Research article

Stability analysis of general delayed HTLV-I dynamics model with mitosis and CTL immunity

  • Received: 08 June 2022 Revised: 14 July 2022 Accepted: 01 August 2022 Published: 31 August 2022
  • This paper formulates and analyzes a general delayed mathematical model which describe the within-host dynamics of Human T-cell lymphotropic virus class I (HTLV-I) under the effect Cytotoxic T Lymphocyte (CTL) immunity. The models consist of four components: uninfected CD4+T cells, latently infected cells, actively infected cells and CTLs. The mitotic division of actively infected cells are modeled. We consider general nonlinear functions for the generation, proliferation and clearance rates for all types of cells. The incidence rate of infection is also modeled by a general nonlinear function. These general functions are assumed to be satisfy some suitable conditions. To account for series of events in the infection process and activation of latently infected cells, we introduce two intracellular distributed-time delays into the models: (ⅰ) delay in the formation of latently infected cells, (ⅱ) delay in the activation of latently infected cells. We determine a bounded domain for the system's solutions. We calculate two threshold numbers, the basic reproductive number R0 and the CTL immunity stimulation number R1. We determine the conditions for the existence and global stability of the equilibrium points. We study the global stability of all equilibrium points using Lyapunov method. We prove the following: (a) if R01, then the infection-free equilibrium point is globally asymptotically stable (GAS), (b) if R11<R0, then the infected equilibrium point without CTL immunity is GAS, (c) if R1>1, then the infected equilibrium point with CTL immunity is GAS. We present numerical simulations for the system by choosing special shapes of the general functions. The effects of proliferation of CTLs and time delay on the HTLV-I progression is investigated. We noted that the CTL immunity does not play the role in clearing the HTLV-I from the body, but it has an important role in controlling and suppressing the viral infection. On the other hand, we observed that, increasing the time delay intervals can have similar influences as drug therapies in removing viruses from the body. This gives some impression to develop two types of treatments, the first type aims to extend the intracellular delay periods, while the second type aims to activate and stimulate the CTL immune response.

    Citation: A. M. Elaiw, A. S. Shflot, A. D. Hobiny. Stability analysis of general delayed HTLV-I dynamics model with mitosis and CTL immunity[J]. Mathematical Biosciences and Engineering, 2022, 19(12): 12693-12729. doi: 10.3934/mbe.2022593

    Related Papers:

    [1] Xuejuan Lu, Lulu Hui, Shengqiang Liu, Jia Li . A mathematical model of HTLV-I infection with two time delays. Mathematical Biosciences and Engineering, 2015, 12(3): 431-449. doi: 10.3934/mbe.2015.12.431
    [2] A. M. Elaiw, N. H. AlShamrani . Analysis of an HTLV/HIV dual infection model with diffusion. Mathematical Biosciences and Engineering, 2021, 18(6): 9430-9473. doi: 10.3934/mbe.2021464
    [3] A. M. Elaiw, N. H. AlShamrani . Stability of HTLV/HIV dual infection model with mitosis and latency. Mathematical Biosciences and Engineering, 2021, 18(2): 1077-1120. doi: 10.3934/mbe.2021059
    [4] 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
    [5] 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
    [6] Pensiri Yosyingyong, Ratchada Viriyapong . Global dynamics of multiple delays within-host model for a hepatitis B virus infection of hepatocytes with immune response and drug therapy. Mathematical Biosciences and Engineering, 2023, 20(4): 7349-7386. doi: 10.3934/mbe.2023319
    [7] 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
    [8] Yan Wang, Tingting Zhao, Jun Liu . Viral dynamics of an HIV stochastic model with cell-to-cell infection, CTL immune response and distributed delays. Mathematical Biosciences and Engineering, 2019, 16(6): 7126-7154. doi: 10.3934/mbe.2019358
    [9] Ning Bai, Rui Xu . Mathematical analysis of an HIV model with latent reservoir, delayed CTL immune response and immune impairment. Mathematical Biosciences and Engineering, 2021, 18(2): 1689-1707. doi: 10.3934/mbe.2021087
    [10] Abdessamad Tridane, Yang Kuang . Modeling the interaction of cytotoxic T lymphocytes and influenza virus infected epithelial cells. Mathematical Biosciences and Engineering, 2010, 7(1): 171-185. doi: 10.3934/mbe.2010.7.171
  • This paper formulates and analyzes a general delayed mathematical model which describe the within-host dynamics of Human T-cell lymphotropic virus class I (HTLV-I) under the effect Cytotoxic T Lymphocyte (CTL) immunity. The models consist of four components: uninfected CD4+T cells, latently infected cells, actively infected cells and CTLs. The mitotic division of actively infected cells are modeled. We consider general nonlinear functions for the generation, proliferation and clearance rates for all types of cells. The incidence rate of infection is also modeled by a general nonlinear function. These general functions are assumed to be satisfy some suitable conditions. To account for series of events in the infection process and activation of latently infected cells, we introduce two intracellular distributed-time delays into the models: (ⅰ) delay in the formation of latently infected cells, (ⅱ) delay in the activation of latently infected cells. We determine a bounded domain for the system's solutions. We calculate two threshold numbers, the basic reproductive number R0 and the CTL immunity stimulation number R1. We determine the conditions for the existence and global stability of the equilibrium points. We study the global stability of all equilibrium points using Lyapunov method. We prove the following: (a) if R01, then the infection-free equilibrium point is globally asymptotically stable (GAS), (b) if R11<R0, then the infected equilibrium point without CTL immunity is GAS, (c) if R1>1, then the infected equilibrium point with CTL immunity is GAS. We present numerical simulations for the system by choosing special shapes of the general functions. The effects of proliferation of CTLs and time delay on the HTLV-I progression is investigated. We noted that the CTL immunity does not play the role in clearing the HTLV-I from the body, but it has an important role in controlling and suppressing the viral infection. On the other hand, we observed that, increasing the time delay intervals can have similar influences as drug therapies in removing viruses from the body. This gives some impression to develop two types of treatments, the first type aims to extend the intracellular delay periods, while the second type aims to activate and stimulate the CTL immune response.



    Human T-cell lymphotropic virus class I (HTLV-I) is one of the most dangerous viruses that infect the human body and causes inflammatory and malignant diseases. It was estimated that about 10 to 25 million humans infected with HTLV-I worldwide [1]. HTLV-I can be transferred via bodily fluids including blood and semen. Further, HTLV-I can be transferred from mother to child via breast-feeding [2]. The high transmission of HTLV-I was reported in Caribbean Islands, Central Africa, Northeast South America and Southwestern Japan [3]. HTLV-I can progress to serious diseases, adult T-cell leukemia (ATL) and HTLV-I-associated myelopathy/tropical spastic paraparesis (HAM/TSP) [4]. HTLV-I is a single-stranded RNA virus which infects the most important component of the immune system, CD4+T cells [5]. There are two modes of within-host transmission, vertical or mitotic transmission of actively HTLV-I infected cells and horizontal by infectious transmission through direct cell-to-cell touch [6].

    Mathematical modeling of HTLV-I was established since 1990 to understand the pathogenesis of the HTLV-I infection inside the host cell that has a critical ramifications for the movement of remedial measures and for the acknowledgment of hazard elements of the advancement of HAM/TSP. A lot of mathematical models were planned by numerous scientists to epitomize the intuitive dynamics within-host among HTLV-I infection to depict the pathogenesis of HTLV-I related infection [7,8,9,10,11,12].

    Stilianakis and Seydel [7] constructed a four-dimensional system of ODEs which describes the within-host HTLV-I dynamics as:

    ˙U(t)=νˉνU(t)βU(t)A(t), (1.1)
    ˙L(t)=βU(t)A(t)(π+μ)L(t), (1.2)
    ˙A(t)=πL(t)(δ+ϰ)A(t), (1.3)
    ˙Q(t)=ϰA(t)+λQ(t)(1Q(t)Qmax)Q(t), (1.4)

    where U, L, A and Q (cells mm3) are the concentrations of healthy or uninfected CD4+T cells, latently CD4+T infected cells, actively CD4+T infected cells and leukemia cells (ATL cells), respectively. The variable t denotes to the time (days). Uninfected cells are created at a constant rate ν. The bilinear incidence term βUA accounts for the rate of infectious via horizontal transmission appears by cell-to-cell touch between uninfected and actively infected cells. πL is the activation rate of latently infected cells, and ϰA is the conversion rate of actively infected cells to ATL cells. Qmax is the maximal growth of ATL cells and λ is the maximum proliferation rate of ATL cells. The death rates of the uninfected cells, latently infected cells, actively infected cells and ATL cells are indicated by ˉνU, μL, δA and Q, respectively.

    As mentioned before HTLV-I has two modes of transmission, horizontal and vertical. Actively infected cells propagate faster than uninfected and latently infected cells. This causes an increasing in the proviral load. As a consequence, mitotic transmission plays an instrumental role in the persistence of HTLV-I infection [3]. Li and Lim [10] developed an HTLV-I infection model by considering both horizontal and vertical transmissions as:

    ˙U(t)=νˉνU(t)βU(t)A(t), (1.5)
    ˙L(t)=ϵβU(t)A(t)+ωεA(t)(1U(t)+L(t)Umax)(π+μ)L(t), (1.6)
    ˙A(t)=πL(t)δA(t), (1.7)

    where ϵ(0,1) is the probability of new HTLV-I infections via horizontal route become latently infected. The vertical transmission is represented in proliferating the actively infected cells at rate εA(1U+LUmax), where Umax is the CD4+T cells carrying capacity. The actively infected cells become latent at rate ωεA(1U+LUmax), where ω(0,1). It was shown in [13] that U+L<Umax. Therefore, in [14], the logistic term εA(1U+LUmax) was replaced by an exponential growth term εA.

    Cytotoxic T lymphocyte (CTL) is identified as the tremendous element of human immunity in killing viral infections. It represses viral reproduction and kills the cells which can be tainted by infections. It was mentioned in [13,15] that the CTLs assume a strong part in controlling the HTLV-I related diseases. CTLs can realize and kill the actively infected CD4+T cells, further, they able to lessen the proviral load. The effect of CTL immunity was included in several models for HTLV-I infection (e.g., [4,16,17,18,19,20,21,22,23,24,25]).

    To incorporate both CTL immunity and mitosis into the HTLV-I dynamics model, Lim and Maini [14] proposed the following system:

    ˙U(t)=νˉνU(t)βU(t)A(t), (1.8)
    ˙L(t)=βU(t)A(t)+εA(t)(π+μ)L(t), (1.9)
    ˙A(t)=πL(t)δA(t)αA(t)C(t), (1.10)
    ˙C(t)=ρA(t)γC(t), (1.11)

    where C denotes the concentration of CTLs. The removal rate of actively infected by CTLs is represented by αAC, where α represents the CTL-mediated lysis constant [5]. The term ρA indicates the proliferation rate of CTLs. The death rate of the CTLs is denoted by γC. The authors studied the global stability of equilibrium points by utilizing Lyapunov approach. Models (1.8)–(1.11) was developed and modified in [26,27,28]. Recently, Khajanchi et al. [1] modified models (1.8)–(1.11) by replacing Eq (1.11) by

    ˙C(t)=ρA(t)C(t)γC(t).

    The authors investigated the global stability for chronic infection equilibrium point via two approaches Lyapunov and geometric.

    In the above models, it was assumed that the uninfected CD4+T cells become latently infected instantaneously. Moreover, the transition from latently infected to actively infected is also instantaneously. However, there exist delays in the infection process and activation of latently infected cells. Therefore, ignoring these time delays is biologically unrealistic. In addition, these time intracellular delays have an important effect on the stability of system [29,30]. HTLV-I infection models have been incorporated with intracellular delay [23,24,31,32,33] or immune response delay [16,20,21,34,35]. Both intracellular and immune response delays were considered in [25,29,36]. Katri and Ruan [31] modified models (1.1)–(1.4) by including a discrete-time intracellular delay τ1 which describes the time between the initial infection of uninfected CD4+T cells to become latently CD4+T infected cells. Equation (1.2) was replaced by

    ˙L(t)=βU(tτ1)A(tτ1)(π+μ)L(t).

    The delayed HTLV-I infection models presented in the literature included either discrete-time delay [16,20,21,24,25,32,33,34,35,36], or distributed-time delay [23]. All these models neglected the latently infected cells.

    In the literature of HTLV-I infection models, the intrinsic growth rate of uninfected cells was given by two main forms, linear form Φ(U)=νˉνU [7,11,23,24,25], and logistic function Φ(U)=νˉνU+ςU(1UUmax) [16], where ς>0 is the maximum propagation rate of uninfected cells and Umax>0 is maximal population level of uninfected cells.

    The incidence rate that models the touch between the uninfected cells and actively infected cells presented in models (1.1)–(1.4), (1.5)–(1.7) and (1.8)–(1.11) was given by bilinear incidence βUA. This bilinear form may not be completely characterize the incidence between uninfected cells and actively infected cells [37]. Several HTLV-I infection models were introduced in the literature by considering different incidence rate forms such as: (ⅰ) saturated incidence βUA1+υA [11], βUA1+ηU [27], standard incidence βUAU+A [38], Beddington-DeAngelis incidence βUA1+ηU+υA [29], Crowley Martin incidence βUA(1+ηU)(1+υA) [23], and nonlinear incidence in the form βUqAp [28]; βUΛ(A) [24] and [39]; Λ(U,A)A [26], where β,υ,η,p and q are positive constants and Λ is a general function satisfies some conditions.

    In models (1.1)–(1.4), (1.5)–(1.7) and (1.8)–(1.11) and most of the HTLV-I models introduced in the literature, the death rate of the cells is given by linear function of their concentration. However, the death rate of the cells is generally not known.

    Our aim of the present paper is to formulate and analyze a general delayed HTLV-I mathematical model under the following generalizations:

    G1: We assume that the intrinsic growth rate of the uninfected CD4+T cells is given by a general function Φ(U).

    G2: We consider a general incidence rate in the form Λ(U,A).

    G3: We assume that the death rates of the latently infected cells, actively infected cells and CTLs are given by general nonlinear functions μΨ1(L), δΨ2(A) and γΨ3(C), respectively.

    G4: We assume that the transition rate from latently infected phase to actively infected phase is given by a general nonlinear function in the form πΨ1(L).

    G5: We assume that actively infected cells proliferate at rate εΨ2(A), with (1ω)εΨ2(A), ω(0,1) staying in the actively infected cells compartment, while ωεΨ2(A) being latent and therefore escaping from the CTL immunity.

    G6: We assume that the replenishment rate of CTLs and the elimination of actively infected cells by CTLs are given by ρΨ2(A)Ψ3(C) and αΨ2(A)Ψ3(C), respectively.

    G7: We include distributed intracellular delay which describes the cell mutates from uninfected to latent.

    G8: We include distributed delay in the activation of latently infected.

    With the above generalizations G1–G8 we propose the following HTLV-I dynamics model with two types of distributed-time delays:

    ˙U(t)=Φ(U(t))Λ(U(t),A(t)), (2.1)
    ˙L(t)=κ10f1(ϖ)eξ1ϖΛ(U(tϖ),A(tϖ))dϖ+ωεΨ2(A(t))(π+μ)Ψ1(L(t)), (2.2)
    ˙A(t)=πκ20f2(ϖ)eξ2ϖΨ1(L(tϖ))dϖ+(1ω)εΨ2(A(t))δΨ2(A(t))αΨ2(A(t))Ψ3(C(t)), (2.3)
    ˙C(t)=ρΨ2(A(t))Ψ3(C(t))γΨ3(C(t)), (2.4)

    Here, ϖ is a random variable generated from probability distribution functions fi(ϖ), i=1,2 over the time interval [0,κi], i=1,2 where κi is the upper limit of the delay period. Here f1(ϖ)eξ1ϖ represents the probability of surviving the cells from time tϖ to time t. The factor f2(ϖ)eξ2ϖ represents the probability that latently infected cells is survived ϖ time units to become actively infected cells. Functions fi(ϖ)i=1,2 satisfy the following conditions:

    (a) fi(ϖ)>0,i=1,2,

    (b) κi0fi(ϖ)dϖ=1,i=1,2,

    (c) κi0eξiϖfi(ϖ)dϖ<, ξi>0, i=1,2.

    Here, Φ, Λ, Ψi,  i=1,2,3 are continuously differentiable functions, in addition they satisfy a set of conditions:

    Condition(C1):(a) there exists a U0>0 such that Φ(U0)=0 and Φ(U)>0 for U[0,U0).

    (b) Φ(U)<0 for all U>0.

    (c) there exists ν>0 and ˉν>0 such that Φ(U)νˉνU, for all U0.

    Condition(C2):(a) Λ(U,A)>0 and Λ(0,A)=Λ(U,0)=0, for all U>0,A>0.

    (b) Λ(U,A)U>0, Λ(U,A)A>0 and Λ(U,0)A>0, for all U>0, A>0.

    (c) ddU(Λ(U,0)A)>0, for all U>0.

    Condition(C3): (a) Ψi(s)>0 for all s>0, Ψi(0)=0, i=1,2,3.

    (b) Ψi(s)>0, for all s>0, i=1,3,  Ψ2(s)>0, for all s0.

    (c) there is αi>0, i=1,2,3 such that Ψi(s)αis, for all s0.

    Condition(C4): Λ(U,A)Ψ2(A) is decreasing w.r.t A, for all A>0.

    Let κ=max{κ1,κ2} and assume the initial conditions of systems (2.1)–(2.4) having the following form:

    U(θ)=ϕ1(θ),L(θ)=ϕ2(θ),A(θ)=ϕ3(θ),C(θ)=ϕ4(θ);ϕi(θ)0,θ[κ,0],i=1,2,3,4, (2.5)

    where ϕ C and C=C([κ,0],R4+) is the Banach space of continuous mapping the interval [κ,0] into R4+ with the norm ϕi=supκθ0|ϕi(θ)| for ϕiC, i=1,2,3,4, where R4+={(U,L,A,C):U0,L0,A0,C0}. We note that the systems (2.1)–(2.4) with initial conditions (2.5) by the fundamental theory of functional differential equations has a unique solution [40].

    The parameters presented in models (2.1)–(2.4) are positive. In [14,41,42], it was assumed that ε<min{ˉν,μ,δ}. Since ε<δ and 0<ω<1, so (1ω)ε<δ. Let's assume that δ=δ(1ω)ε>0 and ε=ωε. Hence, δε=δε>0. Then the systems (2.1)–(2.4) will take the following form:

    ˙U=Φ(U(t))Λ(U(t),A(t)), (2.6)
    ˙L=κ10f1(ϖ)eξ1ϖΛ(U(tϖ),A(tϖ))dϖ+εΨ2(A(t))(π+μ)Ψ1(L(t)), (2.7)
    ˙A=πκ20f2(ϖ)eξ2ϖΨ1(L(tϖ))dϖδΨ2(A(t))αΨ2(A(t))Ψ3(C(t)), (2.8)
    ˙C=ρΨ2(A(t))Ψ3(C(t))γΨ3(C(t)). (2.9)

    Let us define ϝi as:

    ϝi=κi0eξiϖfi(ϖ)dϖ,i=1,2.

    Clearly 0<ϝi1, i=1,2.

    Remark 1. If the time delay is not considered, ϖ=0, all proliferated actively infected by mitosis become latent, ω=1, the intrinsic growth rate of the uninfected CD4+T cells is given by linear function, Φ(U)=νˉνU, the infection rate is given by bilinear incidence, Λ(U,A)=βUA, the death rates of the compartments are given by linear functions Ψi(x)=x, i=1,2,3, and the CTL immune response is given by linear form, ρA, then systems (2.1)–(2.4) will lead to the ODEs systems (1.8)–(1.11). In addition, if the CTL immune response is given by a nonlinear form, ρAC, then systems (2.1)–(2.4) will reduce to the model presented in [1]. Thus, systems (2.1)–(2.4) can be considered a generalization of the systems presented in [1,14]. We mention that functions Φ(U), Λ(U,A) and Ψi, i=1,2,3, include and generalize many forms presented in the literature. The general distribution functions fi, i=1,2 considered in our work allows us to include several special forms of intracellular delays existing in the literature (see [24,31,32,33]). Finally, models (2.1)–(2.4) can be considered as a generalization of the model presented in [23] which neglected the latently infected cells.

    Proposition 1. Suppose that Conditions C1–C3 are valid. Then thesolutions of (2.6)–(2.9) are non-negative and ultimately bounded.

    Proof. We assume that systems (2.6)–(2.9) can be written in matrix form ˙X=Y(X(t)), where

    Y(X(t))=(Y1(X(t))Y2(X(t))Y3(X(t))Y4(X(t)))=(Φ(U(t))Λ(U(t),A(t))κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εΨ2(A(t))(π+μ)Ψ1(L(t))πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδΨ2(A(t))αΨ2(A(t))Ψ3(C(t))ρΨ2(A(t))Ψ3(C(t))γΨ3(C(t))).

    Clearly, Y(X(t)) satisfies the following: Yi(X(t))|Xi=0,X(t)R4+0,i=1,2,3,4. From Lemma 2 in [43] we get that the solutions of systems (2.6)–(2.9) with initial conditions (2.5) satisfies (U(t),L(t),A(t),C(t))0 for all t0. Hence, the R40 is positively invariant for systems (2.6)–(2.9). Now, we show that the solutions of the system are bounded. The non-negativity of the system's solution implies that lim suptU(t)νˉν=M1. We define

    Θ(t)=κ10eξ1ϖf1(ϖ)U(tϖ)dϖ+L(t)+A(t)+πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖ+αρC(t).

    Then

    ˙Θ(t)=κ10eξ1ϖf1(ϖ)˙U(tϖ)dϖ+˙L(t)+˙A(t)ξ2πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖ+πΨ1(L(t))κ20f2(ϖ)dϖπκ20f2(ϖ)eξ2ϖΨ1(L(tϖ))dϖ+αρ˙C(t)=κ10eξ1ϖf1(ϖ)[Φ(U(tϖ))Λ(U(tϖ),A(tϖ))]dϖ+κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εΨ2(A(t))(π+μ)Ψ1(L(t))+πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδΨ2(A(t))αΨ2(A(t))Ψ3(C(t))ξ2πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖ+πΨ1(L(t))πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖ+αρ(ρΨ2(A(t))Ψ3(C(t))γΨ3(C(t)))=κ10eξ1ϖf1(ϖ)Φ(U(tϖ))dϖμΨ1(L(t))(δε)Ψ2(A(t))αγρΨ3(C(t))ξ2πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖ.

    By using Conditions C1 and C3, we get

    ˙Θ(t)νκ10eξ1ϖf1(ϖ)dϖˉνκ10eξ1ϖf1(ϖ)U(tϖ)dϖμα1L(t)(δε)α2A(t)ξ2πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖαγρα3C(t).

    It follows that

    ˙Θ(t)νˉνκ10eξ1ϖf1(ϖ)U(tϖ)dϖμα1L(t)(δε)α2A(t)αγρα3C(t)ξ2πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖνσ(κ10eξ1ϖf1(ϖ)U(tϖ)dϖ+L(t)+A(t)+πκ20f2(ϖ)ttϖeξ2(tθ)Ψ1(L(θ))dθdϖ+αρC(t))=νσΘ(t),

    where σ=min{ˉν,μα1,(δε)α2,ξ2,γα3}. The non-negativity of the model's solutions imply that lim suptL(t)M1, lim suptA(t)M1 and lim suptC(t)M2, where M2=ραM1. Therefore, U(t),L(t),A(t) and C(t) are all ultimately bounded.

    According to Proposition 1, we can establish that the region Ω={(U,L,A,C)C40:UM1, LM1, AM1, CM2} is positively invariant for systems (2.6)–(2.9).

    Here, we look at the model's equilibrium points and deduce the conditions that allow them to exist.

    Lemma 1. Suppose that C1–C4 are valid, then there exist twopositive threshold numbers R0 and R1 with R0>R1 such that:

    (a) if R01, then there exists a single equilibrium point EP0=(U0,0,0,0).

    (b) if R11<R0, then there exist two equilibrium points EP0 and EP1=(U1,L1,A1,0).

    (c) if R1>1, then there exist three equilibrium points EP0, EP1 and EP2=(U2,L2,A2,C2).

    Proof. Any equilibrium point EP=(U,L,A,C) satisfies

    0=Φ(U)Λ(U,A), (2.10)
    0=ϝ1Λ(U,A)+εΨ2(A)(π+μ)Ψ1(L), (2.11)
    0=πϝ2Ψ1(L)δΨ2(A)αΨ2(A)Ψ3(C), (2.12)
    0=ρΨ2(A)Ψ3(C)γΨ3(C). (2.13)

    Equation (2.13) has two possibilities Ψ3(C)=0 and Ψ2(A)=γρ.

    If Ψ3(C)=0, then from Condition C3 we obtain C=0. From Condition C3 we have Ψ1i, i=1, 2 exist, moreover, these functions are strictly increasing. Define

    Ξ(U)=Ψ11(δϝ1π(δεϝ2)+δμΦ(U)), Π(U)=Ψ12(πϝ1ϝ2π(δεϝ2)+δμΦ(U)).

    Therefore, Eqs (2.11) and (2.12) imply that

    L=Ξ(U),A=Π(U). (2.14)

    Obviously, from Condition C1 we have Ξ(0),Π(0)>0 and Ξ(U0)=Π(U0)=0.

    From Eqs (2.10)–(2.12) we get

    ϱΛ(U,Π(U))Ψ2(Π(U))=0, (2.15)

    where ϱ=πϝ1ϝ2π(δεϝ2)+δμ.

    Equation (2.15) admits a solution U=U0 which yields L=A=C=0 and provides the infection-free equilibrium point EP0=(U0,0,0,0). Let

    Γ1(U)=ϱΛ(U,Π(U))Ψ2(Π(U)).

    It is clear from Conditions C1–C3 that,

    Γ1(0)=Ψ2(Π(0))<0.
    Γ1(U0)=ϱΛ(U0,0)Ψ2(0)=0.

    Moreover,

    Γ1(U)=ϱ(Λ(U,A)U+Π(U)Λ(U,A)A)Ψ2(Π(U))Π(U).
    Γ1(U0)=ϱ(Λ(U0,0)U+Π(U0)Λ(U0,0)A)Ψ2(0)Π(U0).

    We note from Condition C1 that, Λ(U0,0)U=0. Then

    Γ1(U0)=Ψ2(0)Π(U0)(ϱΨ2(0)Λ(U0,0)A1).

    Since Ψ2(Π(U))=ϱΦ(U) then Ψ2(Π(U))Π(U)=ϱΦ(U). Substituting by U=U0, we get Ψ2(0)Π(U0)=ϱΦ(U0). Therefore,

    Γ1(U0)=ϱΦ(U0)(ϱΨ2(0)Λ(U0,0)A1).

    Condition C1 implies that Φ(U0)<0, thus if ϱΨ2(0)Λ(U0,0)A>1, then Γ1(U0)<0 and there exists U1(0,U0) such that Γ1(U1)=0. It follows from Eq (2.14) that, L1=Ξ(U1)>0 and A1=Π(U1)>0. This shows that, an infected equilibrium point without CTL immunity EP1=(U1,L1,A1,0) exists when ϱΨ2(0)Λ(U0,0)A>1. We denote

    R0=ϱϱΨ2(0)Λ(U0,0)A,

    which defines the basic infection reproductive number and decides when the HTLV-I infection will be chronic. The second possibility of Eq (2.13) is Ψ3(C)0 and Ψ2(A2)=γρ. This yields that A2=Ψ12(γρ)>0. We insert the value of A2 in Eq (2.10) and let

    Γ2(U)=Φ(U)Λ(U,A2)=0.

    Based on Conditions C1 and C2, Γ2 is a strictly decreasing function of U. Clearly, Γ2(0)=Φ(0)>0 and Γ2(U0)=Λ(U0,A2)<0. Thus, there exists a unique U2(0,U0) such that Γ2(U2)=0. From Eq (2.11) we get

    Ψ1(L2)=1π+μ(ϝ1Λ(U2,A2)+εΨ2(A2))=1π+μ(ϝ1Λ(U2,A2)+εγρ)L2=Ψ11(1π+μ(ϝ1Λ(U2,A2)+εγρ))>0.

    From Eq (2.12) we get

    C2=Ψ13(π(δεϝ2)+δμα(π+μ)(ϱΛ(U2,A2)Ψ2(A2)1)).

    Clearly, C2>0 when ϱΛ(U2,A2)Ψ2(A2)>1. Let us denote

    R1=ϱΛ(U2,A2)Ψ2(A2),

    which represents the CTL immunity stimulation number and determines when CTL immunity can be activated. Therefore, C2 can be written as:

    C2=Ψ13(π(δεϝ2)+δμα(π+μ)(R11)).

    Thus, there exists an infected equilibrium point with CTL immunity EP2=(U2,L2,A2,C2) when R1>1.

    Conditions C2 and C4 imply that

    R1=ϱΛ(U2,A2)Ψ2(A2)ϱlimA0+Λ(U2,A)Ψ2(A)=ϱΨ2(0)Λ(U2,0)AϱΨ2(0)Λ(U0,0)A=R0.

    Stability analysis of equilibrium points is at the heart of dynamical system analysis. Stable solutions can only be observed experimentally. Thus, we demonstrate the global asymptotic stability of all equilibrium points in this section by establishing appropriate Lyapunov functions [44,45,46] and applying Lyapunov-LaSalle asymptotic stability theorem (L-LAST) [47,48,49]. We define the function H(x)0 for all x>0 as: H(x)=x1lnx, where H(x)=0 if and only if x=1. Hence,

    lnxx1.

    Denote (U,L,A,C)=(U(t),L(t),A(t),C(t)).

    We will use the follow equations:

    ln(Λ(U(tϖ),A(tϖ))Λ(U,A))=ln(Λ(Ui,Ai)Λ(U,Ai))+ln(Λ(U(tϖ),A(tϖ))Ψ1(Li)Λ(Ui,Ai)Ψ1(L))+ln(Ψ1(L)Λ(U,Ai)Ψ1(Li)Λ(U,A)),i=1,2 (2.16)
    ln(Ψ1(L(tϖ))Ψ1(L))=ln(Ψ1(L(tϖ))Ψ2(Ai)Ψ1(Li)Ψ2(A))+ln(Ψ2(A)Ψ1(Li)Ψ2(Ai)Ψ1(L)),i=1,2 (2.17)
    ln(Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A))=ln(Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L))+ln(Ψ1(L)Λ(U,A1)Ψ1(L1)Λ(U,A)) (2.18)

    Let Tj(U,L,A,C) be a Lyapunov function candidate and define a set

    Υj={(U,L,A,C):dTjdt=0},j=0,1,2.

    and Υj be the largest invariant subset of Υj.

    Theorem 1. If R01 and Conditions C1–C4 are satisfied, then the equilibrium point EP0 is globally asymptotically stable (GAS).

    Proof. Let T0 be defined as:

    T0=UU0UU0limA0+Λ(U0,A)Λ(η,A)dη+1ϝ1L+(π+μ)πϝ1ϝ2A+α(π+μ)ρπϝ1ϝ2C+1ϝ1κ10eξ1ϖf1(ϖ)ttϖΛ(U(θ),A(θ))dθdϖ+(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)ttϖΨ1(L(θ))dθdϖ.

    Obviously, T0(U,L,A,C)>0 for all U,L,A,C>0, while T0(U0,0,0,0)=0. The time derivative of T0 along the trajectories of systems (2.6)–(2.9) is calculated as follows:

    dT0dt=(1limA0+Λ(U0,A)Λ(U,A))˙U+1ϝ1˙L+(π+μ)πϝ1ϝ2˙A+α(π+μ)ρπϝ1ϝ2˙C+1ϝ1κ10eξ1ϖf1(ϖ)(Λ(U,A)Λ(U(tϖ),A(tϖ)))dϖ+(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)(Ψ1(L)Ψ1(L(tϖ)))dϖ=(1limA0+Λ(U0,A)Λ(U,A))(Φ(U)Λ(U,A))+1ϝ1(κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εΨ2(A)(π+μ)Ψ1(L))+(π+μ)πϝ1ϝ2(πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδΨ2(A)αΨ2(A)Ψ3(C))+α(π+μ)ρπϝ1ϝ2(ρΨ2(A)Ψ3(C)γΨ3(C))+1ϝ1(ϝ1Λ(U,A)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ)+(π+μ)ϝ1ϝ2(ϝ2Ψ1(L)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖ)=(1limA0+Λ(U0,A)Λ(U,A))Φ(U)+Λ(U,A)limA0+Λ(U0,A)Λ(U,A)+1ϝ1(εδ(π+μ)πϝ2)Ψ2(A)αγ(π+μ)ρπϝ1ϝ2Ψ3(C).

    Since Φ(U0)=0, then

    dT0dt=(1limA0+Λ(U0,A)Λ(U,A))(Φ(U)Φ(U0))+(Λ(U,A)Ψ2(A)limA0+Λ(U0,A)Λ(U,A)π(δεϝ2)+δμπϝ1ϝ2)Ψ2(A)αγ(π+μ)ρπϝ1ϝ2Ψ3(C).

    Using Condition C4 we obtain

    dT0dt(1limA0+Λ(U0,A)Λ(U,A))(Φ(U)Φ(U0))+(limA0+Λ(U,A)Ψ2(A)limA0+Λ(U0,A)Λ(U,A)π(δεϝ2)+δμπϝ1ϝ2)Ψ2(A)αγ(π+μ)ρπϝ1ϝ2Ψ3(C)=(1Λ(U0,0)/AΛ(U,0)/A)(Φ(U)Φ(U0))+(Λ(U0,0)/AΨ2(0)π(δεϝ2)+δμπϝ1ϝ2)Ψ2(A)αγ(π+μ)ρπϝ1ϝ2Ψ3(C)=(1Λ(U0,0)/AΛ(U,0)/A)(Φ(U)Φ(U0))+π(δεϝ2)+δμπϝ1ϝ2(R01)Ψ2(A)αγ(π+μ)ρπϝ1ϝ2Ψ3(C).

    Therefore, if R01, then dT0dt0 and dT0dt=0 when U=U0 and A=C=0. The solutions of systems (2.6)–(2.9) are limited Υ0 which contains elements with A(t)=0 and C(t)=0. Then from Eq (2.8)

    ˙A(t)=0=πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖL(t)=0,forallt.

    It follows that Υ0={EP0} and using L-LAST, we derive that EP0 is GAS [47,48,49].

    Remark 2. We know from Conditions C2 and C4 that

    (Λ(U,A)Λ(U,Ai))(Λ(U,A)Ψ2(A)Λ(U,Ai)Ψ2(Ai))0, U,Ai>0, i=1,2.

    which provides us

    (1Λ(U,Ai)Λ(U,A))(Λ(U,A)Λ(U,Ai)Ψ2(A)Ψ2(Ai))0, U,Ai>0, i=1,2.

    Lemma 2. Let R0>1 and Conditions C1C4 aresatisfied, then

    sgn(U2U1)=sgn(A1A2)=sgn(R11).

    Proof. From C1 and C2, for U1,U2,A1,A2>0, we obtain

    (Φ(U2)Φ(U1))(U1U2)>0, (2.19)
    (Λ(U2,A2)Λ(U1,A2))(U2U1)>0, (2.20)
    (Λ(U1,A2)Λ(U1,A1))(A2A1)>0. (2.21)

    Using Condition C4, we get

    (Λ(U1,A2)Ψ2(A2)Λ(U1,A1)Ψ2(A1))(A1A2)>0. (2.22)

    First, assume that sgn(U2U1)=sgn(A2A1). We have

    Φ(U2)Φ(U1)=Λ(U2,A2)Λ(U1,A1)=(Λ(U2,A2)Λ(U1,A2))+(Λ(U1,A2)Λ(U1,A1)).

    Hence, from Inequalities (2.19)–(2.21) we get:

    sgn(U1U2)=sgn(U2U1),

    which leads to a contradiction and hence, sgn(U2U1)=sgn(A1A2). Utilizing the equilibrium pointconditions for EP1 we obtain ϱΛ(U1,A1)Ψ2(A1)=1, then

    R11=ϱΛ(U2,A2)Ψ2(A2)ϱΛ(U1,A1)Ψ2(A1).=ϱ(Λ(U2,A2)Ψ2(A2)Λ(U1,A1)Ψ2(A1)).=ϱ(1Ψ2(A2)(Λ(U2,A2)Λ(U1,A2))+Λ(U1,A2)Ψ2(A2)Λ(U1,A1)Ψ2(A1)).

    Thus, from Inequalities (2.20) and (2.22) we obtain sgn(R11)=sgn(U2U1)=sgn(A1A2).

    Theorem 2. If R11<R0 and Conditions C1-C4 are satisfied, then the infected equilibrium point without CTL immunity EP1 is GAS.

    Proof. Let T1 be defined as:

    T1=UU1UU1Λ(U1,A1)Λ(η,A1)dη+1ϝ1(LL1LL1Ψ1(L1)Ψ1(η)dη)+(π+μ)πϝ1ϝ2(AA1AA1Ψ2(A1)Ψ2(η)dη)+α(π+μ)ρπϝ1ϝ2C+1ϝ1Λ(U1,A1)κ10eξ1ϖf1(ϖ)ttϖH(Λ(U(θ),A(θ))Λ(U1,A1))dθdϖ+(π+μ)ϝ1ϝ2Ψ1(L1)κ20eξ2ϖf2(ϖ)ttϖH(Ψ1(L(θ))Ψ1(L1))dθdϖ.

    Clearly T1(U,L,A,C)>0 for all U,L,A,C>0 and T1(U1,L1,A1,0)=0. Calculate dT1dt as

    dT1dt=(1Λ(U1,A1)Λ(U,A1))˙U+1ϝ1(1Ψ1(L1)Ψ1(L))˙L+(π+μ)πϝ1ϝ2(1Ψ2(A1)Ψ2(A))˙A+α(π+μ)ρπϝ1ϝ2˙C+1ϝ1κ10eξ1ϖf1(ϖ)[Λ(U,A)Λ(U(tϖ),A(tϖ))+Λ(U1,A1)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ]+(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)[Ψ1(L)Ψ1(L(tϖ))+Ψ1(L1)ln(Ψ1(L(tϖ))Ψ1(L))]dϖ=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Λ(U,A))+1ϝ1(1Ψ1(L1)Ψ1(L))(κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εΨ2(A)(π+μ)Ψ1(L))+(π+μ)πϝ1ϝ2(1Ψ2(A1)Ψ2(A))(πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδΨ2(A)αΨ2(A)Ψ3(C))+α(π+μ)ρπϝ1ϝ2(ρΨ2(A)Ψ3(C)γΨ3(C))+Λ(U,A)+(π+μ)ϝ1Ψ1(L)1ϝ1κ10eξ1ϖf1(ϖ)(Λ(U(tϖ),A(tϖ))Λ(U1,A1)ln(Λ(U(tϖ),A(tϖ))Λ(U,A)))dϖ(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)(Ψ1(L(tϖ))Ψ1(L1)ln(Ψ1(L(tϖ))Ψ1(L)))dϖ.

    Collecting terms we get

    dT1dt=(1Λ(U1,A1)Λ(U,A1))Φ(U)+Λ(U1,A1)Λ(U,A)Λ(U,A1)1ϝ1Ψ1(L1)Ψ1(L)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εϝ1Ψ2(A)εϝ1Ψ2(A)Ψ1(L1)Ψ1(L)+(π+μ)ϝ1Ψ1(L1)(π+μ)ϝ1ϝ2Ψ2(A1)Ψ2(A)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδ(π+μ)πϝ1ϝ2Ψ2(A)+δ(π+μ)πϝ1ϝ2Ψ2(A1)+α(π+μ)πϝ1ϝ2Ψ2(A1)Ψ3(C)αγ(π+μ)ρπϝ1ϝ2Ψ3(C)+1ϝ1Λ(U1,A1)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+(π+μ)ϝ1ϝ2Ψ1(L1)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ.

    Using Φ(U1)=Λ(U1,A1) we obtain

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(1Λ(U1,A1)Λ(U,A1))+Λ(U1,A1)Λ(U,A)Λ(U,A1)1ϝ1Ψ1(L1)Ψ1(L)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖπ(δεϝ2)+δμπϝ1ϝ2Ψ2(A)εϝ1Ψ2(A)Ψ1(L1)Ψ1(L)+(π+μ)ϝ1Ψ1(L1)(π+μ)ϝ1ϝ2Ψ2(A1)Ψ2(A)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖ+δ(π+μ)πϝ1ϝ2Ψ2(A1)+α(π+μ)πϝ1ϝ2(Ψ2(A1)γρ)Ψ3(C)+1ϝ1Λ(U1,A1)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+(π+μ)ϝ1ϝ2Ψ1(L1)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ.

    Utilizing the equilibrium point conditions for EP1:

    (π+μ)Ψ1(L1)=δ(π+μ)ϝ1π(δεϝ2)+δμΛ(U1,A1),
    δ(π+μ)πΨ2(A1)=δ(π+μ)ϝ1ϝ2π(δεϝ2)+δμΛ(U1,A1),

    we obtain

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(1Λ(U1,A1)Λ(U,A1))+Λ(U1,A1)Λ(U,A)Λ(U,A1)1ϝ1Λ(U1,A1)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L1)Λ(U1,A1)Ψ1(L)dϖΛ(U1,A1)Ψ2(A)Ψ2(A1)πεϝ2π(δεϝ2)+δμΛ(U1,A1)Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L)+2δ(π+μ)π(δεϝ2)+δμΛ(U1,A1)δ(π+μ)ϝ2[π(δεϝ2)+δμ]Λ(U1,A1)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ+δ(π+μ)ϝ2[π(δεϝ2)+δμ]Λ(U1,A1)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ+1ϝ1Λ(U1,A1)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+α(π+μ)πϝ1ϝ2(Ψ2(A1)γρ)Ψ3(C).

    Therefore dT1dt can be written as:

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(Λ(U,A)Λ(U,A1)Ψ2(A)Ψ2(A1))+Λ(U1,A1)[1Λ(U1,A1)Λ(U,A1)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L1)Λ(U1,A1)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ]πεϝ2π(δεϝ2)+δμΛ(U1,A1)Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L)+δ(π+μ)π(δεϝ2)+δμΛ(U1,A1)[21ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]+α(π+μ)πϝ1ϝ2(Ψ2(A1)γρ)Ψ3(C)=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(Λ(U,A)Λ(U,A1)1Ψ2(A)Ψ2(A1)+Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A))+Λ(U1,A1)[4Λ(U1,A1)Λ(U,A1)Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L1)Λ(U1,A1)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ1ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]πεϝ2π(δεϝ2)+δμΛ(U1,A1)Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L)2Λ(U1,A1)+1ϝ2Λ(U1,A1)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ1ϝ2Λ(U1,A1)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ+α(π+μ)πϝ1ϝ2(Ψ2(A1)γρ)Ψ3(C)+δ(π+μ)π(δεϝ2)+δμΛ(U1,A1)[21ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]. (2.23)

    Equation (2.23) can be written as:

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(Λ(U,A)Λ(U,A1)Ψ2(A)Ψ2(A1))(1Λ(U,A1)Λ(U,A))+Λ(U1,A1)[4Λ(U1,A1)Λ(U,A1)Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L1)Λ(U1,A1)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ1ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]+πεπ(δεϝ2)+δμΛ(U1,A1)×κ20eξ2ϖf2(ϖ)[2Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L)Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A)+ln(Ψ1(L(tϖ))Ψ1(L))]dϖ+α(π+μ)πϝ1ϝ2(Ψ2(A1)γρ)Ψ3(C). (2.24)

    Equation (2.24) can be simplified as:

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(Λ(U,A)Λ(U,A1)Ψ2(A)Ψ2(A1))(1Λ(U,A1)Λ(U,A))Λ(U1,A1)[H(Λ(U1,A1)Λ(U,A1))+1ϝ1κ10eξ1ϖf1(ϖ)H(Λ(U(tϖ),A(tϖ))Ψ1(L1)Λ(U1,A1)Ψ1(L))dϖ+H(Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A))+1ϝ2κ20eξ2ϖf2(ϖ)H(Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A))dϖ]πεϝ2π(δεϝ2)+δμΛ(U1,A1)[H(Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L))+1ϝ2κ20eξ2ϖf2(ϖ)H(Ψ1(L(tϖ))Ψ2(A1)Ψ1(L1)Ψ2(A))dϖ]+α(π+μ)πϝ1ϝ2(Ψ2(A1)Ψ2(A2))Ψ3(C).

    Lemma 2 implies that, if R11, then A1A2 and from Condition C3 we get Ψ2(A1)Ψ2(A2). Therefore, if R11, then dT1dt0, where dT1dt=0 occurs at the equilibrium point EP1. Hence Υ1={EP1} and then L-LAST guarantees the global asymptotic stability of EP1.

    Theorem 3. If R1>1 and Conditions C1–C4 are satisfied, thenthe infected equilibrium point with CTL immunity EP2 is GAS.

    Proof. We define T2 as:

    T2=UU2UU2Λ(U2,A2)Λ(η,A2)dη+1ϝ1(LL2LL2Ψ1(L2)Ψ1(η)dη)+(π+μ)πϝ1ϝ2(AA2AA2Ψ2(A2)Ψ2(η)dη)+α(π+μ)ρπϝ1ϝ2(CC2CC2Ψ3(C2)Ψ3(η)dη)+1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)ttϖH(Λ(U(θ),A(θ))Λ(U2,A2))dθdϖ+(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)ttϖH(Ψ1(L(θ))Ψ1(L2))dθdϖ.

    Clearly, T2(U,L,A,C)>0 for all U,L,A,C>0 and T2(U2,L2,A2,C2)=0. Calculate dT2dt as:

    dT2dt=(1Λ(U2,A2)Λ(U,A2))˙U+1ϝ1(1Ψ1(L2)Ψ1(L))˙L+(π+μ)πϝ1ϝ2(1Ψ2(A2)Ψ2(A))˙A+α(π+μ)ρπϝ1ϝ2(1Ψ3(C2)Ψ3(C))˙C+1ϝ1κ10eξ1ϖf1(ϖ)(Λ(U,A)Λ(U(tϖ),A(tϖ))+Λ(U2,A2)ln(Λ(U(tϖ),A(tϖ))Λ(U,A)))dϖ+(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)×(Ψ1(L)Ψ1(L(tϖ))+Ψ1(L2)ln(Ψ1(L(tϖ))Ψ1(L)))dϖ=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Λ(U,A))+1ϝ1(1Ψ1(L2)Ψ1(L))(κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εΨ2(A)(π+μ)Ψ1(L))+(π+μ)πϝ1ϝ2(1Ψ2(A2)Ψ2(A))(πκ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδΨ2(A)αΨ2(A)Ψ3(C))+α(π+μ)ρπϝ1ϝ2(1Ψ3(C2)Ψ3(C))(ρΨ2(A)Ψ3(C)γΨ3(C))+Λ(U,A)+(π+μ)ϝ1Ψ1(L)1ϝ1κ10eξ1ϖf1(ϖ)(Λ(U(tϖ),A(tϖ))dϖΛ(U2,A2)ln(Λ(U(tϖ),A(tϖ))Λ(U,A)))dϖ(π+μ)ϝ1ϝ2κ20eξ2ϖf2(ϖ)(Ψ1(L(tϖ))dϖΨ1(L2)ln(Ψ1(L(tϖ))Ψ1(L)))dϖ.

    Collecting terms we get

    dT2dt=(1Λ(U2,A2)Λ(U,A2))Φ(U)+Λ(U2,A2)Λ(U,A)Λ(U,A2)1ϝ1Ψ1(L2)Ψ1(L)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))dϖ+εϝ1Ψ2(A)εϝ1Ψ2(A)Ψ1(L2)Ψ1(L)+(π+μ)ϝ1Ψ1(L2)(π+μ)ϝ1ϝ2Ψ2(A2)Ψ2(A)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))dϖδ(π+μ)πϝ1ϝ2Ψ2(A)+δ(π+μ)πϝ1ϝ2Ψ2(A2)+α(π+μ)πϝ1ϝ2Ψ2(A2)Ψ3(C)α(π+μ)πϝ1ϝ2Ψ2(A)Ψ3(C2)αγ(π+μ)ρπϝ1ϝ2Ψ3(C)+αγ(π+μ)ρπϝ1ϝ2Ψ3(C2)+1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ.

    Using Φ(U2)=Λ(U2,A2) we obtain

    dT2dt=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(1Λ(U2,A2)Λ(U,A2))+Λ(U2,A2)Λ(U,A)Λ(U,A2)+1ϝ1(εΨ2(A2)(π+μ)πϝ2(δΨ2(A2)+αΨ2(A2)Ψ3(C2)))Ψ2(A)Ψ2(A2)+(π+μ)ϝ1Ψ1(L2)1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L)dϖεϝ1Ψ2(A2)Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L)(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+(π+μ)πϝ1ϝ2(δΨ2(A2)+αγρΨ3(C2))+α(π+μ)πϝ1ϝ2(Ψ2(A2)γρ)Ψ3(C2)+1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ.

    Using the equilibrium point conditions for EP2:

    Λ(U2,A2)=1ϝ1((π+μ)Ψ1(L2)εΨ2(A2)),πΨ1(L2)=1ϝ2(δΨ2(A2)+αΨ2(A2)Ψ3(C2)),Ψ2(A2)=γρ,

    we obtain

    dT2dt=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(1Λ(U2,A2)Λ(U,A2))+Λ(U2,A2)Λ(U,A)Λ(U,A2)+1ϝ1(εΨ2(A2)(π+μ)Ψ1(L2))Ψ2(A)Ψ2(A2)+2(π+μ)ϝ1Ψ1(L2)εϝ1Ψ2(A2)Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L)1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L)dϖ(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+1ϝ1Λ(U2,A2)κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ+(π+μ)ϝ1ϝ2Ψ1(L2)κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(Λ(U,A)Λ(U,A2)Ψ2(A)Ψ2(A2))+Λ(U2,A2)[1Λ(U2,A2)Λ(U,A2)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ]εϝ1Ψ2(A2)Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L)+(π+μ)ϝ1Ψ1(L2)[21ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]. (2.26)

    Equation (2.26) can be written as:

    dT2dt=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(Λ(U,A)Λ(U,A2)1Ψ2(A)Ψ2(A2)+Ψ2(A)Λ(U,A2)Ψ2(A2)Λ(U,A))+Λ(U2,A2)[4Λ(U2,A2)Λ(U,A2)Ψ2(A)Λ(U,A2)Ψ2(A2)Λ(U,A)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ1ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]εϝ1Ψ2(A2)Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L)2Λ(U2,A2)+1ϝ2Λ(U2,A2)κ20eξ2ϖf2(ϖ)(Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)ln(Ψ1(L(tϖ))Ψ1(L)))dϖ+(π+μ)ϝ1Ψ1(L2)[21ϝ2κ20eξ2ϖf2(ϖ)(Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)ln(Ψ1(L(tϖ))Ψ1(L)))dϖ]=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(Λ(U,A)Λ(U,A2)Ψ2(A)Ψ2(A2))(1Λ(U,A2)Λ(U,A))+Λ(U2,A2)[4Λ(U2,A2)Λ(U,A2)Ψ2(A)Λ(U,A2)Ψ2(A2)Λ(U,A)1ϝ1κ10eξ1ϖf1(ϖ)Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L)dϖ+1ϝ1κ10eξ1ϖf1(ϖ)ln(Λ(U(tϖ),A(tϖ))Λ(U,A))dϖ1ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ]+εϝ1Ψ2(A2)[2Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L)1ϝ2κ20eξ2ϖf2(ϖ)Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A)dϖ+1ϝ2κ20eξ2ϖf2(ϖ)ln(Ψ1(L(tϖ))Ψ1(L))dϖ].

    Using Eqs (2.16) and (2.18) we get

    dT2dt=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(Λ(U,A)Λ(U,A2)Ψ2(A)Ψ2(A2))(1Λ(U,A2)Λ(U,A))Λ(U2,A2)[H(Λ(U2,A2)Λ(U,A2))+1ϝ1κ10eξ1ϖf1(ϖ)H(Λ(U(tϖ),A(tϖ))Ψ1(L2)Λ(U2,A2)Ψ1(L))dϖ+H(Ψ2(A)Λ(U,A2)Ψ2(A2)Λ(U,A))+1ϝ2κ20eξ2ϖf2(ϖ)H(Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A))dϖ]εϝ1Ψ2(A2)[H(Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L))+1ϝ2κ20eξ2ϖf2(ϖ)H(Ψ1(L(tϖ))Ψ2(A2)Ψ1(L2)Ψ2(A))dϖ].

    Hence, if R1>1, then U2, L2, A2, C2>0. From Conditions C1, C2 and C4, we get that dT2dt0 and dT2dt=0 when U=U2,L=L2 and A=A2. The solutions of the system converge to Υ2 which has elements with U=U2, L=L2 and A=A2. It follows that ˙A=0 and Eq (2.8) becomes

    0=πϝ2Ψ1(L2)δΨ2(A2)αΨ2(A2)Ψ3(C(t))C(t)=C2forallt,

    and hence Υ2={EP2}. L-LAST implies the global asymptotic stability of EP2.

    In this section, we take a particular form of the probability distributed functions as:

    fi(ϖ)=Δ(ϖτi), i=1,2,

    where Δ(.) is the Dirac delta function and τi, i=1,2  are constants. When κi,i=1,2, we have

    0fi(ϖ)dϖ=1,0Δ(ϖτi)eξiϖdϖ=eξiτi.

    Moreover,

    0Δ(ϖτ1)eξ1ϖΛ(U(tϖ),A(tϖ))dϖ=eξ1τ1Λ(U(tτ1),A(tτ1)),0Δ(ϖτ2)eξ2ϖΨ1(L(tϖ))dϖ=eξ2τ2Ψ1(L(tτ2)),

    Hence, models (2.6)–(2.9) will reduce to the following model with discrete-time delays:

    ˙U=Φ(U(t))Λ(U(t),A(t)), (3.1)
    ˙L=eξ1τ1Λ(U(tτ1),A(tτ1))+εΨ2(A(t))(π+μ)Ψ1(L(t)), (3.2)
    ˙A=πeξ2τ2Ψ1(L(tτ2))δΨ2(A(t))αΨ2(A(t))Ψ3(C(t)), (3.3)
    ˙C=ρΨ2(A(t))Ψ3(C(t))γΨ3(C(t)). (3.4)

    Similar to the proof of Proposition 1, one can show that the solutions of systems (3.1)–(3.4) are non-negative and ultimately bounded.

    Corollary 1. Consider systems (3.1)–(3.4) and suppose that C1C4 are satisfied, then there exist two positive threshold numbers R0 and R1 with R0>R1 such that:

    (a) If R01, then there exists a single equilibrium point EP0.

    (b) If R11<R0, then there exist two equilibrium points EP0 and EP1.

    (c) If R1>1, then there exist three equilibrium points EP0, EP1 and EP2.

    The proof can be completed as the same as given in Lemma 1 by replacing ϝi by eξiτi,i=1,2. The parameters R0 and R1 for system (3.1)-(3.4) are given by

    R0=ϱSΨ2(0)Λ(U0,0)A,   R1=ϱSΛ(U2,A2)Ψ2(A2),

    where ϱS=πeξ1τ1ξ2τ2π(δεeξ2τ2)+δμ.

    We can obtain the global dynamics of models (3.1)–(3.4) as follows:

    Corollary 2. Consider systems (3.1)–(3.4) and suppose that Conditions C1C4 are satisfied.

    (ⅰ) If R01, then the equilibrium point EP0 is GAS.

    (ⅱ) If R11<R0, then the infected equilibrium point without CTL immunity EP1 is GAS.

    (ⅲ) If R1>1, then the infected equilibrium point with CTL immunity EP2 is GAS.

    Proof. (ⅰ) We consider the Lyapunov function T0 as:

    T0=UU0UU0limA0+Λ(U0,A)Λ(η,A)dη+eξ1τ1L+(π+μ)eξ1τ1+ξ2τ2πA+α(π+μ)eξ1τ1+ξ2τ2ρπC+ttτ1Λ(U(θ),A(θ))dθ+(π+μ)eξ1τ1ttτ2Ψ1(L(θ))dθ.

    Following the proof of Theorem 1 we obtain dT0dt as:

    dT0dt(1Λ(U0,0)/AΛ(U,0)/A)(Φ(U)Φ(U0))+(π(δεeξ2τ2)+δμ)eξ1τ1+ξ2τ2π(R01)Ψ2(A)αγ(π+μ)eξ1τ1+ξ2τ2ρπΨ3(C).

    (ⅱ) We define the Lyapunov function T1 as:

    T1=UU1UU1Λ(U1,A1)Λ(η,A1)dη+eξ1τ1(LL1LL1Ψ1(L1)Ψ1(η)dη)+(π+μ)eξ1τ1+ξ2τ2π(AA1AA1Ψ2(A1)Ψ2(η)dη)+α(π+μ)eξ1τ1+ξ2τ2ρπC+Λ(U1,A1)ttτ1H(Λ(U(θ),A(θ))Λ(U1,A1))dθ+(π+μ)eξ1τ1Ψ1(L1)ttτ2H(Ψ1(L(θ))Ψ1(L1))dθ.

    Then, we follow the proof of Theorem 2 to get

    dT1dt=(1Λ(U1,A1)Λ(U,A1))(Φ(U)Φ(U1))+Λ(U1,A1)(Λ(U,A)Λ(U,A1)Ψ2(A)Ψ2(A1))(1Λ(U,A1)Λ(U,A))Λ(U1,A1)[H(Λ(U1,A1)Λ(U,A1))+H(Ψ2(A)Λ(U,A1)Ψ2(A1)Λ(U,A))+H(Λ(U(tτ1),A(tτ1))Ψ1(L1)Λ(U1,A1)Ψ1(L))+H(Ψ1(L(tτ2))Ψ2(A1)Ψ1(L1)Ψ2(A))]πεeξ2τ2(π(δεeξ2τ2)+δμ)Λ(U1,A1)[H(Ψ2(A)Ψ1(L1)Ψ2(A1)Ψ1(L))+H(Ψ1(L(tτ2))Ψ2(A1)Ψ1(L1)Ψ2(A))]+α(π+μ)eξ1τ1+ξ2τ2π(Ψ2(A1)γρ)Ψ3(C).

    (ⅲ) We define the Lyapunov function T2 as:

    T2=UU2UU2Λ(U2,A2)Λ(η,A2)dη+eξ1τ1(LL2LL2Ψ1(L2)Ψ1(η)dη)+(π+μ)eξ1τ1+ξ2τ2π(AA2AA2Ψ2(A2)Ψ2(η)dη)+α(π+μ)eξ1τ1+ξ2τ2ρπ(CC2CC2Ψ3(C2)Ψ3(η)dη)+Λ(U2,A2)ttτ1H(Λ(U(θ),A(θ))Λ(U2,A2))dθ+(π+μ)eξ1τ1Ψ1(L2)ttτ2H(Ψ1(L(θ))Ψ1(L2))dθ.

    Following the proof of Theorem 3 we obtain dT2dt as:

    dT2dt=(1Λ(U2,A2)Λ(U,A2))(Φ(U)Φ(U2))+Λ(U2,A2)(Λ(U,A)Λ(U,A2)Ψ2(A)Ψ2(A2))(1Λ(U,A2)Λ(U,A))Λ(U2,A2)[H(Λ(U2,A2)Λ(U,A2))+H(Ψ2(A)Λ(U,A2)Ψ2(A2)Λ(U,A))+H(Λ(U(tτ1),A(tτ1))Ψ1(L2)Λ(U2,A2)Ψ1(L))+H(Ψ1(L(tτ2))Ψ2(A2)Ψ1(L2)Ψ2(A))]εeξ1τ1Ψ2(A2)[H(Ψ2(A)Ψ1(L2)Ψ2(A2)Ψ1(L))+H(Ψ1(L(tτ2))Ψ2(A2)Ψ1(L2)Ψ2(A))]. (3.5)

    Similar to the proofs of Theorems 1–3 one can complete the proof.

    To illustrate and validate our theoretical results we present some numerical simulations for systems (3.1)–(3.4) with the following specific shapes of the general functions:

    ● Logistic function for the intrinsic growth rate of uninfected cells

    Φ(U)=νˉνU+ςU(1UUmax),

    ● Crowley-Martin incidence rate

    Λ(U,A)=βUA(1+η1U)(1+η2A),

    ● Linear death rate of the cells

    Ψi(s)=s,i=1,2,3.

    Thus, systems (3.1)–(3.4) become

    ˙U(t)=νˉνU(t)+ςU(t)(1U(t)Umax)βU(t)A(t)(1+η1U(t))(1+η2A(t)), (4.1)
    ˙L(t)=eξ1τ1βU(tτ1)A(tτ1)(1+η1U(tτ1))(1+η2A(tτ1))+ωεA(t)(π+μ)L(t), (4.2)
    ˙A(t)=πeξ2τ2L(tτ2)+(1ω)εA(t)δA(t)αA(t)C(t), (4.3)
    ˙C(t)=ρA(t)C(t)γC(t). (4.4)

    Clearly Φ(0)=ν>0 and Φ(U0)=0, where U0=Umax2ς(ςˉν+(ςˉν)2+4νςUmax). It was mentioned in [16] that the term ςˉν is not necessary to be positive, therefore we assume that ς<ˉν. We see that Φ(U)=ˉν+ς2ςUUmax<0.

    Obviously, Φ(U)>0 for all U[0,U0). Hence, Condition C1 is satisfied. We note that Λ(U,A)>0 for all U,A>0 and Λ(U,0)=Λ(0,A)=0. Moreover, for all U>0,A>0 we have

    Λ(U,A)U=βA(1+η1U)2(1+η2A)>0,
    Λ(U,A)A=βU(1+η1U)(1+η2A)2>0,
    Λ(U,0)A=βU1+η1U>0,
    ddU(Λ(U,0)A)=β(1+η1U)2>0.

    Then Condition C2 is satisfied.

    Also, we can see that Ψi(s)>0, for all s>0 and Ψi(0)=0,i=1,2,3. Moreover, Ψi(s)=1>0, i=1,2,3 for all s0. Thus, Condition C3 is satisfied. Finally, we have

    Λ(U,A)Ψ2(A)=βU(1+η1U)(1+η2A),
    A(Λ(U,A)Ψ2(A))=η2βU(1+η1U)(1+η2A)2<0,forallU,A>0.

    Thus, Condition C4 is satisfied as well. We can also see that our example satisfied the global stability results (Theorems 1–3) with, R0 and R1 given by

    R0=βπeξ1τ1ξ2τ2U0[π(δ(1ω)εωεeξ2τ2)+(δ(1ω)ε)μ](1+η1U0),R1=βπeξ1τ1ξ2τ2U2ρ[π(δ(1ω)εωεeξ2τ2)+(δ(1ω)ε)μ](1+η1U2)(ρ+η2γ),

    where U2 is the positive root of the equation

    νˉνU+ςU(1UUmax)βUγ(1+η1U)(ρ+η2γ)=0.

    We present some numerical results for models (4.1)–(4.4) to address the stability of equilibrium points. Further, we address the effect of time delays on the dynamical behavior of the system. All the simulations are done using the solver dde23 MATLAB.

    In this subsection, we illustrate our global stability results given in Theorems 1–3 by showing that from any selected initial conditions, the solution of the HTLV-I infection model will tend to one of the three equilibrium points. Let us select three initial conditions as:

    IC1:(U(θ),L(θ),A(θ),C(θ))=(600,150,15,4),IC2:(U(θ),L(θ),A(θ),C(θ))=(500,200,20,3),IC3:(U(θ),L(θ),A(θ),C(θ))=(400,250,25,2),whereθ[max{τ1,τ2},0].

    We fix the time delays τ1 and τ2 to become τ1=τ2=0.1. We use the values of some parameters of models (4.1)–(4.4) given in Table 1 and establish the effect of the other parameters, β and ρ on the stability of the three equilibrium points. We investigate three cases as follows:

    Table 1.  The data of models (4.1)–(4.4).
    P Value Sources P Value Sources
    ν 10 cells mm3 day1 [10,14,50] μ 0.03 day1 [1,10,14,27]
    ˉν 0.012 day1 [10,14,26] δ 0.03 day1 [10,14,26]
    ς 0.03 day1 [51,52] α 0.029 cells1 mm3 day1 [14,26,54]
    Umax 1500 cells mm3 [51,52] ρ Varies Assumed
    β Varies Assumed γ 0.03 day1 [10,14,26]
    η1 0.0013 cells1 mm3 [53] ξ1 1 Assumed
    η2 0.009 cells1 mm3 [53] ξ2 1 Assumed
    ε 0.011 day1 [27] τ1 Varies Assumed
    ω 0.9 [27] τ2 Varies Assumed
    π 0.003 day1 [10,13,14,26]

     | Show Table
    DownLoad: CSV

    Case 1. (R01): Choosing β=0.0001 and ρ=0.0001 which gives R0=0.1276<1 and R1=0.0332<1. Lemma 1 and Theorem 1 prove that the system has a single equilibrium point, EP0 and it is GAS. Figure 1 shows that for all the threeinitials IC1-IC3, the concentrations of latently infected cells, activelyinfected cells and CTLs are tending to zero. In the mean time, theconcentration of the uninfected cells tends to its healthy value U0=1288.15. This case means that there is no HTLV-I infection in thebody.

    Figure 1.  Solutions of systems (4.1)–(4.4) with initial conditions IC1-IC3 in case of R01, (a) Uninfected cells; (b) Latently infected cells; (c) Actively infected cells; (d) CTLs.

    Case 2. (R11<R0): We take β=0.004 and ρ=0.0001. So, we get R0=5.1054>1 and R1=0.0936<1. Based on Lemma 1 and Theorem 2, the infected equilibrium point withoutCTL immunity EP1 exists. In such case, EP1 is GAS. Figure 2 shows that the numerical simulations illustrate our theoreticalresult given in Theorem 2. The system'ssolutions converge to the equilibrium point EP1=(144.65,343.80,32.29,0) for all the three initials IC1-IC3. In this situation, the HTLV-I-patienthas non-active CTL immunity.

    Figure 2.  Solutions of systems (4.1)–(4.4) with initial conditions IC1-IC3 in case of R11<R0, (a) Uninfected cells; (b) Latently infected cells; (c) Actively infected cells; (d) CTLs.

    Case 3. (R1>1): We choose β=0.004 and ρ=0.005. Then, we calculate R0=5.1054>1 and R1=4.2216>1. Lemma 1 and Theorem 3 state that the infected equilibrium pointwith CTL immunity EP2=(923.92,263.87,6,3.12) exists and isGAS. Figure 3 displays the numerical solutions of the systemconverge to EP2 for all the three initials IC1-IC3. The results supportthe theoretical results presented in Theorem 3. This case corresponds to an HTLV-I chronic infection with active CTLimmunity.

    Figure 3.  Solutions of systems (4.1)–(4.4) with initial conditions IC1-IC3 in case of R1>1, (a) Uninfected cells; (b) Latently infected cells; (c) Actively infected cells; (d) CTLs.

    We observe that the values of the parameter α,ρ and γ have no effect on the value of R0. It means that, CTL immunity has no effect on the stability of the infection-free equilibrium point EP0. Now we address the effect of the proliferation rate constant of CTL ρ on the HTLV-I dynamics. Using the values of the parameters in Table 1 to solve the model (4.1)-(4.4) with initial condition IC2. We choose β=0.004, τ1=τ2=0.1. We can see from Table 2 that R1 can be increased by increasing the value of ρ. We can see from Figure 4 when ρ is increased, the concentrations of uninfected cells and CTLs are increased, while the concentration of actively infected cells is decreased. Therefore, CTL immunity does not play the role in deleting the virus, but it has an important role in controlling the HTLV-I infection.

    Table 2.  The values of R1 for models (4.1)–(4.4) with selected values of ρ.
    ρ 0.0001 0.0005 0.001 0.002 0.005 0.01
    R1 0.094 0.50 1.09 2.47 4.22 4.69

     | Show Table
    DownLoad: CSV
    Figure 4.  Impact of the CTL immunity ρ on the dynamics of HTLV-I given by systems (4.1)–(4.4) with initial condition IC2, (a) Uninfected cells; (b) Latently infected cells; (c) Actively infected cells; (d) CTLs.

    In this part, we study the effect of vary the time delays τ1 and τ2 on the HTLV-1 dynamics. Using the values of the parameters in Table 1 and considering β=0.004 and ρ=0.001 to solve the models (4.1)–(4.4) with initial condition:

    IC4:U(θ)=500,L(θ)=200,A(θ)=20andC(θ)=0.2whereθ[max{τ1,τ2},0].

    Consider τ=τ1=τ2, then R0 and R1 can be given as functions of τ as follows:

    R0(τ)=βπe(ξ1+ξ2)τU0[πωεeξ2τ(π+μ)(δ(1ω)ε)](1+η1U0).
    R1(τ)=βπe(ξ1+ξ2)τU2ρ[πωεeξ2τ(π+μ)(δ(1ω)ε)](1+η1U2)(ρ+η2γ).

    We note that R0 and R1 are decreasing functions of τ. Let τCR0 and τCR1 be such that R0(τCR0)=1 and R1(τCR1)=1. Using the values of the parameters in Table 1 and considering β=0.004 and ρ=0.001, we obtain τCR0=0.907187 and τCR1=0.142356. Therefore, we have the following cases:

    (a) if τ0.907187, then R0(τ)1 and EP0 is GAS;

    (b) if 0.142356τ<0.907187, then R1(τ)1<R0(τ) and EP1 is GAS;

    (c) if τ<0.142356, then R1(τ)>1 and EP2 is GAS.

    It is clear from Figure 5 that as τ increases, the concentration of uninfected CD4+T cells is increased and the other concentrations are decreased. In Table 3 we calculate the values of R0 and R1 as functions of τ. We note that, when τ is increased, the values of both R0 and R1 are decreased. We can see from the above discussion that increasing time delays values can have similar effect as antiviral treatments.

    Figure 5.  Impact of time delays τ1, τ2 on the HTLV-I dynamics given by systems (4.1)–(4.4) with initial condition IC4 (a) Uninfected cells; (b) Latently infected cells; (c) Actively infected cells; (d) CTLs.
    Table 3.  The values of R0 and R1 for models (4.1)–(4.4) with selected values of τ.
    τ R0 R1
    0 6.25 1.34
    0.142356 4.69 1
    0.7 1.52 0.32
    0.907187 1 0.21
    1 0.83 0.18
    1.5 0.30 0.06

     | Show Table
    DownLoad: CSV

    In this work, we proposed and examined a general HTLV-I infection models with CTL immunity and mitotic transmission of actively infected cells. We introduce two intracellular distributed-time delays into the models: (ⅰ) delay in the formation of latently infected cells, (ⅱ) delay in the activation of latently infected cells. We constructed the model with more general nonlinear functions for the generation/stimulation and clearance rates of all compartments as well as the incidence rate of infection. We established some conditions on such general functions and determined two threshold numbers R0 and R1 which control the existence and global stability of the three equilibrium points. We formulated Lyapunov functions and used L-LAST to establish the global asymptotic stability of the equilibrium points. We proved that

    ● If R01, then the infection-free equilibrium point EP0 is GAS. This result suggests that when R01, the HTLV-I infection is predicted to die out regardless of the initial conditions. From a control viewpoint, making R01 will be an ideal way but HTLV-I infection is lifelong, and the viruses are rarely cleared.

    ● If R11<R0, then the infected equilibrium point without CTL immunity EP1 is GAS. This result establishes that when R11<R0, the HTLV-I infection with inactive CTL immunity is always established regardless of the initial conditions.

    ● If R1>1, then the infected equilibrium point with CTL immunity EP2 is GAS. This result illustrates that when R1>1, the HTLV-I infection with active CTL immunity is always established regardless of the initial conditions. This implies that the infection becomes chronic and the CTL immune response is persistent. In cases R11<R0 and R1>1, the individuals have a high risk of developing HAM/TSP and can be characterized as the HAM/TSP patients.

    We studied the global stability of a general HTLV-I model with discrete-time delays as a special case of the model with distributed-time delays when the probability distributed function is given by Dirac delta function. To support the obtained theoretical results, we conducted numerical simulations for the model with discrete-time delays, logistic form for the intrinsic growth rate of uninfected cells, Crowley-Martin incidence rate and linear death rate of the cells. We discussed the impacts of the proliferation of CTLs on the HTLV-I dynamics. We noted that the CTL immunity does not play the role in deleting the virus from the body, but it has an important role in controlling and suppressing the HTLV-I infection. We showed that the parameter ρ has a significant effect of the dynamical behavior of the HTLV-I. The effect of the time delays on the HTLV-I dynamics was discussed. We established that, when all other parameters are fixed and delays are sufficiently large, R0 becomes less than one, which makes the infection-free equilibrium point EP0 globally asymptotically stable. From a biological viewpoint, time delays play positive roles in the HTLV-I infection process in order to eliminate the virus. Sufficiently large time delays makes the HTLV-I development slower, and the HTLV-I is controlled and disappears. This gives us some suggestions on new drugs to prolong the time of formation of latent infected cells, or the time of activation of latent infected cells.

    We mention that models (1.8)–(1.11) presented by Lim and Maini [14] considered a linear proliferation rate of the CTLs, ρA. In comparison with our proposed model, models (1.8)–(1.11) admits only two equilibrium points, infection-free equilibrium point ~EP0=(νˉν,0,0,0) and infected equilibrium point ~EP1=(˜U1,˜L1,˜A1,˜C1). The existence and stability of ~EP0 and ~EP1 are determined by the basic reproductive number

    ˜R0=πδ(π+μ)(βνˉν+ε).

    We note that, ˜R0 does not depends on the proliferation rate constant of the CTLs, ρ. As a result, ρ does not play any role on the existence and stability of both equilibria. In this case, the effect of the CTL immune response on the dynamical behavior of the HTLV-I does not appear explicitly. In our model, by considering a nonlinear proliferation rate of the CTLs, the role of the CTLs in controlling the HTLV-I infection is well presented. In [1], an HTLV-I infection model was presented with a nonlinear proliferation rate of the CTLs, ρAC, however, the intracellular time delay was neglected. In addition, the infection rate was given by bilinear incidence which may not be completely characterize the incidence between uninfected cells and actively infected cells [37]. Overall, our proposed model helps to better understand the dynamical behavior of HTLV-I infection with CTL immunity and time delays.

    The model presented in this paper can be extended in many directions such as (ⅰ) including the reaction diffusion [55], (ⅱ) considering the stochastic interactions [56], (ⅲ) modeling the coinfection with HTLV-I and other types of viruses.

    This project was funded by the Deanship of Scientific Research (DSR), King Abdulaziz University, Jeddah, Saudi Arabia under grant no. (KEP-PHD-38-130-43). The authors, therefore, acknowledge with thanks DSR technical and financial support.

    The authors declare there is no conflict of interest.



    [1] S. Khajanchi, S. Bera, T. K. Roy, Mathematical analysis of the global dynamics of a HTLV-I infection model, considering the role of cytotoxic T-lymphocytes, Math. Comput. Simul., 180 (2021), 354–378. https://doi.org/10.1016/j.matcom.2020.09.009 doi: 10.1016/j.matcom.2020.09.009
    [2] F. A. Proietti, A. B. F. Carneiro-Proietti, B. C. Catalan-Soares, E. L. Murphy, Global epidemiology of HTLV-I infection and associated diseases, Oncogene, 24 (2005), 6058–6068. https://doi.org/10.1038/sj.onc.1208968 doi: 10.1038/sj.onc.1208968
    [3] C. R. M. Bangham, HTLV-I infections, J. Clin. Pathol., 53 (2000), 581–586. http://dx.doi.org/10.1136/jcp.53.8.581 doi: 10.1136/jcp.53.8.581
    [4] D. Wodarz, C. R. M. Bangham, Evolutionary dynamics of HTLV-I, J. Mol. Evol., 50 (2000), 448–455. https://doi.org/10.1007/s002390010047 doi: 10.1007/s002390010047
    [5] B. Asquith, C. R. M. Bangham, How does HTLV-I persist despite a strong cell-mediated immune response?, Trends Immunol., 29 (2008), 4–11. https://doi.org/10.1016/j.it.2007.09.006 doi: 10.1016/j.it.2007.09.006
    [6] H. Shiraki, Y. Sagara, Y. Inoue, Cell-to-cell transmission of HTLV-I, Gann Monogr. Cancer Res., 50 (2003), 303–316.
    [7] N. I. Stilianakis, J. Seydel, Modeling the T-cell dynamics and pathogenesis of HTLV-I infection, Bull. Math. Biol., 61 (1999), 935–947. https://doi.org/10.1006/bulm.1999.0117 doi: 10.1006/bulm.1999.0117
    [8] H. Gomez-Acevedo, M. Y. Li, Backward bifurcation in a model for HTLV-I infection of CD4+T cells, Bull. Math. Biol., 67 (2005), 101–114. https://doi.org/10.1016/j.bulm.2004.06.004 doi: 10.1016/j.bulm.2004.06.004
    [9] C. Vargas-De-Leon, The complete classification for global dynamics of a model for the persistence of HTLV-1 infection, Appl. Math. Comput., 237 (2014), 489–493. https://doi.org/10.1016/j.amc.2014.03.138 doi: 10.1016/j.amc.2014.03.138
    [10] M. Y. Li, A. G. Lim, Modelling the role of tax expression in HTLV-I persistence in vivo, Bull. Math. Biol., 73 (2011), 3008–3029. https://doi.org/10.1007/s11538-011-9657-1 doi: 10.1007/s11538-011-9657-1
    [11] X. Song, Y. Li, Global stability and periodic solution of a model for HTLV-1 infection and ATL progression, Appl. Math. Comput., 180 (2006), 401–410. https://doi.org/10.1016/j.amc.2005.12.022 doi: 10.1016/j.amc.2005.12.022
    [12] L. Wang, M. Y. Li, D. Kirschner, Mathematical analysis of the global dynamics of a model for HTLV-I infection and ATL progression, Math. Biosci., 179 (2002), 207–217. https://doi.org/10.1016/S0025-5564(02)00103-7 doi: 10.1016/S0025-5564(02)00103-7
    [13] B. Asquith, C. R. M. Bangham, Quantifying HTLV-I dynamics, Immunol. Cell Biol., 85 (2007), 280–286. https://doi.org/10.1038/sj.icb.7100050
    [14] A. G. Lim, P. K. Maini, HTLV-I infection: A dynamic struggle between viral persistence and host immunity, J. Theor. Biol., 352 (2014), 92–108. https://doi.org/10.1016/j.jtbi.2014.02.022 doi: 10.1016/j.jtbi.2014.02.022
    [15] C. R. M. Bangham, CTL quality and the control of human retroviral infections, Eur. J. Immunol., 39 (2009), 1700–1712. https://doi.org/10.1002/eji.200939451 doi: 10.1002/eji.200939451
    [16] X. Pan, Y. Chen, H. Shu, Rich dynamics in a delayed HTLV-I infection model: Stability switch, multiple stable cycles, and torus, J. Math. Anal. Appl., 479 (2019), 2214–2235. https://doi.org/10.1016/j.jmaa.2019.07.051 doi: 10.1016/j.jmaa.2019.07.051
    [17] C. Bartholdy, J. P. Christensen, D. Wodarz, A. R. Thomsen, Persistent virus infection despite chronic cytotoxic T-lymphocyte activation in gamma interferon-deficient mice infected with lymphocytic choriomeningitis virus, J. Virol., 74 (2000), 10304–10311. https://doi.org/10.1128/JVI.74.22.10304-10311.2000 doi: 10.1128/JVI.74.22.10304-10311.2000
    [18] H. Gomez-Acevedo, M. Y. Li, S. Jacobson, Multi-stability in a model for CTL response to HTLV-I infection and its implications to HAM/TSP development and prevention, Bull. Math. Biol., 72 (2010), 681–696. https://doi.org/10.1007/s11538-009-9465-z doi: 10.1007/s11538-009-9465-z
    [19] J. Lang, M. Y. Li, Stable and transient periodic oscillations in a mathematical model for CTL response to HTLV-I infection, J. Math. Biol., 65 (2012), 181–199. https://doi.org/10.1007/s00285-011-0455-z doi: 10.1007/s00285-011-0455-z
    [20] M. Y. Li, H. Shu, Multiple stable periodic oscillations in a mathematical model of CTL response to HTLV-I infection, Bull. Math. Biol., 73 (2011), 1774–1793. https://doi.org/10.1007/s11538-010-9591-7 doi: 10.1007/s11538-010-9591-7
    [21] M. Y. Li, H. Shu, Global dynamics of a mathematical model for HTLV-I infection of CD4+ T cells with delayed CTL response, Nonlinear Anal. Real World Appl., 13 (2012), 1080–1092. https://doi.org/10.1016/j.nonrwa.2011.02.026 doi: 10.1016/j.nonrwa.2011.02.026
    [22] D. Wodarz, M. A. Nowak, C. R. M. Bangham, The dynamics of HTLV-I and the CTL response, Immunol. Today, 20 (1999), 220–227. https://doi.org/10.1016/S0167-5699(99)01446-2 doi: 10.1016/S0167-5699(99)01446-2
    [23] L. Wang, Z. Liu, Y. Li, D. Xu, Complete dynamical analysis for a nonlinear HTLV-I infection model with distributed delay, CTL response and immune impairment, Discrete Contin. Dyn. Syst., 25 (2020), 917–933. http://dx.doi.org/10.3934/dcdsb.2019196 doi: 10.3934/dcdsb.2019196
    [24] Y. Muroya, Y. Enatsu, H. Li, Global stability of a delayed HTLV-I infection model with a class of nonlinear incidence rates and CTLs immune response, Appl. Math. Comput., 219 (2013), 10559–10573. https://doi.org/10.1016/j.amc.2013.03.081 doi: 10.1016/j.amc.2013.03.081
    [25] Y. Wang, J. Liu, J. M. Heffernan, Viral dynamics of an HTLV-I infection model with intracellular delay and CTL immune response delay, J. Math. Anal. Appl., 459 (2018), 506–527. https://doi.org/10.1016/j.jmaa.2017.10.027 doi: 10.1016/j.jmaa.2017.10.027
    [26] F. Li, W. Ma, Dynamics analysis of an HTLV-1 infection model with mitotic division of actively infected cells and delayed CTL immune response, Math. Methods Appl. Sci., 41 (2018), 3000–3017. https://doi.org/10.1002/mma.4797 doi: 10.1002/mma.4797
    [27] S. Li, Y. Zhou, Backward bifurcation of an HTLV-I model with immune response, Discrete Contin. Dyn. Syst., 21 (2016), 863–881. http://dx.doi.org/10.3934/dcdsb.2016.21.863 doi: 10.3934/dcdsb.2016.21.863
    [28] W. Wang, W. Ma, Global dynamics of a reaction and diffusion model for an HTLV-I infection with mitotic division of actively infected cells, J. Appl. Anal. Comput., 7 (2017), 899–930. http://dx.doi.org/10.11948/2017057 doi: 10.11948/2017057
    [29] X. Jia, R. Xu, Global dynamics of a delayed HTLV-I infection model with Beddington-DeAngelis incidence and immune impairment, Chaos Solitons Fractals, 155 (2022), 111733. https://doi.org/10.1016/j.chaos.2021.111733 doi: 10.1016/j.chaos.2021.111733
    [30] A. M. Elaiw, N. H. AlShamrani, Stability of HIV/HTLV-I co-infection model with delays, Math. Methods Appl. Sci., 45 (2022), 238–300. https://doi.org/10.1002/mma.7775 doi: 10.1002/mma.7775
    [31] P. Katri, S. Ruan, Dynamics of human T-cell lymphotropic virus I (HTLV-I) infection of CD4+T cells, C. R. Biol., 327 (2004), 1009–1016. https://doi.org/10.1016/j.crvi.2004.05.011 doi: 10.1016/j.crvi.2004.05.011
    [32] Y. Wang, J. Liu, Global stability for delay-dependent HTLV-I model with CTL immune response, AIP Conf. Proc., 1738 (2016), 480074. https://doi.org/10.1063/1.4952310 doi: 10.1063/1.4952310
    [33] X. Sun, J. Wei, Global dynamics of a HTLV-I infection model with CTL response, Elec. J. Qual. Theory Diff. Equations, 2013 (2013), 1–15.
    [34] M. Y. Li, X. Lin, H. Wang, Global Hopf branches in a delayed model for immune response to HTLV-1 infections: coexistence of multiple limit cycles, Can. Appl. Math. Q., 20 (2012), 39–50.
    [35] S. Bera, S. Khajanchi, T. K. Roy, Dynamics of an HTLV-I infection model with delayed CTLs immune response, Appl. Math. Comput., 430 (2022), 127206. https://doi.org/10.1016/j.amc.2022.127206 doi: 10.1016/j.amc.2022.127206
    [36] X. Lu, L. Hui, S. Liu, J. Li, A mathematical model of HTLV-I infection with two time delays, Math. Biosci. Eng., 12 (2015), 431–449. http://dx.doi.org/10.3934/mbe.2015.12.431 doi: 10.3934/mbe.2015.12.431
    [37] A. Korobeinikov, Global asymptotic properties of virus dynamics models with dose-dependent parasite reproduction and virulence and non-linear incidence rate, Math. Med. Biol. J. IMA, 26 (2009), 225–239. https://doi.org/10.1093/imammb/dqp006 doi: 10.1093/imammb/dqp006
    [38] K. Qi, D. Jiang, Threshold behavior in a stochastic HTLV-I infection model with CTL immune response and regime switching, Math. Methods Appl. Sci., 41 (2018), 6866–6882. https://doi.org/10.1002/mma.5198 doi: 10.1002/mma.5198
    [39] L. Cai, X. Li, M. Ghosh, Global dynamics of a mathematical model for HTLV-I infection of CD4+ T-cells, Appl. Math. Modell., 35 (2011), 3587–3595. https://doi.org/10.1016/j.apm.2011.01.033 doi: 10.1016/j.apm.2011.01.033
    [40] J. Hale, S. M. V. Lunel, Introduction to Functional Differential Equations, Applied Mathematical Science, Springer Verlag, New York, 1993.
    [41] A. M. Elaiw, N. H. AlShamrani, Modeling and analysis of a within-host HIV/HTLV-I co-infection, Bol. Soc. Mat. Mex., 27 (2021), 27–38. https://doi.org/10.1007/s40590-021-00330-6 doi: 10.1007/s40590-021-00330-6
    [42] A. M. Elaiw, N. H. AlShamrani, Analysis of a within-host HIV/HTLV-I co-infection model with immunity, Virus Res., 295 (2021), 198204. https://doi.org/10.1016/j.virusres.2020.198204 doi: 10.1016/j.virusres.2020.198204
    [43] X. Yang, L. Chen, J. Chen, Permanence and positive periodic solution for the single-species nonautonomous delay diffusive models, Comput. Math. Appl., 32 (1996), 109–116. https://doi.org/10.1016/0898-1221(96)00129-0 doi: 10.1016/0898-1221(96)00129-0
    [44] A. Korobeinikov, Global properties of basic virus dynamics models, Bull. Math. Biol., 66 (2004), 879–883. https://doi.org/10.1016/j.bulm.2004.02.001 doi: 10.1016/j.bulm.2004.02.001
    [45] A. Korobeinikov, Global properties of infectious disease models with nonlinear incidence, Bull. Math. Biol., 69 (2007), 1871–1886. https://doi.org/10.1007/s11538-007-9196-y doi: 10.1007/s11538-007-9196-y
    [46] A. M. Elaiw, N. H. AlShamrani, Global stability of humoral immunity virus dynamics models with nonlinear infection rate and removal, Nonlinear Anal. Real World Appl., 26 (2015), 161–190. https://doi.org/10.1016/j.nonrwa.2015.05.007 doi: 10.1016/j.nonrwa.2015.05.007
    [47] E. A. Barbashin, Introduction to the theory of stability, Wolters-Noordhoff, Groningen, 1970.
    [48] J. P. LaSalle, The Stability of Dynamical Systems, SIAM, Philadelphia, 1976.
    [49] A. M. Lyapunov, The general problem of the stability of motion, Int. J. Control, 55 (1992), 531–534. https://doi.org/10.1080/00207179208934253 doi: 10.1080/00207179208934253
    [50] A. S. Perelson, D. E. Kirschner, R. de boer, Dynamics of HIV Infection of CD4+ T cells, Math. Biosci., 114 (1993), 81–125. https://doi.org/10.1016/0025-5564(93)90043-A doi: 10.1016/0025-5564(93)90043-A
    [51] R. V. Culshaw, S. Ruan, A delay-differential equation model of HIV infection of CD4+ T-cells, Math. Biosci., 165 (2000), 27–39. https://doi.org/10.1016/S0025-5564(00)00006-7 doi: 10.1016/S0025-5564(00)00006-7
    [52] Y. Wang, Y. Zhou, J. Wu, J. Heffernan, Oscillatory viral dynamics in a delayed HIV pathogenesis model, Math. Biosci., 219 (2009), 104–112. https://doi.org/10.1016/j.mbs.2009.03.003 doi: 10.1016/j.mbs.2009.03.003
    [53] M. N. Jan, N. Ali, G. Zaman, I. Ahmad, Z. Shah, P. Kumam, HIV-1 infection dynamics and optimal control with Crowley-Martin function response, Comput. Methods Prog. Biomed., 193 (2020), 105503. https://doi.org/10.1016/j.cmpb.2020.105503 doi: 10.1016/j.cmpb.2020.105503
    [54] B. Asquith, A. J. Mosley, A. Barfield, S. E. F. Marshall, A. Heaps, P. Goon, et al., A functional CD8+ cell assay reveals individual variation in CD8 + cell antiviral efficacy and explains differences in human T-lymphotropic virus type 1 proviral load, J. Gener. Virol., 86 (2005), 1515–1523. https://doi.org/10.1099/vir.0.80766-0 doi: 10.1099/vir.0.80766-0
    [55] N. Bellomo, N. Outada, J. Soler, Y. Tao, M. Winkler, Chemotaxis and cross diffusion models in complex environments: Models and analytic problems toward a multiscale vision, Math. Models Methods Appl. Sci., 32 (2022), 713–792. https://doi.org/10.1142/S0218202522500166 doi: 10.1142/S0218202522500166
    [56] L. Gibelli, A. M. Elaiw, M. A. Alghamdi, A. M. Althiabi, Heterogeneous population dynamics of active particles: Progression, mutations, and selection dynamics, Math. Models Methods Appl. Sci., 27 (2017), 617–640. https://doi.org/10.1142/S0218202517500117 doi: 10.1142/S0218202517500117
  • This article has been cited by:

    1. Ahmed M. Elaiw, Abdulsalam S. Shflot, Aatef D. Hobiny, Shaban A. Aly, Global Dynamics of an HTLV-I and SARS-CoV-2 Co-Infection Model with Diffusion, 2023, 11, 2227-7390, 688, 10.3390/math11030688
    2. A. M. Elaiw, A. S. Shflot, A. D. Hobiny, Stability analysis of SARS-CoV-2/HTLV-I coinfection dynamics model, 2022, 8, 2473-6988, 6136, 10.3934/math.2023310
    3. M.A. Alshaikh, A.K. Aljahdali, Stability of a discrete HTLV-1/SARS-CoV-2 dual infection model, 2024, 10, 24058440, e28178, 10.1016/j.heliyon.2024.e28178
  • 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(2092) PDF downloads(109) Cited by(3)

Figures and Tables

Figures(5)  /  Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog