Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article Special Issues

Global dynamics of an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immunity response

  • In this paper, an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immune response is investigated. We give a rigorous mathematical analysis on some necessary technical materials, including the relative compactness and persistence of the solution semiflow, and existence of a global attractor. By subtle construction and estimates of a Lyapunov functional, we show that the global dynamics is determined by two sharp thresholds, namely, basic reproduction number 0 and immune-response reproduction number 1. When 0<1, the virus-free steady state is globally asymptotically stable, which means that the viruses are cleared and immune-response is not active; when 1<1<0, the immune-inactivated infection steady state exists and is globally asymptotically stable; and when 1>1, which implies that 0>1, the immune-activated infection steady state exists and is globally asymptotically stable. Numerical simulations are given to support our theoretical results.

    Citation: Ran Zhang, Shengqiang Liu. Global dynamics of an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immunity response[J]. Mathematical Biosciences and Engineering, 2020, 17(2): 1450-1478. doi: 10.3934/mbe.2020075

    Related Papers:

    [1] Cameron Browne . Immune response in virus model structured by cell infection-age. Mathematical Biosciences and Engineering, 2016, 13(5): 887-909. doi: 10.3934/mbe.2016022
    [2] Ting Guo, Zhipeng Qiu . The effects of CTL immune response on HIV infection model with potent therapy, latently infected cells and cell-to-cell viral transmission. Mathematical Biosciences and Engineering, 2019, 16(6): 6822-6841. doi: 10.3934/mbe.2019341
    [3] Jiazhe Lin, Rui Xu, Xiaohong Tian . Threshold dynamics of an HIV-1 model with both viral and cellular infections, cell-mediated and humoral immune responses. Mathematical Biosciences and Engineering, 2019, 16(1): 292-319. doi: 10.3934/mbe.2019015
    [4] Jinhu Xu, Yicang Zhou . Bifurcation analysis of HIV-1 infection model with cell-to-cell transmission and immune response delay. Mathematical Biosciences and Engineering, 2016, 13(2): 343-367. doi: 10.3934/mbe.2015006
    [5] Ali Moussaoui, Vitaly Volpert . The impact of immune cell interactions on virus quasi-species formation. Mathematical Biosciences and Engineering, 2024, 21(11): 7530-7553. doi: 10.3934/mbe.2024331
    [6] Jiawei Deng, Ping Jiang, Hongying Shu . Viral infection dynamics with mitosis, intracellular delays and immune response. Mathematical Biosciences and Engineering, 2023, 20(2): 2937-2963. doi: 10.3934/mbe.2023139
    [7] Shaoli Wang, Jianhong Wu, Libin Rong . A note on the global properties of an age-structured viral dynamic model with multiple target cell populations. Mathematical Biosciences and Engineering, 2017, 14(3): 805-820. doi: 10.3934/mbe.2017044
    [8] Patrick W. Nelson, Michael A. Gilchrist, Daniel Coombs, James M. Hyman, Alan 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. Mathematical Biosciences and Engineering, 2004, 1(2): 267-288. doi: 10.3934/mbe.2004.1.267
    [9] Xiaohong Tian, Rui Xu, Jiazhe Lin . Mathematical analysis of an age-structured HIV-1 infection model with CTL immune response. Mathematical Biosciences and Engineering, 2019, 16(6): 7850-7882. doi: 10.3934/mbe.2019395
    [10] Cuicui Jiang, Kaifa Wang, Lijuan Song . Global dynamics of a delay virus model with recruitment and saturation effects of immune responses. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1233-1246. doi: 10.3934/mbe.2017063
  • In this paper, an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immune response is investigated. We give a rigorous mathematical analysis on some necessary technical materials, including the relative compactness and persistence of the solution semiflow, and existence of a global attractor. By subtle construction and estimates of a Lyapunov functional, we show that the global dynamics is determined by two sharp thresholds, namely, basic reproduction number 0 and immune-response reproduction number 1. When 0<1, the virus-free steady state is globally asymptotically stable, which means that the viruses are cleared and immune-response is not active; when 1<1<0, the immune-inactivated infection steady state exists and is globally asymptotically stable; and when 1>1, which implies that 0>1, the immune-activated infection steady state exists and is globally asymptotically stable. Numerical simulations are given to support our theoretical results.


    Since a basic within-host viral infection model introduced by Nowak et al. [1], the dynamics of viral infection such as hepatitis B virus (HBV), hepatitis C virus (HCV) and human immunodeficiency virus (HIV) infection models have been widely studied by incorporating various biological factors. Consider age as a continuous variable, writing the production rate of viral particles and the death rate of productively infected cells as two continuous functions of age, Nelson et al. [2] studied a HIV infection model with infection-age, the model is described as follows:

    {dT(t)dt=Λμ1T(t)βT(t)V(t),(t+a)i(t,a)=δ(a)i(t,a),dV(t)dt=0p(a)i(t,a)daμ2V(t) (1.1)

    with the boundary and initial condition

    {i(t,0)=βT(t)V(t),T(0)=T0>0,  V(0)=V0>0  and  i(0,a)=i0(a)L1+(0,), (1.2)

    where T(t) and V(t) denote the densities of uninfected target cell and free viruses at time t, respectively; i(t,a) denote the density of infected cells at time t with infection-age a. The parameters of model (1.1) are biologically explained in Table 1.

    Table 1.  Parameters and their biological meaning in model (1.1). All these parameters are assumed to be positive.
    Parameter Interpretation
    Λ Constant recruitment rate;
    β Virus infection rate;
    μ1 Mortality rate of uninfected target cell;
    μ2 Mortality rate of free viruses;
    δ(a) Mortality rate of infected cell with age a;
    p(a) Production rate of viral particles.

     | Show Table
    DownLoad: CSV

    Nelson et al. analyzed the local stability of the model by evaluating eigenvalues and its related characteristic equation. In [3], Rong et al. extended the model with combination antiretroviral therapy, and analyzed the local stability of the model. Huang et al. [4] have been further investigated the global stability of the model (1.1) with (1.2) by using Lyapunov direct method and LaSalle invariance principle. For some recent works on viral models with age structure, we refer readers to the papers[5,6,7,8,9,10,11,12,13].

    Recently, experimental work [14] shows that direct cell-to-cell transmission also contributes to the viral persistence. In a more recent work [15], the authors reveals that environmental restrictions limit infection by cell-free virions but promote cell-associated HIV-1 transmission. In fact, cell-to-cell transmission could be also found in other viral infection for human and animals. For example, hepatitis C virus [16]; bovine viral diarrhea virus [17]; vaccinia virus [18]. Due to this fact, Lai and Zou [19] formulated a HIV-1 viral model with direct cell-to-cell transmission and studied the global threshold dynamics. Yang et al. [20] studied a cell-to-cell virus model with three distributed delays, they also obtained the global stability of each equilibrium for the model. Wang et al. [21] investigated an age-structured HIV model with virus-to-cell infection and cell-to-cell transmission, the model takes the following form:

    {dT(t)dt=Λμ1T(t)βT(t)V(t)0k(a)T(t)i(t,a)da,(t+a)i(t,a)=δ(a)i(t,a),dV(t)dt=0p(a)i(t,a)daμ2V(t), (1.3)

    with the boundary and initial condition

    {i(t,0)=βT(t)V(t)+0k(a)T(t)i(t,a)da,T(0)=T0>0,  V(0)=V0>0  and  i(0,a)=i0(a)L1+(0,). (1.4)

    By constructing suitable Lyapunov functional, Wang et al. were able to complete a global analysis for the model (1.3). In [22], Zhang and Liu studied the Hopf bifurcation of an age-structured HIV model with cell-to-cell transmission and logistic growth.

    In viral infection, the host immune system play a critical part on the progress of the infection. The role of the immune system is to fight off pathogenic organisms within the host, for example, cytotoxic T lymphocyte cells (CTLs) attack infected cells, and antibody cells attack viruses (humoral immunity response). In [23], Murase et al. studied an viral infection model with humoral immunity response:

    {dT(t)dt=Λμ1T(t)βT(t)V(t),dI(t)dt=βT(t)V(t)aI(t),dV(t)dt=arI(t)μ2V(t)kV(t)Z(t),dZ(t)dt=hV(t)Z(t)μ3Z(t), (1.5)

    where T(t), I(t), V(t) and Z(t) denote the densities of uninfected cells, infected cells, free viruses and humoral immunity response released by B cells, respectively; the viruses are removed at rate kZ by the humoral immunity response; the humoral immunity response are activated in proportion to hV(t) and removed at rate μ3. The global dynamics of model (1.5) were obtain in [23]. Consider the delay between viral entry into a cell and the maturation delay of the newly produced viruses, Wang et al. [24] studied a virus model with two delays and humoral immunity response. They established the global dynamics based on two threshold parameters, and they found that the three equilibria are globally asymptotically stable under some conditions. For another delay, which is the time that antigenic stimulation needs for generating immunity response, Wang et al. [25] considered another virus model with delay and humoral immunity response, they found that this delay could lead to a Hopf bifurcation at the infected equilibrium with immunity. In [26], Kajiwara et al. proposed a age-structured viral infection model contains humoral immunity response and the effect of absorption of pathogens into uninfected cells, they also proved the global stability of each equilibria. Duan and Yuan [30] considered an infection-age viral model with saturation humoral immune response, the local and global stability of this model are obtained. Additionally, for the virus model with CTL immune response, we refer readers to the papers [27,28,29,31,32,33,34] and the reference therein.

    Based on the above facts, we propose an age-structured viral infection model with cell-to-cell transmission and general humoral immune response in this paper. Precisely, we study the following model:

    {dT(t)dt=Λμ1T(t)βT(t)V(t)0k(a)T(t)i(t,a)da,(t+a)i(t,a)=δ(a)i(t,a),dV(t)dt=0p(a)i(t,a)daμ2V(t)qV(t)f(Z(t)),dZ(t)dt=cV(t)f(Z(t))μ3Z(t) (1.6)

    with the boundary and initial condition

    {i(t,0)=βT(t)V(t)+0k(a)T(t)i(t,a)da,T(0)=T0>0,  V(0)=V0>0,  Z(0)=Z0>0  and  i(0,a)=i0(a)L1+(0,), (1.7)

    where L1+ is the set of integrable functions from (0,+) into [0,+). T(t), V(t) and Z(t) denote the densities of uninfected target cell, free viruses and antibody responses released from B cells at time t, respectively; i(t,a) denotes the density of infected cells at time t with infection-age a; k(a) denote the infection rate of productively infected cells with age a; qV(t)f(Z(t)) is the neutralization rate of viruses and cV(t)f(Z(t)) is the activation rate of antibody responses. The antibody responses vanish at rate μ3. Other parameters of model (1.6) have the same biological meaning in the Table 1.

    We made the following assumption on parameters and nonlinear function f:RR.

    (A1) k(a), δ(a), θ(a), p(a), c(a)L+(0,), with respective essential supremums ˉk, ˉδ,ˉθ, ˉp, ˉc and respective essential infimums ˜k, ˜δ, ˜θ, ˜p, ˜c.

    (A2) f(Z)0 for Z0, f(Z)=0 if and only if Z=0; f is Lipschitz continuous on R+.

    (A3) f(Z) is differentiable such that f(Z)>0 and f(Z) is concave down on R+.

    Here are some examples on function f(Z) satisfies (A2) and (A3):

    (ⅰ) f(Z(t))=Z(t) which is the bilinear function (see [23]);

    (ⅱ) Saturation immune response function f(Z(t))=Z(t)h+Z(t) (see [30]).

    The paper is organized as follows. In Section 2, we introduce the existence and uniqueness of the solutions to system (1.6), the steady state and reproduction numbers are also determined in this section; In Section 3, we show that system (1.6) is asymptotically smooth; Section 4 is devoted to proving the local stability of each steady state; uniform persistence and global stability of each steady state is considered in Section 5; We perform a numerical simulation of a special case in Section 6; Section 7 provide some brief discussions.

    In this section, we show the existence and uniqueness of the solutions to system (1.6) by a standard method [36] (see also [37,38]), which is to rewrite system (1.6) as an abstract Cauchy problem.

    For convenience, we first denote the following notations.

    Γ(a)=ea0δ(τ)dτ,  P=0p(a)Γ(a)da,  K=0k(a)Γ(a)da.

    It is easy to see that

    Γ(0)=1   and   Γ(a)=δ(a)Γ(a).

    Set the following spaces:

    X:=R×L1(R+,R)×R×R, X+:=R+×L1+(R+,R)×R+×R+,

    with the following norm

    φ(),ϕ1,ϕ2,ϕ3X=φL1+|ϕ1|+|ϕ2|+|ϕ3|,

    Furthermore, define

    X0:={0}×L1(R+,R)×R×R×R, X0+:={0}×L1+(R+,R)×R×R+×R+,

    Let A:Dom(A)XX be the following linear operator:

    A((0φ)ϕ1ϕ2ϕ3)=((φ(0)φδφ)μ1ϕ1μ2ϕ2μ3ϕ3) (2.1)

    with Dom(A)=R×{0}×W1,1(0,+)×R×R. In the following, we apply the method in [36] since ¯Dom(A)=X0 is not dense in X. Consider the nonlinear map F:Dom(A)X defined by

    F((0φ)ϕ1ϕ2ϕ3)=((βϕ1ϕ2+0k(a)ϕ1φ(a)da0)Λβϕ1ϕ20k(a)ϕ1φ(a)da0p(a)φ(a)daqϕ2f(ϕ3)cϕ2f(ϕ3)).

    One can see that F is Lipschitz continuous on bounded sets. Let

    u(t)=(T(t),(0i(t,)),V(t),Z(t))T,

    where T represents transposition. Then we can rewrite system (1.6) as the following abstract Cauchy problem:

    {du(t)dt=Au(t)+F(u(t)),  t0,u(0)=u0X0+. (2.2)

    In order to use the method in [36], we need to show that A is a Hille-Yosida operator. Denote ρ(A) be the resolvent set of A. The definition of Hille-Yosida operator is:

    Definition 2.1. (See [36,Definition 2.4.1]) A linear operator A:Dom(A)XX on a Banach space (X,) (densely defined or not) is called a Hille-Yosida operator if there exist real constants M1, and ωR, such that (ω,+)ρ(A), and

    (λA)nM(λω)n,    for  nN+   and all  λ>ω.

    Now, we prove the following lemma.

    Lemma 2.1. The operator A defined in (2.1) is a Hille-Yosida operator.

    Proof. Let

    (λIA)1((ˆφ0ˆφ(a))ˆϕ1ˆϕ2ˆϕ3)=((0φ)ϕ1ϕ2ϕ3),

    by some simple calculations, we have

    ϕ1=^ϕ1λ+μ1,  ϕ2=^ϕ2λ+μ2,  ϕ3=^ϕ3λ+μ3

    and

    φ(a)=ˆφ0ea0(λ+δ(s))ds+0ˆφ(τ)eaτ(λ+δ(s))dsdτ.

    Denote ζ=((ˆφ0ˆφ(a)),ˆϕ1,ˆϕ2,ˆϕ3)T, then

    (λIA)1ζX=|0|+0φ(a)da+|ϕ1|+|ϕ2|+|ϕ3|=0φ(a)da+|ˆϕ1||λ+μ1|+|ˆϕ2||λ+μ2|+|ˆϕ3||λ+μ3||ˆφ0||λ+μ|+ˆφ(a)L1|λ+μ|+|ˆϕ1||λ+μ|+|ˆϕ2||λ+μ|+|ˆϕ3||λ+μ|=1λ+μζX.

    where μ=min. By the Definition 2.1, the operator A is a Hille-Yosida operator. This ends the proof.

    Let X_0 = \left(T_0, \left(\begin{array}{c} 0 \\ i_0 \\ \end{array} \right), V_0, Z_0\right)^ \text{T}\in \mathcal{X}_{0+}, by using [36,Theorem 5.2.7] (see also in [37,39]), we have the following theorem.

    Theorem 2.1. There exists a uniquely determined semi-flow \{U(t)\}_{t\geqslant 0} on \mathcal{X}_{0+} such that for each X_0 , there exists a unique continuous map U\in C([0, +\infty), \mathcal{X}_{0+}) which is an integrated solution of Cauchy problem (2.2), that is

    \begin{equation} \left\{ \begin{array}{l} \int_0^t U(s) X_0 \text{d} s\in \text{Dom} (A), \ \ \forall t \geqslant 0, \\ U(t)X_0 = X_0 + A \int_0^tU(s)X_0 \text{d} s +\int_0^\infty F(U(s)X_0) \text{d} s, \ \ \forall t \geqslant 0. \end{array}\right. \end{equation} (2.3)

    Let

    \begin{equation} \Omega = \left\{(T, (0, i(\cdot)), V, Z)\in \mathcal{X}_{0+}\ \bigg|\ T(t) + \int_0^{+\infty}i(t, a) \text{d} a \leqslant \frac{\Lambda}{\mu_0}, \ V(t) + \frac{q}{c}Z(t)\leqslant \frac{\Lambda \bar{p}}{\mu_0\hat{\mu}}\right\}, \end{equation} (2.4)

    where \mu_0 = \min\{\mu_1, \tilde{\delta}\} and \hat{\mu} = \min\left\{\mu_2, \mu_3\right\} . We show that \Omega is a positively invariant set under semi-flow \{U(t)\}_{t\geqslant 0}.

    Theorem 2.2. \Omega is a positively invariant set under semi-flow \{U(t)\}_{t\geqslant 0}. Moreover the semi-flow \{U(t)\}_{t\geqslant 0} is point dissipative and \Omega attracts all positive solutions of (2.2) in \mathcal{X}_{0+} .

    Proof. Integrating the second equation of (1.6) along the characteristic line t-a = constant, yields

    \begin{equation} i(t, a) = \left\{ \begin{array}{l} i(t-a, 0)\Gamma(a), \ \ t \gt a \gt 0, \\ i_0(a-t)\frac{\Gamma(a)}{\Gamma(a-t)}, \ \ a \gt t \gt 0. \end{array}\right. \end{equation} (2.5)

    Then

    \begin{align*} \int_0^\infty i(t, a) \text{d} a & = \int_0^t i(t-a, 0)\Gamma(a) \text{d} a + \int_t^\infty i_0(a-t)\frac{\Gamma(a)}{\Gamma(a-t)} \text{d} a\\ & = \int_0^t i(\sigma, 0)\Gamma(t-\sigma) \text{d} \sigma + \int_0^\infty i_0(a)\frac{\Gamma(t+a)}{\Gamma(a)} \text{d} a. \end{align*}

    Note that \Gamma(0) = 1 and \Gamma'(a) = -\delta(a)\Gamma(a) , thus

    \begin{align*} \frac{ \text{d}}{ \text{d} t}\int_0^\infty i(t, a) \text{d} a & = \frac{\partial}{\partial t}\int_0^t i(\sigma, 0)\Gamma(t-\sigma) \text{d} \sigma + \frac{ \text{d}}{ \text{d} t}\int_0^\infty i_0(a)\frac{\Gamma(t+a)}{\Gamma(a)} \text{d} a\\ & = i(t, 0) - \int_0^\infty \delta(a)i(t, a) \text{d}a. \end{align*}

    One has that

    \begin{align*} \frac{ \text{d}}{ \text{d}t} \left(T(t)+\int_0^{+\infty}i(t, a) \text{d}a\right) = &\Lambda -\mu_1 T(t) - \int_0^\infty \delta(a) i(t, a) \text{d}a\\ \leqslant &\Lambda -\mu_0\left( T(t) - \int_0^\infty \delta(a)i(t, a) \text{d}a\right). \end{align*}

    We have

    \limsup\limits_{t\rightarrow\infty} \left\{T(t) + \int_0^{\infty}i(t, a) \text{d}a\right\} \leqslant \frac{\Lambda}{\mu_0}, \ \ t\geqslant0.

    From the third and forth equations of (1.6), it is easy to check

    \limsup\limits_{t\rightarrow\infty} \left(V(t) + \frac{q}{c}Z(t)\right)\leqslant \frac{\Lambda \bar{p}}{\mu_0\hat{\mu}}, \ \ t\geqslant0.

    Hence

    \|U(t)X_0\|_{\mathcal{X}_+}\leqslant \Pi,

    where \Pi = \frac{\Lambda}{\mu_0}\left(1 + \frac{\bar{p}}{\hat{\mu}} + \frac{c\bar{P}}{q\hat{\mu}}\right) . Therefore, for any solution of (2.2) satisfying X_0\in \Omega and U(t)X_0\in \Omega for all t\geqslant0 , \Omega is a positively invariant set under semi-flow \{U(t)\}_{t\geqslant 0}. Moreover the semi-flow \{U(t)\}_{t\geqslant 0} is point dissipative and \Omega attracts all positive solutions of (2.2) in \mathcal{X}_{0+} .

    In this subsection, we concern with the existence of steady states for system (1.6). Obviously, the system (1.6) always has a virus-free steady state E_0 = (T_0, 0, 0, 0) = (\frac{\Lambda}{\mu_1}, 0, 0, 0) . E^0 is the unique equilibrium if \Re_0 \leqslant 1 , where \Re_0 = \frac{\beta T_0 \mathcal{P}}{\mu_2} + T_0 \mathcal{K} is the basic reproduction number of system (1.6). If \Re_0 > 1 , there exists an immune-inactivated infection steady state E_1 = (T_1^*, i_1^*(a), V_1^*, 0), which is the same situation in [21], that is,

    \begin{equation} T_1^* = \frac{T_0}{\Re_0}, \ \ \ i_1^*(a) = \Lambda\left(1-\frac{1}{\Re_0}\right)\Gamma(a), \ \ \ V_1^* = \frac{1}{\mu_2}\int_0^\infty p(a)i_1^*(a) \text{d}a. \end{equation} (2.6)

    There also exists another immune-activated infection steady state E_2 = (T_2^*, i_2^*(a), V_2^*, Z_2^*), which is satisfies

    \begin{equation} \left\{ \begin{array}{l} \Lambda - \mu_1 T_2^* = i_2^*(0) = \beta T_2^* V_2^* + \int_0^\infty k(a)T_2^*i_2^*(a) \text{d}a, \\ \frac{ \text{d} i_2^*(a)}{ \text{d} {a}} = -\delta(a) i_2^*(a), \\ \int_0^\infty p(a) i_2^*(a) \text{d}a - \varpi(Z_2^*) V_2^* = 0, \\ c V_2^* f(Z_2^*) - \mu_3 Z_2^* = 0, \end{array}\right. \end{equation} (2.7)

    where

    \varpi(Z_2^*) = \mu_2 + {q}f(Z_2^*).

    By some calculations, we have

    \begin{align} &i_2^*(a) = i_2^*(0) \Gamma(a). \end{align} (2.8)

    From the third equation of (2.7), yields

    \begin{equation} V_2^* = \frac{\int_0^\infty p(a) i_2^*(a) \text{d}a}{\varpi(Z_2^*)}. \end{equation} (2.9)

    Substituting (2.8) and (2.9) into the first equation of (2.7) one has that

    \begin{equation} T_2^* = \frac{\varpi(Z_2^*)}{\beta \mathcal{P} + \varpi(Z_2^*) \mathcal{K}} \end{equation} (2.10)

    and

    \begin{equation} i_2^*(0) = \Lambda - \mu_1 T_2^* = \Lambda - \frac{\varpi(Z_2^*)}{\beta \mathcal{P} + \varpi(Z_2^*) \mathcal{K}}. \end{equation} (2.11)

    In the following, we show that Z_2^* > 0 . In fact, combining the last two equations of (2.7) give us

    \begin{equation} \int_0^\infty p(a)\left(\Lambda - \frac{\mu_1\varpi(Z_2^*)}{\beta \mathcal{P} + \varpi(Z_2^*) \mathcal{K}}\right)\Gamma(a) \text{d}a - \frac{\mu_3 Z_2^* \varpi(Z_2^*)}{c f(Z_2^*)} = 0. \end{equation} (2.12)

    Denote

    \Phi(Z^*) = \int_0^\infty p(a)\left(\Lambda - \frac{\mu_1\varpi(Z_2^*)}{\beta \mathcal{P} + \varpi(Z_2^*) \mathcal{K}}\right)\Gamma(a) \text{d}a - \frac{\mu_3 Z_2^* \varpi(Z_2^*)}{c f(Z_2^*)}

    and

    \Re_1 : = \frac{\Lambda c \mathcal{P} f'(0)}{\mu_2\mu_3}\left(1-\frac{1}{\Re_0}\right).

    Then

    \lim\limits_{Z_2^*\rightarrow0}\Phi(Z_2^*) \gt 0 \Leftrightarrow \Re_1 \gt 1.

    It is easy to check \frac{ \text{d} \Phi(Z^*)}{ \text{d} Z^*} < 0 and \lim_{Z^* \rightarrow +\infty}\Phi(Z^*) \rightarrow -\infty . Hence, there is only one positive root for (2.12) if \Re_1 > 1 . By the expressions of \Re_0 and \Re_1 , we have \Re_1 > 0 \Rightarrow \Re_0 > 1, then there is the following theorem on the existence of steady states.

    Theorem 2.3. For system (1.6), there are two threshold parameters \Re_0 and \Re_1 such that

    (ⅰ) if \Re_0 \leq 1, there exists only one positive steady state E_0 ;

    (ⅱ) if \Re_1 < 1 < \Re_0, there exists two positive steady states E_0 and E_1^* ;

    (ⅲ) if \Re_1 > 1, there exists three positive steady states E_0, E_1^* and E_2^* .

    The following lemma on immune-inactivated infection steady state and immune-activated infection steady state will be used in the proof of global stability.

    Lemma 2.2. The immune-inactivated infection steady state (T_1^*, i_1^*(a), V_1^*, 0) satisfies

    \begin{align} \int_0^\infty \left[\frac{\beta T_1^*}{\mu_2}p(a)i_1^*(a)\left(1-\frac{i_1^*(0)T(t)V(t)}{i(t, 0)T_1^*V_1^*}\right) + T_1^*k(a)i_1^*(a)\left(1-\frac{i_1^*(0)T(t)i(t, a)}{i_1(t, 0)T_1^*i_1^*(a)}\right)\right] \text{d}a = 0, \end{align} (2.13)

    and immune-activated infection steady state (T_2^*, i_2^*(a), V_2^*, Z_2^*) satisfies

    \begin{align} \int_0^\infty \left[\frac{\beta T_2^*}{\mu_2+qf(Z_2^*)}p(a)i_2^*(a)\left(1-\frac{i_2^*(0)T(t)V(t)}{i(t, 0)T_2^*V_2^*}\right) + T_2^*k(a)i_2^*(a)\left(1-\frac{i_2^*(0)T(t)i(t, a)}{i_2(t, 0)T_2^*i_2^*(a)}\right)\right] \text{d}a = 0. \end{align} (2.14)

    Proof. For the immune-inactivated infection steady state (T_1^*, i_1^*(a), V_1^*, 0) , it follows from the third equation of (2.6), we have

    \begin{align*} \int_0^\infty \frac{\beta T_1^*}{\mu_2}p(a)i_1^*(a)\frac{i_1^*(0)T(t)V(t)}{i(t, 0)T_1^*V_1^*} \text{d}a = &\ \frac{\beta i_1^*(0)T(t)V(t)}{i(t, 0)\mu_2 V_1^*}\int_0^\infty p(a)i_1^*(a) \text{d}a\\ = &\ \beta T(t)V(t) \frac{i^*(0)}{i(t, 0)}. \end{align*}

    Recall that i(t, 0) = \beta T(t) V(t) + \int_0^\infty k(a)T(t)i(t, a) \text{d}a in (1.7), hence

    \begin{align*} &\ \int_0^\infty \frac{\beta T_1^*}{\mu}p(a)i_1^*(a)\frac{i_1^*(0)T(t)V(t)}{i(t, 0)T_1^*V_1^*} \text{d}a + \int_0^\infty T_1^*k(a)i_1^*(a)\frac{i_1^*(0)T(t)i(t, a)}{i_i(t, 0)T_1^*i_1^*(a)} \text{d}a\\ = &\ \beta T(t)V(t) \frac{i_1^*(0)}{i(t, 0)} + T(t) \int_0^\infty k(a) i(t, a) \text{d}a \frac{i_1^*(0)}{i(t, 0)}\\ = &\ i_1^*(0)\\ = &\ \beta T_1^* V_1^* + \int_0^\infty k(a)T_1^*i_1^*(a) \text{d}a, \end{align*}

    Thus, (2.13) holds true. The proof of Eq (2.14) is similar to (2.13), so we omitted it. This ends the proof.

    In this section, we show that the semi-flow \{U(t)\}_{t\geqslant0} is asymptotically smooth. Since the state space \mathcal{X}_{0+} is the infinite dimensional Banach space, we need the semi-flow \{U(t)\}_{t\geqslant0} is asymptotically smooth to proof the global stability of each steady states. Rewrite U: = \Phi+\Psi , where

    \begin{align} \Phi(t)X_0:& = (0, \varpi_1(\cdot, t), 0, 0), \end{align} (3.1)
    \begin{align} \Psi(t)X_0:& = (T(t), \varpi_2(\cdot, t), V(t), Z(t)), \end{align} (3.2)

    with

    \begin{equation*} \varpi_1(\cdot, t) = \left\{ \begin{array}{l} 0, \ \ \ \ \ \ \ \ \ t \gt a\geqslant0, \\ i(t, a), \ \ \ a\geqslant t\geqslant0, \end{array}\right. \ \ \ \text{and}\ \ \ \varpi_2(\cdot, t) = \left\{ \begin{array}{l} i(t, a), \ \ \ t \gt a\geqslant0, \\ 0, \ \ \ \ \ \ \ \ \ a\geqslant t\geqslant0. \end{array}\right. \end{equation*}

    We are now in the position to prove the following theorem.

    Theorem 3.1. For any X_0\in \Omega , \{U(t)X_0: t\geqslant0\} has compact closure in \mathcal{X} if the following two conditions hold:

    (ⅰ) There exists a function \Delta: \mathbb{R}_+\times\mathbb{R}_+\rightarrow\mathbb{R}_+ such that for any r > 0 , \lim_{t\rightarrow\infty}\Delta(t, r) = 0 , and if X_0\in\Omega with \|X_0\|_{\mathcal{X}}\leqslant r , then \|\Phi(t)X_0\|_{\mathcal{X}}\leqslant \Delta (t, r) for t\geqslant0 ;

    (ⅱ) For t\geqslant0 , \Psi(t)X_0 maps any bounded sets of \Omega into sets with compact closure in \mathcal{X} .

    Proof. Step Ⅰ, to show that (ⅰ) holds.

    Let \Delta(t, r): = e^{-\tilde{\delta}t}r , it is obvious that \lim_{t\rightarrow\infty}\Delta(t, r) = 0 . Then for X_0\in\Omega satisfying \|X_0\|_{\mathcal{X}}\leqslant r , we have

    \begin{align*} \|\Phi(t)X_0\|_{\mathcal{X}}& = |0| + \int_0^\infty|\varpi_1(a, t) \text{d}a| + |0| + |0|\\ & = \int_t^\infty\left|i_0(a-t)\frac{\Gamma(a)}{\Gamma(a-t)}\right| \text{d}a\\ & = \int_0^\infty\left|i_0(s)\frac{\Gamma(s+t)}{\Gamma(s)}\right| \text{d}s\\ &\leqslant e^{-\tilde{\delta}t}\int_0^\infty |i_0(s)| \text{d}s\\ &\leqslant e^{-\tilde{\delta}t} \|X_0\|_{\mathcal{X}}\\ &\leqslant \Delta(t, r), \ \ t\geqslant0. \end{align*}

    This completes the proof of condition (ⅰ).

    Step Ⅱ, to show that (ⅱ) holds.

    We just have to show that \varpi_2(t, a) remains in a precompact subset of L_+^1(0, \infty). In order to prove it, we should show the following conditions hold [40,Theorem B.2]:

    (a) The supremum of \int_0^\infty \varpi_2(t, a) \text{d}a with respect to X_0\in \Omega is finite;

    (b) \lim_{u\rightarrow\infty} \int_u^\infty \varpi_2(t, a) \text{d}a = 0 uniformly with respect to X_0\in \Omega ;

    (c) \lim_{u\rightarrow0^+} \int_0^\infty (\varpi_2(t, a+u)-\varpi_2(t, a) \text{d}a = 0 uniformly with respect to X_0\in \Omega ;

    (d) \lim_{u\rightarrow0^+} \int_u^\infty \varpi_2(t, a) \text{d}a = 0 uniformly with respect to X_0\in \Omega .

    Conditions (a), (b) and (d) hold since \varpi_2(t, a)\leqslant \left(\frac{\beta \bar{p}}{\hat{\mu}}+\bar{k}\right)\frac{\Lambda^2}{\mu_0^2} . Next, we verify condition (c). For sufficiently small u\in(0, t), set K(t) = \int_0^\infty k(a)i(t, a) \text{d}a , we have

    \begin{align*} &\int_0^\infty |\varpi_2(t, a+u)-\varpi_2(t, a)| \text{d}a\\ = &\int_0^{t-u}|(\beta T(t-a-u)V(t-a-u)+K(t-a-u)T(t-a-u))\Gamma(a+u)\\ &-(\beta T(t-a)V(t-a)+K(t-a)T(t-a))\Gamma(a)| \text{d}a\\ &+\int_{t-u}^t|0-\beta T(t-a)V(t-a)+K(t-a)T(t-a))\Gamma(a)| \text{d}a\\ \leqslant&\int_0^{t-u}(\beta T(t-a-u)V(t-a-u)+K(t-a-u)T(t-a-u))|\Gamma(a+u)-\Gamma(a)| \text{d}a\\ &+\int_0^{t-u}|\beta T(t-a-u)V(t-a-u)-\beta T(t-a)V(t-a)|\Gamma(a) \text{d}a\\ &+\int_0^{t-u}|K(t-a-u)T(t-a-u)-K(t-a)T(t-a)|\Gamma(a) \text{d}a + u\left(\frac{\Lambda}{\mu_0}\right)^2\left(\frac{\beta \bar{p}}{\hat{\mu}}+\tilde{k}\right). \end{align*}

    Since \Gamma(a) is non-increasing function with respect to a and 0\leqslant\Gamma(a)\leqslant1 , we have

    \begin{align*} \int_0^{t-u} |\Gamma(a+u)-\Gamma(a)| \text{d}a & = \int_0^{t-u} \left(\Gamma(a) - \Gamma(a+u)\right) \text{d}a\\ & = \int_0^{t-u} \Gamma(a) \text{d}a - \int_u^t \Gamma(a) \text{d}a\\ &\leqslant\int_0^{t-u} \Gamma(a) \text{d}a + \int_{t-u}^u \Gamma(a) \text{d}a\leqslant u. \end{align*}

    Then

    \int_0^\infty |\varpi_2(t, a+u)-\varpi_2(t, a)| \text{d}a \leqslant 2u\left(\frac{\Lambda}{\mu_0}\right)^2\left(\frac{\beta \bar{p}}{\hat{\mu}}+\bar{k}\right) + \Xi,

    where

    \begin{align*} \Xi = &\int_0^{t-u}|\beta T(t-a-u)V(t-a-u)-\beta T(t-a)V(t-a)|\Gamma(a) \text{d}a\\ &+\int_0^{t-u}|K(t-a-u)T(t-a-u)-K(t-a)T(t-a)|\Gamma(a) \text{d}a. \end{align*}

    Thanks to the argument in [41,Proposition 6], T(\cdot)V(\cdot) and T(\cdot)K(\cdot) are Lipschitz on \mathbb{R}_+ . Let M_1 and M_2 be the Lipschitz coefficients of T(\cdot)V(\cdot) and T(\cdot)K(\cdot) respectively. Then

    \Xi\leqslant (\beta M_1+M_2)u\int_0^{t-u}\Gamma(a) \text{d}a\leqslant(\beta M_1+M_2)u\int_0^{t-u}\Gamma(a) \text{d}a\leqslant \frac{u(\beta M_1+M_2)}{\tilde{\delta}}.

    Hence

    \int_0^\infty |\varpi_2(t, a+u)-\varpi_2(t, a)| \text{d}a \leqslant 2u\left(\frac{\Lambda}{\mu_0}\right)^2\left(\frac{\beta \bar{p}}{\mu_2}+\bar{k}\right) + \frac{u(\beta M_1+M_2)}{\tilde{\delta}},

    which converges to 0 as u\rightarrow0^+ , the condition (c) holds. Let \mathcal{Y}\subset\mathcal{X} be a bounded closed set and B be a bound for \mathcal{Y} , where B\geqslant A . It is easy to check M_1 and M_2 are only depend on A , that is M_1 and M_2 are independent on \mathcal{X} . Consequently, \varpi_2(t, a) remains in a precompact subset \mathcal{Y} of L_1^+(0, +\infty) and thus

    \Psi(t, \mathcal{Y})\subseteq [0, B] \times Y \times [0, B] \times [0, B],

    which has compact closure in \mathcal{X} . The proof is completed.

    In this section, we show the local stability of system (1.6) at each steady states.

    Theorem 4.1. If \Re_0 < 1 , then the virus-free steady state E_0 of system (1.6) is locally asymptotically stable.

    Proof. Denote \bar{T}_1(t) = T(t) - T_0 , \bar{i}_1(t, a) = i(t, a) , \bar{V}_1(t) = V(t) and \bar{Z}_1(t) = Z(t) , the linearized equation of (1.6) at E_0 as follows:

    \begin{equation} \left\{ \begin{array}{l} \frac{ \text{d} \bar{T}_1(t)}{ \text{d}t} = - \mu_1 \bar{T}_1(t) - \beta T_0 \bar{V}_1(t) - \int_0^\infty T_0 k(a)\bar{i}_1(t, a) \text{d}a, \\ \left(\frac{\partial}{\partial t} + \frac{\partial}{\partial a}\right) \bar{i}_1(t, a) = -\delta(a) \bar{i}_1(t, a), \\ \frac{ \text{d} \bar{V}_1(t)}{ \text{d}t} = \int_0^\infty p(a) \bar{i}_1(t, a) \text{d}a - \mu_2 \bar{V}_1(t), \\ \frac{ \text{d} \bar{Z}_1(t)}{ \text{d}t} = -\mu_3\bar{Z}_1(t), \\ {\bar{i}_1(t, 0)} = \beta T_0 \bar{V}_1(t) + \int_0^\infty T_0 k(a)\bar{i}_1(t, a) \text{d}a. \end{array}\right. \end{equation} (4.1)

    Let the solution of (4.1) has the following exponential form:

    \bar{T}_1(t) = \bar{T}_1 e^{\lambda t}, \ \ \bar{V}_1(t) = \bar{V}_1 e^{\lambda t}, \ \ \bar{Z}_1(t) = \bar{Z}_1 e^{\lambda t}\ \ \text{and}\ \ \bar{i}_1(t, a) = \bar{i}_1(a) e^{\lambda t},

    then

    \begin{equation} \left\{ \begin{array}{l} \lambda \bar{T}_1 = - \mu_1 \bar{T}_1 - \beta T_0 \bar{V}_1 - \int_0^\infty T_0 k(a)\bar{i}_1(a) \text{d}a, \\ \lambda \bar{i}_1(a) + \frac{ \text{d} \bar{i}_1(a)}{ \text{d}a} = -\delta(a) \bar{i}_1(a), \\ \lambda \bar{V}_1 = \int_0^\infty p(a) \bar{i}_1(a) \text{d}a - \mu_2 \bar{V}_1, \\ \lambda \bar{Z}_1 = -\mu_3\bar{Z}_1, \\ \bar{i}_1(0) = \beta T_0 \bar{V}_1 + \int_0^\infty T_0 k(a)\bar{i}_1(a) \text{d}a. \end{array}\right. \end{equation} (4.2)

    Solve the second equation of (4.2) yields

    \bar{i}_1(a) = \bar{i}_1(0) e^{-\lambda a} \Gamma(a).

    We can write the characteristic equation as following

    \begin{align*} &\left| \begin{array}{ccc} \lambda+\mu_1 & T_0\int_0^\infty k(a) e^{-\lambda a} \Gamma(a) \text{d}a & \beta T_0 \\ 0 & 1-T_0\int_0^\infty k(a) e^{-\lambda a} \Gamma(a) \text{d}a & -\beta T_0 \\ 0 & -\int_0^\infty p(a) e^{-\lambda a} \Gamma(a) \text{d}a & \lambda+\mu_2 \\ \end{array} \right|\\ = &\Delta(\lambda)(\lambda+\mu_1)\\ = &0, \end{align*}

    where

    \Delta(\lambda): = \lambda+\mu_2-\lambda T_0\int_0^\infty k(a) e^{-\lambda a} \Gamma(a) \text{d}a-\mu_2T_0\int_0^\infty k(a) e^{-\lambda a} \Gamma(a) \text{d}a -\beta T_0\int_0^\infty p(a) e^{-\lambda a} \Gamma(a) \text{d}a.

    Since \lambda = - \mu_1 < 0 , then we only need to consider the root of \Delta(\lambda) = 0 . By way of contradiction, we assume that it has an eigenvalue \lambda_0 with Re(\lambda_0)\geqslant 0. We have

    \begin{align*} |\lambda + \mu_2| = \ &\left|(\lambda+\mu_2)T_0\int_0^\infty k(a)e^{-\lambda a}\Gamma(a) \text{d}a + \beta T_0 \int_0^\infty p(a)e^{-\lambda a}\Gamma(a) \text{d}a\right|\\ \leqslant\ & |\lambda + \mu_2| \left|T_0\int_0^\infty k(a)e^{-\lambda a}\Gamma(a) \text{d}a + \frac{\beta T_0 \int_0^\infty p(a)e^{-\lambda a}\Gamma(a) \text{d}a}{\lambda + \mu_2}\right|\\ \leqslant\ & |\lambda + \mu_2| \left(T_0\int_0^\infty k(a)\Gamma(a) \text{d}a + \frac{\beta T_0 \int_0^\infty p(a)\Gamma(a) \text{d}a}{\mu_2}\right). \end{align*}

    Hence,

    T_0\int_0^\infty k(a)\Gamma(a) \text{d}a + \frac{\beta T_0 \int_0^\infty p(a)\Gamma(a) \text{d}a}{\mu_2}\geqslant1,

    which is impossible because \Re_0 = \frac{\beta T_0 \mathcal{P}}{\mu_2} + T_0 \mathcal{K} < 1. This completes the proof.

    Theorem 4.2. If \Re_1 < 1 < \Re_0 , then the immune-inactivated steady state E^*_1 of system (1.6) is locally asymptotically stable.

    Proof. Denote \bar{T}_2(t) = T(t) - T_1^* , \bar{i}_2(t, a) = i(t, a)-i^*_1(a) , \bar{V}_1(t) = V(t)-V_1^* and \bar{Z}_2(t) = Z(t) , the linearized equation of (1.6) at E_1^* as follows:

    \begin{equation} \left\{ \begin{array}{l} \frac{ \text{d} \bar{T}_2(t)}{ \text{d}t} = - \beta T_1^* \bar{V}_2(t) - \int_0^\infty T_1^* k(a)\bar{i}_2(t, a) \text{d}a - \left(\beta V_1^* + \mu_1 + \int_0^\infty i_1^*(a) k(a) \text{d}a\right)\bar{T}_2(t), \\ \left(\frac{\partial}{\partial t} + \frac{\partial}{\partial a}\right) \bar{i}_2(t, a) = -\delta(a) \bar{i}_1(t, a), \\ \frac{ \text{d} \bar{V}_2(t)}{ \text{d}t} = \int_0^\infty p(a) \bar{i}_2(t, a) \text{d}a - \mu_2 \bar{V}_2(t) - qf'(0)V_1^*\bar{Z}_2(t), \\ \frac{ \text{d} \bar{Z}_2(t)}{ \text{d}t} = cf'(0)V_1^*\bar{Z}_2(t) - \mu_3\bar{Z}_2(t), \\ \bar{i}_2(t, 0) = \beta T_1^* \bar{V}_2(t) + \beta V_1^* \bar{T}_2(t) + \int_0^\infty T_1^* k(a)\bar{i}_2(t, a) \text{d}a + \int_0^\infty i_1^*(a) k(a)\bar{T}_2(t) \text{d}a. \end{array}\right. \end{equation} (4.3)

    Let \bar{T}_2(t) = \bar{T}_2 e^{\lambda t} , \bar{V}_2(t) = \bar{V}_2 e^{\lambda t} , \bar{Z}_2(t) = \bar{Z}_2 e^{\lambda t} and \bar{i}_2(t, a) = \bar{i}_2(a) e^{\lambda t} , thus we have the following characteristic equation:

    \begin{align*} 0 = &\ (\lambda-cf'(0)V_1^*+\mu_3)(\lambda+\mu_1)(\lambda+\mu_2)\left(1-T_1^*\int_0^\infty k(a)e^{-\lambda a}\Gamma(a) \text{d}a\right)\\ &\ -(\lambda-cf'(0)V_1^*+\mu_3)(\lambda+\mu_1)\beta T_1^*\int_0^\infty p(a)e^{-\lambda a}\Gamma(a) \text{d}a\\ &\ +(\lambda-cf'(0)V_1^*+\mu_3)(\lambda+\mu_2)\left(\beta V_1^* + \int_0^\infty k(a) i_1^*(a) \text{d}a\right). \end{align*}

    Note that \Re_1 < 1 and using (2.6), we can obtain that \mu_3-cf'(0)V_1^* > 0 , then the characteristic equation is equivalent to

    \begin{align} 0 = &(\lambda+\mu_1)\left[(\lambda+\mu_2)\left(1-T_1^*\int_0^\infty k(a)e^{-\lambda a}\Gamma(a) \text{d}a\right)-\beta T_1^*\int_0^\infty p(a)e^{-\lambda a}\Gamma(a) \text{d}a\right]\\ &+(\lambda+\mu_2)\left(\beta V_1^* + \int_0^\infty k(a) i_1^*(a) \text{d}a\right). \end{align} (4.4)

    By way of contradiction, we assume that it has an eigenvalue \lambda_0 with Re(\lambda_0)\geqslant 0. Obviously, \lambda = -\mu_1 and \lambda = -\mu_2 are not the roots of (4.4) and note that

    \begin{align*} i_1^*(0) = \ &T_1^*\int_0^\infty k(a)i_1^*(a) \text{d}a + \frac{\beta T_1^*}{\mu_2}\int_0^\infty p(a)i^*_1(a) \text{d}a\\ = \ & T_1^*\int_0^\infty k(a)i_1^*(0)\Gamma(a) \text{d}a + \frac{\beta T_1^*}{\mu_2}\int_0^\infty p(a)i_1^*(0)\Gamma(a) \text{d}a. \end{align*}

    Then we have

    \begin{align*} &\left|1 + \frac{\beta V_1^* + \int_0^\infty k(a)i_1^*(a) \text{d}a}{\lambda_0 + \mu_1}\right|\\ = & \left|T_1^*\int_0^\infty k(a)e^{-\lambda_0 a}\Gamma(a) \text{d}a + \frac{1}{\lambda_0 + \mu_2}\beta T_1^*\int_0^\infty p(a)e^{-\lambda_0 a}\Gamma(a) \text{d}a\right|\\ \leqslant & \left|T_1^*\int_0^\infty k(a)\Gamma(a) \text{d}a + \frac{\beta T_1^*}{\mu_2}\int_0^\infty p(a)\Gamma(a) \text{d}a\right| = 1, \end{align*}

    which is impossible since V_1^* > 0 and i_1^*(a) > 0 . Accordingly, the immune-inactivated steady state E^*_1 of system (1.6) is local asymptotically stable if \Re_1 < 1 < \Re_0 .

    Theorem 4.3. If \Re_1 > 1 , then the immune-activated steady state E^*_2 of system (1.6) is locally asymptotically stable.

    Proof. Applying similar method in the proof of the Theorem (4.2), we have the characteristic equation as following:

    \begin{align} & (\lambda + \mu_1)\left[(\lambda + \mu_2 + qf(Z_2^*))(\lambda - cV_2^*f'(Z_2^*)+\mu_3)+qcV_2^*f(Z_2^*)f'(Z_2^*)\right]\left(1-\int_0^\infty T_2^* k(a)e^{-\lambda}\Gamma(a) \text{d}a\right)\\ &+ (\lambda + \mu_1)(\lambda - cV_2^*f'(Z_2^*)+\mu_3) \beta T_2^* \int_0^\infty p(a) e^{-\lambda a}\Gamma(a) \text{d}a\\ = & -\left[(\lambda + \mu_2 + qf(Z_2^*))(\lambda - cV_2^*f'(Z_2^*)+\mu_3)+qcV_2^*f(Z_2^*)f'(Z_2^*)\right]\left(\beta V_2^* + \int_0^\infty k(a)i_2^*(a) \text{d}a\right). \end{align} (4.5)

    By way of contradiction, we assume that it has an eigenvalue \lambda_0 with Re(\lambda_0)\geqslant 0. From the concavity of function f in (A2), we have \mu_3 - cV_2^*f'(Z_2^*) \geqslant 0 since \mu_3Z_2^* - cV_2^*f(Z_2^*) = 0 . Note that \lambda = -\mu_1 , \lambda = -(\mu_2+qf(Z_2^*)) and \lambda = cV_2^*f'(Z_2^*)-\mu_3 are not the roots of (4.5), then we can rewrite (4.5) as

    \begin{equation} 1 + \Xi = \int_0^\infty T_2^*k(a) e^{-\lambda a}\Gamma(a) \text{d}a. \end{equation} (4.6)

    where

    \begin{align*} \Xi = & \frac{\left(1-\int_0^\infty T_2^* k(a)e^{-\lambda a}\Gamma(a) \text{d}a\right)q c V_2^* f(Z_2^*)f'(Z_2^*)}{(\lambda + \mu_2 + qf(Z_2^*))(\lambda - cV_2^*f'(Z_2^*)+\mu_3)} + \frac{\beta T_2^* \int_0^\infty p(a) e^{-\lambda a}\Gamma(a) \text{d}a}{\lambda + \mu_2 + qf(Z_2^*)}\\ & +\frac{\beta V_2^* + \int_0^\infty k(a)i_2^*(a) \text{d}a}{\lambda + \mu_1} + \frac{q c V_2^* f(Z_2^*)f'(Z_2^*)\left(\beta V_2^* + \int_0^\infty k(a)i_2^*(a) \text{d}a\right)}{(\lambda + \mu_2 + qf(Z_2^*))(\lambda - cV_2^*f'(Z_2^*)+\mu_3)}. \end{align*}

    Then

    \begin{align*} |1 + \Xi| = &\left|\int_0^\infty T_2^*k(a) e^{-\lambda a}\Gamma(a) \text{d}a\right|\\ \lt & \left|\frac{1}{i_2^*(0)}\left(\int_0^\infty i_2^*(0)T_2^*k(a) e^{-\lambda a}\Gamma(a) \text{d}a + \frac{\beta T_2^*}{\mu_2+q f(Z_2^*)}\int_0^\infty i_2^*(0)p(a)\Gamma(a) \text{d}a\right)\right|\\ \leqslant &1, \end{align*}

    which is contradictory. Here we use the fact:

    1-\int_0^\infty T_2^* k(a)e^{-\lambda a}\Gamma(a) \text{d}a = \frac{1}{i_2^*(0)}\left(\beta T_2^* V_2^* + \int_0^\infty T^*_2k(a)\left[1-e^{-\lambda a}\right]i^*_2(a)\right) \gt 0.

    This completes the proof.

    In this section, we discuss the global stability of system (1.6) by using Lyapunov direct method and LaSalle invariance principle. We first give the result on uniform persistence.

    Theorem 5.1. Assume that \Re_0 > 1 . Then there exists a constant \zeta > 0 such that

    \liminf\limits_{t\rightarrow+\infty}T(t)\geqslant\zeta, \ \ \liminf\limits_{t\rightarrow+\infty}\|i(\cdot, t)\|_{L^1}\geqslant\zeta, \ \ \liminf\limits_{t\rightarrow+\infty}V(t)\geqslant\zeta

    for each X_0\in\mathcal{X} .

    The proof of Theorem 5.1 is similar with that in [21,Section 4] or [30], so we omit the details. In the following, we proof the global stability of each steady states.

    Theorem 5.2. The virus-free steady state E_0 of system (1.6) is globally asymptotically stable if \Re_0 < 1 .

    Proof. Let

    \begin{align} \alpha_1(a):& = \int_a^\infty \left(\frac{\beta T^0}{\mu_2}p(\epsilon) + T^0 k(\epsilon)\right) {e^{-\int_a^\epsilon \delta(s) \text{d} s} \text{d} \epsilon}; \end{align} (5.1)
    \begin{align} g(x):& = x-1-\ln x. \end{align} (5.2)

    By direct calculations, we have \alpha_1(0) = \Re_0 and \alpha_1'(a) = \delta(a)\alpha(a) - \left(\frac{\beta T^0}{\mu_2}p(a) + T^0 k(a)\right) . Define the following Lyapunov functional:

    H(t) = H_1(t) + H_2(t),

    where

    \begin{align} H_1(t) & = T_0 g\left(\frac{T(t)}{T_0}\right) + \frac{\beta T_0}{\mu_2}V(t) + \frac{q \beta T_0}{c \mu_2}Z(t), \end{align} (5.3)
    \begin{align} H_2(t) & = \int_0^\infty \alpha_1(a) i(t, a) \text{d}a. \end{align} (5.4)

    Calculating the derivatives of H_1(t) and H_2(t) along system (1.6), we have

    \begin{align*} \frac{ \text{d}H_1(t)}{ \text{d}t} = &\left(1-\frac{T_0}{T(t)}\right)\frac{ \text{d}T(t)}{ \text{d}t} + \frac{\beta T_0}{\mu_2}\frac{ \text{d}V(t)}{ \text{d}t} + \frac{q \beta T_0}{c \mu_2}\frac{ \text{d}Z(t)}{ \text{d}t}\\ = & {\mu_1T_0\left(2-\frac{T_0}{T(t)}-\frac{T(t)}{T_0}\right) - i(t, 0) - \frac{q \beta T_0\mu_3}{c\mu_2}Z(t)}\\ & + \int_0^\infty k(a) T_0i(t, a) \text{d}a + \frac{\beta T_0}{\mu_2} \int_0^\infty p(a)i(t, a) \text{d}a, \end{align*}

    and

    \begin{align*} \frac{ \text{d}H_2(t)}{ \text{d}t} = & \int_0^\infty \alpha_1(a) \frac{\partial i(t, a)}{\partial t} \text{d}a\\ = & -\int_0^\infty \alpha_1(a) \frac{\partial i(t, a)}{\partial a} \text{d}a - \int_0^\infty \alpha_1(a) \delta(a)i(t, a) \text{d}a\\ = & \Re_0 i(t, 0) - \int_0^\infty \left(\frac{\beta T_0}{\mu_2}p(a) + T_0k(a)\right)i(t, a) \text{d}a, \end{align*}

    thus

    \begin{align} \frac{ \text{d}H(t)}{ \text{d}t} = {\mu_1T_0\left(2-\frac{T_0}{T(t)}-\frac{T(t)}{T_0}\right) + (\Re_0 - 1)i(t, 0) - \frac{q \beta T_0\mu_3}{c\mu_2}Z(t)} \leqslant 0 \end{align} (5.5)

    if \Re_0 < 1 . Note that \frac{ \text{d}H(t)}{ \text{d}t}|_{(1.6)} = 0 implies that T(t) = T_0 , i(t, 0) = 0 and Z(t) = 0 , then the largest invariant subset of \left\{\frac{ \text{d}H(t)}{ \text{d}t}|_{(1.6)} = 0\right\} is \{E_0\} . Therefore, the virus-free steady state E_0 of system (1.6) is global asymptotically stable if \Re_0 < 1 by Lyapunov-LaSalle theorem. This ends the proof.

    Theorem 5.3. The immune-inactivated steady state E_1^* = (T_1^*, i_1^*(a), V_1^*, 0) of system (1.6) is globally asymptotically stable if \Re_1 < 1 < \Re_0 .

    Proof. Let

    \begin{align} \alpha_2(a):& = \int_a^\infty \left(\frac{\beta T_1^*}{\mu_2}p(\epsilon) + T_1^* k(\epsilon)\right) e^{-\int_a^\epsilon \delta(s) \text{d}s} \text{d} \epsilon. \end{align} (5.6)

    Define the following Lyapunov functional:

    W(t): = W_1(t)+W_2(t)+W_3(t),

    where

    \begin{align} W_1(t): & = T_1^* g\left(\frac{T(t)}{T_1^*}\right); \end{align} (5.7)
    \begin{align} W_2(t): & = \int_0^\infty \alpha_2(a) i_1^*(a) g\left(\frac{i(t, a)}{i_1^*(a)}\right) \text{d}a; \end{align} (5.8)
    \begin{align} W_3(t): & = \frac{\beta T_1^*}{\mu_2} V_1^* g\left(\frac{V(t)}{V_1^*}\right) + \frac{q\beta T_1^*}{c\mu_2}Z(t). \end{align} (5.9)

    The derivative of W_1(t) is calculated as follows:

    \begin{align} \frac{ \text{d} W_1(t)}{ \text{d}t} = &\left(1-\frac{T_1^*}{T(t)}\right) \frac{ \text{d} T_1(t)}{ \text{d}t}\\ = &\left(1-\frac{T_1^*}{T(t)}\right) \left(\Lambda-\mu_1 T(t) - \beta T(t)V(t) - \int_0^\infty k(a)T(t)i(t, a) \text{d}a\right)\\ = &\mu_1T_1^*\left(2 - \frac{T_1^*}{T(t)} - \frac{T(t)}{T_1^*}\right) + (i_1^*(0)-i(t, 0))\left(1-\frac{T_1^*}{T(t)}\right). \end{align}

    Note that

    \begin{align} &i_1^*(a)\frac{ \text{d}}{ \text{d}a}\left(\frac{i(t, a)}{i_1^*(a)} - 1 - \ln\frac{i(t, a)}{i_1^*(a)}\right)\\ = & \left(1-\frac{i_1^*(a)}{i(t, a)}\right)\frac{\partial i(t, a)}{\partial a} + \delta(a)i(t, a)\left(1-\frac{i_1^*(a)}{i(t, a)}\right), \end{align} (5.10)

    which leads to

    \begin{align} &\int_0^\infty \alpha_2(a)\left(1-\frac{i_1^*(a)}{i(t, a)}\right)\frac{\partial i(t, a)}{\partial a} \text{d}a\\ = & \alpha_2(a) i_1^*(a) \left(\frac{i(t, a)}{i_1^*(a)} - 1 - \ln \frac{i(t, a)}{i_1^*(a)}\right)\bigg{|}_{a = 0}^{a = \infty}+\int_0^\infty \alpha(a)\delta(a)[i_1^*(a)-i(t, a))] \text{d}a\\ &-\int_0^\infty \left(\frac{i(t, a)}{i_1^*(a)} - 1 - \ln \frac{i(t, a)}{i_1^*(a)}\right) \left(\frac{ \text{d} \alpha_2(a)}{ \text{d}a}i_1^*(a) + \alpha_2(a) \frac{\partial i_1^*(a)}{\partial a}\right) \text{d}a\\ = & \lim\limits_{a\rightarrow\infty} \alpha_2(a) i^*(a) g\left(\frac{i(t, a)}{i^*(a)}\right) - \alpha_2(0) i_1^*(0) g\left(\frac{i(t, 0)}{i^*(0)}\right)\\ & +\int_0^\infty \alpha_2(a)\delta(a)[i_1^*(a)-i(t, a))] \text{d}a -\int_0^\infty g\left(\frac{i(t, a)}{i^*(a)}\right) \left(\frac{ \text{d} \alpha_2(a)}{ \text{d}a}i_1^*(a) + \alpha_2(a) \frac{ \text{d} i_1^*(a)}{ \text{d}a}\right) \text{d}a. \end{align}

    By some calculations, we have

    \alpha_2(0) = 1, \ \ \ \alpha'_2(a) = \delta(a)\alpha_2(a) -\left(\frac{\beta T^*_1}{\mu_2}p(a) + T^*_1k(a)\right),

    and using the fact that

    \frac{ \text{d}i_1^*(a)}{ \text{d}a} = -\delta(a) i_1^*(a).

    Hence

    \begin{align*} \nonumber &\int_0^\infty \alpha_2(a)\left(1-\frac{i_1^*(a)}{i(t, a)}\right)\frac{\partial i(t, a)}{\partial a} \text{d}a\\ \nonumber = & \lim\limits_{a\rightarrow\infty} \alpha_2(a) i^*(a) g\left(\frac{i(t, a)}{i^*(a)}\right) - i_1^*(0) g\left(\frac{i(t, 0)}{i^*(0)}\right)+\int_0^\infty \alpha_2(a)\delta(a)[i_1^*(a)-i(t, a))] \text{d}a\\ \nonumber & +\int_0^\infty \left(T_1^*k(a)i^*(a)+\frac{\beta T_1^*}{\mu_2}p(a)i^*(a)\right)g\left(\frac{i(t, a)}{i^*(a)}\right) \text{d}a. \end{align*}

    Then we have the derivative of W_2(t) as follows:

    \begin{align} \frac{ \text{d}W_2(t)}{ \text{d}t} = &\int_0^\infty \alpha(a)\left(1-\frac{i^*(a)}{i(t, a)}\right)\frac{\partial i(t, a)}{\partial t} \text{d}a\\ = & -\int_0^\infty \alpha(a)\left(1-\frac{i^*(a)}{i(t, a)}\right)\left(\frac{\partial i(t, a)}{\partial a} + \delta(a)i(t, a)\right) \text{d}a\\ = & -\lim\limits_{a\rightarrow\infty} \alpha(a) i^*(a)g\left(\frac{i(t, a)}{i^*(a)}\right) + i^*(0) g\left(\frac{i(t, 0)}{i^*(0)}\right)\\ & - \int_0^\infty T_1^*k(a)i^*(a) g\left(\frac{i(t, a)}{i^*(a)}\right) \text{d}a - \int_0^\infty \frac{\beta T_1^*}{\mu_2}p(a)i^*(a) g\left(\frac{i(t, a)}{i^*(a)}\right) \text{d}a. \end{align}

    For W_3(t) , we have

    \begin{align} \frac{ \text{d}W_3(t)}{ \text{d}t} = &\frac{\beta T_1^*}{\mu_2} \left(1-\frac{V_1^*}{V(t)}\right)\frac{ \text{d}V(t)}{ \text{d}t} + \frac{q\beta T_1^*}{c\mu_2}\frac{ \text{d}Z(t)}{ \text{d}t}\\ = & \frac{\beta T_1^*}{\mu_2}\int_0^\infty p(a)i(t, a) \text{d}a - \frac{\beta T_1^*}{\mu_2} \mu_2 V(t) - \frac{\beta T_1^*}{\mu_2} \frac{V_1^*}{V(t)}\int_0^\infty p(a)i(t, a) \text{d}a + \frac{\beta T_1^*}{\mu_2} \mu_2 V_1^*\\ &+ \frac{q \beta T_1^*}{\mu_2} V_1^* f(Z(t)) - \frac{q\beta T_1^*}{c\mu_2}\mu_3 Z(t). \end{align}

    Hence,

    \begin{align} \frac{ \text{d}W(t)}{ \text{d}t} \nonumber = & \mu_1T_1^*\left(2 - \frac{T_1^*}{T(t)} - \frac{T(t)}{T_1^*}\right)\\ &-\beta T_1^*V_1^*\frac{T_1^*}{T(t)} - \int_0^\infty k(a)T_1^*i_1^*(a)\frac{T_1^*}{T(t)} \text{d}a + \int_0^\infty k(a)T_1^*i(t, a) \text{d}a\\ & -\lim\limits_{a\rightarrow\infty} \alpha(a) i_1^*(a)g\left(\frac{i(t, a)}{i_1^*(a)}\right) - i_1^*(0) \ln \frac{i(t, 0)}{i_1^*(0)} - \int_0^\infty p(a)i_1^*(a) g\left(\frac{i(t, a)}{i_1^*(a)}\right) \text{d}a\\ & +\frac{\beta T_1^*}{\mu_2}\int_0^\infty p(a)i(t, a) \text{d}a - \frac{\beta T_1^*}{\mu_2} \frac{V_1^*}{V(t)}\int_0^\infty p(a)i(t, a) \text{d}a + \frac{\beta T_1^*}{\mu_2} \mu_2 V_1^*\\ &+ \frac{q \beta T_1^*}{\mu_2} V_1^* f(Z(t)) - \frac{q\beta T_1^*}{c\mu_2}\mu_3 Z(t). \end{align} (5.11)

    Recalling that V_1^* = \int_0^\infty \frac{p(a)}{\mu_2}i_1^*(a) \text{d}a and i_1^*(0) = \beta T_1^* V_1^* + \int_0^\infty k(a)T_1^*i_1^*(a) \text{d}a . Substituting (2.13) into (5.11), after some calculations and rearranging the equation yield

    \begin{align*} \label{W2} \frac{ \text{d}W(t)}{ \text{d}t} = & \mu_1T_1^*\left(2 - \frac{T_1^*}{T(t)} - \frac{T(t)}{T_1^*}\right) - \lim\limits_{a\rightarrow\infty} \alpha(a) i_1^*(a)g\left(\frac{i(t, a)}{i_1^*(a)}\right)\\ &+\int_0^\infty \frac{\beta T_1^*}{\mu_2} p(a) i_1 ^*(a) \left( 2 - \frac{T_1^*}{T(t)} - \frac{V_1^* i(t, a)}{V(t)i_1^*(a)} - \ln \frac{i(t, 0)}{i_1^*(0)} + \ln \frac{i(t, a)}{i_1^*(a)}\right) \text{d}a\\ &+\int_0^\infty T_1^*k(a)i_1^*(a)\left( 1 - \frac{T_1^*}{T(t)} - \ln \frac{i(t, 0)}{i_1^*(0)} + \ln \frac{i(t, a)}{i_1^*(a)}\right) \text{d}a\\ &+\int_0^\infty \frac{\beta T_1^*}{\mu_2}p(a)i_1^*(a)\left(1-\frac{i_1^*(0)T(t)V(t)}{i(t, 0)T_1^*V_1^*}\right) \text{d}a\\ & + \int_0^\infty T_1^*k(a)i_1^*(a)\left(1-\frac{i_1^*(0)T(t)i(t, a)}{i(t, 0)T_1^*i_1^*(a)}\right) \text{d}a\\ = & \mu_1T_1^*\left(2 - \frac{T_1^*}{T(t)} - \frac{T(t)}{T_1^*}\right) - \lim\limits_{a\rightarrow\infty} \alpha(a) i_1^*(a)g\left(\frac{i(t, a)}{i_1^*(a)}\right)\\ &- \int_0^\infty \frac{\beta T_1^*}{\mu_2} p(a) i^*(a)\left\{g\left(\frac{T_1^*}{T(t)}\right)+g\left(\frac{V_1^*i(t, a)}{V(t)i_1^*(a)}\right)+g\left(\frac{i_1^*(0)T(t)V(t)}{i(t, 0)T_1^*V_1^*}\right)\right\} \text{d}a\\ &- \int_0^\infty T_1^*k(a) i_1^*(a)\left\{g\left(\frac{T_1^*}{T(t)}\right)+g\left(\frac{i_1^*(0)T(t)i(t, a)}{i(t, 0)T_1^*i_1^*(a)}\right)\right\} \text{d}a\\ &+ \frac{q \beta T_1^*}{\mu_2} V_1^* f(Z(t)) - \frac{q\beta T_1^*}{c\mu_2}\mu_3 Z(t). \end{align*}

    Note that

    \begin{align*} \frac{q \beta T_1^*}{\mu_2} V_1^* f(Z(t)) - \frac{q\beta T_1^*}{c\mu_2}\mu_3 Z(t)\leqslant& \frac{q \beta T_1^*}{c\mu_2} \left[ \frac{\Lambda c \mathcal{P} f'(0)}{\mu_2\mu_3}\left(1-\frac{1}{\Re_0}\right) - 1\right]\\ = &\frac{q \beta T_1^*}{c\mu_2}(\Re_1-1). \end{align*}

    Thus \frac{ \text{d}W(t)}{ \text{d}t}|_{(1.6)} \leqslant 0 when \Re_1 < 1 < \Re_0 , \frac{ \text{d}W(t)}{ \text{d}t} = 0 if and only if (T(t), i(t, a), V(t), Z(t)) = (T_1^*, i_1^*(a), V_1^*, 0). Applying Lyapunov-LaSalle theorem, the immune-inactivated steady state E_1^* = (T_1^*, i_1^*(a), V_1^*, 0) of system (1.6) is globally asymptotically stable if \Re_1 < 1 < \Re_0 .

    Theorem 5.4. The immune-activated steady state E_2^* = (T_2^*, i_2^*(a), V_2^*, Z_2^*) of system (1.6) is globally asymptotically stable if \Re_1 > 1 .

    Proof. Let

    \begin{align} \alpha_3(a):& = \int_a^\infty \left(\frac{\beta T_2^*}{\varpi(Z_2^*)}p(\epsilon) + T_2^* k(\epsilon)\right) e^{-\int_a^\epsilon \delta(s) \text{d}s} \text{d} \epsilon. \end{align} (5.12)

    Define the Lyapunov functional as follows

    L(t): = L_1(t)+L_2(t)+L_3(t)+L_4(t),

    where

    \begin{align*} L_1(t): & = T_2^* g\left(\frac{T(t)}{T_2^*}\right);\\ L_2(t): & = \int_0^\infty \alpha_3(a) i_2^*(a) g\left(\frac{i(t, a)}{i_2^*(a)}\right) \text{d}a;\\ L_3(t): & = \frac{\beta T_2^*}{\mu_2} V_2^* g\left(\frac{V(t)}{V_2^*}\right) + \frac{q\beta T_2^*}{c\mu_2} \left(Z(t) - Z_2^* - \int_{Z_2^*}^{Z(t)}\frac{f(Z_2^*)}{f(\tau)} \text{d} \tau\right). \end{align*}

    Using the results in the proof of Theorem (5.3), after some calculations, we have the derivative of L(t) as follows:

    \begin{align*} \label{L2} \frac{ \text{d}L(t)}{ \text{d}t} = & \mu_1T_2^*\left(2 - \frac{T_2^*}{T(t)} - \frac{T(t)}{T_2^*}\right) - \lim\limits_{a\rightarrow\infty} \alpha_3(a) i_2^*(a)g\left(\frac{i(t, a)}{i_2^*(a)}\right)\\ &- \int_0^\infty \frac{\beta T_2^*}{\varpi(Z_2^*)} p(a) i_2^*(a)\left\{g\left(\frac{T_2^*}{T(t)}\right)+g\left(\frac{V_2^*i(t, a)}{V(t)i_2^*(a)}\right)+g\left(\frac{i_2^*(0)T(t)V(t)}{i(t, 0)T_2^*V_2^*}\right)\right\} \text{d}a\\ &- \int_0^\infty T_2^*k(a) i_2^*(a)\left\{g\left(\frac{T_2^*}{T(t)}\right)+g\left(\frac{i_2^*(0)T(t)i(t, a)}{i(t, 0)T_2^*i_2^*(a)}\right)\right\} \text{d}a\\ &+ \frac{q\beta T_2^*V_2^*f(Z_2^*)}{\varpi(Z_2^*)}\left(\frac{Z(t)}{Z_2^*}-\frac{f(Z(t))}{f(Z_2^*)}\right)\left(\frac{f(Z_2^*)}{f(Z(t))}-1\right). \end{align*}

    It follows from follows (A2) and (A3) that \left(\frac{Z(t)}{Z_2^*}-\frac{f(Z(t))}{f(Z_2^*)}\right)\left(\frac{f(Z_2^*)}{f(Z(t))}-1\right)\leqslant0 , thus \frac{ \text{d}L(t)}{ \text{d}t}|_{(1.6)} \leqslant 0 and \frac{ \text{d}L(t)}{ \text{d}t} = 0 if and only if (T(t), i(t, a), V(t), Z(t)) = (T_2^*, i_2^*(a), V_2^*, Z_2^*). Therefore, the immune-activated steady state E_2 of system (1.6) is global asymptotically stable if \Re_1 > 1 by Lyapunov-LaSalle theorem.

    In this subsection, as special case for the age-infection model (1.6) and (1.7) with general nonlinear immune response f(Z) , we introduce the following age-infection model with saturation immune response function, which have been used for modeling HIV infection in [30,42].

    \begin{equation} \left\{ \begin{array}{l} \frac{ \text{d}T(t)}{ \text{d}t} = \Lambda - \mu_1 T(t) - \beta T(t) V(t) - \int_0^\infty k(a)T(t)i(t, a) \text{d}a, \\ \left(\frac{\partial}{\partial t} + \frac{\partial}{\partial a}\right) i(t, a) = -\delta(a) i(t, a), \\ \frac{ \text{d}V(t)}{ \text{d}t} = \int_0^\infty p(a) i(t, a) \text{d}a - \mu_2 V(t) - \frac{qV(t)Z(t)}{h+Z(t)}, \\ \frac{ \text{d}Z(t)}{ \text{d}t} = \frac{cV(t)Z(t)}{h+Z(t)} -\mu_3Z(t) \end{array}\right. \end{equation} (6.1)

    with the boundary and initial condition

    \begin{equation} \left\{ \begin{array}{l} i(t, 0) = \beta T(t) V(t) + \int_0^\infty k(a)T(t)i(t, a) \text{d}a, \\ T(0) = T_0 \gt 0, \ \ V(0) = V_0 \gt 0, \ \ Z(0) = Z_0 \gt 0\ \ \text{and} \ \ i(0, a) = i_0(a)\in L^1_+(0, \infty). \end{array}\right. \end{equation} (6.2)

    The model without cell-to-cell transmission of system (6.1) with (6.2) has been studied in [30]. It is easy to see that f(Z) = \frac{Z(t)}{h+Z(t)} satisfy (A2) and (A3). System (6.1) with (6.2) is a special case of the original system (1.6) and (1.7).

    The virus-free equilibrium of system (6.1) with (6.2) is similar to the previous one, E_{01} = (T_0, 0, 0, 0) , where T_0 = \frac{\Lambda}{\mu_1} . By some calculation, we obtain the basic reproduction number and immune-activated reproduction number of system (6.1) with (6.2) as \Re_{01} = \frac{\beta T_0 \mathcal{P}}{\mu_2} + T_0 \mathcal{K} and \Re_{11} = \frac{\Lambda c \mathcal{P}}{h\mu_2\mu_3}\left(1-\frac{1}{\Re_0}\right) , respectively. We have the following corollary:

    Corollary 6.1. For system (6.1) with (6.2), there are two threshold parameters \Re_{01} and \Re_{11} with \Re_{01} > \Re_{11} such that

    (ⅰ) If \Re_{01} < 1 , there exists a virus-free steady state E_{01} , and E_{01} is globally asymptotically stable;

    (ⅱ) If \Re_{11} < 1 < \Re_{01} , there exists a immune-inactivated steady state E_{11} which is globally asymptotically stable;

    (ⅲ) If \Re_{11} > 1 , there exists a immune-activated steady state E_{21} which is globally asymptotically stable.

    In this subsection, we perform some numerical simulations to the validity of the theoretical result of this paper. Specifically, we focus on the age-infection model with saturation immune response function (see model (6.1)).

    The parameter values will be used in numerical simulation are listed in Table 2. Furthermore, we set the maximum age for the viral production as a^† = 10 and we set

    \delta(a) = 0.03\left(1+\sin\frac{(a-5)\pi}{10}\right), \ \ \ p(a) = 2.9\left(1+\sin\frac{(a-5)\pi}{10}\right)
    Table 2.  Parameter values for numerical simulations.
    Parameter Value Unite Case 1 Case 2 Case 3 Ref.
    \Lambda 0\sim100 cells ml-day ^{-1} 1.2 8 100 [44]
    \beta 5 \times 10^{-7}\sim0.5 ml virion-day ^{-1} 0.001 0.001 0.001 [42]
    \mu_1 0.007\sim0.1 day ^{-1} 0.01 0.01 0.01 [44]
    \mu_2 2.4\sim3 day ^{-1} 6 6 6 [44]
    \mu_3 0.3 day ^{-1} 0.3 0.3 0.3 [45]
    q 0.006 ml cell ^{-1} day ^{-1} 0.006 0.006 0.006 [45]
    c 0.1 ml virion ^{-1} day ^{-1} 0.1 0.1 0.1 [45]
    h 1\sim100 Saturation constant 10 10 100 Assumed

     | Show Table
    DownLoad: CSV

    and

    k(a) = 0.0003\left(1+\sin\frac{(a-5)\pi}{10}\right).

    Thus, the averages of \delta(a) , p(a) and k(a) are 0.03 , 2.9 and 0.0003 , which are the same in [43,20].

    Numerical simulation shows the following three cases:

    Case 1: Choose parameter values as in Case 1 of Table 2, then we can calculate the basic reproduction number as \Re_0\approx0.8093 < 1 . Corollary 1 asserts that the virus-free steady state of system (6.1) with (6.2) is globally asymptotically stable. From Figure 1, one can observe that the levels of all compartmental individuals tend to stable values, where T(t) , V(t) , i(t, a) and Z(t) converge to a virus-free steady states (100, 0, 0, 0).

    Figure 1.  The long time dynamical behaviors of system (6.1) with (6.2) for \Re_0 = 0.8093 < 1 , that is, the virus-free steady state of system (6.1) with (6.2) is globally asymptotically stable.

    Case 2: Choose parameter values as in Case 2 of Table 2. By some computing, we can obtain that \Re_0 \approx 5.3952 > 1 > \Re_1\approx 0.9040 . From Corollary 1, we derive that immune-inactivated infection steady state is globally asymptotically stable. Numerical simulation illustrates this fact (see Figure 2).

    Figure 2.  The long time dynamical behaviors of system (6.1) with (6.2) for \Re_0 \approx 5.3952 and \Re_1\approx 0.9040 , that is, the immune-activated steady state of system (6.1) with (6.2) is globally asymptotically stable.

    Case 3: Choose parameter values as in Case 3 of Table 2. Similarly, we can obtain that \Re_0 \approx 29.9893 > 1 and \Re_1\approx 1.3408 > 1 . Numerical simulation shows that the levels of all compartmental individuals tend to stable values (see Figure 3), that is, immune-inactivated infection steady state is globally asymptotically stable.

    Figure 3.  The long time dynamical behaviors of system (6.1) with (6.2) for \Re_0 \approx 29.9893 > 1 and \Re_1\approx 1.3408 > 1 , that is, the immune-activated steady state of system (6.1) with (6.2) is globally asymptotically stable.

    In this paper, we proposed and investigated an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immunity response. We have shown that the global stability of equilibria of model (1.6) are determined by the corresponding basic reproduction numbers \Re_0 and the basic immune reproductive number \Re_1 . That is, when \Re_0 < 1 , the virus-free steady state is globally asymptotically stable, which means that the viruses are cleared and immune response is not active; when \Re_1 < 1 < \Re_0 , the immune-inactivated infection steady state exists and is globally asymptotically stable, which means that viral infection becomes chronic and humoral immune response would not be activated; and when \Re_1 > 1 , the immune-activated infection steady state exists and is globally asymptotically stable, in this case the infection causes a persistent humoral immune response and is chronic.

    Now, we show the relevance of model formulations between our age-structured model (1.6) and the standard ODE models. We consider \delta(a)\equiv\delta , k(a)\equiv k and p(a)\equiv p in model (1.6). Letting

    I(t) = \int_0^\infty i(t, a) \text{d}a.

    Recall that

    i(t, 0) = \beta T(t)V(t) + kT(t)I(t),

    then we have

    \begin{align*} \frac{ \text{d} I(t)}{ \text{d}t} = \int_0^\infty \frac{\partial i(t, a)}{\partial t} \text{d}a = &\ -\int_0^\infty \left(\frac{\partial i(t, a)}{\partial a} + \delta i(t, a)\right) \text{d}a\\ = &\ i(t, 0) - \int_0^\infty \delta i(t, a) \text{d}a\\ = &\ \beta T(t)V(t) + kT(t)I(t) -\delta I(t), \end{align*}

    here we assume that \lim_{a\rightarrow\infty}i(t, a) = 0 , which means that there is no biological individual can live forever. Thus, system (1.6) is equivalent to the following ODE model as

    \begin{equation} \left\{ \begin{array}{l} \frac{ \text{d}T(t)}{ \text{d}t} = \Lambda - \mu_1 T(t) - \beta T(t) V(t) - kT(t)I(t), \\ \frac{ \text{d}I(t)}{ \text{d}t} = \beta T(t)V(t) + kT(t)I(t) -\delta I(t), \\ \frac{ \text{d}V(t)}{ \text{d}t} = p I(t) - \mu_2 V(t) - q V(t)f(Z(t)), \\ \frac{ \text{d}Z(t)}{ \text{d}t} = c V(t)f(Z(t)) -\mu_3Z(t), \end{array}\right. \end{equation} (7.1)

    which is the model studied by [23] when f(Z) = Z and k = 0 . In fact, we have not found the above model in any existing literatures, but we think it has the same dynamic behavior with (1.6).

    It is necessary to mention it here, in the proof of Lemma 4.1, there may exists zero eigenvalue if R_0 = 1 , and it may lead to more complex dynamic behavior. For example, Qesmi et al. [46] propose a mathematical model describing the dynamics of hepatitis B or C virus infection with age-structure, and they found that when \Re_0 = 1 , the system may undergo a backward bifurcation. In a recent work [22], Zhang and Liu studied an age-structured HIV model with cell-to-cell transmission and logistic growth in uninfected cells. They have shown that there exists Hopf bifurcation of the model by using the Hopf bifurcation theory for semilinear equations with non-dense domain. Introducing logistic growth in uninfected cells to model (1.6), it will be interesting to investigate the existence of a Hopf bifurcation. We leave the above two studies for future consideration.

    The authors are very grateful to the editors and two reviewers for their valuable comments and suggestions that have helped us improving the presentation of this paper. The authors were supported by Natural Science Foundation of China (11871179; 11771374).

    All authors declare no conflicts of interest in this paper.



    [1] M. A. Nowak, S. Bonhoeffer, A. M. Hill, R. Boehme, H. C. Thomas, H. McDade, Viral dynamics in hepatitis B virus infection, Proc. Natl. Acad. Sci. USA, 93 (1996), 4398-4402.
    [2] P. Nelson, M. Gilchrist, D. Coombs, J. M. Hyman, A. S. perelson, An age-structured model of HIV infection that allow for variations in the production rate of viral particles and the death rate of productively infected cells, Math. Biosci. Eng., 1 (2004), 267-288.
    [3] L. Rong, Z. Feng, A. S. Perelson, Mathematical analysis of age-structured HIV-1 dynamics with combination antiretroviral therapy, SIAM J. Appl. Math., 67 (2007), 731-756.
    [4] G. Huang, X. Liu, Y. Takeuchi, Lyapunov functions and global stability for age-structured HIV infection model, SIAM J. Appl. Math., 72 (2012), 25-38.
    [5] L. Zou, S. Ruan, W. Zhang, An age-structured model for the transmission dynamics of hepatitis B, SIAM J. Appl. Math., 70 (2010), 3121-3139.
    [6] J. Wang, R. Zhang, T. Kuniya, Global dynamics for a class of age-infection HIV models with nonlinear infection rate, J. Math. Anal. Appl., 432 (2015), 289-313.
    [7] M. Shen, Y. Xiao, L. Rong, Global stability of an infection-age structured HIV-1 model linking within-host and between-host dynamics, Math. Biosci., 263 (2015), 37-50.
    [8] Y. Wang, K. Liu, Y. Lou, An age-structured within-host HIV model with T-cell competition, Nonlinear Anal. Real World Appl., 38 (2017), 1-20.
    [9] C.-Y. Cheng, Y. Dong, Y. Takeuchi, An age-structured virus model with two routes of infection in heterogeneous environments, Nonlinear Anal. Real World Appl., 39 (2018), 464-491.
    [10] Z. Liu, Y. Rong, Zero-Hopf bifurcation for an infection-age structured epidemic model with a nonlinear incidence rate, Sci. China Math., 60 (2017), 1371-1398.
    [11] J. Wang, X. Dong, Analysis of an HIV infection model incorporating latency age and infection age, Math. Biosci. Eng., 15 (2018), 569-594.
    [12] J. Pang, J. Chen, Z. Liu, P. Bi, S. Ruan, Local and global stabilities of a viral dynamics model with infection-age and immune response, J. Dyn. Diff. Equat., 31 (2019), 793-813.
    [13] K. Hattaf, Y. Yang, Global dynamics of an age-structured viral infection model with general incidence function and absorption, Int. J. Biomath., 11 (2018), 1850065.
    [14] W. Hübner, G. McNerney, P. Chen, B. M. Dale, R. E. Gordon, F. Y. Chuang, et al., Quantitative 3D video microscopy of HIV transfer across T cell virological synapses, Science, 323 (2009), 1743-1747.
    [15] A. Imle, P. Kumberger, N. D. Schnellbächer, J. Fehr, P. Carrillo-Bustamante, J. Ales, et al., Experimental and computational analyses reveal that environmental restrictions shape HIV-1 spread in 3D cultures, Nat. Commun., 10 (2019), 2144.
    [16] N. Barretto, B. Sainz, S. Hussain, S. L. Uprichard, Determining the involvement and therapeutic implications of host cellular factors in hepatitis C virus cell-to-cell spread, J. Virol., 88 (2014), 5050-5061.
    [17] F. Merwaiss, C. Czibener, D. E. Alvarez, Cell-to-cell transmission is the main mechanism supporting bovine viral diarrhea virus spread in cell culture, J. Virol., 93 (2019), e01776.
    [18] G. L. Smith, B. J. Murphy, M. Law, Vaccinia virus motility, Annu. Rev. Microbiol., 57 (2003), 323-342.
    [19] X. Lai, X. Zou, Modeling HIV-1 virus dynamics with both virus-to-cell infection and cell-to-cell transmission, SIAM J. Appl. Math., 74 (2014), 898-917.
    [20] Y. Yang, L. Zou, S. Ruan, Global dynamics of a delayed within-host viral infection model with both virus-to-cell and cell-to-cell transmissions, Math. Biosci., 270 (2015), 183-191.
    [21] J. Wang, J. Lang, X. Zou, Analysis of an age structured HIV infection model with virus-to-cell infection and cell-to-cell transmission, Nonlinear Anal. Real World Appl., 34 (2017), 75-96.
    [22] X. Zhang, Z. Liu, Bifurcation analysis of an age structured HIV infection model with both virusto-cell and cell-to-cell transmissions, Int. J. Bifurcation Chaos, 28 (2018), 1850109.
    [23] A. Murase, T. Sasaki, T. Kajiwara, Stability analysis of pathogen-immune interaction dynamics, J. Math. Biol., 51 (2005), 247-267.
    [24] S. Wang, D. Zou, Global stability of in-host viral models with humoral immunity and intracellular delays, Appl. Math. Model, 51 (2012), 1313-1322.
    [25] T. Wang, Z. Hu, F. Liao, Stability and Hopf bifurcation for a virus infection model with delayed humoral immunity response, J. Math. Appl. Anal., 411 (2014), 63-74.
    [26] T. Kajiwara, T. Sasaki, Y. Otani, Global stability of an age-structured model for pathogen-immune interaction, J. Appl. Math. Comput., 59 (2019), 631-660.
    [27] M. A. Nowak, C. R. M. Bangham, Population dynamics of immune responses to persistent viruses, Science, 272 (1996), 74-79.
    [28] Y. Wang, Y. Zhou, F. Brauer, J. M. Heffernan, Viral dynamics model with CTL immune response incorporating antiretroviral therapy, J. Math. Biol., 67 (2013), 901-934.
    [29] C. Browne, Immune response in virus model structured by cell infection-age, Math. Biosci. Eng., 13 (2016), 887-909.
    [30] X. Duan, S. Yuan, Global dynamics of an age-structured virus model with saturation effects, Math. Methods Appl. Sci., 40 (2017), 1851-1864.
    [31] X. Wang, S. Liu, A class of delayed viral models with saturation infection rate and immune response, Math. Methods Appl. Sci., 36 (2013), 125-142.
    [32] H. Shu, L. Wang, J. Watmough, Global stability of a nonlinear viral infection model with infinitely distributed intracellular delays and CTL immune responses, SIAM J. Appl. Math., 73 (2013), 1280-1302.
    [33] X. Tian, R. Xu, J. Lin, Mathematical analysis of an age-structured HIV-1 infection model with CTL immune response, Math. Biosci. Eng., 16 (2019), 7850-7882.
    [34] K. Hattaf, Spatiotemporal dynamics of a generalized viral infection model with distributed delays and CTL immune response, Computation, 7 (2019), 21.
    [35] X. Tian, R. Xu, Global stability of a delayed HIV-1 infection model with absorption and CTL immune response, IMA J. Appl. Math., 79 (2014), 347-359.
    [36] P. Magal, S. Ruan, Theory and Applications of Abstract Semilinear Cauchy Problems, Applied Mathematical Sciences Vol. 201, Springer, Cham, 2018.
    [37] P. Magal, Compact attractors for time-periodic age structured population models, Elect. J. Differ. Eqs., 65 (2001), 65.
    [38] P. Magal, C. C. McCluskey, G. F. Webb, Lyapunov functional and global asymptotic stability for an infection-age model, Appl. Anal., 89 (2010), 1109-1140.
    [39] P. Magal, H. Thieme, Eventual compactness for a semiflow generated by an age-structured models, Commun. Pure Appl. Anal., 3 (2004), 695-727.
    [40] H. L. Smith, H. R. Thieme, Dynamical Systems and Population Persistence, Graduate Studies in Mathematics Vol. 118, American Mathematical Society, Providence, RI, 2011.
    [41] C. C. McCluskey, Global stability for an SEI epidemiological model with continuous age-structure in the exposed and infectious classes, Math. Biosci. Eng., 9 (2012), 819-841.
    [42] C. Jiang, K. Wang, L. Song, Global dynamics of a delay virus model with recruitment and saturation effects of immune responses, Math. Biosci. Eng., 14 (2017), 1233-1246.
    [43] H. Dahari, A. Lo, R. M. Ribeiro, A. S. Perelson, Modeling hepatitis C virus dynamics: liver regeneration and critical drug efficacy, J. Theor. Biol., 247 (2007), 371-381.
    [44] Y. Wang, Y. Zhou, J. Wu, J. Heffernan, Oscillatory viral dynamics in a delayed HIV pathogenesis model, Math. Biosci., 219 (2009), 104-112.
    [45] J. Reyes-Silveyra, A. R. Mikler, Modeling immune response and its effect on infectious disease outbreak dynamics, Theor. Biol. Med. Model., 13 (2016), 10.
    [46] R. Qesmi, S. ElSaadany, J. M. Heffernan, J. Wu, A hepatitis B and C virus model with age since infection that exhibits backward bifurcation, SIAM J. Appl. Math., 71 (2011), 1509-1530.
  • This article has been cited by:

    1. Maoxin Liao, Yanjin Liu, Shinan Liu, Ali M. Meyad, Stability and Hopf bifurcation of HIV-1 model with Holling II infection rate and immune delay, 2021, 1751-3758, 1, 10.1080/17513758.2021.1895334
    2. Lili Han, Qiuhui Pan, Baolin Kang, Mingfeng He, Effects of masks on the transmission of infectious diseases, 2021, 2021, 1687-1847, 10.1186/s13662-021-03321-z
    3. Zhongzhong Xie, Xiuxiang Liu, Global dynamics in an age-structured HIV model with humoral immunity, 2021, 14, 1793-5245, 2150047, 10.1142/S1793524521500479
    4. Lili Liu, Jian Zhang, Ran Zhang, Hongquan Sun, Hopf Bifurcation of an Age-Structured Epidemic Model with Quarantine and Temporary Immunity Effects, 2021, 31, 0218-1274, 2150183, 10.1142/S0218127421501832
    5. Jinhu Xu, Dynamic analysis of a cytokine-enhanced viral infection model with infection age, 2023, 20, 1551-0018, 8666, 10.3934/mbe.2023380
    6. Xin Jiang, Global Dynamics for an Age-Structured Cholera Infection Model with General Infection Rates, 2021, 9, 2227-7390, 2993, 10.3390/math9232993
    7. Zakaria Yaagoub, Karam Allali, Fractional HBV infection model with both cell-to-cell and virus-to-cell transmissions and adaptive immunity, 2022, 165, 09600779, 112855, 10.1016/j.chaos.2022.112855
    8. Zakaria Yaagoub, Marya Sadki, Karam Allali, A generalized fractional hepatitis B virus infection model with both cell-to-cell and virus-to-cell transmissions, 2024, 112, 0924-090X, 16559, 10.1007/s11071-024-09867-3
    9. Yuequn Gao, Ning Li, Fractional order PD control of the Hopf bifurcation of HBV viral systems with multiple time delays, 2023, 83, 11100168, 1, 10.1016/j.aej.2023.10.032
    10. Tingting Zheng, Yantao Luo, Zhidong Teng, Spatial dynamics of a viral infection model with immune response and nonlinear incidence, 2023, 74, 0044-2275, 10.1007/s00033-023-02015-8
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(4209) PDF downloads(486) Cited by(10)

Figures and Tables

Figures(3)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog