Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Global dynamics of a delayed diffusive virus infection model with cell-mediated immunity and cell-to-cell transmission

  • In this paper, we propose and analyze a delayed diffusive viral dynamic model incorporating cell-mediated immunity and both cell-free and cell-to-cell transmission. After discussing the well-posedness, we provide some preliminary results on solutions. Then we study the existence and uniqueness of homogeneous steady states, which turned out to be completely determined by the basic reproduction number of infection R0 and the basic reproduction number of immunity R1. Note that when R1 is defined, it is necessary that R0 > 1. The main result is a threefold dynamics. Roughly speaking, when R0 < 1 the infection-free steady state is globally asymptotically stable; when R1 ≤ 1 < R0 the immunity-free infected steady state is globally asymptotically stable; when R1 > 1 the infected-immune steady state is globally asymptotically stable. The approaches are linearization technique and the Lyapunov functional method. The theoretical results are also illustrated with numerical simulations.

    Citation: Chunyang Qin, Yuming Chen, Xia Wang. Global dynamics of a delayed diffusive virus infection model with cell-mediated immunity and cell-to-cell transmission[J]. Mathematical Biosciences and Engineering, 2020, 17(5): 4678-4705. doi: 10.3934/mbe.2020257

    Related Papers:

    [1] Jiazhe Lin, Rui Xu, Xiaohong Tian . Threshold dynamics of an HIV-1 model with both viral and cellular infections, cell-mediated and humoral immune responses. Mathematical Biosciences and Engineering, 2019, 16(1): 292-319. doi: 10.3934/mbe.2019015
    [2] Yan Wang, Minmin Lu, Daqing Jiang . Viral dynamics of a latent HIV infection model with Beddington-DeAngelis incidence function, B-cell immune response and multiple delays. Mathematical Biosciences and Engineering, 2021, 18(1): 274-299. doi: 10.3934/mbe.2021014
    [3] 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
    [4] 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
    [5] Jinhu Xu, Yicang Zhou . Bifurcation analysis of HIV-1 infection model with cell-to-cell transmission and immune response delay. Mathematical Biosciences and Engineering, 2016, 13(2): 343-367. doi: 10.3934/mbe.2015006
    [6] 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
    [7] A. M. Elaiw, N. H. AlShamrani, A. D. Hobiny . Stability of an adaptive immunity delayed HIV infection model with active and silent cell-to-cell spread. Mathematical Biosciences and Engineering, 2020, 17(6): 6401-6458. doi: 10.3934/mbe.2020337
    [8] A. M. Elaiw, A. S. Shflot, A. D. Hobiny . Stability analysis of general delayed HTLV-I dynamics model with mitosis and CTL immunity. Mathematical Biosciences and Engineering, 2022, 19(12): 12693-12729. doi: 10.3934/mbe.2022593
    [9] 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
    [10] Farzad Fatehi, Yuliya N. Kyrychko, Konstantin B. Blyuss . Time-delayed model of autoimmune dynamics. Mathematical Biosciences and Engineering, 2019, 16(5): 5613-5639. doi: 10.3934/mbe.2019279
  • In this paper, we propose and analyze a delayed diffusive viral dynamic model incorporating cell-mediated immunity and both cell-free and cell-to-cell transmission. After discussing the well-posedness, we provide some preliminary results on solutions. Then we study the existence and uniqueness of homogeneous steady states, which turned out to be completely determined by the basic reproduction number of infection R0 and the basic reproduction number of immunity R1. Note that when R1 is defined, it is necessary that R0 > 1. The main result is a threefold dynamics. Roughly speaking, when R0 < 1 the infection-free steady state is globally asymptotically stable; when R1 ≤ 1 < R0 the immunity-free infected steady state is globally asymptotically stable; when R1 > 1 the infected-immune steady state is globally asymptotically stable. The approaches are linearization technique and the Lyapunov functional method. The theoretical results are also illustrated with numerical simulations.


    Viral dynamics is a field of applied mathematics which employs mathematical models to describe the changes over time of infected cells and the viral load. Nowak et al. [1] and Nowak and May [2] proposed the following basic viral dynamic model,

    {du(t)dt=sdu(t)βu(t)v(t),dw(t)dt=βu(t)v(t)δw(t),dv(t)dt=Nδw(t)cv(t), (1.1)

    where u(t), w(t), and v(t) are the numbers of uninfected cells, productively infected cells, and virus particles at time t, respectively. See the references for the biological meanings of the parameters. This basic model has been modified to study different viral infections, which include hepatitis C virus (HCV) [3,4], human immunodeficiency virus (HIV) [5,6,7,8], human T-cell leukemia virus (HTLV) [9,10,11], and so on.

    During the process of viral infection, specific immune response plays an important role. Specific immune response includes cell-mediated immunity (which depends on cytotoxic T lymphocytes response (CTLs)) and humoral immunity (which depends on antibody response). Since the work of Nowak and Bangham [12], much has been done on mathematical models on immune response against infected cells [13,14,15,16].

    Nowadays, time delays have been taken into account in order to better understand viral dynamics. Usually, distributed time delays [17,18,19] and discrete time delays [20,21,22] have been incorporated into viral dynamic models. In particular, based on (1.1), Zhu and Zou [20] proposed the following viral dynamic model with time delay and CTL immune response,

    {du(t)dt=sdu(t)βu(t)v(t),dw(t)dt=βemˆτu(tˆτ)v(tˆτ)δw(t)pw(t)z(t),dv(t)dt=Nδw(t)cv(t),dz(t)dt=qw(t)z(t)bz(t), (1.2)

    where z(t) denotes the density of immune effectors at time t. Here the delay ˆτ represents the time from a virus entering a target cell to the production of new free virus particles. We refer to [20] for the meanings of the other parameters.

    Note that both models (1.1) and (1.2) and most existing ones are described by ordinary differential equations. The cells and free virus particles are assumed to be uniform in location. In other words, the effect of spatial heterogeneity is ignored. For example, the lymphoid tissues are among the primary sites of HIV infection and replication. The lymphoid tissues consist of many lymph nodes with different sizes. The different tissue architecture and composition and biophysical parameters can influence the spread and replication of the virus [23]. To understand the viral pathogenesis better, it is necessary to consider the spatial aspects of the tissues. In [24], Wang and Wang proposed the following mathematical model of HBV infection with spatial dependence,

    {u(x,t)t=sdu(x,t)βu(x,t)v(x,t),w(x,t)t=βu(x,t)v(x,t)δw(x,t),v(x,t)t=DΔv(x,t)+Nδw(x,t)cv(x,t), (1.3)

    where u(x,t), w(x,t), and v(x,t) are the densities of uninfected cells, productively infected cells, and free virus particles at spatial position x and time t, respectively. D is the diffusion coefficient and Δ is the Laplace operator. Using the geometric singular perturbation method, they studied the existence of traveling waves. Since then a lot of works have followed in this direction (see, for example, [25,26,27,28,29]).

    In (1.1), (1.2), and (1.3), only the cell-free transmission (newly released free virus particles infect uninfected cells [2]) is considered. Recent experimental studies [30,31] prove that a healthy cell can be infected when it comes with close contact of an infected cell (cell-to-cell transmission [32,33]). Sigal et al. [34] found that the cell-to-cell spread of HIV can still permit ongoing replication even with an antiretroviral therapy. Consequently, viral dynamic models incorporating both transmission modes have been formulated and studied (to name a few, see [35,36,37,38,39]). We should mention that the incidences in these works are bilinear. Incidence is the number of new infections per unit of time. It depends on the infectivity of viruses and behavior of cells. Thus it is reasonable to be nonlinear in general. For example, the saturated incidence rate βuv1+αv is used in [40] and the Beddington-DeAngelis incidence function is used in [41]. In a recent work, Sun and Wang [42] also used a general incidence f(u,v) in a diffusive viral dynamic model.

    Based on the above discussion, in this paper, we propose and study the following delayed diffusive viral dynamic model with cell-mediated immunity, cell-to-cell transmission, and general incidences,

    {u(x,t)t=sdu(x,t)f(u(x,t),v(x,t))g(u(x,t),w(x,t)),w(x,t)t=emτf(u(x,tτ),v(x,tτ))+emτg(u(x,tτ),w(x,tτ))δw(x,t)pw(x,t)z(x,t),v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t), xΩ, t>0,z(x,t)t=D2Δz(x,t)+qw(x,t)z(x,t)bz(x,t), xΩ, t>0, (1.4)

    where z(x,t) denotes the densities of immune effectors at spatial position x and time t. Ω is a general open bounded domain in Rn with smooth boundary Ω. We consider model (1.4) with the homogeneous Neumann boundary conditions

    vn=0, zn=0, xΩ, t>0, (1.5)

    where n denotes the outward normal derivative on Ω. We also assume the initial conditions

    u(x,θ)=ϕ1(x,θ)0,w(x,θ)=ϕ2(x,θ)0,v(x,θ)=ϕ3(x,θ)0,z(x,θ)=ϕ4(x,θ)0,(x,θ)¯Ω×[τ,0], (1.6)

    where ϕi's (i=1, 2, 3, 4) are bounded and uniformly continuous functions on ¯Ω×[τ,0].

    In (1.4), intracellular delays for both transmission modes are assumed to be the same. In general, the intracellular delay in the cell-to-cell transmission is less than that in the cell-free infection [30,39]. However, the difference is not large enough. As for some existing studies (for example, [35,38]), for simplicity of presentation, we make the above assumption on the delays.

    In (1.4), the incidences due to the cell-free transmission and the cell-to-cell transmission are given by the nonlinear functions f(u,v) and g(u,w), respectively. As in [26], we always make the following assumption on them in the sequel.

    (A1) The nonlinear incidence functions f and g satisfy the following properties.

    (i) f(u,v)0 and g(u,w)0 for u0, v0, and w0, and the equalities hold if and only if uv=0 and uw=0;

    (ii) There exists η1>0 and η2>0 such that f(u,v)η1u and g(u,w)η2u for u0, v0, and w0;

    (iii) f(u,v)u and g(u,w)u are continuous with f(u,v)u>0 and g(u,w)u>0 for u0, v>0, and w>0;

    (iv) f(u,v)v and g(u,w)w are continuous with f(u,v)v0 and g(u,w)w0 for u0, v0, and w0;

    (v) vf(u,v)vf(u,v)0 and wg(u,w)wg(u,w)0 for u0, v0, and w0.

    Note that, by Assumption (A1), for any u>0,

    (f(u,v)v)v=f(u,v)vvf(u,v)v2<0,

    which implies that f(u,v)v is decreasing on (0,). In particular, for any u>0 and v>0,

    f(u,v)vlimv0+f(u,v)v=f(u,0)v.

    An analog also holds for g. Thus we have

    f(u,v)f(u,0)vv and g(u,w)g(u,0)ww for u0, v0, and w0. (1.7)

    The rest of the paper is organized as follows. In section 2, we consider the existence, uniqueness, positivity, and boundedness of solutions to system (1.4)–(1.6). Then we study the existence of homogeneous steady states in section 3, which depend on the basic reproduction number of infection and the basic reproduction number of immunity. The main part is section 4, where we discuss the local and global dynamics of system (1.4)–(1.6) by analyzing the characteristic equations and constructing suitable Lyapunov functionals. These results are supported with numerical simulations in section 5. The paper ends with a brief conclusion.

    Let X:=C(¯Ω,R4) be the Banach space equipped with the supremum norm X. For τ0, define C=C([τ,0],X), which is a Banach space equipped with the norm ϕ=maxθ[τ,0]ϕ(θ)X. If σ>0 and U:[τ,σ)X, then for t[0,σ), UtC is defined by Ut(θ)=U(t+θ) for θ[τ,0]. Denote X+=C(¯Ω,R4+) and C+=C([τ,0],X+). Then both (X,X+) and (C,C+) are strongly ordered spaces. According to Corollary 4 in [43], we have the following result on the well-posedness. The arguments are standard and hence are omitted here. Interested readers can refer to, for example, a recent paper by Gao and Wang [44].

    Theorem 2.1. For each ϕ=(ϕ1,ϕ2,ϕ3,ϕ4)C+, system (1.4)–(1.6) has a unique solution U(,t,ϕ)=(u(,t,ϕ),w(,t,ϕ),v(,t,ϕ),z(,t,ϕ)) on [0,) with U0(,ϕ)=ϕ. Moreover, Ut(,ϕ)C+ for t0 and U(,t,ϕ) is a classical solution.

    Let Φ(t):C+C+ be the solution semiflow associated with (1.4)–(1.6), that is, Φ(t,ϕ)=Ut(,ϕ), where U(,t,ϕ) is the solution of (1.4)–(1.6) with the initial condition ϕC+.

    The following result gives some properties of solutions.

    Lemma 2.2. For ϕC+, the following statements hold for the solution U(,t,ϕ) of (1.4)–(1.6).

    (i) lim suptu(x,t,ϕ)sd, lim suptw(x,t,ϕ)emτ(η1+η2)sdδ, lim suptv(x,t,ϕ)Nemτ(η1+η2)sdc, and lim suptz(x,t,ϕ)emτ(η1+η2)sdmin{δ,b} uniformly for all xΩ.

    (ii) u(,t,ϕ)>0 for t>0 and lim inftu(x,t,ϕ)sd+η1+η2 uniformly for all xΩ.

    (iii) If w(,t0,ϕ)0 or v(,t0,ϕ)0 for some t00, then w(x,t,ϕ)>0 and v(x,t,ϕ)>0 for all xΩ and t>t0+τ.

    (iv) If z(,t0,ϕ)0 for some t00, then z(x,t,ϕ)>0 for all xΩ and t>t0.

    Proof. For simplicity of notation, in the proof here we omit ϕ from the expressions of the solution.

    (ⅰ) First, we have

    u(x,t)tsdu(x,t),

    which implies that lim suptu(x,t)sd uniformly for all xΩ. Next, by Assumption (A1) and the second equation in (1.4), we have

    w(x,t)temτ(η1+η2)u(x,t)δw(x,t).

    Then lim suptw(x,t)emτ(η1+η2)sdδ uniformly for xΩ follows easily from this and lim suptu(x,t)sd uniformly for xΩ. Similarly, adding the second and fourth equations of (1.4) yields

    (w(x,t)+z(x,t))temτ(η1+η2)u(x,t)min{δ,b}(w(x,t)+z(x,t)).

    It follows that lim supt(w(x,t)+z(x,t))emτ(η1+η2)sdmin{δ,b} uniformly for xΩ and hence lim suptz(x,t)emτ(η1+η2)sdmin{δ,b} uniformly for xΩ. Now, lim suptw(x,t)emτ(η1+η2)sdδ uniformly for xΩ together with the third equation of (1.4) (Lemma 1 in [45]), and comparison theorem, gives lim suptv(x,t)Nemτ(η1+η2)sdc uniformly for xΩ.

    (ⅱ) Noting that u(x,t)ts(d+η1+η2)u(x,t), one can easily get

    u(x,t)e(d+η1+η2)tu(x,0)+sd+η1+η2e(d+η1+η2)tsd+η1+η2

    for xΩ and t0. Then (ⅱ) follows immediately.

    (ⅲ) Note that the operator D1ΔcId generates a positive semigroup on C(¯Ω,R), where Id is the identity operator. Thus if w(,t0)0, then from the third equation of (1.4), we see that v(,t)0 for t>t0. Without loss of generality, we assume that v(,t0)0. We first show that v(,t)>0 for t>t0. By Theorem 2.1, v(x,t) satisfies

    {v(x,t)tD1Δv(x,t)cv(x,t),t>t0, xΩ,v(x,t)n=0,t>t0, xΩ.

    Let ˉv(x,t) be the solution of

    {ˉv(x,t)t=D1Δˉv(x,t)cˉv(x,t),t>t0, xΩ,ˉv(x,t)n=0,t>t0, xΩ,ˉv(x,t0)=v(x,t0), x¯Ω.

    Then ˉv(x,t)>0 for xΩ and t>t0. In fact, suppose, by contradiction, there exist x0Ω and ˆt>t0 such that ˉv(x0,ˆt)=0. Then, according to the strong maximum principle [46], ˉv(x,t)0 for each tt0, contradicting with ˉv(,t0)0. Applying the comparison theorem, we know that v(x,t)ˉv(x,t)>0 for t>t0 and xΩ. We now prove that w(x,t)>0 for xΩ and t>t0+τ. Otherwise, there exist ˉxΩ and ˉt>t0+τ such that w(ˉx,ˉt)=0. As w(x,t)0, we have w(ˉx,ˉt)t=0. This is impossible since

    w(ˉx,ˉt)t=emτf(u(ˉx,ˉtτ),v(ˉx,ˉtτ))+emτg(u(ˉx,ˉtτ),w(ˉx,ˉtτ))>0

    by Assumption (A1) (ⅱ) due to u(ˉx,ˉtτ)>0 and v(ˉx,ˉtτ)>0. This proves statement (ⅲ).

    (ⅳ) The proof is similar to that of (ⅲ) on v(x,t)>0 for xΩ and t>t0 and hence we omit it here. This completes the proof.

    Lemma 2.2 tells us that Φ is point dissipative. Then it follows from Theorem 2.1.8 in [47] that Φ(t) is compact for all t>τ. This, together with Theorem 3.4.8 in [48], gives the following result.

    Theorem 2.3. The semiflow Φ has a global compact attractor A in C+. Moreover, u(x,t,ϕ)sd for all x¯Ω, t0, and ϕA.

    System (1.4) with (1.5) always has a unique infection-free steady state P0=(u0,0,0,0), where u0=s/d. Applying the result of Wang and Zhao (Theorem 3.4 in [49]), we can obtain the expression of the basic reproduction number of infection, R0, which is given by

    R0=Nemτcf(sd,0)v+emτδg(sd,0)w.

    Denote

    R01=Nemτcf(sd,0)vandR02=emτδg(sd,0)w.

    Then R01 is the number of secondly infected cells through the cell-free transmission and it is referred to as the basic reproduction number from the cell-free transmission; while R02 is the number of secondly infected cells through the cell-to-cell transmission and it is referred to as the basic reproduction number from the cell-to-cell transmission [38].

    In the following, we discuss the existence of homogeneous steady states for (1.4) (the stability results in section 4 indicate that they are the only possible steady states). Clearly, a homogeneous steady state P=(u,w,v,z) satisfies

    sduf(u,v)g(u,w)=0, (3.1a)
    f(u,v)emτ+g(u,w)emτδwpwz=0, (3.1b)
    Nδwcv=0, (3.1c)
    qwzbz=0. (3.1d)

    It follows from (3.1d) that z=0 (which corresponds to the immunity-free infected steady states) or w=bq (which, when z0, corresponds to the infected-immune steady states).

    We firstly consider the case where z=0. It follows from (3.1c) that v=Nδwc. Multiplying both sides of (3.1b) by emτ and then adding up the resultant and (3.1a) to get u=sδwemτd. It is necessary that w(0,sδemτ). Substituting u=sδwemτd and v=Nδwc into (3.1a), we see that w is a positive zero of H1, where

    H1(w)=f(sδwemτd,Nδwc)+g(sδwemτd,w)δwemτ. (3.2)

    According to Assumption (A1), we have H1(0)=0, H1(sδemτ)s<0, and

    H1(0)=δemτ(Nemτcf(sd,0)v+emτδg(sd,0)w)δemτ=δemτ(R01).

    If R0>1, then H1(0)>0. This, together with H1(0)=0, implies that H1(w) is positive for all sufficiently small w>0. By the Intermediate Value Theorem, H1 has at least one zero in (0,sδemτ) and hence (1.4) has at least one immunity-free infected steady state. In fact, there is only one such steady state by the claim that H1(w1)<0 for any immunity-free infected steady state, which is proved as follows. Note that δemτ=f(u1,v1)w1+g(u1,w1)w1 and w1=cv1Nδ. By Assumption (A1),

    H1(w1)=δemτdf(u1,v1)u+Nδcf(u1,v1)vδemτdg(u1,w1)u+ g(u1,w1)wNδcv1f(u1,v1)1w1g(u1,w1)=δemτdf(u1,v1)uδemτdg(u1,w1)u+ Nδcv1(v1f(u1,v1)vf(u1,v1))+ 1w1(w1g(u1,w1)wg(u1,w1))<0.

    This proves the claim. Next, we assume that R0<1. Then H1(0)=δemτ(R01)<0, which combined with H1(0)=0 implies that H1(w)<0 for w>0 sufficiently small. Using the above claim, we can easily see that there is no immunity-free infected steady state when R0<1. Moreover, H1(w)<0 for w(0,sδemτ]. Finally, we assume that R0=1. We use contradictive arguments to show that there is no immunity-free infected steady state in this case. Otherwise, assume that H1(w) has a positive zero say w. Then from the above claim H1(w)>0 for w<w and closely enough to w. Note that H1(w) depends continuously on the parameters and H1(w)<0 for w(0,sδemτ] when R0<1. Fix w(0,w). Choose a sequence of parameters such that the basic reproduction number R0<1 and tends to 1. Then H1(w) tends to H1(w)>0, a contradiction to the fact that the limit is less than or equal to 0. This proves that there is no immunity-free infected steady state when R0=1.

    Now we study the case where w=bq. This, together with (3.1c), yields v=Nδbcq. As before, add up (3.1a) and (3.1b) multiplied by emτ to get z=sduδemτwpemτw, which necessarily requires u(0,sdδbdqemτ). Substituting w=bq and z=sduδemτwpemτw into (3.1a), we see that u is a positive zero of H2, where

    H2(u)=f(u,Nδbcq)+g(u,bq)s+du. (3.3)

    With Assumption (A1), we have H2(0)=s<0 and

    H2(u)=f(u,Nδbcq)u+g(u,bq)u+d>0.

    Therefore, in order for model (1.4) to have an infected-immune steady state (if exists there is a unique one), it is necessary and sufficient that H2(sdδbdqemτ)=H1(bq)>0. Recall that when R01, H1(w)<0 for w(0,sδemτ]; while when R0>1, H1(w)>0 for w(0,w1) and H1(w)<0 for w(w1,sδemτ). It follows that H1(bq)>0 if and only if R0>1 and bq<w1. Denote,

    R1=qw1b.

    As q is the average number of immune effectors produced from contacting with a productively infected cell and 1b is the average life of an immune effector, it follows that R1 is the total number of immune effectors produced at the immunity-free infected steady state. Thus R1 is called the basic reproduction number of immunity.

    Summarizing the above discussion, we have obtained the following result on the existence of homogeneous steady states.

    Theorem 3.1. For model (1.4) with (1.5), the following statements on the existence of homogeneous steady states are true.

    (i) If R01, then there is only the infection-free steady state P0.

    (ii) If R11<R0, then besides P0, there is also a unique immunity-free infected steady state P1=(u1,w1,v1,0), where w1 is the only positive zero of H1 defined by (3.2), u1=sδw1emτd and v1=Nδw1c.

    (iii) If R1>1 (it is necessary that R0>1), then in addition to P0 and P1, there is also a unique infected-immune steady state P2=(u2,w2,v2,z2), where u2 is the only positive zero of H2 defined by (3.3), w2=bq, v2=Nδbcq, and z2=sdu2δw2emτpw2emτ.

    In the main part of this paper, we establish the stability of each steady state obtained in Theorem 3.1.

    Let P=(u,w,v,z) be an arbitrary homogeneous steady state. The linearization of (1.4) at P is

    Qt=LΔQ+AQ+BQτ, (4.1)

    where

    L=diag(0,0,D1,D2),Q=(u(x,t),w(x,t),v(x,t),z(x,t)),Qτ=(uτ,wτ,vτ,zτ)=(u(x,tτ),w(x,tτ),v(x,tτ),z(x,tτ)),A=(df(u,v)ug(u,w)ug(u,w)wf(u,v)v00δpz0pw0Nδc00qz0qwb),B=(0000(f(u,v)u+g(u,w)u)emτg(u,w)wemτf(u,v)vemτ000000000).

    Denote 0=μ0<μ1<μ2<<μn< to be all the eigenvalues of the operator Δ on Ω with the homogeneous Neumann boundary condition. Then P is locally asymptotically stable if, for any iN={0,1,2,}, every solution of the characteristic equation

    |λE+μiLABeλτ|=0 (4.2)

    has a negative real part and P is unstable if there exists i0N such that (4.2) has a solution with a positive real part.

    We first study the local stability of P0.

    Proposition 4.1. The infection-free steady state P0 of (1.4) is locally asymptotically stable if R0<1 and unstable if R0>1.

    Proof. By (4.2), the characteristic equation at P0 is

    (λ+b+μiD2)(λ+d)[(λ+c+μiD1)(λ+δg(sd,0)we(λ+m)τ)Nδf(sd,0)ve(λ+m)τ]=0.

    Obviously, the stability of P0 is determined by

    (λ+c+μiD1)(λ+δg(sd,0)wemτeλτ)Nδf(sd,0)vemτeλτ=0. (4.3)

    Firstly, suppose that R0<1. We claim that all roots of (4.3) have negative real parts. Otherwise, there exists i0N such that (4.3) has a root λ0 with Re(λ0)0. Then

    1=1λ0+δg(sd,0)wemτeλ0τ+Nδ(λ0+δ)(λ0+c+μi0D1)f(sd,0)vemτeλ0τ.

    It follows that

    1=|1λ0+δg(sd,0)wemτeλ0τ+Nδ(λ0+δ)(λ0+c+μi0D1)f(sd,0)vemτeλ0τ||1λ0+δg(sd,0)wemτeλ0τ|+|Nδ(λ0+δ)(λ0+c+μi0D1)f(sd,0)vemτeλ0τ|1δg(sd,0)wemτ+Nc+μi0D1f(sd,0)vemτ1δg(sd,0)wemτ+Ncf(sd,0)vemτ=R0,

    a contradiction to R0<1. This proves the claim and hence P0 is locally asymptotically stable when R0<1.

    Secondly, assume R0>1. For iN, denote

    F(λ,i)=(λ+c+μiD1)(λ+δg(sd,0)wemτeλτ)Nδf(sd,0)vemτeλτ.

    Recall that μ0=0. We have

    F(0,0)=cδ(1R0)<0

    and

    F(λ,0)=λ2+(c+δ)λ+cδ(cδR0+g(sd,0)wemτ)eλτ as λ.

    By the Intermediate Value Theorem, F(λ,0) has a positive zero and hence (4.3) has at least one positive zero for i=0. This means that P0 is unstable when R0>1.

    In fact, P0 is globally stable if it is locally stable.

    Theorem 4.2. If R01, then the infection-free steady state P0 of (1.4) is globally attractive. In particular, P0 is globally asymptotically stable when R0<1.

    Proof. It suffices to show that P0 is globally attractive in A. For this purpose, we consider the Lyapunov functional

    W(t)=Ω(emτw(x,t)+1Nemτv(x,t)+pqemτz(x,t))dx+Ω(ttτf(u(x,θ),v(x,θ))dθ+tτg(u(x,θ),w(x,θ))dθ)dx.

    Calculating the time derivative of W(t) along solutions of model (1.4), we have

    dW(t)dt=Ω(f(u(x,t),v(x,t))cNemτv(x,t)pbqemτz(x,t))dx+1NemτΩD1Δv(x,t)dx.

    It follows from the homogeneous Neumann boundary condition (1.5) and the Divergence Theorem that

    ΩΔv(x,t)dx=Ωv(x,t)ndx=0.

    Moreover, by Theorem 2.3, u(x,t)sd for xΩ and t0. With the help of (1.7), for v(x,t)>0, we have

    f(u(x,t),v(x,t))cNemτv(x,t)=cNemτv(x,t)(Ncemτf(u(x,t),v(x,t))v(x,t)1)cNemτv(x,t)(Ncemτf(sd,v(x,t))v(x,t)1)cNemτv(x,t)(Ncemτf(sd,0)v1)cNemτv(x,t)(R01).

    The above inequality holds automatically for v(x,t)=0 and also observe that the inequality is strict for u(x,t)<sd and v(x,t)>0. Therefore,

    dW(t)dtΩ(cNemτv(x,t)(R01)pbqemτz(x,t))dx0.

    Moreover, dW(t)dt=0 if and only if v(x,t)=0 and z(x,t)=0. In fact, if v(x0,t0)0, then there exists a neighborhood N(x0,t0) of (x0,t0) such that v(x,t)0 for (x,t)N(x0,t0). Then by the observation, u(x,t)=sd for (x,t)N(x0,t0). This, together with the first equation of (1.4) and Assumption (A1), implies that v(x,t)=0 for (x,t)N(x0,t0), a contradiction. Then it is easy to see that the largest invariant subset of dW(t)dt=0 is {P0}. By LaSalle's Invariance Principle (see Theorem 5.3.1 in [50] or Theorem 3.4.7 in [51]), the infection-free steady state P0 is globally attractive. In particular, this together with Proposition 4.1, tells us that P0 is globally asymptotically stable when R0<1.

    Next we consider the stability of the immunity-free infected steady state P1. For convenience of notations, denote

    f1=f(u1,v1),g1=g(u1,w1),f1u=f(u1,v1)u,f1v=f(u1,v1)v,g1u=g(u1,w1)u,g1w=g(u1,w1)w.

    Theorem 4.3. Suppose R0>1. Then the immunity-free infected steady state P1 of (1.4) is locally asymptotically stable if R1<1 and unstable if R1>1.

    Proof. From (4.2), we know that the characteristic equation at P1 is given by

    (λqw1+b+μiD2)ρi(λ)=0, iN,

    where

    ρi(λ)=(λ+d+f1u+g1u)(λ+δ)(λ+c+μiD1)[(λ+d)(λ+c+μiD1)g1w+(λ+d)Nδf1v]e(λ+m)τ.

    Clearly, the eigenvalue λ=b(R11)μiD2<0 for iN when R1<1 but when R1>1, with i=0, we have a positive eigenvalue λ=b(R11). Thus P1 is unstable if R1>1. Now, we assume that R1<1. Then the stability of P1 is determined by the roots of ρi(λ)=0, which is equivalent to

    1=λ+dλ+d+f1u+g1u(g1wλ+δe(λ+m)τ+Nδf1v(λ+δ)(λ+c+μiD1)e(λ+m)τ). (4.4)

    We claim that all solutions of (4.4) have negative real parts. Otherwise, suppose that there exists i1N such that (4.4) has a solution λ1 with Re(λ1)0. Then

    1=|λ1+dλ1+d+f1u+g1u(g1wλ1+δemτeλ1τ+Nδf1v(λ1+δ)(λ1+c+μi1D1)emτeλ1τ)|<|g1wλ1+δemτeλ1τ|+|Nδf1v(λ1+δ)(λ1+c+μi1D1)emτeλ1τ|<g1wδemτ+Nf1vcemτ. (4.5)

    However, from the steady state Eqs (3.1b) and (3.1c), we have

    g(u1,w1)δw1emτ+Nf(u1,v1)cv1emτ=1.

    This and Assumption (A1) (v) together give us

    g1wδemτ+Nf1vcemτg(u1,w1)δw1emτ+Nf(u1,v1)cv1emτ=1,

    which is a contradiction with (4.5). This proves the claim and hence P1 is locally asymptotically stable when R1<1<R0.

    Before studying the global stability of P1, we establish the persistence of infection.

    From the linearized system at P0 (see (4.1)), we have the following cooperative system for (w,v),

    {w(x,t)t=emτf(u0,0)vv(x,tτ)+emτg(u0,0)ww(x,tτ)δw(x,t),v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t). (4.6)

    With similar arguments as those for Lemma 3 and Lemma 4 in Lou and Zhao [45], we can obtain the following results.

    Lemma 4.4. There exists a principal eigenvalue ˉλ(u0,τ)ˉλ(P0,τ) of (4.6) associated with a strongly positive eigenvector. Moreover, ˉλ(u0,τ) has the same sign as λ(u0)ˉλ(u0,0).

    Lemma 4.5. R01 and λ(u0) have the same sign.

    Theorem 4.6. Suppose R11<R0. Then the infection is persistent, that is, there exists ε>0 such that

    lim inftu(x,t,ϕ)ε,lim inftw(x,t,ϕ)ε,lim inftv(x,t,ϕ)ε

    uniformly for all x¯Ω, where ϕW1:={ϕC+:w(,0)0 and v(,0)0}.

    Proof. Define

    W1:=C+W1={ϕC+:w(,0)0 or v(,0)0}.

    By Lemma 2.2 and the second equation of (1.4), we know that Φ(t)W1W1 for all t0. Denote

    M:={ϕW1:Φ(t)ϕW1 for t0}.

    Claim 1. ω(ϕ)={(u0,0,0,0)} for ϕM, where ω(ϕ) is the omega limit set of the orbit O+(ϕ):={Φ(t)ϕ:t0}.

    Since ϕM, for all t0, either w(x,t,ϕ)0 or v(x,t,ϕ)0. If w(x,t,ϕ)0 for all t0, then limtv(x,t,ϕ)=0 uniformly for x¯Ω from the third equation of (1.4). Now, suppose that w(x,t1,ϕ)0 for some t10. Then by Lemma 2.2, w(x,t,ϕ)>0 for all tt1+τ and xΩ. Thus v(x,t,ϕ)0 for all tt1+τ. This, combined with the third equation of (1.4), implies that w(x,t,ϕ)0 for x¯Ω and tt1+τ. Then, in either case, limtv(x,t,ϕ)=limtw(x,t,ϕ)=0 uniformly for x¯Ω. Thus u is asymptotic to

    u(x,t)t=sdu(x,t).

    By Corollary 4.3 in [52], we get limtu(x,t,ϕ)=u0 uniformly for x¯Ω. The above discussion tells us that w(x,t,ϕ)0 for all t large enough. Then we can easily see from the fourth equation of (1.4) that limtz(x,t,ϕ)=0 uniformly for x¯Ω. This proves ω(ϕ)={(u0,0,0,0)}.

    Since R11<R0, by Lemma 4.4 and Lemma 4.5, there exists a sufficiently small ε0>0 such that the following linear system

    {w(x,t)t=emτ(f(u0,0)vε0)v(x,tτ)+emτ(g(u0,0)wε0)w(x,tτ)(δ+pε0)w(x,t),v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t)

    has a positive principal eigenvalue ˉλ(u0ε0) with positive eigenfunction (wε0,vε0). By the continuity in Assumption (A1), there exists δ0(0,ε0] such that

    f(u,v)vf(u0,0)vε0andg(u,w)wg(u0,0)wε0

    for all u0δ0uu0+δ0, 0vδ0, and 0wδ0.

    Claim 2. {(u0,0,0,0)} is a uniform weak repeller for W1 in the sense that

    lim suptΦ(t)ϕ(u0,0,0,0)δ0 for ϕW1.

    Suppose, by contradiction, there exists ϕ1W1 such that lim suptΦ(t)ϕ1(u0,0,0,0)<δ0. Then there exists t2>0 such that u(x,t,ϕ1)>u0δ0u0ε0, w(x,t,ϕ1)δ0, and v(x,t,ϕ1)δ0 for tt2 and x¯Ω. It follows from Assumption (A1) and the choice of δ0 that w and v satisfy

    {w(x,t)temτ(f(u0,0)vε0)v(x,tτ)+emτ(g(u0,0)wε0)w(x,tτ)(δ+pε0)w(x,t),tt2, xΩ,v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t),tt2, xΩ,v(x,t)n=0,tt2,xΩ. (4.7)

    Due to w(x,t,ϕ1)>0 and v(x,t,ϕ1)>0 for t>0 and xΩ, there exists κ1>0 such that (w(x,t2+θ,ϕ1),v(x,t2+θ,ϕ1))κ1eˉλ(u0ε0)(t2+θ)(wε0(x),vε0(x)) for all xΩ and θ[τ,0]. Then it follows from the comparison principle that w(x,t,ϕ1)κ1eˉλ(u0ε0)twε0(x) and v(x,t,ϕ1)κ1eˉλ(u0ε0)tvε0(x) for all xΩ and tt2, a contradiction to the fact that both w(x,t,ϕ1) and v(x,t,ϕ1) are bounded. This proves Claim 2.

    Define a continuous function P1:C+[0,) by

    P1(ϕ):=min{minx¯Ωϕ2(x,0), minx¯Ωϕ3(x,0)} for all ϕC+.

    Clearly, P11(0,)W1, and P1 has the property that if P1(ϕ)=0 and ϕW1 or P1(ϕ)>0, then P1(Φ(t)ϕ)>0 for all t>0. Hence, P1 is a generalized distance function for the semiflow Φ(t) [53]. According to the above discussions, we obtain that any forward orbit of Φ(t) in M converges to (u0,0,0,0), which is isolated in C+ and Ws(u0,0,0,0)W1=, where Ws(u0,0,0,0) is the stable manifold of (u0,0,0,0). Moreover, there is no cycle in W1 from (u0,0,0,0) to (u0,0,0,0). Applying Theorem 3 in [53], we know that there exists an ˉε>0 such that min{P1(ϕ)}>ˉε for any ϕW1. It follows that

    lim inftw(x,t)ˉε and lim inftv(x,t)ˉε uniformly for all x¯Ω.

    This combined with Lemma 2.2 finishes the proof with ε=min{ˉε,sd+η1+η2}.

    In order to study the global stability of P1, define G:(0,)xx1lnx. Obviously, G(x)>0 for x(0,) and G attains its global minimum only at x=1. We also need the following assumption.

    (A2) The nonlinear incidence functions f(u,v) and g(u,w) satisfy the following conditions.

    (i) For any u>0,

    {vv1u1f(u,v)uf(u1,v1)<1if  0<v<v1,1u1f(u,v)uf(u1,v1)<vv1if  v1<v.

    (ii) For any u>0,

    {ww1u1g(u,w)ug(u1,w1)<1if  0<w<w1,1u1g(u,w)ug(u1,w1)<ww1if  w1<w.

    Theorem 4.7. Suppose that R11<R0 and Assumption (A2) are satisfied. Then the immunity-free infected steady state P1 is globally attractive in

    C+1={ϕC+|there exists t3R+ such that w(,t3,ϕ)0 or v(,t3,ϕ)0}.

    In particular, it is globally asymptotically stable in C+1 if further R1<1.

    Proof. According to Lemma 2.2 and Theorem 4.6, we know that there exists ε>0 such that lim inftu(x,t,ϕ)ε, lim inftw(x,t,ϕ)ε, and lim inftv(x,t,ϕ)ε for ϕC+1. Without loss of generality, we define a Lyapunov functional

    L(t)=ΩL(x,t)dx,

    where

    L(x,t)=u1G(u(x,t)u1)+emτw1G(w(x,t)w1)+f(u1,v1)cv1v1G(v(x,t)v1)+pqemτz(x,t)+f(u1,v1)ttτG(f(u(x,θ),v(x,θ))f(u1,v1))dθ+g(u1,w1)ttτG(g(u(x,θ),w(x,θ))g(u1,w1))dθ.

    Calculating the time derivative of L(x,t) along solutions of (1.4) yields

    L(x,t)t=(1u1u(x,t))[sdu(x,t)f(u,v)g(u,w)]+pqemτ[qw(x,t)z(x,t)bz(x,t)]+emτ(1w1w(x,t))[emτf(uτ,vτ)+emτg(uτ,wτ)δw(x,t)pw(x,t)z(x,t)]+f(u1,v1)cv1(1v1v(x,t))[D1Δv(x,t)+Nδw(x,t)cv(x,t)]+pqemτD2Δz(x,t)+f(u,v)f(uτ,vτ)+f(u1,v1)lnf(uτ,vτ)f(u,v)+g(u,w)g(uτ,wτ)+g(u1,w1)lng(uτ,wτ)g(u,w).

    With

    s=du1+f(u1,v1)+g(u1,w1), δ=emτf(u1,v1)+emτg(u1,w1)w1, c=Nδw1v1,

    we have

    L(x,t)t=du1(2u1u(x,t)u(x,t)u1)+pqemτ(R11)z(x,t)+f(u1,v1)cv1(1v1v(x,t))D1Δv(x,t)+pqemτD2Δz(x,t)f(u1,v1)w(x,t)v1v(x,t)w1+f(u1,v1)[3u1u(x,t)f(uτ,vτ)w1f(u1,v1)w(x,t)+f(u,v)u1f(u1,v1)u(x,t)v(x,t)v1+lnf(uτ,vτ)f(u,v)]+g(u1,w1)[2u1u(x,t)g(uτ,wτ)w1g(u1,w1)w(x,t)+g(u,w)u1g(u1,w1)u(x,t)w(x,t)w1+lng(uτ,wτ)g(u,w)]=du1(2u1u(x,t)u(x,t)u1)+pqemτ(R11)z(x,t)+f(u1,v1)cv1(1v1v(x,t))D1Δv(x,t)+pqemτD2Δz(x,t)f(u1,v1)G(w(x,t)v1v(x,t)w1)+f(u1,v1)[G(f(u,v)u1f(u1,v1)u(x,t))G(v(x,t)v1)G(u1u(x,t))G(f(uτ,vτ)w1f(u1,v1)w(x,t))]+g(u1,w1)[G(g(u,w)u1g(u1,w1)u(x,t))G(w(x,t)w1)G(u1u(x,t))G(g(uτ,wτ)w1g(u1,w1)w(x,t))].

    Clearly,

    Ωdu1(2u(x,t)u1u1u(x,t))dx0

    and

    Ω(pbqemτ(R11)z(x,t))dx0 since R11.

    Using the Divergence Theorem and the homogeneous Neumann boundary conditions of (1.5), we have

    ΩΔv(x,t)dx=Ωv(x,t)ndx=0,ΩΔz(x,t)dx=Ωz(x,t)ndx=0,

    and

    0=Ω1v(x,t)v(x,t)  n dx=Ω(1v(x,t)v(x,t))dx=Ω(1v(x,t)Δv(x,t)1v2(x,t)v(x,t)2)dx.

    The latter gives

    Ω1v(x,t)Δv(x,t)dx=Ω1v2(x,t)v(x,t)2dx0

    and hence

    Ωf(u1,v1)cv1(1v1v(x,t))D1Δv(x,t)dx0.

    To summarize, we have obtained

    dL(t)dtf(u1,v1)Ω[G(f(u,v)u1f(u1,v1)u(x,t))G(v(x,t)v1)]dx+g(u1,w1)Ω[G(g(u,w)u1g(u1,w1)u(x,t))G(w(x,t)w1)]dxg(u1,w1)Ω[G(u1u(x,t))+G(g(uτ,wτ)w1g(u1,w1)w(x,t))]dxf(u1,v1)Ω[G(u1u(x,t))+G(f(uτ,vτ)w1f(u1,v1)w(x,t))+G(w(x,t)v1v(x,t)w1)]dx.

    Note that the monotonicity of G(x) on each side of x=1 and Assumption (A2) give us

    G(f(u,v)u1f(u1,v1)u(x,t))G(v(x,t)v1)andG(g(u,w)u1g(u1,w1)u(x,t))G(w(x,t)w1).

    Thus

    dL(t)dtg(u1,w1)Ω[G(u1u(x,t))+G(g(uτ,wτ)w1g(u1,w1)w(x,t))]dxf(u1,v1)Ω[G(u1u(x,t))+G(f(uτ,vτ)w1f(u1,v1)w(x,t))+G(w(x,t)v1v(x,t)w1)]dx0.

    Moreover, dL(t)dt=0 if and only if u(x,t)=u1, w(x,t)=w1, v(x,t)=v1, and z(x,t)=0. Then the largest invariant subset of dL(t)dt=0 is {P1}. By LaSalle's Invariance Principle (see Theorem 5.3.1 in [50] or Theorem 3.4.7 in [51]), the immunity-free infected steady state P1 is globally attractive in C+1 when R11<R0. This, together with Theorem 4.3, implies that P1 is globally asymptotically stable in C+1 if further R1<1.

    For convenience of notations, denote

    f2=f(u2,v2),g2=g(u2,w2),f2u=f(u2,v2)u,f2v=f(u2,v2)v,g2u=g(u2,w2)u,g2w=g(u2,w2)w.

    Theorem 4.8. If R1>1, then the infected-immune steady state P2 is locally asymptotically stable.

    Proof. According to (4.2), the characteristic equation at P2 is

    0=(λ+d+f2u+g2u)(λ+c+μiD1)(λ+bqw2+μiD2)(λ+δ+pz2)+qw2pz2(λ+c+μiD1)(λ+d+f2u+g2u)(λ+d)(λ+c+μiD1)(λ+bqw2+μiD2)g2we(λ+m)τ(λ+d)(λ+bqw2+μiD2)Nδf2ve(λ+m)τ. (4.8)

    We claim that all roots of (4.8) have negative real parts. Otherwise, suppose for some i2N, it has a root λ2 with Re(λ2)0. Since w2=bq, we have

    1=(λ2+μi2D2)(λ2+d)(λ2+d+f2u+g2u)[(λ2+μi2D2)(λ2+δ+pz2)+pbz2](g2we(λ2+m)τ+Nδf2ve(λ2+m)τλ2+c+μi2D1),

    which implies

    1<|λ2+μi2D2(λ2+μi2D2)(λ2+δ+pz2)+pbz2|×(g2wemτ+Nδf2vemτc).

    With similar arguments as those in the proof of Theorem 4.3, we can obtain

    g2wemτ+Nδf2vemτcδ+pz2.

    Thus we have arrived at

    1<|λ2+μi2D2(λ2+μi2D2)(λ2+δ+pz2)+pbz2|×(δ+pz2),

    which is impossible as one can check that |(λ2+μi2D2)(λ2+δ+pz2)+pbz2|>|(λ2+μi2D2)(δ+pz2)|.

    This completes the proof.

    To establish the global stability of P2, we need the persistence of immunity.

    From the linearized system at P1 (see (4.1)), we have the following cooperative system for (w,v,z),

    {w(x,t)t=emτf(u1,v1)vv(x,tτ)+emτg(u1,w1)ww(x,tτ)(δ+pz1)w(x,t)pw1z(x,t),v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t),z(x,t)t=D2Δz(x,t)+qz1w(x,t)+(qw1b)z(x,t). (4.9)

    With similar arguments as those for Lemma 3 and Lemma 4 in Lou and Zhao [45], we can obtain the following results.

    Lemma 4.9. There exists a principal eigenvalue ˆλ(P1,τ) of (4.9) associated with a strongly positive eigenvector. Moreover, ˆλ(P1,τ) has the same sign as ˆλ(P1,0).

    Lemma 4.10. R11 and ˆλ(P1,0) have the same sign.

    Theorem 4.11. Suppose that R1>1 (it is necessary that R0>1) and (A2) holds. Then the immunity is persistent, that is, there exists ϵ>0 such that

    lim inftu(x,t,ϕ)ϵ,lim inftw(x,t,ϕ)ϵ,lim inftv(x,t,ϕ)ϵ,lim inftz(x,t,ϕ)ϵ

    uniformly for all x¯Ω, where ϕW2:={ϕC+:w(,0)0, v(,0)0, and z(,0)0}.

    Proof. The proof is quite similar to that of Theorem 4.6. Denote

    W2:=C+W2={ϕC+:w(,0)0, or v(,0)0, or z(,0)0}.

    Set M0={P0} and M1={P1}.

    According to Lemma 2.2, we know that w(x,t,ϕ)>0, v(x,t,ϕ)>0, and z(x,t,ϕ)>0 for all t>0 and xΩ, ϕW2, which implies that Φ(t)W2W2 for all t0. Define

    M:={ϕW2:Φ(t)ϕW2 for t0}.

    Claim 3. Let ϕM. Then ω(ϕ)=M0 or M1.

    Sine ϕM, for any t0, we have either w(x,t,ϕ)0, or v(x,t,ϕ)0, or z(x,t,ϕ)0. If z(x,t4,ϕ)0 for some t40, then by Lemma 2.2, z(x,t,ϕ)>0 for t>t4 and xΩ. Then either w(x,t,ϕ)0 or v(x,t,ϕ)0 for each t>t4. By the proof of Claim 1, we know that ω(ϕ)=M0. Now, suppose that z(x,t,ϕ)0 for all t0. If for each t0, either w(x,t,ϕ)0 or v(x,t,ϕ)0, then by Claim 1, ω(ϕ)=M0. If there exists ˜t0 such that w(x,˜t,ϕ)0 and v(x,˜t,ϕ)0. Then by Theorem 4.6, there exists ξ>0 such that

    lim inftw(x,t,ϕ)ξ and lim inftv(x,t,ϕ)ξ uniformly in ¯Ω.

    Now consider the reduced system of (1.4) with z=0. Modifying the Lyapunov functional L(t) in the proof of Theorem 4.7 by ignoring the term pqemτz(x,t) in L(x,t), we can show that the solution of the reduced system converges to (u1,w1,v1) and hence ω(ϕ)=M1. This proves Claim 3.

    Claim 4. Both M0 and M1 are uniform weak repellers for W2. Since W2W1, by Claim 2, M0 is a uniform repeller for W2. The proof of M1 being a uniform repeller of W2 is similar as that of Claim 2 by using Lemma 4.9 and Lemma 4.10. Therefore, we omit the detail here.

    Define a continuous function P2:C+[0,) by

    P2(ϕ):=min{minx¯Ωϕ2(x,0), minx¯Ωϕ3(x,0), minx¯Ωϕ4(x,0)} for ϕC+.

    It is easy to see that P12(0,)W2, and P2 has the property that if P2(ϕ)=0 and ϕW2 or P2(ϕ)>0, then P2(Φ(t)ϕ)>0 for all t>0. Thus P2 is a generalized distance function for the semiflow Φ(t). As M0 and M1 are repellers, we know that both M0 and M1 are isolated, and Ws(Mi)W2= for i=0 and 1. Moreover, no subset of {M0,M1} forms a cycle in W2. By Smith and Zhao [53,Theorem 3], there exists a ˉϵ>0 such that min{P2(ϕ)}>ˉϵ for any ϕW2. Then as for Theorem 4.6, with ε=min{ˉϵ,sd+η1+η2} finishes the proof.

    As for the global stability of P2, we make the following assumption to establish the global stability of P2.

    (A3) The nonlinear incidence functions f(u,v) and g(u,w) satisfy the following conditions.

    (i) For any u>0,

    {vv2u2f(u,v)uf(u2,v2)<1if  0<v<v2,1u2f(u,v)uf(u2,v2)<vv2if  v2<v.

    (ii) For any u>0,

    {ww2u2g(u,w)ug(u2,w2)<1if  0<w<w2,1u2g(u,w)ug(u2,w2)<ww2if  w2<w.

    Theorem 4.12. Suppose that R1>1 and Assumptions (A2) and (A3) are satisfied. Then the infected-immune steady state P2 is globally asymptotically stable in

    C+2={ϕC+|there exists t5R+ such that either w(,t5,ϕ)0 or v(,t5,ϕ)0there exists t6R+ such that z(,t6,ϕ)0}.

    Proof. It follows from Lemma 2.2 and Theorem 4.11 that there exists an ε>0 such that

    lim inftu(x,t,ϕ)ε, lim inftw(x,t,ϕ)ε, lim inftv(x,t,ϕ)ε, lim inftz(x,t,ϕ)ε

    unfiormly in ¯Ω and ϕC+2. Without loss of generality, we define a Lyapunov functional

    I(t)=ΩI(x,t)dx,

    where

    I(x,t)=u2G(u(x,t)u2)+emτw2G(w(x,t)w2)+f(u2,v2)cv2v2G(v(x,t)v2)+pqemτz2G(z(x,t)z2)+f(u2,v2)ttτG(f(u(x,θ),v(x,θ))f(u2,v2))dθ+g(u2,w2)ttτG(g(u(x,θ),w(x,θ))g(u2,w2))dθ.

    Calculate the time derivative of I(x,t) along the solutions of (1.4) to get

    I(x,t)t=(1u2u(x,t))[sdu(x,t)f(u,v)g(u,w)]+pqemτ(1z2z(x,t))[qw(x,t)z(x,t)bz(x,t)]+emτ(1w2w(x,t))[emτf(uτ,vτ)+emτg(uτ,wτ)δw(x,t)pw(x,t)z(x,t)]+f(u2,v2)cv2(1v2v(x,t))[D1Δv(x,t)+Nδw(x,t)cv(x,t)]+pqemτ(1z2z(x,t))D2Δz(x,t)+f(u,v)f(uτ,vτ)+f(u2,v2)lnf(uτ,vτ)f(u,v)+g(u,w)g(uτ,wτ)+g(u2,w2)lng(uτ,wτ)g(u,w).

    With the following relations,

    s=du2+f(u2,v2)+g(u2,w2),δ=emτf(u2,v2)+emτg(u2,w2)pw2z2w2,c=Nδw2v2,w2=bq,

    we get

    I(x,t)t=du2(2u2u(x,t)u(x,t)u2)+f(u2,v2)cv2(1v2v(x,t))D1Δv(x,t)+pqemτ(1z2z(x,t))D2Δz(x,t)f(u2,v2)w(x,t)v2v(x,t)w2+f(u2,v2)[3u2u(x,t)f(uτ,vτ)w2f(u2,v2)w(x,t)+f(u,v)u2f(u2,v2)u(x,t)v(x,t)v2+lnf(uτ,vτ)f(u,v)]+g(u2,w2)[2u2u(x,t)g(uτ,wτ)w2g(u2,w2)w(x,t)+g(u,w)u2g(u2,w2)u(x,t)w(x,t)w2+lng(uτ,wτ)g(u,w)].

    Then

    dI(t)dt=Ωdu2(2u2u(x,t)u(x,t)u2)dx+Ωf(u2,v2)cv2(1v2v(x,t))D1Δv(x,t)dx+Ωpqemτ(1z2z(x,t))D2Δz(x,t)dxf(u2,v2)ΩG(w(x,t)v2v(x,t)w2)dx+f(u2,v2)Ω[G(f(u,v)u2f(u2,v2)u(x,t))G(v(x,t)v2)G(u2u(x,t))G(f(uτ,vτ)w2f(u2,v2)w(x,t))]dx+g(u2,w2)Ω[G(g(u,w)u2g(u2,w2)u(x,t))G(w(x,t)w2)G(u2u(x,t))G(g(uτ,wτ)w2g(u2,w2)w(x,t))]dx.

    Similarly as in the proof of Theorem 4.7, we can show

    Ωdu2(2u(x,t)u2u2u(x,t))dx0,Ωf(u2,v2)cv2(1v2v(x,t))D1Δv(x,t)dx0,Ωpqemτ(1z2z(x,t))D2Δz(x,t)dx0,G(f(u,v)u2f(u2,v2)u(x,t))G(v(x,t)v2),G(g(u,w)u2g(u2,w2)u(x,t))G(w(x,t)w2).

    Therefore, we have dI(t)dt0. Moreover, dI(t)dt=0 if and only if u(x,t)=u2, w(x,t)=w2, v(x,t)=v2, z(x,t)=z2. Then the largest invariant subset of dI(t)dt=0 is {P2}. By LaSalle's Invariance Principle (see Theorem 5.3.1 in [50] or Theorem 3.4.7 in [51]), the infected-immune steady state P2 is globally attractive in C+2. This, together with Theorem 4.8, implies the global asymptotic stability of P2 in C+2.

    In this section, we perform some numerical simulations to illustrate the results obtained in section 4. Let f(u,v)=β1uv1+α1v and g(u,w)=β2uw1+α2w. One can easily verify that f and g satisfy (A1)–(A3). Then the model (1.4) becomes

    {u(x,t)t=sdu(x,t)β1u(x,t)v(x,t)1+α1v(x,t)β2u(x,t)w(x,t)1+α2w(x,t), xΩ, t>0,w(x,t)t=emτ(β1u(x,tτ)v(x,tτ)1+α1v(x,tτ)+β2u(x,tτ)w(x,tτ)1+α2w(x,tτ))δw(x,t)pw(x,t)z(x,t), xΩ, t>0,v(x,t)t=D1Δv(x,t)+Nδw(x,t)cv(x,t), xΩ, t>0,z(x,t)t=D2Δz(x,t)+qw(x,t)z(x,t)bz(x,t), xΩ, t>0, (5.1)

    subject to the homogeneous Neumann boundary conditions

    vn=0, zn=0, xΩ, t>0.

    For (5.1), the basic reproduction number of infection is given by

    R0=Nβ1scdemτ+β2sδdemτ

    and the basic reproduction number of immunity is given by

    R1=qw1b=q(B2+B224B1B3)2bB1,

    where

    B1=Nδ2emτ(β1α2+β2α1+α1α2d)>0,B2=δdemτ(Nδα1+cα2)Nδβ1α2s+Nδ2β1emτNδβ2α1s+δemτβ2c,B3=dcδemτ(R01)<0 since R0>1.

    For simulations, we take α1=0.01, α2=0.01, D1=0.0017, D2=0.0001, and the values of the other parameters are summarized in Table 1. Moreover, Ω=[0,4] and the initial condition used is

    u(x,θ)=23+0.2cosπx2,w(x,θ)=0.7+0.2cosπx2,v(x,θ)=3+0.2cosπx2,z(x,θ)=2+0.2cosπx2
    Table 1.  Parameter values for simulation.
    Parameters Ranges value Units References
    s 010 10 cellsml1day1 [54]
    d 0.00010.2 0.01 day1 [22]
    β1 4.6×1080.5 variable ml1day1 [55]
    β2 1×1050.7 2.4×105 ml1day1 [39]
    m α[d,δ] 0.05 day1 [55]
    τ 01.5 0.5 days [55]
    δ 0.000191.4 1 day1 [22]
    p 0.00014.048 0.024 ml1day1 [22,55]
    N 6.2523599.9 2000 vironcells1 [22]
    c 236 23 day1 [39,22]
    q 0.00513.912 0.15 day1 [22]
    b 0.0048.087 0.5 day1 [39,22]

     | Show Table
    DownLoad: CSV

    for x[0,4] and θ[0.5,0].

    Firstly, we take β1=1×105. Then R0=0.8715<1. By Theorem 4.2, the infection-free steady state P0=(1000,0,0,0) is globally asymptotically stable (see Figure 1).

    Figure 1.  When R0<1, the infection-free steady state P0 is globally asymptotically stable. Parameter values are given in the text.

    Next, we choose β1=2.4×105. Then R0=2.0588>1 and R1=0.2989<1. From Theorem 4.7, the immunity-free infected steady state P1=(897.8483,0.9963,86.6344,0) is globally asymptotically stable (see Figure 2).

    Figure 2.  When R0=2.0588>1 and R1=0.2989<1, the immunity-free infected steady state P1 is globally asymptotically stable. See the text for the parameter values.

    Finally, with β1=8.4×105, we get R0=7.1474>1 and R1=1.1594>1. By Theorem 4.12, the infected-immune steady state P2=(609.8631,3.3333,289.8550,5.8964) is globally asymptotically stable (see Figure 3).

    Figure 3.  When R1>1, the infected-immune steady state P2 is globally asymptotically stable. See the text for the parameter values.

    In this paper, we have proposed and studied a reaction-diffusion virus infection model by incorporating time delays, general incidence functions, and cell-to-cell transmission.

    We have proved that the global dynamics of system (1.4)–(1.6) is determined by the basic reproduction number of infection R0 and the basic reproduction number of immunity R1. By analyzing the characteristic equations and constructing Lyapunov functionals, we have obtained the following conclusions: if R0<1, then the infected-free steady state P0 is globally asymptotically stable; if R11<R0, then the immunity-free infected steady state P1 is globally asymptotically stable under additional Assumption (A2); if R1>1, then the infected-immune steady state P2 is globally asymptotically stable under additional Assumptions (A2) and (A3). We mention that most commonly used incidences satisfy (A1)–(A3). Some examples are the Holling type Ⅱ incidence f(u,v)=βuv1+αv [40], Beddington-DeAnglis incidence [41], and f(u,v)=kuln(1+βvk) [56].

    Chen is supported by NSERC of Canada. Wang is supported by the NSFC (No. 11771374), the Nanhu Scholars Program for Young Scholars of Xinyang Normal University.

    The authors declare that there is no conflict of interest.



    [1] M. A. Nowak, S. Bonhoeffer, A. M. Hill, R. Boehme, H. C. Thomas, H. McDade, Viral dynamics in hepatitis B virus infection, Proc. Natl. Acad. Sci. USA, 93 (1996), 4398-4402.
    [2] M. A. Nowak, R. M. May, Virus dynamics: Mathematical principles of immunology and virology, Oxford University Press, 2000.
    [3] J. Guedj, A. U. Neumann, Understanding hepatitis C viral dynamics with direct-acting antiviral agents due to the interplay between intracellular replication and cellular infection dynamics, J. Theoret. Biol., 267 (2010), 330-340.
    [4] L. Rong, A. S. Pereslon, Mathematical analysis of multiscale models for hepatitis C virus dynamics under therapy with direct-acting antiviral agents, Math. Biosci., 245 (2013), 22-30.
    [5] R. V. Culshaw, S. Ruan, G. Webb, A mathematical model of cell-to-cell spread of HIV-1 that includes a time delay, J. Math. Biol., 46 (2003), 425-444.
    [6] P. K. Srivastava, M. Banerjee, P. Chandra, A primary infection model for HIV and immune response with two discrete time delays, Differ. Equ. Dyn. Syst., 18 (2010), 385-399.
    [7] X. Wang, X. Song, S. Tang, L. Rong, Dynamics of an HIV model with multiple infection stages and treatment with different drug classes, Bull. Math. Biol., 78 (2016), 322-349.
    [8] X. Wang, G. Mink, D. Lin, X. Song, L. Rong, Influence of raltegravir intensification on viral load and 2-LTR dynamics in HIV patients on suppressive antiretroviral therapy, J. Theoret. Biol., 416 (2017), 16-27.
    [9] M. Y. Li, H. Shu, Impact of intracellular delays and target-cell dynamics on in vivo viral infections, SIAM J. Appl. Math., 70 (2010), 2434-2448.
    [10] 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.
    [11] H. Zhao, S. Liu, A mathematical model of HTLV-I infection with nonlinear incidence and two time delays, Commun. Math. Biol. Neurosci., 2016 (2016).
    [12] M. A. Nowak, C. R. M. Bangham, Population dynamics of immune responses to persistent viruses, Science, 272 (1996), 74-79.
    [13] K. Allali, S. Harroudi, D. F. M. Torres, Analysis and optimal control of an intracellular delayed HIV model with CTL immune response, Math. Comput. Sci., 12 (2018), 111-127.
    [14] X. Wang, Y. Tao, Lyapunov function and global properties of virus dynamics with CTL immune response, Int. J. Biomath., 1 (2008), 443-448.
    [15] J. Li, K. Men, Y. Yang, D. Li, Dynamical analysis on a chronic hepatitis C virus infection model with immune response, J. Theoret. Biol., 365 (2015), 337-346.
    [16] A. M. Elaiw, A. A. Raezah, K. Hattaf, Stability of HIV-1 infection with saturated virus-target and infected-target incidences and CTL immune response, Int. J. Biomath., 10 (2017), 1750070.
    [17] A. V. M. Herz, S. Bonhoeffer, R. M. Anderson, R. M. May, M. A. Nowak, Viral dynamics in vivo: limitations on estimates of intracellular delay and virus decay, Proc. Natl. Acad. Sci. USA, 93 (1996), 7247-7251.
    [18] H. Shu, L. Wang, J. Watmough, Global stability of a nonlinear viral infection model with infinitely distributed intracellular delays and CTL immune responses, SIAM J. Appl. Math., 73 (2013), 1280-1302.
    [19] Y. Liu, C. Wu, Global dynamics for an HIV infection model with Crowley-Martin functional response and two distributed delays, J. Syst. Sci. Complex., 31 (2018), 385-395.
    [20] H. Zhu, X. Zou, Dynamics of a HIV-1 infection model with cell-mediated immune response and intracellular delay, Discrete Contin. Dyn. Syst. Ser. B, 12 (2009), 511-524.
    [21] X. Wang, S. Liu, A class of delayed viral models with saturation infection rate and immune response, Math. Methods Appl. Sci., 36 (2013), 125-142.
    [22] B. Li, Y. Chen, X. Lu, S. Liu, A delayed HIV-1 model with virus waning term, Math. Biosci. Eng., 13 (2016), 135-157.
    [23] O. T. Fackler, T. T. Murooka, A. Imle, T. R. Mempel, Adding new dimensions: towards an integrative understanding of HIV-1 spread, Nat. Rev. Microbiol., 12 (2014), 563-574.
    [24] K. Wang, W. Wang, Propagation of HBV with spatial dependence, Math. Biosci., 210 (2007), 78-95.
    [25] R. Xu, Z. Ma, An HBV model with diffusion and time delay, J. Theoret. Biol., 257 (2009), 499-509.
    [26] 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.
    [27] C. M. Brauner, D. Jolly, L. Lorenzi, R. Thiebaut, Heterogeneous viral environment in an HIV spatial model, Discrete Contin. Dyn. Syst. Ser. B, 15 (2011), 545-572.
    [28] Y. Zhang, Z. Xu, Dynamics of a diffusive HBV model with delayed Beddington-DeAngelis response, Nonlinear Anal. Real World Appl., 15 (2014), 118-139.
    [29] Y. Yang, Y. Xu, Global stability of a diffusive and delayed virus dynamics model with BeddingtonDeAngelis incidence function and CTL immune response, Comput. Math. Appl., 71 (2016), 922-930.
    [30] W. Hübner, G. P. McNerney, P. Chen, B. M. Dale1, 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.
    [31] P. Zhong, L. M. Agosto, J. B. Munro, W. Mothes, Cell-to-cell transmission of viruses, Curr. Opin. Virol., 3 (2013), 44-50.
    [32] N. Martin, Q. Sattentau, Cell-to-cell HIV-1 spread and its implications for immune evasion, Curr. Opin. HIV AIDS, 4 (2009), 143-149.
    [33] C. Zhang, S. Zhou, E. Groppelli, P. Pellegrino, I. Williams, P. Borrow, Hybrid spreading mechanisms and T cell activation shape the dynamics of HIV-1 infection, PLoS Comput. Biol., 11 (2015), e1004179.
    [34] 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.
    [35] X. Lai, X. Zou, Modeling HIV-1 virus dynamics with both virus-to-cell infection and cell-to-cell transmission, SIAM J. Appl. Math., 74 (2014), 898-917.
    [36] X. Lai, X. Zou, Modeling cell-to-cell spread of HIV-1 with logistic target cell growth, J. Math. Anal. Appl., 426 (2015), 563-584.
    [37] X. Wang, S. Tang, X. Song, L. Rong, Mathematical analysis of an HIV latent infection model including both virus-to-cell infection and cell-to-cell transmission, J. Biol. Dyn., 11 (2017), 455-483.
    [38] H. Shu, Y. Chen, L. Wang, Impacts of the cell-free and cell-to-cell infection modes on viral dynamics, J. Dyn. Differ. Equ., 30 (2018), 1817-1836.
    [39] A. Debadatta, B. Nandadulal, Analysis and computation of multi-pathways and multi-delays HIV-1 infection model, Appl. Math. Model., 54 (2018), 517-536.
    [40] X. Song, A. U. Neumann, Global stability and periodic solution of the viral dynamics, J. Math. Anal. Appl., 329 (2007), 281-297.
    [41] X. Wang, Y. Tao, X. Song, Global stability of a virus dynamics model with Beddington-DeAngelis incidence rate and CTL immune response, Nonlinear Dynam., 66 (2011), 825-830.
    [42] 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.
    [43] R. H. Martin, H. L. Smith, Abstract functional-differential equations and reaction-diffusion systems, Trans. Amer. Math. Soc., 321 (1990), 1-44.
    [44] Y. Gao, J. Wang, Threshold dynamics of a delayed nonlocal reaction-diffusion HIV infection model with both cell-free and cell-to-cell transmissions, J. Math. Anal. Appl., 488 (2020), 124047.
    [45] Y. Lou, X. Zhao, A reaction-diffusion malaria model with incubation period in the vector population, J. Math. Biol., 62 (2011), 543-568.
    [46] M. H. Protter, H. F. Weinberger, Maximum principles in differential equations, Springer-Verlag, 1984.
    [47] J. Wu, Theory and applications of partial functional differential equations, Springer, New York, 1996.
    [48] J. K. Hale, Asymptotic behavior of dissipative systems, American Mathematical Society, Providence, 1988.
    [49] W. Wang, X. Zhao, Basic reproduction numbers for reaction-diffusion epidemic models, SIAM J. Appl. Dyn. Syst., 11 (2012), 1652-1673.
    [50] J. K. Hale, S. M. Verduyn Lunel, Introduction to functional differential equations, Springer-Verlag, 1993.
    [51] J. LaSalle, S. Lefschetz, Stability by Lyapunov's direct method, with applications, in Mathematics in Science and Engineering, Academic Press, 1961.
    [52] H. R. Thieme, Convergence results and a Poincaré-Bendixson trichotomy for asymptotically autonomous differential equations, J. Math. Biol., 30 (1992), 755-763.
    [53] H. L. Smith, X. Zhao, Robust persistence for semidynamical systems, Nonlinear Anal., 47 (2001), 6169-6179.
    [54] P. W. Nelson, J. D. Murray, A. S. Perelson, A model of HIV-1 pathogenesis that includes an intracellular delay, Math. Biosci., 163 (2000), 201-215.
    [55] K. A. Pawelek, S. Liu, F. Pahlevani, L. Rong, A model of HIV-1 infection with two time delays: mathematical analysis and comparison with patient data, Math. Biosci., 235 (2012), 98-109.
    [56] H. McCallum, N. Barlow, J. Hone, How should pathogen transmission be modelled? Trends Ecol. Evol., 16 (2001), 295-300.
  • This article has been cited by:

    1. A. M. Elaiw, N. H. AlShamrani, Analysis of an HTLV/HIV dual infection model with diffusion, 2021, 18, 1551-0018, 9430, 10.3934/mbe.2021464
    2. Qing Ge, Xia Wang, Libin Rong, A delayed reaction–diffusion viral infection model with nonlinear incidences and cell-to-cell transmission, 2021, 14, 1793-5245, 10.1142/S179352452150100X
    3. Baolin Li, Fengqin Zhang, Xia Wang, A delayed diffusive HBV model with nonlinear incidence and CTL immune response, 2022, 45, 0170-4214, 11930, 10.1002/mma.8547
    4. Zakaria Yaagoub, Karam Allali, Fractional HBV infection model with both cell-to-cell and virus-to-cell transmissions and adaptive immunity, 2022, 165, 09600779, 112855, 10.1016/j.chaos.2022.112855
    5. Weixin Wu, Zengyun Hu, Long Zhang, Zhidong Teng, Traveling waves for a diffusive virus infection model with humoral immunity, cell‐to‐cell transmission, and nonlinear incidence, 2023, 0170-4214, 10.1002/mma.9291
    6. Yuequn Gao, Ning Li, Fractional order PD control of the Hopf bifurcation of HBV viral systems with multiple time delays, 2023, 83, 11100168, 1, 10.1016/j.aej.2023.10.032
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(3956) PDF downloads(243) Cited by(6)

Figures and Tables

Figures(3)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog