Loading [MathJax]/jax/element/mml/optable/GreekAndCoptic.js
Research article

Threshold dynamics of a general delayed HIV model with double transmission modes and latent viral infection

  • Received: 09 August 2021 Accepted: 03 November 2021 Published: 12 November 2021
  • MSC : 92D30, 34K20

  • In this paper, a general HIV model incorporating intracellular time delay is investigated. Taking the latent virus infection, both virus-to-cell and cell-to-cell transmissions into consideration, the model exhibits threshold dynamics with respect to the basic reproduction number R0. If R0<1, then there exists a unique infection-free equilibrium E0, which is globally asymptotically stable. If R0>1, then there exists E0 and a globally asymptotically stable infected equilibrium E. When R0=1, E0 is linearly neutrally stable and a forward bifurcation takes place without time delay around R0=1. The theoretical results and corresponding numerical simulations show that the existence of latently infected cells and the intracellular time delay have vital effect on the global dynamics of the general virus model.

    Citation: Xin Jiang. Threshold dynamics of a general delayed HIV model with double transmission modes and latent viral infection[J]. AIMS Mathematics, 2022, 7(2): 2456-2478. doi: 10.3934/math.2022138

    Related Papers:

    [1] Liang Hong, Jie Li, Libin Rong, Xia Wang . Global dynamics of a delayed model with cytokine-enhanced viral infection and cell-to-cell transmission. AIMS Mathematics, 2024, 9(6): 16280-16296. doi: 10.3934/math.2024788
    [2] Ahmed M. Elaiw, Ghadeer S. Alsaadi, Aatef D. Hobiny . Global co-dynamics of viral infections with saturated incidence. AIMS Mathematics, 2024, 9(6): 13770-13818. doi: 10.3934/math.2024671
    [3] A. M. Elaiw, E. A. Almohaimeed, A. D. Hobiny . Stability of HHV-8 and HIV-1 co-infection model with latent reservoirs and multiple distributed delays. AIMS Mathematics, 2024, 9(7): 19195-19239. doi: 10.3934/math.2024936
    [4] A. M. Elaiw, N. H. AlShamrani, A. D. Hobiny . Mathematical modeling of HIV/HTLV co-infection with CTL-mediated immunity. AIMS Mathematics, 2021, 6(2): 1634-1676. doi: 10.3934/math.2021098
    [5] Xiong Zhang, Zhongyi Xiang . Piecewise immunosuppressive infection model with viral logistic growth and effector cell-guided therapy. AIMS Mathematics, 2024, 9(5): 11596-11621. doi: 10.3934/math.2024569
    [6] Ru Meng, Yantao Luo, Tingting Zheng . Stability analysis for a HIV model with cell-to-cell transmission, two immune responses and induced apoptosis. AIMS Mathematics, 2024, 9(6): 14786-14806. doi: 10.3934/math.2024719
    [7] Mohammed H. Alharbi . HIV dynamics in a periodic environment with general transmission rates. AIMS Mathematics, 2024, 9(11): 31393-31413. doi: 10.3934/math.20241512
    [8] E. A. Almohaimeed, A. M. Elaiw, A. D. Hobiny . Modeling HTLV-1 and HTLV-2 co-infection dynamics. AIMS Mathematics, 2025, 10(3): 5696-5730. doi: 10.3934/math.2025263
    [9] Chang Hou, Qiubao Wang . The influence of an appropriate reporting time and publicity intensity on the spread of infectious diseases. AIMS Mathematics, 2023, 8(10): 23578-23602. doi: 10.3934/math.20231199
    [10] Abdul Qadeer Khan, Ayesha Yaqoob, Ateq Alsaadi . Discrete Hepatitis C virus model with local dynamics, chaos and bifurcations. AIMS Mathematics, 2024, 9(10): 28643-28670. doi: 10.3934/math.20241390
  • In this paper, a general HIV model incorporating intracellular time delay is investigated. Taking the latent virus infection, both virus-to-cell and cell-to-cell transmissions into consideration, the model exhibits threshold dynamics with respect to the basic reproduction number R0. If R0<1, then there exists a unique infection-free equilibrium E0, which is globally asymptotically stable. If R0>1, then there exists E0 and a globally asymptotically stable infected equilibrium E. When R0=1, E0 is linearly neutrally stable and a forward bifurcation takes place without time delay around R0=1. The theoretical results and corresponding numerical simulations show that the existence of latently infected cells and the intracellular time delay have vital effect on the global dynamics of the general virus model.



    Human immunodeficiency virus (HIV), the causative agent of AIDS, remains one of the biggest causes of morbidity and mortality, infecting around 37.7 million people worldwide and bringing about nearly 0.68 million AIDS-related death in 2020 [42]. HIV attacks cells of the immune system, primarily macrophages and CD4+T cells, leading to immunodeficiency of individual. If untreated, individual has more difficulty fighting againt some opportunistic diseases and eventually results to death. Considerable efforts on immunity to HIV and drug development have been taken since the discovery of AIDS, which have brought about quite a few approved antiretroviral drugs [43]. Antiretroviral drugs, which can prevent the replication of HIV and restrain the transmission and progression to AIDS, are currently the only approved therapies specifically targeting HIV, and there were 27.5 million or so patients receiving antiretroviral therapies (ART) in 2020 [42]. However, ATR can not clear HIV due to the persistence of low-level viraemia from virus reservoirs which are insensitive to ATR [33]. Despite intensive research and tremendous progress, there is still not an explicit explanation for the pathogenesis of HIV infection. Effective vaccines and permanent cure therapies for HIV have not been obtained and thus life-long treatment is necessary.

    In the past decades, based on ongoing technology development and extensive clinical trials, a new research field, viral dynamics, arose. Various mathematical models on HIV are common and applicable to interpret and predict the time-course of viral levels during HIV infection process. Understanding threshold dynamics of virus models can be significant to design preventive measures, intervention means and treatment strategies for the infectious disease control and help to take more effective drug therapies in clinical practice [2,6,12,21]. A basic within-host virus dynamics model consists of three compartments: uninfected target cells T(t), infected cells I(t) and free virus V(t), constituting the following virus model [2,13]:

    {dTdt=ΛdT(t)βT(t)V(t),dIdt=βT(t)V(t)δII(t),dVdt=NδII(t)kV(t), (1.1)

    where the target cells are produced at the constant rate Λ and lost to intrinsic mortality rate d. The viral infection on susceptible cells is modelled by a bilinear incidence function βT(t)V(t). δI and k denote the corresponding death rates of infected cells and free virus, respectively. Per infected cell releases N particles in its lifespan.

    In virus model (1.1), uninfected healthy cells are assumed to be infected only directly by free viruses, that is, through the virus-to-cell transmission in the bloodstream, in which virions are released from infected cells and then move randomly around to find a new uninfected target cell to infect. 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 consisting of attachment of viruses to T cells, fusion of viruses into the T cells, replication and assembling of viruses inside the infected T cells, release of newly produced viral particles from the infected cells, and diffusion of the released viral particles to catch other T cells. However, recent studies have revealed that the cell-to-cell transmission also has a significant impact on the virus infection [31] and cell-to-cell spread mode may be more effective than virus-to-cell spread model in transmitting HIV-1 [7,8,24,30]. A large number of viral particles can also be simultaneously translocated from infected cells to uninfected cells through the structures termed virological synapses [4,11]. Sigal et al. [33] investigated that despite ART, cell-to-cell transmission of HIV permits ongoing replication.

    Besides, it is not instantaneous to produce new virus particles in the process of viral spread. It is delayed by the time for virions entering into cells and the replication of new virions, including the process of the transcription and integration of RNA and the production of the capsid proteins. Since the maturation time of a virus (0.3 days for HIV) is much shorter than the life-span of infected cell [10,23], researchers usually ignored the virus maturation time in viral infection dynamics. Recently, researchers have constructed and analyzed delayed intra-host viral infection models incorporating both virus-to-cell and cell-to-cell transmissions of virions [18,32,34,40]. Spatial diffusion is also significant to describe viral transmission. In Wang and Wang [36], a model incorporating the random movement of viruses was proposed and analyzed while the motion of corresponding cells was ignored, based on which McCluskey and Yang [20] and Sun and Wang [34] further considered general incidence rates and two transmission routes. In Teng et al. [35], a reaction-diffusion virus infection model with nonlinear incidence and humoral immunity was proposed, in which the random movement of individuals was also denoted by Fickian diffusions. In Wang et al. [37], extinction and persistence for a spatially heterogeneous HIV model and the global stability of steady states for corresponding spatially homogeneous model were explored. Further, in view of the propagation mechanism for the virus, age structure is a helpful tool to optimize the model [14,41].

    In addition, effective medical therapies can limit viral replication to a low level while virus particles cannot be eliminated. A critical factor is the existence of latently infected cells, which can not be ignored in clinic treatment and mathematical modelling. The latent reservoir persists mainly as proviruses integrated into the genomes of infected resting memory CD4+ T cells [5,39]. Latently infected CD4+ T cells live long and can not be affected by antiretroviral drugs or immune responses, while they can be activated by relevant antigens to produce virus. Latently infected cells have been a topic of great interest since they were subsequently shown to persist even in individuals on highly active antiretroviral therapy (HAART) [17].

    Motivated by above works, in this paper, we focus on the following delayed virus model with general functions, incorporating latent viral infection, both virus-to-cell transmission and cell-to-cell transmission

    (1.2)

    In model (1.2), L(t) represents the concentration of latently infected cells at time t with the mortality rate δL. Latently infected cells can be activated by relevant antigens to become productively infected cells at the rate α. Here we introduce the constant ξ(0,1) to specify the proportion of infection that lead target cells into latency stage. The parameter τ10 and τ20 represent the intracellular latency for the virus-to-cell infection and the cell-to-cell infection. m1 and m2 denote the constant death rates of latently infected cells and infected cells which have not produced viruses. Then emiτi is the probability for infected cells to survive from time tτi to t, i=1,2. In system (1.2), the virus-to-cell spread and the cell-to-cell spread are expressed by the incidence functions f(T,V) and g(T,I), which are assumed to satisfy the following assumptions:

    (i) f,gC1(R2+,R+) are differentiable; f(T,0)=f(0,V)=g(T,0)=g(0,I)=0 for all T,I,V0, f(T,V)>0 and g(T,I)>0 for all T,I,V>0.

    (ii) f(T,V)T>0 and g(T,I)T>0 for all T0 and V,I>0; f(T,V)V0 and g(T,I)I0 for all T,V,I0.

    (iii) 2f(T,V)TV0, 2g(T,I)TI0, f(T,V)Vf(T,V)V and g(T,I)Ig(T,I)I for all T,V,I0. This indicates that V(f(T,V)V)0 and I(g(T,I)I)0.

    Here, we are concerned with the general interaction functions to express the two modes of virus transmission. That is, the contribution of the interaction between uninfected target cells T and free viruses V (infected cells I) to the growth rate of the infected cells is represented by a general functional response term f(T,V) (g(T,I)), no longer accounted for by some specific function. Through this approach, we establish a unified theoretical framework to describe the HIV propagation process.

    Now define the following Banach space

    C+={ϕC([τ,0],R+)ϕ(θ) is uniformly continuous for θ[τ,0] and ϕ<},

    with norm ϕ=supτθ0|ϕ(θ)|. Then, we consider system (1.2) with the following initial conditions

    T(θ)=ϕ1(θ),L(θ)=ϕ2(θ),I(θ)=ϕ3(θ),V(θ)=ϕ4(θ),θ[τ,0], (1.3)

    where ϕi(θ)C+ satisfies ϕi(θ),0 and ϕi(0)>0,i=1,2,3,4.

    In this paper, we consider the existence, local and global asymptotical stability of the infection-free equilibrium and the infected equilibrium in terms of the basic reproduction number R0. For each equilibrium, we first explore the local asymptotical stability by considering the corresponding characteristic equations, and then discuss their global attractiveness by constructing corresponding Liapunov functionals, arriving at the global asymptotical stability of the equilibria.

    The paper is organized as follows. In Section 2, we study the boundedness and positivity of solutions and the existence of equilibria for system (1.2). In Section 3, we explore the dynamics of system without time delay. In Sections 4 and 5, we explore the corresponding local and global stability of infection-free equilibrium and infected equilibrium, respectively. In Section 6, some specific examples and applications are presented to illustrate the theoretical results. Conclusions and discussions can be found in Section 7.

    Through the fundamental theory analysis on functional differential equations [9], system (1.2) with initial conditions (1.3) admits one unique solution. In this section, we first explore the positivity and boundedness of the solutions of system (1.2).

    Theorem 2.1. Solutions of system (1.2) are positive and ultimately uniformly bounded for all t>0.

    Proof. For T(t), suppose that there exists t1>0 such that T(t1)=0 and T(t)>0 for t[0,t1). Then we have T(t1)0 while the first equation of (1.2) implies that T(t1)=Λ>0. This is a contradiction. Thus, T(t)>0 for t>0.

    Let t2>0 be the first time such that min{L(t2),I(t2),V(t2)}=0. In the following, we verify the non-existence of t2 to ensure the positivity of L(t), I(t) and V(t).

    (I) If L(t2)=0 and L(t)>0 for t[0,t2) (I(t2)0,V(t2)0 and I(t)>0,V(t)>0 for t[0,t2)), then L(t2)0. From the second equation of (1.2), we have

    L(t2)=ξem1τ1f(T(t2τ1),V(t2τ1))+ξem1τ1g(T(t2τ1),I(t2τ1))>0,

    which contradicts with L(t2)0. Thus, L(t)>0 for t>0.

    (II) Similarly, if I(t2)=0 and I(t)>0 for t[0,t2) (L(t2)0,V(t2)0 and L(t)>0,V(t)>0 for t[0,t2)), then I(t2)0. From the third equation of (1.2), we have

    I(t2)=(1ξ)em2τ2f(T(t2τ2),V(t2τ2))+(1ξ)em2τ2g(T(t2τ2),I(t2τ2))+αL(t2)>0,

    which contradicts with I(t2)0. Thus, I(t)>0 for t>0.

    (III) Again, if V(t2)=0 and V(t)>0 for t[0,t2) (L(t2)0,I(t2)>0 and L(t)>0,I(t)>0 for t[0,t2)), then V(t2)0. From the fourth equation of (1.2), we have

    V(t2)=NδII(t2)>0,

    which contradicts with V(t2)0. Thus, V(t)>0 for t>0.

    Hence, from above discussion, T(t)>0,L(t)>0,I(t)>0 and V(t)>0 for all t>0. In the following, we verify the boundedness.

    From the first equation of (1.2), we obtain T(t)ΛdT(t). This implies that lim supt+T(t)Λd.

    Let W1(t)=ξem1τ1T(tτ1)+L(t). Then

    W1(t)Λξem1τ1min{d,α+δL}W1.

    This implies that lim supt+W1(t)Λξem1τ1min{d,α+δL}. Thus, we have lim supt+L(t)Λξem1τ1min{d,α+δL}.

    Let W2(t)=(1ξ)em2τ2T(tτ2)+I(t). Then

    W2(t)[Λ(1ξ)em2τ2+Λξem1τ1min{d,α+δL}]min{d,δI}W2.

    This implies that lim supt+W2(t)Λ(1ξ)em2τ2min{d,δI}+αΛξem1τ1min{d,δI}min{d,α+δL}:=˜K. Thus, lim supt+I(t)˜K.

    From the fourth equation of (1.2), we have V(t)NδI˜KcV(t). Thus, we have lim supt+V(t)NδI˜Kc.

    Hence, solutions of system (1.2) are positive and ultimately uniformly bounded for all t>0.

    Define the following bounded feasible region

    Ω={(T,L,I,V)C4+TΛd,LΛξem1τ1min{d,α+δL},I˜K,VNδI˜Kc}.

    Then from Theorem 2.1, Ω is a positively invariant set for system (1.2).

    Secondly, we explore the existence of equilibria for system (1.2). To this end, we define

    K:=αξα+δLem1τ1+(1ξ)em2τ2

    and the following basic reproduction number

    R0:=K(Ncf(Λd,0)V+1δIg(Λd,0)I).

    Then, the existence of equilibria of system (1.2) is determined by the sign of R01.

    Theorem 2.2. If R01, then system (1.2) only has an infection-free equilibrium E0; whereas, if R0>1, then there exist E0 and an infected equilibrium E.

    Proof. The infection-free equilibrium E0=(Λd,0,0,0) always exists. In order to explore the existence of infected equilibrium E=(T,L,I,V), we need to discuss the following equations

    {ΛdTf(T,V)g(T,I)=0,ξf(T,V)+ξg(T,I)(α+δL)em1τ1L=0,(1ξ)f(T,V)+(1ξ)g(T,I)δIem2τ2I+αem2τ2L=0,NδIIcV=0. (2.1)

    From the first and second equations of (2.1), we obtain

    T=Λdα+δLdξem1τ1L. (2.2)

    Let

    J:=1ξξ(α+δL)em1τ1m2τ2+α=α+δLξem1τ1K. (2.3)

    Then, due to the second and third equations of (2.1), we have

    I=JδIL. (2.4)

    The fourth equation of (2.1) yields

    V=NδIcI=NJcL. (2.5)

    By substituting (2.2), (2.4) and (2.5) into the second and third equations of (2.1), we define the following auxiliary mapping:

    Φ(L):=f(T,V)+g(T,I)δIem2τ2I+αem2τ2L(α+δL)em1τ1L=f(Λdα+δLdξem1τ1L,NJcL)+g(Λdα+δLdξem1τ1L,JδIL)α+δLξem1τ1L.

    Clearly, Φ(0)=f(Λd,0)+g(Λd,0)=0 and Φ(Λξα+δLem1τ1)=Λ<0. Besides, when R0>1, we have

    Φ(0)=J(Ncf(Λd,0)V+1δIg(Λd,0)I)α+δLξem1τ1=α+δLξem1τ1(R01)>0.

    Thus, there exists L(0,Λξα+δLem1τ1), such that E=(T,L,I,V) exists.

    Moreover, we can verify that E is unique. On (T,L,I,V), due to the second equation of (2.1), (2.5) and (2.4), there holds

    α+δLξem1τ1=f(T,V)+g(T,I)L=J(Ncf(T,V)V+1δIg(T,I)I).

    Then under Assumptions (ii) and (iii), we further have

    Φ(L)=α+δLdξem1τ1(f(T,V)T+g(T,I)T)+J(Ncf(T,V)V+1δIg(T,I)I)α+δLξem1τ1=α+δLdξem1τ1(f(T,V)T+g(T,I)T)+NJc(f(T,V)Vf(T,V)V)+JδI(g(T,I)Ig(T,I)I)<0.

    Thus, when R0>1, there exists a unique infected equilibrium E=(T,L,I,V); when R0<1, there exists no infected equilibrium. Whereas, when R0=1, Φ(0)=0 and Φ(0)0. If Φ(0)<0, then there exists no infected equilibrium. If Φ(0)=0, then Φ(l)(0)=0,l=3,4,..., and due to Assumption (ii), for any I(0,ΛKδI),

    Φ(I)=δIKd(f(ΛδIKId,NδIcI)T+g(ΛδIKId,I)T)+NδIcf(ΛδIKId,NδIcI)V+g(ΛδIKId,I)IδIK<NδIcf(Λd,0)V+g(Λd,0)IδIK=δIK(R01)=0.

    Thus, in this case, there exists no infected equilibrium.

    In this section, we consider system (1.2) without time delay. Let T=ς1, L=ς2, I=ς3, V=ς4, then we discuss the following ODE system:

    {dς1dt=Λdς1(t)f(ς1(t),ς4(t))g(ς1(t),ς3(t)):=χ1,dς2dt=ξf(ς1(t),ς4(t))+ξg(ς1(t),ς3(t))(α+δL)ς2(t):=χ2,dς3dt=(1ξ)f(ς1(t),ς4(t))+(1ξ)g(ς1(t),ς3(t))δIς3(t)+ας2:=χ3,dς4dt=NδIς3(t)cς4(t):=χ4. (3.1)

    Through similar analysis as in Section 2, for system (3.1), R0:=K(Ncf(Λd,0)ς4+1δIg(Λd,0)ς3) and we have the following result.

    Theorem 3.1. If R0<1, then the infection-free equilibrium E0 is locally asymptotically stable. If R0>1, then E0 is unstable. If R0=1, then E0 is linearly neutrally stable and there exists a forward bifurcation around E0.

    Proof. The characteristic equation of the linearization for system (3.1) is (λ+d)G0(λ)=0, where

    G0(λ)=(λ+α+δL)(λ+c)(λ+δI)[NδIf(Λd,0)ς4+(λ+c)g(Λd,0)ς3][αξ+(λ+α+δL)(1ξ)].

    Obviously, there is always a negative root λ=d and G0(λ)=0 can be written as

    (3.2)

    By dividing (λ+α+δL)(λ+c)(λ+δI) in both sides of (3.2), we obtain

    1=[NδI(λ+c)(λ+δI)f(Λd,0)ς4+1λ+δIg(Λd,0)ς3][αξλ+α+δL+(1ξ)]. (3.3)

    Suppose that there exists a eigenvalue λ=ι+κi(ι0). Then the modulus of both sides of (3.3) satisfies

    1<(Ncf(Λd,0)ς4+1δIg(Λd,0)ς3)[αξα+δL+(1ξ)]=R0.

    This contradicts with R0<1. Thus, all roots of (3.2) have negative real parts. Hence, when R0<1, E0 is locally asymptotically stable.

    On the other hand, when R0>1, there holds

    G0(0)=cδI(α+δL)(NδIf(Λd,0)ς4+cg(Λd,0)ς3)[αξ+(α+δL)(1ξ)]=cδI(α+δL)(1R0)<0

    and G0(+)=+. Thus, there exists at least one real root λ0>0, such that G0(λ0)=0. Hence, if R0>1, then E0 is unstable.

    When R0=1, characteristic equation (3.2) has one simple zero root λ=0 and two roots with negative real part. Thus E0 is a linearly neutrally stable non-hyperbolic equilibrium as R0=1. We can further verify that system (3.1) exhibits forward bifurcation around E0 at ξ:=(αδL+1)(11ˆK), where ˆK=Ncf(Λd,0)ς4+1δIg(Λd,0)ς3, that is, R0=1.

    Let L0(E0,ξ) be the Jacobian matrix of the system (3.1) at ξ=ξ. The left eigenvector \bf{ \pmb{\mathsf{ μ}}} = (\mu_1, \mu_2, \mu_3, \mu_4) of the Jacobian matrix \mathcal{L}_0(E_0, \xi^*) is given by \bf{ \pmb{\mathsf{ μ}}}\cdot\mathcal{L}_0(E_0, \xi^*) . We obtain

    (\mu_1,\mu_2,\mu_3,\mu_4) = (0,\alpha,\alpha+\delta_L, \frac{\alpha+\delta_L}{c}\left[\frac{\alpha\xi}{\alpha+\delta_L}+(1-\xi)\right]\frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4}).

    The right eigenvector \bf{ \pmb{\mathsf{ ω}}} = (\omega_1, \omega_2, \omega_3, \omega_4) of the Jacobian matrix \mathcal{L}_0(E_0, \xi^*) is given by \mathcal{L}_0(E_0, \xi^*)\cdot \bf{ \pmb{\mathsf{ ω}}} . We obtain

    (\omega_1,\omega_2,\omega_3,\omega_4)^T = ( -\frac{1}{d}\left(N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4} +c\frac{\partial g(\frac{\Lambda}{d},0)}{\partial \varsigma_3}\right), \frac{\xi}{\alpha+\delta_L}\left(N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4} +c\frac{\partial g(\frac{\Lambda}{d},0)}{\partial \varsigma_3}\right), 0,N\delta_I)^T.

    Let \chi_i, \; i = 1, 2, 3, 4, denotes the right-hand side of system (3.1). Then we obtain the following non-zero derivations

    \begin{equation*} \begin{split} &\left(\frac{\partial^2 \chi_1}{\partial\varsigma_3^2}\right)_{E_0} = -\frac{\partial^2 g(\frac{\Lambda}{d},0)}{\partial \varsigma^2_3},\; \; \left(\frac{\partial^2 \chi_1}{\partial\varsigma_4^2}\right)_{E_0} = -\frac{\partial^2 f(\frac{\Lambda}{d},0)}{\partial \varsigma^2_4},\\ &\left(\frac{\partial^2 \chi_2}{\partial\varsigma_3^2}\right)_{E_0} = \xi\frac{\partial^2 g(\frac{\Lambda}{d},0)}{\partial \varsigma^2_3},\; \; \left(\frac{\partial^2 \chi_2}{\partial\varsigma_4^2}\right)_{E_0} = \xi\frac{\partial^2 f(\frac{\Lambda}{d},0)}{\partial \varsigma^2_4},\\ & \left(\frac{\partial^2 \chi_3}{\partial\varsigma_3^2}\right)_{E_0} = (1-\xi)\frac{\partial^2 g(\frac{\Lambda}{d},0)}{\partial \varsigma^2_3},\; \; \left(\frac{\partial^2 \chi_3}{\partial\varsigma_4^2}\right)_{E_0} = (1-\xi)\frac{\partial^2 f(\frac{\Lambda}{d},0)}{\partial \varsigma^2_4},\\ &\left(\frac{\partial^2 \chi_2}{\partial\varsigma_3\partial\xi^*}\right)_{E_0} = \frac{\partial g(\frac{\Lambda}{d},0)}{\partial \varsigma_3},\; \; \left(\frac{\partial^2 \chi_2}{\partial\varsigma_4\partial\xi^*}\right)_{E_0} = \frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4},\\ &\left(\frac{\partial^2 \chi_3}{\partial\varsigma_3\partial\xi^*}\right)_{E_0} = -\frac{\partial g(\frac{\Lambda}{d},0)}{\partial \varsigma_3},\; \; \left(\frac{\partial^2 \chi_3}{\partial\varsigma_4\partial\xi^*}\right)_{E_0} = -\frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4}.\\ \end{split} \end{equation*}

    Due to [3], we yield the following bifurcation constants \varrho and \sigma :

    \begin{equation*} \begin{split} \varrho& = \sum\limits_{i,j,k = 1}^{4} \mu_k\omega_i\omega_j \left(\frac{\partial^2 \chi_k} {\partial\varsigma_i\partial\varsigma_j} \right)_{E_0} = \left[\alpha\xi^2+(\alpha+\delta_L)(1-\xi)^2\right] \left(N^2\delta^2_I\frac{\partial^2 f(\frac{\Lambda}{d},0)}{\partial \varsigma_4^2} +c^2\frac{\partial^2 g(\frac{\Lambda}{d},0)}{\partial \varsigma_3^2} \right)\leq0,\\ \sigma& = \sum\limits_{i,k = 1}^{4} \mu_k\omega_i\left(\frac{\partial^2 \chi_k} {\partial\varsigma_i\partial\xi^*} \right)_{E_0} = -\delta_L \left(N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial \varsigma_4} +c\frac{\partial g(\frac{\Lambda}{d},0)}{\partial \varsigma_3} \right)\leq0.\\ \end{split} \end{equation*}

    Under conditions \frac{\partial^2 f(\frac{\Lambda}{d}, 0)}{\partial \varsigma_4^2} +\frac{\partial^2 g(\frac{\Lambda}{d}, 0)}{\partial \varsigma_3^2}\neq0 and \frac{\partial f(\frac{\Lambda}{d}, 0)}{\partial \varsigma_4} +\frac{\partial g(\frac{\Lambda}{d}, 0)}{\partial \varsigma_3}\neq0 , a_1 < 0 and b_1 < 0 . Thus, due to Theorem 4.1 in [3], there exists a forward bifurcation. When \xi < \xi^* , that is, \mathfrak{R}_0 > 1 , E_0 is unstable and there exists an locally asymptotically stable positive equilibrium E^* ; when \xi > \xi^* , that is, \mathfrak{R}_0 < 1 , E_0 is locally asymptotically stable.

    Note that the presence of forward bifurcation confirms the inhibition of the disease when \mathfrak{R}_0 < 1 . From the forward bifurcation analysis, we know that if there is a stable coexistence equilibrium E^* bifurcating from E_0 , E_0 changes its stability from stable to unstable.

    In this section, we investigate the stability of the infection-free equilibrium E_0 . We first focus on the local asymptotical stability of E_0 by discussing the distribution of the corresponding characteristic values.

    Theorem 4.1. If \mathfrak{R}_0 < 1 , then the infection-free equilibrium E_0 is locally asymptotically stable. If \mathfrak{R}_0 = 1 , then E_0 is linearly neutrally stable. If \mathfrak{R}_0 > 1 , then E_0 is unstable.

    Proof. Let \mathrm{X} = (T(t), L(t), I(t), V(t)) and \mathrm{X_{\tau_i}} = (T(t-{\tau_i}), L(t-{\tau_i}), I(t-{\tau_i}), V(t-{\tau_i})), \; i = 1, 2 . Then the linearization of system (1.2) at E_0 can be expressed by

    \frac{{\rm d} \mathrm{X}}{{\rm d} t} = \mathcal{L}_0\mathrm{X} +\mathcal{M}_{01}\mathrm{X_{\tau_1}} +\mathcal{M}_{02}\mathrm{X_{\tau_2}},

    where

    \begin{gather*} \mathcal{L}_0 = \begin{pmatrix} -d & 0 &-\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I} & -\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V}\\ 0 & -(\alpha+\delta_L) & 0 & 0\\ 0 & \alpha & -\delta_I & 0\\ 0 & 0 & N\delta_I & -c\\ \end{pmatrix},\\ \end{gather*}
    \begin{gather*} \mathcal{M}_{01} = \begin{pmatrix} 0 & 0 & 0 & 0\\ 0 & 0 & \xi e^{-m_1\tau_1}\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I} & \xi e^{-m_1\tau_1}\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V}\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ \end{pmatrix}\\ \end{gather*}

    and

    \begin{gather*} \mathcal{M}_{02} = \begin{pmatrix} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & (1-\xi)e^{-m_2\tau_2}\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I} & (1-\xi)e^{-m_2\tau_2}\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V}\\ 0 & 0 & 0 & 0\\ \end{pmatrix}. \end{gather*}

    Then the characteristic equation is

    \begin{equation*} det(\lambda I-\mathcal{L}_0-\mathcal{M}_{01}e^{-\lambda\tau_{1}} -\mathcal{M}_{02}e^{-\lambda\tau_{2}}) = (\lambda+d)G_0(\lambda) = 0, \end{equation*}

    where

    \begin{equation*} \begin{split} G_0(\lambda) = &(\lambda+\alpha+\delta_L)(\lambda+c)(\lambda+\delta_I)\\ &-\left[N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V} +(\lambda+c)\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I}\right] \left[\alpha\xi e^{-(m_1+\lambda)\tau_1} +(\lambda+\alpha+\delta_L)(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]. \end{split} \end{equation*}

    Obviously, there is always a negative root \lambda = -d and G_0(\lambda) = 0 can be written as

    \begin{equation} \begin{split} &(\lambda+\alpha+\delta_L)(\lambda+c)(\lambda+\delta_I) = \\ &\left[N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V} +(\lambda+c)\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I}\right] \left[\alpha\xi e^{-(m_1+\lambda)\tau_1} +(\lambda+\alpha+\delta_L)(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]. \end{split} \end{equation} (4.1)

    By dividing (\lambda+\alpha+\delta_L)(\lambda+c)(\lambda+\delta_I) in both sides of (4.1), we obtain

    (4.2)

    Suppose that there exists a eigenvalue \lambda = \iota+\kappa i\; (\iota\geq0) . Then the modulus of both sides of (4.2) satisfies

    \begin{equation*} \begin{split} 1 < \left(\frac{N}{c}\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V} +\frac{1}{\delta_I}\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I}\right) \left[\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} +(1-\xi) e^{-m_2\tau_2}\right] = \mathfrak{R}_0. \end{split} \end{equation*}

    This contradicts with \mathfrak{R}_0 < 1 . Thus, all roots of (4.1) have negative real parts. Hence, when \mathfrak{R}_0 < 1 , E_0 is locally asymptotically stable. Further, due to similar analysis, when \mathfrak{R}_0 = 1 , any root of characteristic equation (4.1) has negative real part except a simple zero root \lambda = 0 and thus E_0 is linearly neutrally stable.

    On the other hand, when \mathfrak{R}_0 > 1 , there holds

    \begin{equation*} \begin{split} G_0(0) = &c\delta_I(\alpha+\delta_L) -\left(N\delta_I\frac{\partial f(\frac{\Lambda}{d},0)}{\partial V} +c\frac{\partial g(\frac{\Lambda}{d},0)}{\partial I}\right) \left[\alpha\xi e^{-m_1\tau_1} +(\alpha+\delta_L)(1-\xi) e^{-m_2\tau_2}\right]\\ = &c\delta_I(\alpha+\delta_L)(1-\mathfrak{R}_0) < 0 \end{split} \end{equation*}

    and G_0(+\infty) = +\infty . Thus, there exists at least one real root \lambda_0 > 0 , such that G_0(\lambda_0) = 0 . Hence, if \mathfrak{R}_0 > 1 , then E_0 is unstable.

    Based on the local asymptotical stability analysis, we construct a Lyapunov functional to verify the global asymptotical stability of E_0 . Similar to the local asymptotical stability of E_0 , the following theorem shows that the basic reproduction number \mathfrak{R}_0 acts as the threshold value for the global asymptotical stability of E_0 .

    Theorem 4.2. If \mathfrak{R}_0 < 1 , then the infection-free equilibrium E_0 is globally asymptotically stable; if \mathfrak{R}_0 = 1 , then E_0 is globally attractive.

    Proof. Let

    \begin{equation*} \begin{split} \mathfrak{R}_{01} = \frac{NK}{c}\frac{\partial f(\frac{\Lambda}{d},0)} {\partial V} \; \; \text{and}\; \; \mathfrak{R}_{02} = \frac{K}{\delta_I}\frac{\partial g(\frac{\Lambda}{d},0)} {\partial I}. \end{split} \end{equation*}

    Then \mathfrak{R}_{0} = \mathfrak{R}_{01}+\mathfrak{R}_{02} . Define the following Liapunov functional

    \begin{equation*} \begin{split} \mathcal{V}_0(t) = &\frac{\alpha}{\alpha+\delta_L}L+I+\frac{1-\mathfrak{R}_{02}}{N}V\\ &+ \frac{\alpha\xi}{\alpha+\delta_L}e^{-m_1\tau_1} \int^{\tau_1}_0(f(T(t-\theta),V(t-\theta))+g(T(t-\theta),I(t-\theta))) {\rm d}\theta\\ & + (1-\xi)e^{-m_2\tau_2} \int^{\tau_2}_0(f(T(t-\theta),V(t-\theta))+g(T(t-\theta),I(t-\theta))) {\rm d}\theta. \end{split} \end{equation*}

    Calculating the time derivative of \mathcal{V}_0(t) along (1.2) yields

    \begin{equation*} \begin{split} \frac{{\rm d}\mathcal{V}_0}{{\rm d}t} = &K(f(T,V)+g(T,I))-\delta_I\mathfrak{R}_{02}I-(1-\mathfrak{R}_{02})\frac{c}{N}V\\ = &\left[K\frac{f(T,V)}{V}-(1-\mathfrak{R}_{02})\frac{c}{N}\right]V +\left(K\frac{g(T,I)}{I}-\delta_I\mathfrak{R}_{02}\right)I.\\ \end{split} \end{equation*}

    From Assumption (iii), we know \frac{f(\frac{\Lambda}{d}, V)}{V} is decreasing with respect to V and \frac{g(\frac{\Lambda}{d}, I)}{I} is decreasing with respect to I . Thus, when \mathfrak{R}_0 < 1 ,

    \begin{equation*} \begin{split} \frac{{\rm d}\mathcal{V}_0}{{\rm d}t} \leq&\left[K\lim\limits_{V\rightarrow0}\frac{f(\frac{\lambda}{d},V)}{V}-(1-\mathfrak{R}_{02})\frac{c}{N}\right]V +\left(K\lim\limits_{I\rightarrow0}\frac{g(\frac{\lambda}{d},I)}{I}-\delta_I\mathfrak{R}_{02}\right)I\\ = &\left[\frac{NK}{c}\frac{\partial f(\frac{\lambda}{d},0)}{\partial V} -(1-\mathfrak{R}_{02})\right]\frac{c}{N}V +\left(\frac{K}{\delta_I}\frac{\partial g(\frac{\lambda}{d},0)}{\partial I} -\mathfrak{R}_{02}\right)\delta_II\\ = &[\mathfrak{R}_{01}-(1-\mathfrak{R}_{02})]\frac{c}{N}V+(\mathfrak{R}_{02}-\mathfrak{R}_{02})\delta_II\\ = &(\mathfrak{R}_{0}-1)\frac{c}{N}V\leq0. \end{split} \end{equation*}

    Let \Theta_0 = \left\{(T, L, I, V):\frac{{\rm d}\mathcal{V}_0}{{\rm d}t} = 0\right\} . Then the largest invariant subset of \Theta_0 just consists of E_0 . Thus, due to the LaSalle's invariance principle [15], when \mathfrak{R}_0\leq1 , E_0 is a global attractor. Further, combining the local asymptotical stability of E_0 under the condition \mathfrak{R}_0 < 1 , E_0 is globally asymptotically stable.

    In this section, we analyze the stability of the infected equilibrium E^* . In order to explore the local asymptotical stability of E^* , we need to analyze the characteristic equation of system (1.2) on E^* . The linearization of system (1.2) at E^* can be expressed by

    \frac{{\rm d} \mathrm{X}}{{\rm d} t} = \mathcal{L}_1\mathrm{X} +\mathcal{M}_{11}\mathrm{X_{\tau_1}} +\mathcal{M}_{12}\mathrm{X_{\tau_2}},

    where

    \begin{gather*} \mathcal{L}_1 = \begin{pmatrix} -d-\frac{\partial g(T^*,I^*)}{\partial T}-\frac{\partial f(T^*,V^*)}{\partial T} & 0 & -\frac{\partial g(T^*,I^*)}{\partial I} & -\frac{\partial f(T^*,V^*)}{\partial V} \\ 0 & -(\alpha+\delta_L) & 0 & 0\\ 0 & \alpha & -\delta_I & 0\\ 0 & 0 & N\delta_I & -c\\ \end{pmatrix},\\ \end{gather*}
    \begin{gather*} \mathcal{M}_{11} = \begin{pmatrix} 0 & 0 & 0 & 0\\ \xi e^{-m_1\tau_1}\left(\frac{\partial g(T^*,I^*)}{\partial T} +\frac{\partial f(T^*,V^*)}{\partial T}\right) & 0 & \xi e^{-m_1\tau_1}\frac{\partial g(T^*,I^*)}{\partial I} & \xi e^{-m_1\tau_1}\frac{\partial f(T^*,V^*)}{\partial V}\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ \end{pmatrix}\\ \end{gather*}

    and

    \begin{array}{l} \mathcal{M}_{12} =\\ \begin{pmatrix} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ (1-\xi)e^{-m_2\tau_2}\left(\frac{\partial g(T^*,I^*)}{\partial T} +\frac{\partial f(T^*,V^*)}{\partial T}\right) & 0 & (1-\xi)e^{-m_2\tau_2}\frac{\partial g(T^*,I^*)}{\partial I} & (1-\xi)e^{-m_2\tau_2}\frac{\partial f(T^*,V^*)}{\partial V}\\ 0 & 0 & 0 & 0\\ \end{pmatrix}. \end{array}

    Then the characteristic equation is

    \begin{equation*} \begin{split} &det(\lambda I-\mathcal{L}_1-\mathcal{M}_{11}e^{-\lambda\tau_{1}} -\mathcal{M}_{12}e^{-\lambda\tau_{2}})\\ = &\left(\lambda+d+\frac{\partial g(T^*,I^*)}{\partial T} +\frac{\partial f(T^*,V^*)}{\partial T}\right) (\lambda+\alpha+\delta_L) (\lambda+c)(\lambda+\delta_I)\\ &-(\lambda+d)\left[N\delta_I\frac{\partial f(T^*,V^*)}{\partial V} +(\lambda+c)\frac{\partial g(T^*,I^*)}{\partial I}\right]\\ & \times \left[\alpha\xi e^{-(m_1+\lambda)\tau_1} +(\lambda+\alpha+\delta_L)(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]\\ = &0, \end{split} \end{equation*}

    that is,

    \begin{equation} \begin{split} &\left(\lambda+d+\frac{\partial g(T^*,I^*)}{\partial T} +\frac{\partial f(T^*,V^*)}{\partial T}\right) (\lambda+\alpha+\delta_L) (\lambda+c)(\lambda+\delta_I)\\ = &(\lambda+d)\left[N\delta_I\frac{\partial f(T^*,V^*)}{\partial V} +(\lambda+c)\frac{\partial g(T^*,I^*)}{\partial I}\right]\\ & \times \left[\alpha\xi e^{-(m_1+\lambda)\tau_1} +(\lambda+\alpha+\delta_L)(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]. \end{split} \end{equation} (5.1)

    By dividing (\lambda+\alpha+\delta_L)(\lambda+c)(\lambda+\delta_I)(\lambda+d) in two sides of system (5.1), we obtain

    \begin{equation} \begin{split} &\frac{\lambda+d+\frac{\partial g(T^*,I^*)}{\partial T} +\frac{\partial f(T^*,V^*)}{\partial T}}{\lambda+d}\\ = &\left[\frac{N\delta_I\frac{\partial f(T^*,V^*)}{\partial V}} {(\lambda+c)(\lambda+\delta_I)} +\frac{\frac{\partial g(T^*,I^*)}{\partial I}} {\lambda+\delta_I}\right]\\ & \times \left[\frac{\alpha\xi}{\lambda+\alpha+\delta_L} e^{-(m_1+\lambda)\tau_1} +(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]. \end{split} \end{equation} (5.2)

    Suppose that there exists eigenvalue \lambda = \iota+\kappa i\; (\iota\geq0) . Then, since Assumption (ii), the modulus of the left side of (5.2) is more than one. Note that from the second and third equations of (2.1), we have

    \begin{equation} \begin{split} &\alpha L^* = \frac{\alpha\xi}{\alpha+\delta_L}e^{-m_1\tau_1}(f(T^*,V^*)+g(T^*,I^*)) \end{split} \end{equation} (5.3)

    and

    \begin{equation} \begin{split} &\delta I^* = (1-\xi)e^{-m_2\tau_2}(f(T^*,V^*)+g(T^*,I^*))+\alpha L^*. \end{split} \end{equation} (5.4)

    Then it follows from (5.3) and (5.4) that

    \begin{equation} \begin{split} \delta I^* = K\left(f(T^*,V^*)+g(T^*,I^*)\right). \end{split} \end{equation} (5.5)

    Due to Assumption (iii), V^* = \frac{N\delta_{I}}{c}I^* and (5.5), the other side of (5.2) satisfies

    \begin{equation*} \begin{split} &\left[\frac{N\delta_I\frac{\partial f(T^*,V^*)}{\partial V}} {(\lambda+c)(\lambda+\delta_I)} +\frac{\frac{\partial g(T^*,I^*)}{\partial I}} {\lambda+\delta_I}\right] \left[\frac{\alpha\xi}{\lambda+\alpha+\delta_L} e^{-(m_1+\lambda)\tau_1} +(1-\xi) e^{-(m_2+\lambda)\tau_2}\right]\\ \leq& K\left(\frac{N}{c}\frac{\partial f(T^*,V^*)}{\partial V} +\frac{1}{\delta_I}\frac{\partial g(T^*,I^*)}{\partial I}\right) \leq K\left(\frac{N}{c}\frac{f(T^*,V^*)}{V^*} +\frac{1}{\delta_I}\frac{g(T^*,I^*)}{I^*}\right)\\ \leq& K\left(\frac{N}{c}\frac{f(T^*,V^*)}{\frac{N\delta_{I}}{c}I^*} +\frac{1}{\delta_I}\frac{g(T^*,I^*)}{I^*}\right) = \frac{K}{\delta_{I}I^*}\left(f(T^*,V^*)+g(T^*,I^*)\right) = 1. \end{split} \end{equation*}

    This is a contradiction. Thus, all roots of (5.2) have negative real parts.

    Hence, we obtain the following theorem on local asymptotical stability of E^* .

    Theorem 5.1. If \mathfrak{R}_0 > 1 , then the infected equilibrium E^* is locally asymptotically stable.

    Moreover, using the similar proof of Section 4 in [16] and Theorem 3.1 in [38], we have the following theorem on uniformly persistence.

    Theorem 5.2. If \mathfrak{R}_0 > 1 , then system (1.2) is uniformly persistent, that is, there exists a constant \varpi > 0 such that \limsup\limits_{t\rightarrow +\infty}\min\{T(t), L(t), I(t), V(t)\}\geq \varpi .

    Based on the local asymptotical stability of the infected equilibrium E^* and the uniformly persistence of system (1.2), we further investigate the global asymptotical stability of E^* . For this purpose, we also need to make the following hypothesises on f(T, V) and g(T, I) :

    \bf{(A1):} \left(\frac{T^*f(T, V)}{Tf(T^*, V^*)}-1\right)\left(\frac{V}{V^*}-\frac{T^*f(T, V)}{Tf(T^*, V^*)}\right)\geq0 and \left(\frac{T^*g(T, I)}{Tg(T^*, I^*)}-1\right)\left(\frac{I}{I^*}-\frac{T^*g(T, I)}{Tg(T^*, I^*)}\right)\geq0 .

    Then, by virtue of the Volterra type function [19]

    h(\zeta) = \zeta-1-\ln\zeta,\; \zeta > 0,

    we can construct a Lyapunov functional to ensure E^* to be globally attractive with respect to all the solutions with non-negative initial values, arriving at the following results on global asymptotical stability.

    Theorem 5.3. Suppose that Assumption (A1) holds. If \mathfrak{R}_0 > 1 , then the infected equilibrium E^* is globally asymptotically stable.

    Proof. Define the following Liapunov functional \mathcal{V}_1(t) = \mathcal{V}_{11}(t)+\mathcal{V}_{12}(t) , where

    \begin{equation*} \begin{split} \mathcal{V}_{11}(t) = &K T^*h(\frac{T(t)}{T^*}) +\frac{\alpha L^*}{\alpha+\delta_L}h(\frac{L(t)}{L^*}) +I^*h(\frac{I(t)}{I^*}) +\frac{K}{c}f(T^*,V^*)h(\frac{V(t)}{V^*})\\ \end{split} \end{equation*}

    and

    \begin{equation*} \begin{split} & \mathcal{V}_{12}(t)\\ = &\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}f(T^*,V^*) \int^{\tau_1}_{0}h(\frac{f(T(t-\theta),V(t-\theta))}{f(T^*,V^*)}) {\rm d}\theta\\ &+(1-\xi)e^{-m_2\tau_2}f(T^*,V^*) \int^{\tau_2}_{0}h(\frac{f(T(t-\theta),V(t-\theta))}{f(T^*,V^*)}) {\rm d}\theta\\ & +\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}g(T^*,I^*) \int^{\tau_1}_{0}h(\frac{g(T(t-\theta),I(t-\theta))}{g(T^*,I^*)}) {\rm d}\theta\\ & +(1-\xi)e^{-m_2\tau_2}g(T^*,I^*) \int^{\tau_2}_{0}h(\frac{g(T(t-\theta),I(t-\theta))}{g(T^*,I^*)}) {\rm d}\theta. \end{split} \end{equation*}

    Since \Lambda = dT+f(T^*, V^*)+g(T^*, I^*) and cV^* = N\delta_I I^* , calculating the time derivative of \mathcal{V}_{11}(t) along (1.2) yields

    \begin{align} & \frac{{\rm d}\mathcal{V}_{11}}{{\rm d}t} \\ = &K\left(1-\frac{T^*}{T}\right)(\Lambda-dT-f(T,V)-g(T,I))\\ &+\frac{\alpha}{\alpha+\delta_L} \left(1-\frac{L^*}{L}\right) [\xi f(T(t-\tau_1),V(t-\tau_1))e^{-m_1\tau_1}\\ & +\xi g(T(t-\tau_1),I(t-\tau_1))e^{-m_1\tau_1}-(\alpha+\delta_L)L]\\ &+\left(1-\frac{I^*}{I}\right) [(1-\xi) f(T(t-\tau_2),V(t-\tau_2))e^{-m_2\tau_2}\\ & +(1-\xi) g(T(t-\tau_2), I(t-\tau_2))e^{-m_2\tau_2}-\delta_II+\alpha L]\\ &+K\frac{f(T^*,V^*)}{cV^*} \left(1-\frac{V^*}{V}\right) (N\delta_II-cV)\\ = &-K\frac{d}{T}(T-T^*)^2 +K(-f(T^*,V^*)-g(T^*,I^*)+f(T,V)+g(T,I)) \left(\frac{T^*}{T}-1\right)\\ &-\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} \left(\frac{L^*}{L}-1\right) (f(T(t-\tau_1),V(t-\tau_1))+g(T(t-\tau_1),I(t-\tau_1)))\\ &-(1-\xi) e^{-m_2\tau_2}\left(\frac{I^*}{I}-1\right) (f(T(t-\tau_2),V(t-\tau_2))+g(T(t-\tau_2),I(t-\tau_2)))\\ &+\alpha L^*+\delta_I(I^*-I)-\alpha L\frac{I^*}{I} +Kf(T^*,V^*)\left(1-\frac{V^*}{V}\right) \left(\frac{I}{I^*}-\frac{V}{V^*}\right). \end{align} (5.6)

    Moreover, due to (5.3), \delta_II^* = JL^* and (2.3), we have

    \begin{equation} \begin{split} \delta_I(I^*-I) = -K\left(\frac{I}{I^*}-1\right)(f(T^*,V^*)+g(T^*,I^*)). \end{split} \end{equation} (5.7)

    Then it follows from (5.3), (5.6) and (5.7) that

    \begin{align*} & \frac{{\rm d}\mathcal{V}_1}{{\rm d}t}\\ = &-K\frac{d}{T}(T-T^*)^2 -K(f(T^*,V^*)+g(T^*,I^*))\left(\frac{T^*}{T}-1\right)\\ &+K(f(T,V)+g(T,I))\frac{T^*}{T}\\ &-\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}\frac{L^*}{L} (f(T(t-\tau_1),V(t-\tau_1))+g(T(t-\tau_1),I(t-\tau_1)))\\ & +\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} (f(T^*,V^*)+g(T^*,I^*))\\ &-(1-\xi) e^{-m_2\tau_2}\frac{I^*}{I} (f(T(t-\tau_2),V(t-\tau_2))+g(T(t-\tau_2),I(t-\tau_2)))\\ & -K\left(\frac{I}{I^*}-1\right) (f(T^*,V^*)+g(T^*,I^*))\\ &-\frac{I^*L}{IL^*} \frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} (f(T^*,V^*)+g(T^*,I^*)) + Kf(T^*,V^*) \left(\frac{I}{I^*}-\frac{V}{V^*}-\frac{V^*I}{VI^*}+1\right)\\ &+\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} \left(f(T^*,V^*)\ln{\frac{f(T(t-\tau_1),V(t-\tau_1))}{f(T,V)}} +g(T^*,I^*)\ln{\frac{g(T(t-\tau_1),I(t-\tau_1))}{g(T,I)}}\right)\\ &+(1-\xi) e^{-m_2\tau_2} \left(f(T^*,V^*)\ln{\frac{f(T(t-\tau_2),V(t-\tau_2))}{f(T,V)}} +g(T^*,I^*)\ln{\frac{g(T(t-\tau_2),I(t-\tau_2))}{g(T,I)}}\right), \end{align*}

    that is,

    \begin{equation*} \begin{split} & \frac{{\rm d}\mathcal{V}_1}{{\rm d}t}\\ = &-K\frac{d}{T}(T-T^*)^2 -K(f(T^*,V^*)+g(T^*,I^*))\left(\frac{T^*}{T}-1\right)\\ +&K\left(\frac{T^*}{T}f(T,V)-\frac{I}{I^*}f(T^*,V^*)- \frac{V}{V^*}f(T^*,V^*)+\frac{I}{I^*}f(T^*,V^*)\right)\\ +&K\left(\frac{T^*}{T}g(T,I)-\frac{I}{I^*}g(T^*,I^*)\right) \\ -&\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}f(T^*,V^*) \left(\frac{f(T(t-\tau_1),V(t-\tau_1))L^*}{f(T^*,V^*)L}-1 \right)\\ -&\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}g(T^*,I^*) \left(\frac{g(T(t-\tau_1),I(t-\tau_1))L^*}{g(T^*,I^*)L}-1 \right)\\ -&(1-\xi) e^{-m_2\tau_2}f(T^*,V^*) \left(\frac{f(T(t-\tau_2),V(t-\tau_2))I^*}{f(T^*,V^*)I}-1 \right)\\ -&(1-\xi) e^{-m_2\tau_2}g(T^*,I^*) \left(\frac{g(T(t-\tau_2),I(t-\tau_2))I^*}{g(T^*,I^*)I}-1 \right)\\ -&\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}f(T^*,V^*) \left(\frac{I^*L}{IL^*}-1 \right)\\ -&\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1}g(T^*,I^*) \left(\frac{I^*L}{IL^*}-1 \right) -Kf(T^*,V^*)\left(\frac{V^*I}{VI^*}-1 \right)\\ +&\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} \left(f(T^*,V^*)\ln{\frac{f(T(t-\tau_1),V(t-\tau_1))}{f(T,V)}} +g(T^*,I^*)\ln{\frac{g(T(t-\tau_1),I(t-\tau_1))}{g(T,I)}}\right)\\ +&(1-\xi) e^{-m_2\tau_2} \left(f(T^*,V^*)\ln{\frac{f(T(t-\tau_2),V(t-\tau_2))}{f(T,V)}} +g(T^*,I^*)\ln{\frac{g(T(t-\tau_2),I(t-\tau_2))}{g(T,I)}}\right).\\ \end{split} \end{equation*}

    Moreover, for simplification, let

    \begin{equation*} \begin{split} &f: = f(T,V),f^*: = f(T^*,V^*),f_{\tau_1}: = f(T_{\tau_1},V_{\tau_1}),f_{\tau_2}: = f(T_{\tau_2},V_{\tau_2}),\\ &g: = g(T,I),g^*: = g(T^*,I^*),g_{\tau_1}: = g(T_{\tau_1},I_{\tau_1}),g_{\tau_2}: = g(T_{\tau_2},I_{\tau_2}). \end{split} \end{equation*}

    Then, \frac{{\rm d}\mathcal{V}_1}{{\rm d}t} is converted to

    \begin{equation*} \begin{split} &\frac{{\rm d}\mathcal{V}_1}{{\rm d}t}\\ = &-K\frac{d}{T}(T-T^*)^2 -K(f^*+g^*)h(\frac{T^*}{T})\\ &+Kf^*\left(h(\frac{T^*f}{Tf^*})-h(\frac{V}{V^*})\right) +Kg^*\left(h(\frac{T^*g}{Tg^*})-h(\frac{I}{I^*})\right)\\ &-\frac{\alpha\xi}{\alpha+\delta_L} e^{-m_1\tau_1} \left(f^*h(\frac{f_{\tau_1}L^*}{f^*L}) +g^*h(\frac{g_{\tau_1}L^*}{g^*L}) +f^*h(\frac{I^*L}{IL^*}) +g^*h(\frac{I^*L}{IL^*}) \right)\\ &-(1-\xi) e^{-m_2\tau_2} \left(f^*h(\frac{f_{\tau_2}I^*}{f^*I}) +g^*h(\frac{g_{\tau_2}I^*}{g^*I})\right) -Kf^*h(\frac{V^*I}{VI^*}). \end{split} \end{equation*}

    From Assumption (A1), \frac{fT^*}{f^*T} lies between 1 and \frac{V}{V^*} , \frac{gT^*}{g^*T} lies between 1 and \frac{I}{I^*} . Then, there holds that h(\frac{fT^*}{f^*T})-h(\frac{V}{V^*})\leq0 and h(\frac{gT^*}{g^*T})-h(\frac{I}{I^*})\leq0 . Thus, when \mathfrak{R}_0 > 1 and Assumption (A1) holds, \frac{{\rm d}\mathcal{V}_1}{{\rm d}t}\leq0 . Let \Theta_1 = \left\{(T, L, I, V):\frac{{\rm d}\mathcal{V}_1}{{\rm d}t} = 0\right\} and the largest invariant subset of \Theta_1 just consists of E^* . Thus, due to the LaSalle's invariance principle [15], E^* is a global attractor. Further, combining the local asymptotical stability of E^* under condition \mathfrak{R}_0 > 1 , E^* is globally asymptotically stable.

    In this section, we perform some specific examples to support our main results, verifying the effect of \mathfrak{R}_0 on system (1.2).

    Example 6.1. In [38], X. Wang et al. considered model (1.2) with the virus-to-cell transmission and the cell-to-cell transmission functions as the mass-action infection rates, that is, f(T, V) = \beta TV and g(T, I) = kTI for \beta, k > 0 . Then system (1.2) becomes

    \begin{equation} \left\{\begin{array}{ll} \frac{{\rm d} T}{{\rm d} t} = \Lambda-dT(t) -\beta T(t)V(t))-kT(t)I(t),\\ \frac{{\rm d} L}{{\rm d} t} = \xi e^{-m_1\tau_1}\beta T(t-\tau_1)V(t-\tau_1) +\xi e^{-m_1\tau_1}kT(t-\tau_1),I(t-\tau_1) -(\alpha+\delta_{L})L(t), \\ \frac{{\rm d} I}{{\rm d} t} = (1-\xi)e^{-m_2\tau_2}\beta T(t-\tau_2),V(t-\tau_2) +(1-\xi)e^{-m_2\tau_2}kT(t-\tau_2),I(t-\tau_2) -\delta_{I}I(t)+\alpha L, \\ \frac{{\rm d} V}{{\rm d} t} = N\delta_{I}I(t)-cV(t). \\ \end{array}\right. \end{equation} (6.1)

    Based on the above analysis, we obtain

    \mathfrak{R}_0 = \left[\frac{\alpha\xi}{\alpha+\delta_L}e^{-m_1\tau_1} +(1-\xi)e^{-m_2\tau_2}\right] \left(\frac{N\beta\Lambda}{cd} +\frac{k\Lambda}{\delta_Id}\right).

    If \mathfrak{R}_0 < 1 , then model (6.1) only has an infection-free equilibrium E_0 , which is globally asymptotically stable. If \mathfrak{R}_0 > 1 , then there exist E_0 and an globally asymptotically stable infected equilibrium E^* . The results are the same as those in [38], in which X. Wang et al. analyzed the global dynamics of the HIV latent infection model (6.1).

    Example 6.2. Let f(T, V) = \frac{\beta_1T(t)V(t)}{1+a V(t)} and g(T, I) = \beta_2T(t)I(t) , for \beta_1, \beta_2, a > 0 . We consider the following system

    \begin{equation} \left\{\begin{array}{ll} \frac{{\rm d} T}{{\rm d} t} = \Lambda-dT(t)-\frac{\beta_1T(t)V(t)} {1+aV(t)}-\beta_2T(t)I(t),\\ \frac{{\rm d} L}{{\rm d} t} = \xi e^{-m_1\tau_1}\frac{\beta_1 T(t-\tau_1)V(t-\tau_1)} {1+aV(t-\tau_1)} +\xi e^{-m_1\tau_1}\beta_2 T(t-\tau_1),I(t-\tau_1) -(\alpha+\delta_{L})L(t), \\ \frac{{\rm d} I}{{\rm d} t} = (1-\xi)e^{-m_2\tau_2}\frac{\beta_1 T(t-\tau_2)V(t-\tau_2)} {1+aV(t-\tau_2)} +(1-\xi)e^{-m_2\tau_2}\beta_2T(t-\tau_2)I(t-\tau_2) -\delta_I I(t)+\alpha L(t), \\ \frac{{\rm d} V}{{\rm d} t} = N\delta_{I}I(t)-cV(t). \\ \end{array}\right. \end{equation} (6.2)

    Through our analysis, for system (6.2),

    \mathfrak{R}_0 = \left[\frac{\alpha\xi}{\alpha+\delta_L}e^{-m_1\tau_1} +(1-\xi)e^{-m_2\tau_2}\right] \left(\frac{N\beta_1\Lambda}{cd} +\frac{\beta_2\Lambda}{\delta_Id}\right).

    The model (6.2) presents threshold dynamics with respect to \mathfrak{R}_0 as shown in Theorem 4.2 and Theorem 5.3.

    Note that different values of \tau_1 and \tau_2 can evaluate the effect of time delays on the virus dynamics. Through our analysis, \mathfrak{R}_0 is decreasing with respect to \tau_1 and \tau_2 . Thus, extending the time delay with the aid of drug treatment can help to inhibit viruses. We take numerical simulations to show the effect of time delay \tau_i, \; i = 1, 2 . For system (6.2) with parameters listed in Table 1, the graph trajectories with respect to different values of \tau_1 = \tau_2 = \tau are depicted in Figure 1.

    Table 1.  The parameters of system (6.2).
    Parameter Description Value Source
    \Lambda Recruitment rate of target cells 10000\; cells\cdot ml^{-1}\cdot day^{-1} [1]
    d Death rate of uninfected cells 0.01\; day^{-1} [22]
    \beta_1 Virus-to-cell infection rate 2.4\times10^{-8}\; ml\cdot day^{-1} [27]
    \beta_2 Cell-to-cell infection rate 1\times10^{-6}\; ml\cdot day^{-1} [25]
    \alpha Activation rate of infected cells 0.01\; day^{-1} [28]
    \delta_I Death rate of activated infected cells 0.05\; day^{-1} [1]
    \delta_L Death rate of latently infected cells 0.004\; day^{-1} [1]
    \xi Fraction of latency 0.001 [29]
    N Virus production rate of per infected cell 2000\; per\; cell\; per\; day [27]
    c Death rate of virus 23\; day^{-1} [26]
    m_{1} Death rate of latently infected cells which have not produced viruses 1\; day^{-1} Assume
    m_2 Death rate of infected cells which have not produced viruses 1\; day^{-1} Assume
    a Inhibition taken by virus 0.01\; cells\cdot ml^{-1} Assume

     | Show Table
    DownLoad: CSV
    Figure 1.  Graph trajectories of system (6.2) with respect to different values of \tau_i = \tau, \; i = 1, 2 . The time delay \tau is extended from 0.5\; day (\mathfrak{R}_0 = 8.1268) to 1.75\; days (\mathfrak{R}_0 = 0.6690 ).

    Through \mathfrak{R}_0 , different values of \xi also affect the virus dynamics. Thus, changing the the proportion of latent infection through the aid of drug treatment can help to reduce viruses. We take numerical simulations to show the effect of \xi . For system (6.2) with parameters \tau_1 = \tau_2 = 1.75\; days . The graph trajectories of system (6.2) with respect to different values of \xi are depicted in Figure 2.

    Figure 2.  Graph trajectories of system (6.2) with respect to different values of \xi . The parameter \xi is decreased from 0.05 (\mathfrak{R}_0 = 1.4686) to 0.025 (\mathfrak{R}_0 = 0.7188) .

    In this paper, the global dynamic analysis of a general within-host latent viral infection model with intracellular delays was carried out. By introducing general transmission functions f(T, V) and g(T, I) , which were assumed to satisfy several reasonable biological assumptions, we considered two virus predominant infection modes. The theoretical analysis showed that model (1.2) possesses two equilibria, relying on the basic reproductive number \mathfrak{R}_0 , which consists of two parts: One is the contribution from the virus-to-cell infection and the other is the contribution from the cell-to-cell transmission. Afterwards, through local and global analysis, we verified that the model exhibits threshold dynamics with respect to \mathfrak{R}_0 . When \mathfrak{R}_0 is less than unity, there exists a unique globally asymptotically stable infection-free equilibrium E_0 . In this situation, the disease with virus is inhibited effectively. When \mathfrak{R}_0 is greater than one, E_0 is unstable and there exists a globally asymptotically stable infected equilibrium E^* . For the critical case when \mathfrak{R}_0 equals to one, E_0 is a linearly neutrally stable non-hyperbolic equilibrium. Center manifold theory should be applied to further explore the stability and bifurcation around E_0 when \mathfrak{R}_0 = 1 .

    In our model (1.2), we extend the existing research on HIV transmission process and take into account generalized incidence rates. This not only enables us to establish a unified theoretical framework that can be applied to numerous situations, but also provides deeper insight into the relationship among different transmission dynamics. By considering double virus spread routes, we found that the infection rate tends to increase with higher transmission rates. We also illustrated that the infection can be inhibited through proper drug block, while a proper treatment still needs to be further investigated. According to the results of theoretical analysis and numerical simulations, cellular time lags and latent infection cells have significant effect on the global asymptotical stability. By reducing the proportion of latent infection or extending the intracellular delay, it is possible to inhibit the secondary infection produced by each primary case and block the transmission of HIV.

    Based on our generalized HIV transmission model, further research still need to be conducted. Despite the efforts to explore a general model, there exist several assumptions and some nonlinear incidence rates are not incorporated, such as Beddington-DeAngelis function. Hence, the results could be generalized by adopting more analytical techniques. Spatial diffusion and age structure are also useful to describe viral transmission, which are neglected in model (1.2). Based on the pattern and age structures, more effective strategies can be taken to inhibit the propagation of the virus. Combining the delay effect and PDE structure, it would be more challenging to explore the model, including the well-posedness of solutions, stability and bifurcation analysis, the existence of travelling wave solution. This is a significant work that we will take effort on in the future.

    The research is supported by North China University of Technology Research Fund Program for Young Scholars (No. 110051360002) and Fundamental Research Funds of Beijing Municipal Education Commission (No. 110052972027/141). The author is grateful to the editors and the anonymous referees for helpful comments and suggestions which lead to an improvement of the manuscript.

    The author declares that there are no conflicts of interest regarding the publication of this paper.



    [1] A. Alshorman, X. Wang, M. J. Meyer, L. Rong, Analysis of HIV models with two time delays, J. Biol. Dyn., 11 (2017), 40-64. doi: 10.1080/17513758.2016.1148202. doi: 10.1080/17513758.2016.1148202
    [2] S. Bonhoeffer, R. M. May, G. M. Shaw, M. A. Nowak, Virus dynamics and drug therapy, Proc. Natl. Acad. Sci. USA, 94 (1997), 6971-6976. doi: 10.1073/pnas.94.13.6971. doi: 10.1073/pnas.94.13.6971
    [3] C. Castillo-Chavez, B. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng., 1 (2004), 361-404. doi: 10.3934/mbe.2004.1.361. doi: 10.3934/mbe.2004.1.361
    [4] P. Chen, W. Hübner, M. A. Spinelli, 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), 12582-12595. doi: 10.1128/JVI.00381-07. doi: 10.1128/JVI.00381-07
    [5] T. W. Chun, L. Stuyver, S. B. Mizell, L. A. Ehler, J. A. M. Mican, M. Baseler, et al., Presence of an inducible HIV-1 latent reservoir during highly active antiretroviral therapy, Proc. Natl. Acad. Sci., 94 (1997), 13193-13197. doi: 10.1073/pnas.94.24.13193. doi: 10.1073/pnas.94.24.13193
    [6] S. M. Ciupe, Modeling the dynamics of hepatitis B infection, immunity, and drug therapy, Immunol. Rev., 285 (2018), 38-54. doi: 10.1111/imr.12686. doi: 10.1111/imr.12686
    [7] D. S. Dimitrov, R. L. Willey, H. Sato, L. Chang, R. Blumenthal, M. A. Martin, Quantitation of human immunodeficiency virus type 1 infection kinetics, J. Virol., 67 (1993), 2182-2190. doi: 10.1128/JVI.67.4.2182-2190.1993. doi: 10.1128/JVI.67.4.2182-2190.1993
    [8] S. Gummuluru, C. M. Kinsey, M. Emerman, An in vitro rapid-turnover assay for human immunodeficiency virus type 1 replication selects for cell-to-cell spread of virua, J. Virol., 74 (2000), 10882-10891. doi: 10.1128/jvi.74.23.10882-10891.2000. doi: 10.1128/jvi.74.23.10882-10891.2000
    [9] J. K. Hale, S. M. V. Lunel, Introduction to functional differential equations, New York: Springer-Verlag, 1993. doi: 10.1007/978-1-4612-4342-7.
    [10] A. V. Herz, S. Bonhoeffer, R. M. Anderson, R. M. May, M. A. Nowak, Viral dynamics in vivo: Limitations on estimations on intracellular delay and virus delay, Proc. Natl. Acad. Sci. USA, 93 (1996), 7247-7251. doi: 10.1073/pnas.93.14.7247. doi: 10.1073/pnas.93.14.7247
    [11] W. Hübner, G. P. McNerney, P. Chen, B. M. Dale, R. E. Gordon, F. Y. S. Chuang, et al., Quantitative 3D video microscopy of HIV transfer across T cell virological synapses, Science, 323 (2009), 1743-1747. doi: 10.1126/science.1167525. doi: 10.1126/science.1167525
    [12] D. E. Kirschner, G. F. Webb, A mathematical model of combined drug therapy of HIV infection, Comput. Math. Methods Med., 1 (1997), 293715. doi: 10.1080/10273669708833004. doi: 10.1080/10273669708833004
    [13] A. Korobeinikov, Global properties of basic virus dynamics models, Bull. Math. Biol., 66 (2004), 876-883. doi: 10.1016/j.bulm.2004.02.001. doi: 10.1016/j.bulm.2004.02.001
    [14] T. Kuniya, Hopf bifurcation in an age-structured SIR epidemic model, Appl. Math. Lett., 92 (2019), 22-28. doi: 10.1016/j.aml.2018.12.010. doi: 10.1016/j.aml.2018.12.010
    [15] Y. Kuang, Delay differential equations with applications in population dynamics, New York: Academic Press, 1993.
    [16] X. Lai, X. Zou, Modelling HIV-1 virus dynamics with both virus-to-cell infection and cell-to-cell transmission, SIAM J. Appl. Math., 74 (2014), 898-917. doi: 10.1137/130930145. doi: 10.1137/130930145
    [17] K. Lassen, Y. Han, Y. Zhou, J. Siliciano, R. F. Siliciano, The multifactorial nature of HIV-1 latency, Trends Mol. Med., 11 (2004), 525-531. doi: 10.1016/j.molmed.2004.09.006. doi: 10.1016/j.molmed.2004.09.006
    [18] J. Lin, R. Xu, X. Tian, Threshold dynamics of an HIV-1 virus model with both virus-to-cell and cell-to-cell transmissions, intracellular delay, and humoral immunity, Appl. Math. Comput., 315 (2017), 516-530. doi: 10.1016/j.amc.2017.08.004. doi: 10.1016/j.amc.2017.08.004
    [19] C. C. McCluskey, Using Lyapunov functions to construct Lyapunov functionals for delay differential equations, SIAM J. Appl. Dyn. Syst., 14 (2015), 1-24. doi: 10.1137/140971683. doi: 10.1137/140971683
    [20] C. C. McCluskey, Y. Yang, Global stability of a diffusive virus dynamics model with general incidence function and time delay, Nonlinear Anal.: Real World Appl., 25 (2015), 64-78. doi: 10.1016/j.nonrwa.2015.03.002. doi: 10.1016/j.nonrwa.2015.03.002
    [21] F. E. McKenzie, W. H. Bossert, A target for intervention in plasmodium falciparum infections, Am. J. Trop. Med. Hyg., 58 (1998), 763-767. doi: 10.4269/ajtmh.1998.58.763. doi: 10.4269/ajtmh.1998.58.763
    [22] A. S. Perelson, D. E. Kirschner, R. J. De Boer, Dynamics of HIV infection of CD4+ T cells, Math. Biosci., 114 (1993), 81-125. doi: 10.1016/0025-5564(93)90043-A. doi: 10.1016/0025-5564(93)90043-A
    [23] A. S. Perelson, A. U. Neumann, M. Markowitz, J. M. Leonard, D. D. Ho, HIV-1 dynamics in vivo: Virion clearance rate, infected cell life-span, and viral generation time, Science, 271 (1996), 1582-1586. doi: 10.1126/science.271.5255.1582. doi: 10.1126/science.271.5255.1582
    [24] D. M. Philips, The role of cell-to-cell transmission in HIV infection, AIDS, 8 (1994), 719-731. doi: 10.1097/00002030-199406000-00001. doi: 10.1097/00002030-199406000-00001
    [25] H. Pourbashash, S. S. Pilyugin, P. D. Leenheer, C. McCluskey, Global analysis of within host virus models with cell-to-cell viral transmission, Discrete Contin. Dyn. Syst. Ser. B, 19 (2014), 3341-3357. doi: 10.3934/dcdsb.2014.19.3341. doi: 10.3934/dcdsb.2014.19.3341
    [26] B. Ramratnam, S. Bonhoeffer, J. Binley, A. Hurley, L. Zhang, J. E. Mittler, et al., Rapid production and clearance of HIV-1 and hepatitis C virus assessed by large volume plasma apheresis, Lancet, 354 (1999), 1782-1785. doi: 10.1016/S0140-6736(99)02035-8. doi: 10.1016/S0140-6736(99)02035-8
    [27] L. Rong, Z. Feng, A. S. Perelson, Emergence of HIV-1 drug resistance during antiretroviral treatment, Bull. Math. Biol., 69 (2007), 2027-2060. doi: 10.1007/s11538-007-9203-3. doi: 10.1007/s11538-007-9203-3
    [28] L. Rong, A. S. Perelson, Modeling HIV persistence, the latent reservoir, and viral blips, J. Theoret. Biol., 260 (2009), 308-331. doi: 10.1016/j.jtbi.2009.06.011. doi: 10.1016/j.jtbi.2009.06.011
    [29] L. Rong, A. S. Perelson, Modeling Latently infected cell activation: Viral and latent reservoir persistence, and viral blips in HIV-infected patients on potent therapy, PLoS Comput. Biol., 5 (2009), e1000533. doi: 10.1371/journal.pcbi.1000533. doi: 10.1371/journal.pcbi.1000533
    [30] H. Sato, J. Orenstein, D. S. Dimitrov, M. A. Martin, Cell-to-cell spread of HIV-1 occurs with minutes and may not involve the participation of virus particles, Virology, 186 (1992), 712-724. doi: 10.1016/0042-6822(92)90038-q. doi: 10.1016/0042-6822(92)90038-q
    [31] Q. Sattentau, The direct passage of animal viruses between cells, Curr. Opin. Virol., 1 (2011), 396-402. doi: 10.1016/j.coviro.2011.09.004. doi: 10.1016/j.coviro.2011.09.004
    [32] H. Shu, Y. Chen, L. Wang, Impacts of the cell-free and cell-to-cell infection modes on viral dynamics, J. Dyn. Diff. Equat., 30 (2018), 1817-1836. doi: 10.1007/s10884-017-9622-2. doi: 10.1007/s10884-017-9622-2
    [33] A. Sigal, J. T. Kim, A. B. Balazs, E. Dekel, A. Mayo, R. Milo, et al., Cell-to-cell spread of HIV permits ongoing replication despite antiretroviral therapy, Nature, 477 (2011), 95-98. doi: 10.1038/nature10347. doi: 10.1038/nature10347
    [34] H. Sun, J. Wang, Dynamics of a diffusive virus model with general incidence function, cell-to-cell transmission and time delay, Comput. Math. Appl., 77 (2019), 284-301. doi: 10.1016/j.camwa.2018.09.032. doi: 10.1016/j.camwa.2018.09.032
    [35] S. Tang, Z. Teng, H. Miao, Global dynamics of a reaction-diffusion virus infection model with humoral immunity and nonlinear incidence, Comput. Math. Appl., 78 (2019), 786-806. doi: 10.1016/j.camwa.2019.03.004. doi: 10.1016/j.camwa.2019.03.004
    [36] K. Wang, W. Wang, Propagation of HBV with spatial dependence, Math. Biosci., 210 (2007), 78-95. doi: 10.1016/j.mbs.2007.05.004. doi: 10.1016/j.mbs.2007.05.004
    [37] W. Wang, X. Wang, K. Guo, W. Ma, Global analysis of a diffusive viral model with cell-to-cell infection and incubation period, Math. Meth. Appl. Sci., 43 (2020), 5963-5978. doi: 10.1002/mma.6339. doi: 10.1002/mma.6339
    [38] X. Wang, S. Tang, X. Song, L. Rong, Mathematical analysisi of an HIV latent infection model including both virus-to-cell infection and cell-to-cell transmission, J. Biol. Dyn., 11 (2017), 455-483. doi: 10.1080/17513758.2016.1242784. doi: 10.1080/17513758.2016.1242784
    [39] J. K. Wong, M. Hezareh, H. F. Günthard, D. V. Havlir, C. C. Ignacio, C. A. Spina, et al., Recovery of replication-competent HIV despite prolonged suppression of plasma viremia, Science, 278 (1997), 1291-1295. doi: 10.1126/science.278.5341.1291. doi: 10.1126/science.278.5341.1291
    [40] 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. doi: 10.1016/j.mbs.2015.05.001. doi: 10.1016/j.mbs.2015.05.001
    [41] X. Zhang, Z. Liu, Bifurcation analysis of an age structured HIV infection model with both virus-to-cell and cell-to-cell transmissions, Int. J. Bifurcation and Chaos, 28 (2018), 1850109. doi: 10.1142/S0218127418501092. doi: 10.1142/S0218127418501092
    [42] Fact sheet-Latest statistics on the status of the AIDS epidemic, UNAIDS, 2021. Available from: https://www.unaids.org/sites/default/files/media_asset/UNAIDS_FactSheet_en.pdf.
    [43] FDA approval of HIV medicines, AIDSinfo, 2021. Available from: https://aidsinfo.nih.gov/understanding-hiv-aids/infographics/25/fda-approval-of-hiv-medicines.
  • This article has been cited by:

    1. Prakash Raj Murugadoss, Venkatesh Ambalarajan, Vinoth Sivakumar, Prasantha Bharathi Dhandapani, Dumitru Baleanu, Analysis of Dengue Transmission Dynamic Model by Stability and Hopf Bifurcation with Two-Time Delays, 2023, 28, 2768-6701, 10.31083/j.fbl2806117
  • Reader Comments
  • © 2022 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(1951) PDF downloads(81) Cited by(1)

Figures and Tables

Figures(2)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog