)βx(t)v(t)1+αv(t)x(t)0β1(a)y(a,t)da]+0F(a)(1y(a)y(a,t))y(a,t)tda+k2(1vv(t))[0k(a)y(a,t)dauv(t)].(52) On substituting s=dx+βxv/(1+αv)+x0β1(a)y(a)da and y(a,t)/t=μ(a)y(a,t)y(a,t)/a into Equation (Equation52), it follows that (53) ddtV2(t)=(1xx(t))[d(x(t)x)+βxv1+αv+x0β1(a)y(a)da]βx(t)v(t)1+αv(t)+βxv(t)1+αv(t)x(t)0β1(a)y(a,t)da+x0β1(a)y(a,t)da0F(a)(1y(a)y(a,t))(y(a,t)a+μ(a)y(a,t))da+k2[0k(a)y(a,t)dauv(t)vv(t)0k(a)y(a,t)da+uv].(53) A direct calculation shows that (54) y(a)aG(y(a,t)y(a))=(1y(a)y(a,t))(y(a,t)a+μ(a)y(a,t)).(54) On substituting Equation (Equation54) into Equation (Equation53), one obtains (55) ddtV2(t)=(1xx(t))[d(x(t)x)+βxv1+αv+x0β1(a)y(a)da]βx(t)v(t)1+αv(t)+βxv(t)1+αv(t)x(t)0β1(a)y(a,t)da+x0β1(a)y(a,t)da0F(a)y(a)aG(y(a,t)y(a))da+k2[0k(a)y(a,t)dauv(t)vv(t)0k(a)y(a,t)da+uv].(55) Using integration by parts, it follows from Equation (Equation55) that (56) ddtV2(t)=(1xx(t))[d(x(t)x)+βxv1+αv+x0β1(a)y(a)da]βx(t)v(t)1+αv(t)+βxv(t)1+αv(t)x(t)0β1(a)y(a,t)da+x0β1(a)y(a,t)daF(a)y(a)G(y(a,t)y(a))|0+0G(y(a,t)y(a))[F(a)y(a)+F(a)y(a)]da+k2[0k(a)y(a,t)dauv(t)vv(t)0k(a)y(a,t)da+uv].(56) Noting that y(a)=μ(a)y(a), we obtain from Equation (Equation56) that (57) ddtV2(t)=(1xx(t))[d(x(t)x)+βxv1+αv+x0β1(a)y(a)da]βx(t)v(t)1+αv(t)+βxv(t)1+αv(t)x(t)0β1(a)y(a,t)da+x0β1(a)y(a,t)daF(a)y(a)G(y(a,t)y(a))|0+0(F(a)μ(a)F(a))y(a)G(y(a,t)y(a))da+k2[0k(a)y(a,t)dauv(t)vv(t)0k(a)y(a,t)da+uv].(57) Choose k2=βx/u(1+αv) and F(a)=xaβ1(u)eauμ(s)dsdu+k2ak(u)eauμ(s)dsdu. By calculation, we have (58) F(0)=x0β1(u)e0uμ(s)dsdu+βxu(1+αv)0k(u)e0uμ(s)dsdu=1,(58) and (59) limaF(a)=0,F(a)=xβ1(a)k2k(a)+μ(a)F(a).(59) On substituting Equations (Equation58)–(Equation59) into Equation (Equation57), it follows that (60) ddtV2(t)=(1xx(t))[d(x(t)x)+βxv1+αv+x0β1(a)y(a)da]βx(t)v(t)1+αv(t)+βxv(t)1+αv(t)x(t)0β1(a)y(a,t)da+x0β1(a)y(a,t)day(0)G(y(0,t)y(0))0(xβ1(a)+k2k(a))y(a)G(y(a,t)y(a))da+k2[0k(a)y(a,t)dauv(t)vv(t)0k(a)y(a,t)da+uv].(60) Noting that y(0)=βxv/(1+αv)+x0β1(a)y(a)da and 0k(a)y(a)da=uv, we have from Equations (Equation5) and (Equation60) that (61) ddtV2(t)=(1xx(t))[d(x(t)x)](βxv1+αv+x0β1(a)y(a)da)[xx(t)1lnxx(t)]k20k(a)y(a)[vy(a,t)y(a)v(t)1lnvy(a,t)y(a)v(t)]da+βxv1+αv[(1+αv)v(t)v(1+αv(t))v(t)v1+1+αv(t)1+αv]βxv1+αv[1+αv(t)1+αv1ln1+αv(t)1+αv]βxv1+αvln1+αv(t)1+αvy(0)lny(0,t)y(0)(61) +0(xβ1(a)+k2k(a))y(a)lny(a,t)y(a)da(βxv1+αv+x0β1(a)y(a)da)lnxx(t)k20k(a)y(a)lny(a,t)y(a)vv(t)da=d(x(t)x)2x(t)αβx(v(t)v)2(1+αv)2(1+αv(t))(βxv1+αv+x0β1(a)y(a)da)[xx(t)1lnxx(t)]βxv1+αv[1+αv(t)1+αv1ln1+αv(t)1+αv]k20k(a)y(a)[vy(a,t)y(a)v(t)1lnvy(a,t)y(a)v(t)]da +βxv1+αvln(1+αv)y(0)x(t)v(t)xv(1+αv(t))y(0,t)+x0β1(a)y(a)lny(0)x(t)y(a,t)xy(a)y(0,t)da=d(x(t)x)2x(t)αβx(v(t)v)2(1+αv)2(1+αv(t))(βxv1+αv+x0β1(a)y(a)da)G(xx(t))βxv1+αvG(1+αv(t)1+αv)k20k(a)y(a)G(vy(a,t)y(a)v(t))daβxv1+αvG((1+αv)y(0)x(t)v(t)xv(1+αv(t))y(0,t))x0β1(a)y(a)G(y(0)x(t)y(a,t)xy(a)y(0,t))da. Since the function G(x)=x1lnx0 for all x>0 and G(x)=0 holds iff x=1. Hence, V2(t)0 holds. It is readily seen from Equation (Equation61) that V2(t)=0 if and only if x(t)=x,v(t)=v,y(a,t)y(a)=y(0,t)y(0),for all a0. It is easy to verify that the largest invariant subset of {V2(t)=0} is the singleton E. By Theorem 4.1, we see that if R0>1, E is locally asymptotically stable. Therefore, the global asymptotic stability of of E follows from LaSalle's invariance principle. This completes the proof.

7. Numerical simulations

In this section, we give some numerical simulations to illustrate the theoretical results in Sections 3 and 4. The backward Euler and linearized finite difference method will be used to discretize the ODEs and PDE in system (Equation4), and the integral will be numerically calculated using Simpson's rule.

We choose viral production rate as in [Citation24]: k(a)={0,a<a1,k(1eθ(aa1)),aa1, where the parameter θ determines how quickly k(a) reaches its saturation level k and a1 is the age at which reverse transcription is completed. Here, we choose a1=0.25 days, k=6.4201×103 day−1, θ=1.

The age-dependent per capita death rate of infected cells is chosen as in [Citation15]: μ(a)={δ0,a<a2,δ0+δm(1eγ(aa2)),aa2, where δ0+δm is the maximal death rate, describes the time to saturation and a2 is the delay between infection and the onset of cell-mediated killing. The term δ0 is a background death rate. Here, we choose δ0=0.05 day−1, δm=0.35 day−1, γ=0.5,a2=0, amax=15 days.

The other parameters in system (Equation4) are chosen as follows [Citation24]: s=104 ml−1 day−1, d=0.01 day−1, u=23 day−1, α=0.01.

If we choose β=2.4×108 ml day−1, β1(a)=106 ml day−1, then we have the basic reproduction number R0=21.7534. By Theorem 4.1, we see that in addition to the infection-free steady state E1(106,0,0), system (Equation4) has an endemic steady state E(2.7005×105,7299.5π(a),5.49×106) which is locally asymptotically stable. Numerical simulation illustrates this fact (see Figure ). In Figure , Y(t)=0300y(a,t)da.

Figure 1. The temporal solution found by numerical integration of system (Equation4) with the boundary condition (Equation5) and the initial condition x(0)=106 ml−1, v(0)=106 ml−1, and the parameters β=2.4×108 ml day−1, β1(a)=106 ml day−1.

Figure 1. The temporal solution found by numerical integration of system (Equation4(4) x˙(t)=s−dx(t)−βx(t)v(t)1+αv(t)−x(t)∫0∞β1(a)y(a,t)da,∂y(a,t)∂t+∂y(a,t)∂a=−μ(a)y(a,t),v˙(t)=∫0∞k(a)y(a,t)da−uv(t),(4) ) with the boundary condition (Equation5(5) y(0,t)=βx(t)v(t)1+αv(t)+x(t)∫0∞β1(a)y(a,t)da,t>0,(5) ) and the initial condition x(0)=106 ml−1, v(0)=10−6 ml−1, and the parameters β=2.4×10−8 ml day−1, β1(a)=10−6 ml day−1.

In order to evaluate the effect of cell-to-cell transmission on the virus dynamics, we let β1(a)=0 and other parameters remain unchanged. In this case, a direct calculation shows that the basic reproduction number R0=18.0506, and the endemic steady state becomes E(9.9977×105,2.2665π(a),1704.7). Comparing Figures  and , we see that the cell-to-cell transmission can significantly increase the virus load.

Figure 2. The temporal solution found by numerical integration of system (1.4) with the boundary condition (1.5) and initial condition x(0)=106 ml−1, v(0)=106 ml−1, and the parameters β=2.4×108 ml day−1, β1(a)=0 ml day−1.

Figure 2. The temporal solution found by numerical integration of system (1.4) with the boundary condition (1.5) and initial condition x(0)=106 ml−1, v(0)=10−6 ml−1, and the parameters β=2.4×10−8 ml day−1, β1(a)=0 ml day−1.

If we choose β=2.4×1010 ml day−1, β1(a)=9×109 ml day−1, then we have the basic reproduction number R0=0.2138. By Theorem 4.1, we see that system (Equation4) has only the infection-free steady state E1(106,0,0) which is locally asymptotically stable. Numerical simulation illustrates this fact (see Figure ). In Figure , Y(t)=0300y(a,t)da.

Figure 3. The temporal solution found by numerical integration of system (Equation4) with the boundary condition (Equation5) and the initial condition x(0)=106 ml−1, v(0)=106 ml−1, and the parameters β=2.4×1010 ml day−1, β1(a)=9×109 ml day−1.

Figure 3. The temporal solution found by numerical integration of system (Equation4(4) x˙(t)=s−dx(t)−βx(t)v(t)1+αv(t)−x(t)∫0∞β1(a)y(a,t)da,∂y(a,t)∂t+∂y(a,t)∂a=−μ(a)y(a,t),v˙(t)=∫0∞k(a)y(a,t)da−uv(t),(4) ) with the boundary condition (Equation5(5) y(0,t)=βx(t)v(t)1+αv(t)+x(t)∫0∞β1(a)y(a,t)da,t>0,(5) ) and the initial condition x(0)=106 ml−1, v(0)=10−6 ml−1, and the parameters β=2.4×10−10 ml day−1, β1(a)=9×10−9 ml day−1.

8. Conclusion

In this work, we have investigated an age structured within-host HIV-1 infection model with both virus-to-cell infection and direct cell-to-cell transmission. The model allows the production rate of viral particles and the death rate of productively infected cells to vary and depend on the infection age. By constructing suitable Lyapunov functionals and using LaSalle's invariance principle, it has been shown that global dynamics of system (Equation4) is completely determined by the basic reproduction number. It has been verified that if the basic reproduction number is less than unity, the infection-free steady state is globally asymptotically stable; if the basic reproduction number is greater than unity, the chronic-infection steady state is globally asymptotically stable. The global stability of the chronic-infection steady state rules out any possibility for the existence of Hopf bifurcations and sustained oscillations in system (Equation4).

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by the National Natural Science Foundation of China [No. 11371368] and the Natural Science Foundation of Hebei Province [No. A2014506015].

References

  • S. Bonhoeffer, R.M. May, G.M. Shaw, and M.A. Nowak, Virus dynamics and drug therapy, Proc. Natl. Acad. Sci. USA 94 (1997), pp. 6971–6976. doi: 10.1073/pnas.94.13.6971
  • P. Chen, W. Hübner, M.A. Spinelli, and B.K. Chen, Predominant mode of human immunodeficiency virus transfer between T cells is mediated by sustained Env-dependent neutralization-resistant virological synapses, J. Virol. 81 (2007), pp. 582–595.
  • D. Ebert, C.D. Zschokke-Rohringer, and H.J. Carius, Dose effects and density-dependent regulation of two microparasites of Daphnia magna, Oecologia 122 (2000), pp. 200–209. doi: 10.1007/PL00008847
  • J. Feldmann and O. Schwartz, HIV-1 virological synapse: Live imaging of transmission, Viruses 2 (2010), pp. 1666–1680. doi: 10.3390/v2081666
  • J.K. Hale and P. Waltman, Persistence in infinite dimensional systems, SIAM J. Math. Anal. 20 (1989), pp. 388–395. doi: 10.1137/0520025
  • D.D. Ho, A.U. Neumann, A.S. Perelson, W. Chen, J.M. Leonard, and M. Markowitz, Rapid turnover of plasma virions and CD4+ lymphocytes in HIV-1 infection, Nature 373 (1995), pp. 123–126. doi: 10.1038/373123a0
  • G. Huang, X. Liu, and Y. Takeuchi, Lyapunov functions and global stability for age-structured HIV infection model, SIAM J. Appl. Math. 72 (2012), pp. 25–38. doi: 10.1137/110826588
  • W. Hübner, G.P. McNerney, P. Chen, B.M. Dale, R.E. Gordon, F.Y.S. Chuang, X.-D. Li, D.M. Asmuth, T. Huser, and B.K. Chen, Quantitative 3D video microscopy of HIV transfer across T cell virological synapses, Science 323 (2009), pp. 1743–1747. doi: 10.1126/science.1167525
  • M. Iannelli, Mathematical Theory of Age-Structured Population Dynamics, Applied Mathematics Monographs, Vol. 7, Consiglio Nazionale delle Ricerche (C.N.R), Giardini Pisa, 1995, comitato nazionale per le scienze matematiche.
  • N.L. Komarova, D. Anghelina, I. Voznesensky, B. Trinite, D.N. Levy, and D. Wodarz, Relative contribution of free-virus and synaptic transmission to the spread of HIV-1 through target cell populations, Biol. Lett. 9 (2012), pp. 1049–1055. doi: 10.1098/rsbl.2012.1049
  • X. Lai and X. Zou, Modeling HIV-1 virus dynamics with both virus-to-cell infection and cell-to-cell transmission, SIAM J. Appl. Math. 74 (2014), pp. 898–917. doi: 10.1137/130930145
  • X. Lai and X. Zou, Modeling cell-to-cell spread of HIV-1 with logistic target cell growth, J. Math. Anal. Appl. 426 (2015), pp. 563–584. doi: 10.1016/j.jmaa.2014.10.086
  • P. Magal, Compact attractors for time periodic age-structured population models, Electron. J. Differential Equations 2001(65) (2001), pp. 1–35.
  • N. Martin and Q. Sattentau, Cell-to-cell HIV-1 spread and its implications for immune evasion, Curr. Opin. HIV AIDS 4 (2009), pp. 143–149. doi: 10.1097/COH.0b013e328322f94a
  • P.W. Nelson, M.A. Gilchrist, D. Coombs, J.M. Hyman, and A.S. Perelson, An age-structured model of HIV infection that allows for variations in the production rate of viral particles and the death rate of productively infected cells, Math. Biosci. Eng. 1 (2004), pp. 267–288. doi: 10.3934/mbe.2004.1.267
  • M. Nowak and R. May, Virus Dynamics, Oxford University Press, Oxford, 2000.
  • M.A. Nowak, R.M. Anderson, M.C. Boerlijst, S. Bonhoeffer, R.M. May, A.J. McMichael, S.M. Wolinsky, K.J. Kunstman, J.T. Safrit, R.A. Koup, A.U. Neumann, and B.T.M. Korber, HIV-1 evolution and disease progression, Science 274 (1996), pp. 1008–1011. doi: 10.1126/science.274.5289.1008
  • M.A. Nowak, S. Bonhoeffer, G.M. Shaw, and R.M. May, Anti-viral drug treatment: Dynamics of resistance in free virus and infected cell populations, J. Theoret. Biol. 184 (1997), pp. 203–217. doi: 10.1006/jtbi.1996.0307
  • A.S. Perelson and P.W. Nelson, Mathematical analysis of HIV-1 dynamics in vivo, SIAM Rev. 41 (1999), pp. 3–44. doi: 10.1137/S0036144598335107
  • A.S. Perelson, D.E. Kirschner, and R. De Boer, Dynamics of HIV infection of CD4+ T cells, Math. Biosci. 114 (1993), pp. 81–125. doi: 10.1016/0025-5564(93)90043-A
  • A.S. Perelson, A.U. Neumann, M. Markowitz, J.M. Leonard, and D.D. Ho, HIV-1 dynamics in vivo: Virion clearance rate, infected cell life-span, and viral generation time, Science 271 (1996), pp. 1582–1586. doi: 10.1126/science.271.5255.1582
  • R.R. Regoes, D. Ebert, and S. Bonhoeffer, Dose-dependent infection rates of parasites produce the Allee effect in epidemiology, Proc. R. Soc. Lond. B 269 (2002), pp. 271–279. doi: 10.1098/rspb.2001.1816
  • C. Reilly, S. Wietgrefe, G. Sedgewick, and A. Haase, Determination of simian immunodeficiency virus production by infected activated and resting cells, AIDS 21 (2007), pp. 163–168. doi: 10.1097/QAD.0b013e328012565b
  • L. Rong, Z. Feng, and A.S. Perelson, Mathematical analysis of age-structured HIV-1 dynamics with combination antiretroviral therapy, SIAM J. Appl. Math. 67(3) (2007), pp. 731–756. doi: 10.1137/060663945
  • Q. Sattentau, Avoiding the void: Cell-to-cell spread of human viruses, Nat. Rev. Microbiol. 6 (2008), pp. 815–826. doi: 10.1038/nrmicro1972
  • Q.J. Sattentau, Cell-to-cell spread of retroviruses, Viruses 2 (2010), pp. 1306–1321. doi: 10.3390/v2061306
  • A. Sigal, J.T. Kim, A.B. Balazs, E. Dekel, A. Mayo, R. Milo, and D. Baltimore, Cell-to- cell spread of HIV permits ongoing replication despite antiretroviral therapy, Nature 477 (2011), pp. 95–98. doi: 10.1038/nature10347
  • H.L. Smith and H.R. Thieme, Dynamical Systems and Population Persistence, Amer. Math. Soc., Providence, 2011.
  • M. Sourisseau, N. Sol-Foulon, F. Porrot, F. Blanchet, and O. Schwartz, Inefficient human immunodeficiency virus replication in mobile lymphocytes, J. Virol. 81 (2007), pp. 1000–1012. doi: 10.1128/JVI.01629-06
  • G. Webb, Theory of Nonlinear Age-Dependent Population Dynamics, Marcel Dekker, New York, 1985.