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

Mathematical analysis of an HBV model with antibody and spatial heterogeneity

  • In this paper, we modify the HBV model proposed in [1] to include the spatial variations of free antibody, virus-antibody complexes, and free virus. By using comparison arguments and theory of uniform persistence, we can show that the persistene/extinction of HBV can be determined by the reproduction number(s).

    Citation: Kuo-Sheng Huang, Yu-Chiau Shyu, Chih-Lang Lin, Feng-Bin Wang. Mathematical analysis of an HBV model with antibody and spatial heterogeneity[J]. Mathematical Biosciences and Engineering, 2020, 17(2): 1820-1837. doi: 10.3934/mbe.2020096

    Related Papers:

    [1] 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
    [2] Suxia Zhang, Hongbin Guo, Robert Smith? . Dynamical analysis for a hepatitis B transmission model with immigration and infection age. Mathematical Biosciences and Engineering, 2018, 15(6): 1291-1313. doi: 10.3934/mbe.2018060
    [3] Abulajiang Aili, Zhidong Teng, Long Zhang . Dynamical behavior of a coupling SEIR epidemic model with transmission in body and vitro, incubation and environmental effects. Mathematical Biosciences and Engineering, 2023, 20(1): 505-533. doi: 10.3934/mbe.2023023
    [4] Tingting Xue, Long Zhang, Xiaolin Fan . Dynamic modeling and analysis of Hepatitis B epidemic with general incidence. Mathematical Biosciences and Engineering, 2023, 20(6): 10883-10908. doi: 10.3934/mbe.2023483
    [5] Pengfei Liu, Yantao Luo, Zhidong Teng . Role of media coverage in a SVEIR-I epidemic model with nonlinear incidence and spatial heterogeneous environment. Mathematical Biosciences and Engineering, 2023, 20(9): 15641-15671. doi: 10.3934/mbe.2023698
    [6] Liqiong Pu, Zhigui Lin . A diffusive SIS epidemic model in a heterogeneous and periodically evolvingenvironment. Mathematical Biosciences and Engineering, 2019, 16(4): 3094-3110. doi: 10.3934/mbe.2019153
    [7] Xiaoqin Wang, Yiping Tan, Yongli Cai, Kaifa Wang, Weiming Wang . Dynamics of a stochastic HBV infection model with cell-to-cell transmission and immune response. Mathematical Biosciences and Engineering, 2021, 18(1): 616-642. doi: 10.3934/mbe.2021034
    [8] Zongwei Ma, Hongying Shu . Viral infection dynamics in a spatial heterogeneous environment with cell-free and cell-to-cell transmissions. Mathematical Biosciences and Engineering, 2020, 17(3): 2569-2591. doi: 10.3934/mbe.2020141
    [9] Yongli Cai, Yun Kang, Weiming Wang . Global stability of the steady states of an epidemic model incorporating intervention strategies. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1071-1089. doi: 10.3934/mbe.2017056
    [10] Danfeng Pang, Yanni Xiao . The SIS model with diffusion of virus in the environment. Mathematical Biosciences and Engineering, 2019, 16(4): 2852-2874. doi: 10.3934/mbe.2019141
  • In this paper, we modify the HBV model proposed in [1] to include the spatial variations of free antibody, virus-antibody complexes, and free virus. By using comparison arguments and theory of uniform persistence, we can show that the persistene/extinction of HBV can be determined by the reproduction number(s).


    There has been some work in the study of the relationship between persistent infection with hepatitis B virus and immune responses (see, e.g., [2]). Hepatitis B virus (HBV) is a major cause of various liver diseases around the world. Except acute and chronic hepatitis, it causes liver fibrosis and even hepatocellular carcinoma. When an adult gets first infected with the hepatitis B virus during the early period of six months, it is called an acute infection. On the other hand, innate immune responses on persons may drive huge effector immune cells (CD8 T cells, help T cells, B cells) against infection. It is probably due to such immune system, in clinical observations, only 5–10 percent of healthy adults will develop a chronic hepatitis B infection after they get infection. This motivates researchers to investigate the topic that whether antibodies against hepatitis B play a central role in virus clearance (see, e.g., [1,2]).

    It is practically difficult to obtain experimental results in the study of the antibody response to hepatitis B virus (HBV) infection. Thus, developing suitable mathematical models is an alternative way since it can be used to estimate some crucial factors for the viral infection, and to explore possible mechanisms of protection and viral infection process (see, e.g., [1,2,3,4,5,6,7] and the references therein). We first mention a model of virus infection in the absence of antibody responses, namely, the following model consists of three compartments of populations, corresponding to target hepatocytes (T), infected hepatocytes (I), and virus (V).

    {dT(t)dt=rT(1T+ITm)βVT+ρI,dI(t)dt=βVTδIρI,dV(t)dt=πIcV. (1.1)

    The growth of target cells (T) in system (1.1) is described by a logistic term with carrying capacity Tm and and maximal growth rate r (see, e.g., [8,9]); target cells (T) also get infected at a rate βVT. Infected cells (I) are gained at rate βVT, and die at rate δ. Infected cells (I) produce virus (V) at rate π, and virus clearance rate is denoted by c. Further, system (1.1) also assumes that infected class (I) can get recovery and move back into the target class at rate ρ.

    In order to incorporate antibody response, the authors in [1] ignore the curing of infected cells by setting ρ=0, and introduce two additional classes, free antibody (A) and virus-antibody complexes (X), into system (1.1). Then the governing system takes the following form:

    {dT(t)dt=rT(1T+ITm)βVT,dI(t)dt=βVTδI,dA(t)dt=pA(1+θ)V+rAA(1AAm)+(1+θ)kmX(1+θ)kpAVdAA,dX(t)dt=kmX+kpAVcAVX,dV(t)dt=πIcV+kmXkpAV. (1.2)

    The free antibody (A) is produced at rate pA proportional to the viral and subviral concentrations, and is degraded at rate dA. Without virus, we also introduce a logistic term with maximum growth rate rA and carrying capacity Am for the antibody maintenance. In system (1.2), for simplicity, we have imposed the assumption that the concentration of subviral particles is proportional to the concentration of free virus V, and θ is a constant proportionality. Antigen clearance is caused by the constitution of antigen-antibody complexes. The binding rate with antigen-antibody is kp that causes the free antibody population to descend; km represents the disassociation rate for antibody reacting to viral particles. The complexes (X) are produced by a productive combination rate kp and it decreases at a disassociation rate km and a degradation rate cAV. During infection, free virus (V) are gained at a rate π and binding rate km with complexes, and are degraded by a rate c and binding rate kp with antibody.

    In [1], the authors also mention that it can be a further topic in the investigation of spatial effects in HBV infection. In fact, spatial clustering of infected cells has recently been observed for hepatitis C virus (HCV) infection (see, e.g., [10]). The effects of spatial heterogeneity was also added to within-host HIV models, see [11,12]. Motivated by those previous works, we intend to consider system (1.2) with spatial variations. For this purpose, we add diffusion terms DAΔA, DXΔX and DVΔV into the model, which reflects the spatial variations of free antibody (A), virus-antibody complexes (X) and free virus (V), respectively. Then the modified version of system (1.2) is as follows

    {Tt=rT(1T+ITm)βVT, xΩ, t>0,It=βVTδI, xΩ, t>0,At=DAΔA+pA(1+θ)V+rA(x)A(1AAm)+(1+θ)kmX             (1+θ)kpAVdA(x)A, xΩ, t>0,Xt=DXΔXkmX+kpAVcAVX, xΩ, t>0,Vt=DVΔV+πIcV+kmXkpAV, xΩ, t>0,Aν=Xν=Vν=0, xΩ, t>0,u(x,0)=u0(x), u=T,I,A,X,V, xΩ. (1.3)

    Here, we consider a general bounded domain ΩR3 where virus and cells stay and interact, and pose zero-flux condition on the boundary of Ω (i.e., homogeneous Neumann boundary condition). The notation ν denotes the differentiation along the outward normal ν to Ω. The location dependent parameters are continuous and strictly positive functions on ˉΩ.

    The dynamics of system (1.3) is challenging since there are no diffusion terms in the first two equations, resulting in the loss of compactness of the solution maps. In order to determine the disease-free steady state of system (1.3), we also need to investigate the following system:

    {Tt=rT(1TTm), xΩ, t>0,T(x,0)=T0(x), xΩ. (1.4)

    It is easy to see that T=0 and T=Tm are two steady states of (1.4). However, the global dynamics of system(1.4) is still open to us, due to the loss of compactness of the solution maps. This stops us from using persistence theory in the investigation of the dynamics of system (1.3). Instead, we will focus on the study of the existence of the positive steady states of system (1.3), (ˆT(x),ˆI(x),ˆA(x),ˆX(x),ˆV(x)), which satisfies the following equations:

    {rˆT(1ˆT+ˆITm)βˆVˆT=0, xΩ,βˆVˆTδˆI=0, xΩ,DAΔˆA+pA(1+θ)ˆV+rA(x)ˆA(1ˆAAm)+(1+θ)kmˆX             (1+θ)kpˆAˆVdA(x)ˆA=0, xΩ,DXΔˆXkmˆX+kpˆAˆVcAVˆX=0, xΩ,DVΔˆV+πˆIcˆV+kmˆXkpˆAˆV=0, xΩ,ˆAν=ˆXν=ˆVν=0, xΩ. (1.5)

    In view of the first two equations of (1.5), it follows that

    ˆT+ˆI=Tm(1βrˆV), ˆI=βδˆVˆT. (1.6)

    Then

    {ˆT=Tm1βrˆV1+βδˆV,ˆI=βδTm(1βrˆV)ˆV1+βδˆV. (1.7)

    Substituting the second equality of (1.7) into the fifth equation of (1.5), we arrive at the following elliptic system

    {DAΔˆA+pA(1+θ)ˆV+rA(x)ˆA(1ˆAAm)+(1+θ)kmˆX             (1+θ)kpˆAˆVdA(x)ˆA=0, xΩ,DXΔˆXkmˆX+kpˆAˆVcAVˆX=0, xΩ,DVΔˆV+πβδTm1βrˆV1+βδˆVˆVcˆV+kmˆXkpˆAˆV=0, xΩ,ˆAν=ˆXν=ˆVν=0, xΩ. (1.8)

    The standard approach in seeking for the positive steady states of system (1.8) is the bifurcation argument. Here, we are going to adopt another approach, using the persistence theory, to study the following parabolic system associated with (1.8):

    {At=DAΔA+pA(1+θ)V+rA(x)A(1AAm)+(1+θ)kmX             (1+θ)kpAVdA(x)A, xΩ, t>0,Xt=DXΔXkmX+kpAVcAVX, xΩ, t>0,Vt=DVΔV+πf(V)VcV+kmXkpAV, xΩ, t>0,Aν=Xν=Vν=0, xΩ, t>0,A(x,0)=A0(x), X(x,0)=X0(x), V(x,0)=V0(x), xΩ, (1.9)

    where

    f(V)=βδTm1βrV1+βδV. (1.10)

    If one can show that system (1.9) is uniformly persistent, then (1.9) must admit a positive steady state (see, e.g., [13,CH1]). We point out that the dynamics of systems (1.3) and (1.9) may be different, but they admit the same positive steady states. Thus, we will focus on the search for positive steady state(s) of system (1.9) via the establishment of uniform persistence of system (1.9).

    Let Y:=C(ˉΩ,R3) be the Banach space with the supremum norm Y. Define Y+:=C(ˉΩ,R3+), then (Y,Y+) is a strongly ordered space. By the similar arguments in [14,Lemma 2.2] (see also [15]), together with [16,Corollary 4] (see also [17,Theorem 7.3.1]), we have the following result:

    Lemma 2.1. For every initial value function ϕY+, system (1.9) has a unique mild solution u(x,t,ϕ) on (0,τϕ) with u(,0,ϕ)=ϕ, where τϕ. Furthermore, u(,t,ϕ)Y+, t(0,τϕ) and u(x,t,ϕ) is a classical solution of (1.9).

    Next, we show that solutions of system (1.9) are ultimately bounded, and system (1.9) admits a compact attractor in Y+.

    Lemma 2.2. For every initial value function ϕY+, system (1.9) admits a unique solution u(x,t,ϕ) on [0,) with u(,0,ϕ)=ϕ. Furthermore,

    (ⅰ) u(x,t,ϕ) is ultimately bounded;

    (ⅱ) The semiflow Ψ(t):Y+Y+ generated by (1.9) is defined by Ψ(t)ϕ=u(,t,ϕ), t0, which admits a global compact attractor in Y+,  t0.

    Proof. In view of (1.10), it is not hard to see that

    f(V)VβδTmV1+βδVβδTmVβδV=Tm,  V>0.

    Thus,

    f(V)VTm,  V0. (2.1)

    Setting

    U(t)=Ω[X(x,t)+V(x,t)]dx.

    Then it follows from system (1.9) and (2.1) that

    dU(t)dt=Ωπf(V(x,t))V(x,t)dxΩ[cAVX(x,t)+cV(x,t)]dx       πTm|Ω|cminU(t),

    where cmin:=min{cAV,c}. Thus, we have

    U(t)U(0)ecmin t+πTm|Ω|cmin(1ecmin t). (2.2)

    Using (2.2) and the similar arguments to those in the end of [18,Proposition 2.3], we can show that X(,t,ϕ) and V(,t,ϕ) are ultimately bounded. Therefore, there exists ˆC>0 and t1>0 such that

    pA(1+θ)V(x,t)+(1+θ)kmX(x,t)ˆC,  x¯Ω, tt1. (2.3)

    In view of the first equation of system (1.9) and (2.3), it follows that

    {AtDAΔA+ˆC+rA(x)A(1AAm)dA(x)A,  xΩ, tt1,Aν=0, xΩ, tt1.

    Then

    lim suptA(x,t)ˆA,  x¯Ω,

    where ˆA>0 is a constant such that

    ˆC+rA(x)ˆA(1ˆAAm)dA(x)ˆA0,  xΩ.

    From the above discussions, we see that Ψ(t):Y+Y+ is point dissipative. Obviously, Ψ(t):Y+Y+ is compact,  t>0. It follows from [19,Theorem 3.4.8] that Ψ(t):Y+Y+, t0, admits a global compact attractor.

    Putting X=V=0 into (1.9), we see that

    {At=DAΔA+rA(x)A(1AAm)dA(x)A, xΩ, t>0,Aν=0, xΩ, t>0,A(x,0)=A0(x), xΩ. (2.4)

    It is easy to see that A=0 is the trivial steady state solution of system (2.4). The stability of the trivial steady state solution A=0 is determined by the following eigenvalue problem:

    {μφ(x)=DAΔφ(x)+(rA(x)dA(x))φ(x), xΩ,φ(x)ν=0, xΩ. (2.5)

    Assume that μ0 is the principal eigenvalue of system (2.5). By [20,Proposition 4.4], we see that μ0>0 if the following condition is satisfied

    Ω(rA(x)dA(x))dx>0. (2.6)

    Thus, trivial steady state solution A=0 is unstable for system (2.4) if condition (2.6) holds. If condition (2.6) is true, then one can use [13,Theorem 2.3.2] to show that system (2.4) admits a unique positive steady state A(x) which is globally attractive. Thus, two possible steady states of system (1.9) are as follows:

    E0(x)=(A,X,V)=(0,0,0),

    and

    E1(x)=(A,X,V)=(A(x),0,0).

    Note that E0(x) always exists, and E1(x) exists when (2.6) holds. Linearizing system (1.9) around E1(x), we get the following cooperative system for the infectious compartments:

    {Xt=DXΔXkmX+kpA(x)VcAVX, xΩ, t>0,Vt=DVΔV+πf(0)VcV+kmXkpA(x)V, xΩ, t>0,Xν=Vν=0, xΩ, t>0. (2.7)

    Substituting X(x,t)=eλtψX(x) and V(x,t)=eλtψV(x) into (2.7) and we get the associated eigenvalue problem:

    {λψX(x)=DXΔψX(x)(km+cAV)ψX(x)+kpA(x)ψV(x), xΩ,λψV(x)=DVΔψV(x)+kmψX(x)+(πf(0)ckpA(x))ψV(x), xΩ,ψX(x)ν=ψV(x)ν=0, xΩ. (2.8)

    It is not hard to see that the linear system (2.7) generates a strongly positive semigroup on C(¯Ω,R2+) (see, e.g., Section 4 of CH 7 in [17]). In addition, the semigroup associated with system (2.7) is compact. By a similar argument as in [17,Theorem 7.6.1], we have the following result which is related to the existence of the principal eigenvalue of (2.8):

    Lemma 2.3. The eigenvalue problem (2.8) admits a principal eigenvalue, denoted by λ0, which corresponds a strongly positive eigenfunction.

    Next, we shall adopt the theory developed in [21,Section 3] to define the basic reproduction number for system (1.9). For this purpose, we assume

    F(x)=(0kpA(x)kmπf(0)), (2.9)

    and

    V(x)=(km+cAV00c+kpA(x)). (2.10)

    Let w=(X,V)T, DΔw=(DXΔX,DVΔV)T, and S(t):C(¯Ω,R2)C(¯Ω,R2) be the C0-semigroup generated by the following system

    {wt=DΔwV(x)w, x¯Ω, t>0,Xν=Vν=0, xΩ, t>0. (2.11)

    Assume that the state variables are near the disease-free steady state E1(x) and the distribution of initial infection is described by φC(¯Ω,R2). Then S(t)φ(x) represents the distribution of those infectious cases as time evolves to time t, and hence, the distribution of new infection at time t is F(x)S(t)φ(x). Let L:C(¯Ω,R2)C(¯Ω,R2) be defined by

    L(φ)()=0F()(S(t)φ)()dt.

    It then follows that L(φ)() represents the distribution of accumulated infectious cases during the infection period, and hence, L is the next generation operator. By the idea of next generation operators (see, e.g., [21,22,23]), we define the spectral radius of L as the basic reproduction number for system (1.9), that is,

    R0:=r(L).

    From [24,Theorem 3.5] or [21,Theorem 3.1], the following observation holds.

    Lemma 2.4. R01 and λ0 have the same sign.

    Next, we are going to find an explicit formula for R0 when coefficients of system (1.9) are all positive constants. For this special case, we see that F(x)=F and V(x)=V, for all xˉΩ, and hence, R0=r(FV1) (see e. g., [21,Theorem 3.4]). By direct computations, it follows that

    FV1=(0kpAkmπf(0))(1km+cAV001c+kpA)=(0kpAc+kpAkmkm+cAVπf(0)c+kpA).

    Thus,

    R0=12[πf(0)c+kpA+(πf(0)c+kpA)2+4kpAc+kpAkmkm+cAV ]. (2.12)

    In the establishment of the persistence for (1.9), the following results will be necessary.

    Lemma 2.5. For every initial value function ϕY+, we assume that system (1.9) admits a unique solution u(x,t,ϕ) on [0,) with u(,0,ϕ)=ϕ.

    (ⅰ) If ϕ2()0 and ϕ3()0, then

    ui(x,t,ϕ)>0, for xˉΩ, t>0, and 1i3.

    (ⅱ) Assume that ϕi()0, for i=2,3. If there exists a σ1>0 such that

    lim inftX(x,t,ϕ)σ1 and lim inftV(x,t,ϕ)σ1, uniformly for xˉΩ. (2.13)

    Then there exists a σ>0 such that

    lim inftui(x,t,ϕ)σ, uniformly for xˉΩ, and 1i3. (2.14)

    Proof. Part (ⅰ). By the positivity of solutions (see Lemma 2.1), it follows that X(x,t)0,  x¯Ω, t0. Suppose, by contradiction, there exists x1¯Ω and t1(0,) such that X(x1,t1)=0. Let τ1>0 be such that t1<τ1. Then (x1,t1)¯Ω×[0,τ1] and X attains its minimum on ¯Ω×[0,τ1] at the point (x1,t1). In view of the second equation of (1.9), it follows that

    {XtDXΔX(km+cAV)X, xΩ, t(0,τ1],Xν=0, xΩ, t(0,τ1].

    In case x1Ω, we apply the Hopf boundary lemma (see, e.g., [25,p. 170,Theorem 3]) and we have X(x1,t1,ϕ)ν<0, which is impossible. In case where x1Ω, then the strong maximum principle (see [25,p. 174,Theorem 7]) implies that

    X(x,t,ϕ)X(x1,t1,ϕ)=0,  (x,t)¯Ω×[0,τ1],

    which contradicts the assumption that ϕ2()0. Thus, X(x,t,ϕ)>0,  xˉΩ, t>0. Similarly, we see that V(x,t)0,  x¯Ω, t0 (see Lemma 2.1). Suppose, by contradiction, there exists x2¯Ω and t2(0,) such that V(x2,t2)=0. Let τ2>0 be such that t2<τ2. Then (x2,t2)¯Ω×[0,τ2] and V attains its minimum on ¯Ω×[0,τ2] at the point (x2,t2). Using the third equation of (1.9) and (1.10), it follows that

    {VtDVΔVπ[βδTmβrV1+βδV+c+kpA]V, xΩ, t(0,τ2],Vν=0, xΩ, t(0,τ2]. (2.15)

    In case x2Ω, we apply the Hopf boundary lemma (see, e.g., [25,p. 170,Theorem 3]) and we have V(x2,t2,ϕ)ν<0, which is a contradiction. In case where x2Ω, then the strong maximum principle (see [25,p. 174,Theorem 7]) implies that

    V(x,t,ϕ)V(x2,t2,ϕ)=0,  (x,t)¯Ω×[0,τ2],

    which contradicts the assumption that ϕ3()0. Thus, V(x,t,ϕ)>0,  xˉΩ, t>0.

    Claim. A(x,t,ϕ)>0,  xˉΩ, t>0.

    By Lemma 2.1, it follows that A(x,t)0,  x¯Ω, t0. Suppose, by contradiction, there exists x3¯Ω and t3(0,) such that A(x3,t3)=0. Let τ3>0 be such that t3<τ3. Then (x3,t3)¯Ω×[0,τ3] and A attains its minimum on ¯Ω×[0,τ3] at the point (x3,t3). By the first equation of (1.9), it follows that

    {AtDAΔA[rA(x)AAm+(1+θ)kpV+dA(x)]A, xΩ, t(0,τ3],Aν=0, xΩ, t(0,τ3].

    In case x3Ω, we apply the Hopf boundary lemma (see, e.g., [25,p. 170,Theorem 3]) and we have A(x3,t3,ϕ)ν<0, which is a contradiction. In case where x3Ω, then the strong maximum principle (see [25,p. 174,Theorem 7]) implies that

    A(x,t,ϕ)A(x3,t3,ϕ)=0,  (x,t)¯Ω×[0,τ3].

    This together with the first equation of (1.9) imply that

    X(x,t,ϕ)0 and V(x,t,ϕ)0,  (x,t)¯Ω×[0,τ3],

    which is a contradiction. Thus, A(x,t,ϕ)>0,  xˉΩ, t>0.

    Part (ⅱ). From Lemma 2.2, we see that V(x,t) is ultimately bounded. This together with assumption (2.13) imply that there exists t4>0 and C>0 such that

    12σ1V(x,t)C, and X(x,t)12σ1,  ˉΩ, tt4.

    From the above inequalities and the first equation of (1.9), it follows that

    {AtDAΔA+12(1+θ)(pA+km)σ1+rA(x)A(1AAm)             [(1+θ)kpC+dA(x)]A, xΩ, tt4,Aν=0, xΩ, tt4. (2.16)

    Let A_>0 satisfy the following inequality

    12(1+θ)(pA+km)σ1+rA(x)A_(1A_Am)[(1+θ)kpC+dA(x)]A_0,  xΩ.

    By (2.16) and the standard parabolic comparison theorem (see, e.g., [17,Theorem 7.3.4]), we deduce that

    lim inftA(x,t,ϕ)A_,  xˉΩ.

    Let σ:=min{σ1,A_}. Then (2.14) holds.

    We show that R0 is an important index for the persistence of HBV in system (1.9).

    Theorem 2.1. Assume that (2.6) holds. For every initial value function u0()=(A0,X0,V0)()Y+, we assume that system (1.9) admits a unique solution

    u(x,t,u0):=(A(x,t),X(x,t),V(x,t))

    on [0,) with u(,0,u0)=u0. If R0>1, then system (1.9) admits at least one (componentwise) positive steady state ˆu(x) and there exists a σ>0 such that for any u0()Y+ with X0()0 and V0()0, we have

    lim inft w(x,t,u0())σ, for w=A,X,V, (2.17)

    uniformly for x¯Ω.

    Proof. Let

    W0={u0()=(A0,X0,V0)()Y+:X0()0 and V0()0},

    and

    W0=Y+W0={u0()=(A0,X0,V0)()Y+:X0()0 or V0()0}.

    Recall that the semiflow Ψ(t):Y+Y+ generated by (1.9) is defined in Lemma 2.2. By Lemma 2.5 (ⅰ), it follows that for any u0()W0, we have

    w(x,t,u0())>0, for xˉΩ, t>0, and w=A,X,V.

    In other words, Ψ(t)W0W0,  t0. Let

    M:={u0()W0:Ψ(t)u0()W0, t0},

    and ω(u0()) be the omega limit set of the orbit O+(u0()):={Ψ(t)u0():t0}.

    Claim 1. ω(v0()){E0(x)}{E1(x)},  v0()M.

    Since v0()M, we have Ψ(t)v0()M,  t0, that is, X(,t,v0())0 or V(,t,v0())0,  t0.

    In case where V(,t,v0())0,  t0. Then it follows from the third equation in system (1.9) that X(,t,v0())0,  t0. Thus, X(x,t,v0()) satisfies system (2.4), and hence,

    either limtA(x,t,v0)=0 or limtA(x,t,v0)=A(x), uniformly for xˉΩ.

    Thus,

    either limtu(x,t,v0)=E0(x) or limtu(x,t,v0)=E1(x), uniformly for all xˉΩ.

    In case where V(,ˆt0,v0())0, for some ˆt00. Then we can use similar arguments in Lemma 2.5 to show that V(x,t,v0)>0, for all xˉΩ and t>ˆt0, and hence, X(,t,v0)0, for all t>ˆt0. Then it follows from the second equation in system (1.9) that A(,t,v0())V(,t,v0())0,  t>ˆt0. From the above discussions, it follows that A(,t,v0())0,  t>ˆt0. Thanks to the first equation in system (1.9), it follows that V(,t,v0())0,  t>ˆt0. This is a contradiction, and hence, we cannot allow the possibility that V(,ˆt0,v0())0, for some ˆt00. Therefore, we complete the proof of Claim 1.

    Recall that μ0 is the principal eigenvalue of the eigenvalue problem (2.5), and μ0>0 since (2.6) holds. By continuity, there is a δ0>0 such that μδ0>0, where μδ0>0 is the principal eigenvalue of the following eigenvalue problem:

    {μφ(x)=DAΔφ(x)+[rA(x)(1δ0Am)(1+θ)kpδ0dA(x)]φ(x), xΩ,φ(x)ν=0, xΩ. (2.18)

    Claim 2. E0(x) is a uniform weak repeller for W0 in the sense that

    lim supt

    Suppose, by contradiction, that there exists u^0(\cdot) \in \mathbb{W}_{0} such that

    \limsup\limits_{t\rightarrow\infty}\|\Psi(t)u^0(\cdot)-E_0(\cdot)\| \lt \delta_0.

    Then there exists t_0 > 0 such that

    0\leq w(x, t, u^0) \lt \delta_0, \ \forall \ t \geq t_0, \ x\in\bar{\Omega}, \ w = A, X, V.

    From the first equation of (1.9), we see that

    \begin{equation} \begin{cases} \frac{\partial A}{\partial t}\geq D_A\Delta A+[r_{A}(x)(1-\frac{\delta_0}{A_{m}}) -(1+\theta)k_{p}\delta_0-d_{A}(x)]A, \ x\in \Omega, \ t \geq t_0, \\ \frac{\partial A}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \geq t_0. \end{cases} \end{equation} (2.19)

    Assume that \varphi_{\delta_0}(x) is the positive eigenfunction corresponding to \mu_{\delta_0} , and there exists a C_0 > 0 such that

    A(x, t_0)\geq C_0\varphi_{\delta_0}(x), \ \forall \ x\in\bar{\Omega},

    where we have used the fact that A(x, t_0) > 0, \ \forall \ x\in\bar{\Omega} (see Lemma 2.5). The comparison principle and the inequality (2.19) imply that

    A(x, t)\geq C_0e^{\mu_{\delta_0}(t-t_0)}\varphi_{\delta_0}(x), \ \forall \ t \geq t_0, \ x\in\bar{\Omega}.

    Since \mu_{\delta_0} > 0 , it follows that A(x, t) is unbounded. This contradiction proves the Claim 2.

    Since \mathcal{R}_0 > 1 , it follows from Lemma 2.4 that \lambda^{0} > 0 . By continuity of the principal eigenvalue, we can find an \epsilon_1 > 0 such that \lambda_{\epsilon_1} > 0 , where \lambda_{\epsilon_1} is the principal eigenvalue of the following eigenvalue problem:

    \begin{equation} \begin{cases} \lambda\psi_{X}(x) = D_X\Delta \psi_{X}(x)-(k_{m}+c_{AV})\psi_{X}(x)+k_{p}[A^*(x)-\epsilon_1]\psi_{V}(x), \ x\in \Omega, \\ \lambda\psi_{V}(x) = D_V\Delta \psi_{V}(x)+k_{m}\psi_{X}(x)\\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +[\pi (f(0)-\epsilon_1)-c-k_{p}(A^*(x)+\epsilon_1)]\psi_{V}(x), \ x\in \Omega, \\ \frac{\partial \psi_{X}(x)}{\partial\nu} = \frac{\partial \psi_{V}(x)}{\partial\nu} = 0, \ x\in\partial\Omega. \end{cases} \end{equation} (2.20)

    By continuity of f(V) , we can choose a \delta_1 with 0 < \delta_1\leq\epsilon_1 such that

    \begin{equation} f(V) \gt f(0)-\epsilon_1, \ \forall \ \mid V\mid \lt \delta_1. \end{equation} (2.21)

    Claim 3. E_1(x) is a uniform weak repeller for \mathbb{W}_{0} in the sense that

    \limsup\limits_{t\rightarrow\infty}\|\Psi(t)u^0(\cdot)-E_1(\cdot)\|\geq \frac{1}{2}\delta_1, \ \forall\ u^0(\cdot)\in \mathbb{W}_{0}.

    Suppose, by contradiction, there exists u^0(\cdot) \in \mathbb{W}_{0} such that

    \limsup\limits_{t\rightarrow\infty}\|\Psi(t)u^0(\cdot)-E_1(x)\| \lt \frac{1}{2}\delta_1.

    Then there exists t_1 > 0 such that

    A^*(x)-\epsilon_1 \lt A^*(x)-\frac{1}{2}\delta_1\leq A(x, t, u^0) \lt A^*(x)+\frac{1}{2}\delta_1 \lt A^*(x)+\epsilon_1, \ \forall \ t \geq t_1, \ x\in\bar{\Omega},

    and

    0\leq w(x, t, u^0) \lt \frac{1}{2}\delta_1 \lt \epsilon_1, \ \forall \ t \geq t_1, \ x\in\bar{\Omega}, \ w = X, V.

    From the second and third equations in system (1.9), it follows that

    \begin{equation} \begin{cases} \frac{\partial X}{\partial t}\geq D_X\Delta X-k_{m}X+k_{p}[A^*(x)-\epsilon_1]V-c_{AV}X, \ x\in \Omega, \ t \geq t_1, \\ \frac{\partial V}{\partial t}\geq D_V\Delta V+ \pi [f(0)-\epsilon_1]V-c V+k_{m}X\\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ -k_{p}[A^*(x)+\epsilon_1]V, \ x\in \Omega, \ t \geq t_1, \\ \frac{\partial X}{\partial\nu} = \frac{\partial V}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \geq t_1. \end{cases} \end{equation} (2.22)

    Assume that (\psi_{X}^{\epsilon_1}(x), \psi_{V}^{\epsilon_1}(x)) is the positive eigenfunction corresponding to \lambda_{\epsilon_1} , and there exists a C_1 > 0 such that

    (X(x, t_1), V(x, t_1))\geq C_1(\psi_{X}^{\epsilon_1}(x), \psi_{V}^{\epsilon_1}(x)), \ \forall \ x\in\bar{\Omega},

    where we have used the fact that X(x, t_1) > 0, \ V(x, t_1) > 0, \ \forall \ x\in\bar{\Omega} (see Lemma 2.5). The comparison principle and the inequality (2.22) imply that

    (X(x, t), V(x, t))\geq C_1e^{\lambda_{\epsilon_1}(t-t_1)}(\psi_{X}^{\epsilon_1}(x), \psi_{V}^{\epsilon_1}(x)), \ \forall \ t \geq t_1, \ x\in\bar{\Omega}.

    Since \lambda_{\epsilon_1} > 0 , it follows that (X(x, t), V(x, t)) is unbounded. This contradiction proves Claim 3.

    Define a continuous function \mathbb{P}:\mathbb{Y}^+ \rightarrow [0, \infty) by

    \mathbb{P}(u^0(\cdot)): = \min\{\min\limits_{x\in \bar{\Omega}} X^0(x), \ \min\limits_{x\in \bar{\Omega}} V^0(x)\}, \ \forall \ u^0(\cdot) = (A^0, X^0, V^0)(\cdot)\in \mathbb{Y}^+.

    By Lemma 2.5 (ⅰ), it follows that \mathbb{P}^{-1}(0, \infty)\subseteq \mathbb{W}_{0} and \mathbb{P} has the property that if \mathbb{P}(u^0(\cdot)) > 0 or u^0(\cdot)\in \mathbb{W}_{0} with \mathbb{P}(u^0(\cdot)) = 0 , then \mathbb{P}(\Psi(t)u^0(\cdot)) > 0, \ \forall \ t > 0. That is, \mathbb{P} is a generalized distance function for the semiflow \Psi(t):\mathbb{Y}^+ \rightarrow \mathbb{Y}^+ (see, e.g., [26]).

    From the above claims, it follows that any forward orbit of \Psi(t) in M_{\partial} converges to \{E_0(x)\}\cup \{E_1(x)\} . For i = 0, 1 , \{E_i(x)\} is isolated in \mathbb{Y}^+ and W^{s}(\{E_i(x)\})\cap \mathbb{W}_{0} = \emptyset , where W^{s}(\{E_i(x)\}) is the stable set of \{E_i(x)\} (see [26]). It is obvious that no subset of \{E_0(x)\}\cup \{E_1(x)\} forms a cycle in \partial \mathbb{W}_{0} . By Lemma 2.2, the semiflow \Psi(t):\mathbb{Y}^+\rightarrow\mathbb{Y}^+ has a global compact attractor in \mathbb{Y}^+ , \forall \ t \geq0 . Then it follows from [26,Theorem 3] that there exists a \sigma_1 > 0 such that

    \min\limits_{\psi\in\omega(u^0(\cdot))}p(\psi) \gt \sigma_1, \ \forall \ u^0(\cdot)\in \mathbb{W}_{0}.

    Hence,

    \liminf\limits_{t\rightarrow \infty}X(\cdot, t, u^0(\cdot))\geq \sigma_1\ \mbox{and}\ \liminf\limits_{t\rightarrow \infty}V(\cdot, t, u^0(\cdot))\geq \sigma_1, \ \forall \ u^0(\cdot)\in \mathbb{W}_{0}.

    From Lemma 5 (ⅱ), there exists a \sigma > 0 such that (2.17) is valid. Hence, the uniform persistence stated in the conclusion (ⅱ) hold. By [27,Theorem 3.7 and Remark 3.10], it follows that \Psi(t):\mathbb{W}_{0} \rightarrow \mathbb{W}_{0} has a global attractor \mathcal{A}_0 . Using [27,Theorem 4.7], we deduce that \Psi(t) admits a steady-state \hat{u}(\cdot)\in \mathbb{W}_{0} . By Lemma 2.5 (ⅰ), we can further conclude that \hat{u}(\cdot) is a positive steady state of (1.9). The proof of Part (ⅱ) is finished.

    In this section, we focus on the study of elimination of HBV with antibody. Due to technical reasons, we only consider a special case where we assume k_{m} = 0 in system (1.9), and the coefficients in (1.9) are all positive constants. Then the equation of X in system (1.9) is decoupled from the other equations, and hence, it suffices to investigate the following system:

    \begin{equation} \begin{cases} \frac{\partial A}{\partial t} = D_A\Delta A+p_{A}(1+\theta)V+r_{A}A(1-\frac{A}{A_{m}})\\ \ \ \ \ \ \ \ \ \ \ \ \ \ -(1+\theta)k_{p}AV-d_{A}A, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial V}{\partial t} = D_V\Delta V+ \pi f(V)V-c V-k_{p}AV, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial A}{\partial\nu} = \frac{\partial V}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \gt 0, \\ A(x, 0) = A^0(x), \ V(x, 0) = V^0(x), \ x\in\Omega. \end{cases} \end{equation} (3.1)

    We see that two possible steady states of system (3.1) are as follows:

    \mathcal{E}_0 = (A, V) = (0, 0),

    and

    \mathcal{E}_1 = (A, V) = (A^*, 0),

    where A^*: = A_{m}(1-\frac{d_A}{r_{A}}) > 0 , provided that r_{A} > d_A .

    Linearizing system (3.1) around \mathcal{E}_1 , we get the following scalar system

    \begin{equation} \begin{cases} \frac{\partial V}{\partial t} = D_V\Delta V+ \pi f(0)V-c V-k_{p}A^*V, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial V}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \gt 0. \end{cases} \end{equation} (3.2)

    Substituting V(x, t) = e^{\Lambda t}\psi(x) into (3.2), and we get the associated eigenvalue problem:

    \begin{equation} \begin{cases} \Lambda\psi(x) = D_V\Delta \psi(x)+ (\pi f(0)-c -k_{p}A^*)\psi(x), \ x\in \Omega, \\ \frac{\partial \psi(x)}{\partial\nu} = 0, \ x\in\partial\Omega. \end{cases} \end{equation} (3.3)

    By the same argument in [17,Theorem 7.6.1], we can show that the eigenvalue problem (3.3) admits a principal eigenvalue, denoted by \Lambda^{0} , which corresponds a strongly positive eigenfunction \psi^0(x) . In fact, one can show that \Lambda^0 = \pi f(0)-c -k_{p}A^* and the associated eigenfunction \psi(\cdot)\equiv 1 . Note that one can also adopt the theory developed in [21,Section 3] to define the basic reproduction number, \mathcal{R}_0^0 , for system (3.1). For this purpose, we assume \mathbf{F} = \pi f(0) and \mathbf{V} = c+k_{p}A^* . By [21,Theorem 3.4], it follows that

    \mathcal{R}_0^0 = \mathbf{F}\mathbf{V}^{-1} = \frac{\pi f(0)}{c+k_{p}A^*}.

    Putting k_{m} = 0 in (2.12), and it is easy to see that \mathcal{R}_0^0 = \mathcal{R}_0 when k_{m} = 0 . This is the reason why the reproduction number in this section is denoted by \mathcal{R}_0^0 . Further, it is easy to observe that

    \begin{equation} \mathcal{R}_0^0 \lt 1 \Leftrightarrow \Lambda^{0} \lt 0. \end{equation} (3.4)

    We impose the following condition:

    \begin{equation} \overline{A}: = \frac{p_{A}}{k_{p}}\geq A^*: = A_{m}(1-\frac{d_A}{r_{A}}) \ \mbox{and}\ r_{A} \gt d_A. \end{equation} (3.5)

    Let

    \mathcal{Y}_{P}: = \{(A^0, V^0)\in C(\bar{\Omega}, \mathbb{R}_{+}^2): A^0(x)\leq \overline{A}, \ \forall \ x \in \bar{\Omega}\}.

    Theorem 3.1. Assume that (3.5) holds. For any (A^0(\cdot), V^0(\cdot))\in \mathcal{Y}_{P} with A^0(\cdot)\not\equiv 0 , let (A(\cdot, t), V(\cdot, t)) be the solution of (3.1) with (A(\cdot, 0), V(\cdot, 0)) = (A^0(\cdot), V^0(\cdot)) . If \mathcal{R}_0^0 < 1 , then we have

    \lim\limits _{t \rightarrow \infty}(A(x, t), V(x, t)) = (A^{*}, 0), \ \mathit{\mbox{uniformly for}}\ x \in \overline{\Omega}.

    Proof. Assume \mathcal{R}_0^0 < 1 , that is, \Lambda^{0} < 0 (see (3.4)). Then there exists \xi_0 > 0 such that \Lambda_{\xi_0} < 0 , where \Lambda_{\xi_0} is the principal eigenvalue of the following eigenvalue problem:

    \begin{equation} \begin{cases} \Lambda\psi(x) = D_V\Delta \psi(x)+ [\pi f(0)-c -k_{p}(A^*-\xi_0)]\psi(x), \ x\in \Omega, \\ \frac{\partial \psi(x)}{\partial\nu} = 0, \ x\in\partial\Omega. \end{cases} \end{equation} (3.6)

    The first equation of (3.1) can be rewritten as follows

    \frac{\partial A}{\partial t} = D_A\Delta A+k_{p}[\overline{A}-A](1+\theta)V+\frac{r_{A}}{A_{m}}[A^{*}-A]A.

    From (3.5), we see that

    k_{p}[\overline{A}-A](1+\theta)V+\frac{r_{A}}{A_{m}}[A^{*}-\overline{A}\ ]\overline{A} \lt 0.

    Then it is not hard to show that \mathcal{Y}_{P} is a positively invariant set for system (3.1). Thus,

    [p_{A}-k_{p}A(x, t)](1+\theta)V(x, t)\geq 0, \ \forall \ x\in\Omega, \ t\geq 0.

    In view of the first equation of (3.1), we see that

    \begin{equation} \begin{cases} \frac{\partial A}{\partial t}\geq D_A\Delta A+r_{A}A(1-\frac{A}{A_{m}})-d_{A}A, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial A}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \gt 0, \end{cases} \end{equation} (3.7)

    and hence,

    \liminf\limits_{t\rightarrow \infty}A(x, t)\geq A^{*}, \ \mbox{uniformly for}\ x \in \overline{\Omega}.

    Therefore, we may choose t_1 > 0 such that

    A(x, t)\geq A^{*}(x)-\xi_0, \ \mbox{uniformly for}\ x \in \overline{\Omega}, \ t\geq t_1.

    In view of the second equation of (3.1), we see that

    \begin{equation} \begin{cases} \frac{\partial V}{\partial t}\leq D_V\Delta V+ \pi f(0)V-c V-k_{p}( A^{*}(x)-\xi_0)V, \ x\in \Omega, \ t\geq t_1, \\ \frac{\partial V}{\partial\nu} = 0, \ x\in\partial\Omega, \ t\geq t_1, \end{cases} \end{equation} (3.8)

    where we have used the fact that f(V)\leq f(0), \ \forall \ V \geq 0 . Assume that \psi_{\xi_0}(x) is a strongly positive eigenfunction corresponding to \Lambda_{\xi_0} , and there exists \hat{C} > 0 such that V(x, t_1)\leq\hat{C}\psi_{\xi_0}(x), \ \forall \ x \in \overline{\Omega} . From (3.8), the comparison principle implies that

    V(x, t)\leq \hat{C}e^{\Lambda_{\xi_0}(t-t_1)}\psi_{\xi_0}(x), \ \forall \ t \geq t_1, \ x\in\bar{\Omega}.

    Since \Lambda_{\xi_0} < 0 , it follows that

    \lim\limits_{t\rightarrow \infty}V(x, t) = 0, \ \mbox{uniformly for}\ x \in \overline{\Omega}.

    Then A(x, t) in (3.1) is asymptotic to system (2.4). Using A^0(\cdot)\not\equiv 0 and the theory for asymptotically autonomous semiflows (see, e.g., [28,Corollary 4.3]), we have

    \lim\limits_{t\rightarrow \infty}A(x, t) = A^*, \ \mbox{uniformly for}\ x \in \overline{\Omega}.

    The proof is complete.

    This study presents a reaction-diffusion system (1.3) modeling HBV infection, which consists of five compartments of populations, namely, target cells ( T ), infected cells ( I ), free virus ( V ), free antibody ( A ), and virus-antibody complexes ( X ). In system (1.3), we assume that only free virus ( V ), free antibody ( A ), and virus-antibody complexes ( X ) can diffuse, and the host cells (target and infected cells) do not have the ability to move. Thus, the governing equations are coupled by ODEs and PDEs. Due to the lack of diffusion terms of target cells ( T ) and infected cells ( I ) in (1.3), the steady-state solutions involved T and I can be explicitly expressed by free virus ( V ). Thus, investigating the existence of steady-state solutions of (1.3) is equivalent to the study of steady-state solutions of system (1.9).

    The standard approach in seeking for positive steady-state solutions of system (1.9) is applying theory of bifurcation to the associated elliptic equations of (1.9). Instead, we adopt dynamical approach in the analysis of (1.9) in the current paper. We define an reproduction number, \mathcal{R}_0 , for system (1.9), and we show that system (1.9) is uniformly persistent and it admits at least one (componentwise) positive steady state when \mathcal{R}_0 > 1 (see Theorem 2.1). Mathematically, it is more difficult to investigate the elimination of HBV in system (1.9). Putting k_{m} = 0 in system (1.9), the equation of X in (1.9) is decoupled from the other equations, and we directly study the system (3.1) for the extinction case of HBV. Imposing the assumption (3.5), we can show that HBV will die out for (3.1) if the associated reproduction number \mathcal{R}_0^0 is less than one (Theorem 3.1). Here, we also raise some challenging problems related to system (1.9), which can be future research directions:

    ● The impact of the diffusion coefficients D_X and D_V on the basic reproduction number \mathcal{R}_0 ;

    ● The dynamics of system (1.9) for the critical case when \mathcal{R}_0 = 1 ;

    ● The uniqueness and the global attractiveness of the positive steady state of system (1.9) if it exists;

    ● The asymptotic profile of positive steady state of system (1.9) when the diffusion rates D_X and D_V both tend to zero.

    In order to simplify the modeling in system (1.3), we have ignored two compartments of populations, namely, free subviral particles ( S ) and subviral particles-antibody complexes ( X_{s} ) in [1] by assuming that subviral particles S (resp. subviral particles-antibody complexes X_{s} ) is proportional to the concentration of free virus V (resp. virus-antibody complexes X ) with a constant proportionality \theta . The authors in [1] developed another more complete model about HBV infection with antibody, which includes the interactions of target cells ( T ), infected cells( I ), free subviral particles ( S ), free antibody ( A ), virus-antibody complexes ( X ), subviral particles-antibody complexes ( X_{s} ), and free virus ( V ). After we add spatial variations into such system, we shall investigate the following more realistic and challenging case in the future:

    \begin{equation} \begin{cases} \frac{\partial T}{\partial t} = r T(1-\frac{T+I}{T_{m}})-\beta VT, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial I}{\partial t} = \beta VT-\delta I, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial A}{\partial t} = D_A\Delta A+p_{A}(V+S)+r_{A}(x)A(1-\frac{A}{A_{m}})+k_{m}X\\ \ \ \ \ \ \ \ \ \ \ \ \ -k_{p}AV+k_{m}^{s}X_{S}-k_{p}^{s}AS-d_{A}(x)A , \ x\in \Omega, \ t \gt 0, \\ \frac{\partial X}{\partial t} = D_X\Delta X-k_{m}X+k_{p}AV-c_{AV}X, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial X_{S}}{\partial t} = D_{XS}\Delta X_{S}-k_{m}^{s}X_{S}+k^{s}_{p}AS-c_{AS}X_{S}, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial V}{\partial t} = D_V\Delta V+\pi I-c V+k_{m}X-k_{p}AV, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial S}{\partial t} = D_S\Delta S+\pi\theta I-c_{s}S+k_{m}^{s}X_{s}-k_{p}^{s}AS, \ x\in \Omega, \ t \gt 0, \\ \frac{\partial A}{\partial\nu} = \frac{\partial X}{\partial\nu} = \frac{\partial X_S}{\partial\nu} = \frac{\partial V}{\partial\nu} = \frac{\partial S}{\partial\nu} = 0, \ x\in\partial\Omega, \ t \gt 0, \\ u(x, 0) = u^0(x), \ u = T, I, A, X, X_{S}, V, S, \ x\in\Omega. \end{cases} \end{equation} (4.1)

    The meanings of the parameters in system (4.1) were collected in [1,Table 1].

    We are grateful to three anonymous referees for their careful reading and helpful suggestions which led to significant improvements of our original manuscript. Research of FBW is supported in part by Ministry of Science and Technology, Taiwan; and National Center for Theoretical Sciences (NCTS), National Taiwan University; and Chang Gung Memorial Hospital (BMRPD18, NMRPD5J0201 and CLRPG2H0041). YCS is partially supported by Chang Gung Memorial Hospital (CLRPG2H0041). CLL is partially supported by Chang Gung Memorial Hospital (CRRPG2B0185, CRRPG2H0041, CRRPG2H0081, CLRPG2H0041).

    The authors declare there is no conflicts of interest.



    [1] S. M. Ciupe, R. M. Ribeiro, A. S. Perelson, Antibody responses during hepatitis B viral infection, PLOS Comput. Biol., 10 (2014), e1003730.
    [2] L. G. Guidotti, M. Isogawa, F. V. Chisari, Host-virus interactions in hepatitis B virus infection, Curr. Opin. Immunol., 36 (2015), 61-66.
    [3] S. Bonhoeffer, R. M. May, G. M. Shaw, M. A. Nowak, Virus dynamics and drug therapy, Proc. Natl. Acad. Sci. USA, 94 (1997), 6971-6976.
    [4] M. A. Nowak, S. Bonhoeffer, A. M. Hill, R. Boehme, H. C. Thomas, Viral dynamics in hepatitis B virus infection, Proc. Natl. Acad. Sci. USA, 93 (1996), 4398-4402.
    [5] A. S. Perelson, D. E. Kirschner, R. de Boer, Dynamics of HIV infection of CD4 T cells, Math. Biosci., 114 (1993), 81-125.
    [6] A. S. Perelson, P. W. Nelson, Mathematical analysis of HIV-I dynamics in vivo, SIAM Rev., 41 (1999), 3-44.
    [7] A. S. Perelson, A. U. Neumann, M. Markowitz, J. M. Leonard, D. D. Ho, HIV-1 dynamics in vivo: virion clearance rate, infected cell life-span, and viral generation time, Science, 271 (1996), 1582-1586.
    [8] S. M. Ciupe, R. M. Ribeiro, P. W. Nelson, G. Dusheiko, A. S. Perelson, The role of cells refractory to productive infection in acute hepatitis B viral dynamics, Proc. Natl. Acad. Sci. USA, 104 (2007), 5050-5055.
    [9] S. M. Ciupe, R. M. Ribeiro, A. S. Perelson, Modeling the mechanisms of acute hepatitis B virus infection, J. Theor. Biol., 247 (2007), 23-35.
    [10] A. Kandathil, F. Graw, J. Quinn, H. Hwang, M. Torbenson, A. Perelson, et al., Use of laser capture microdissection to map hepatitis C virus-positive hepatocytes in human liver, Gastroenterol, 145 (2013), 1404-1413.
    [11] X. Ren, Y. Tian, L. Liu, X. Liu, A reaction-diffusion within-host HIV model with cell-to-cell transmission, J. Math. Biol., 76 (2018), 1831-1872.
    [12] W. Wang, T. Zhang, Caspase-1-Mediated Pyroptosis of the predominance for driving CD4+ T cells death: a nonlocal spatial mathematical model, Bull. Math. Biol., 80 (2018), 540-582.
    [13] X.-Q. Zhao, Dynamical Systems in Population Biology, second edition, Springer, New York, 2017.
    [14] T. W. Hwang, F.-B. Wang, Dynamics of a dengue fever transmission model with crowding effect in human population and spatial variation, Discrete Cont Dyn-B, 18 (2013), 147-161.
    [15] Y. Lou, X.-Q. Zhao, A reaction-diffusion malaria model with incubation period in the vector population, J. Math. Biol., 62 (2011), 543-568.
    [16] R. Martin, H. L. Smith, Abstract functional differential equations and reaction-diffusion systems, Trans. Amer. Math. Soc., 321 (1990), 1-44.
    [17] H. L. Smith, Monotone Dynamical Systems:An Introduction to the Theory of Competitive and Cooperative Systems, Math. Surveys Monogr 41, American Mathematical Society Providence, RI, 1995.
    [18] H. C. Li, R. Peng, F.-B. Wang, Varying Total Population Enhances Disease Persistence: Qualitative Analysis on a Diffusive SIS Epidemic Model, J. Differ. Equations, 262 (2017), 885-913.
    [19] J. Hale, Asymptotic behavior of dissipative systems, American Mathematical Society Providence, RI, 1988.
    [20] W.-M. Ni, The mathematics of diffusion, CBMS-NSF Regional Conference Series in Applied Mathematics, 82. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2011.
    [21] W. Wang, X.-Q. Zhao, Basic reproduction numbers for reaction-diffusion epidemic models, SIAM J. Appl. Dyn. Syst., 11 (2012), 1652-1673.
    [22] P. van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29-48.
    [23] O. Diekmann, J. A. P. Heesterbeek, J. A. J. Metz, On the definition and the computation of the basic reproduction ratio R0 in the models for infectious disease in heterogeneous populations, J. Math. Biol., 28 (1990), 365-382.
    [24] H. R. Thieme, Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity, SIAM J. Appl. Math., 70 (2009), 188-211.
    [25] M. H. Protter, H. F. Weinberger, Maximum Principles in Differential Equations, Springer-Verlag, 1984.
    [26] H. L. Smith, X.-Q. Zhao, Robust persistence for semidynamical systems, Nonlinear Anal., 47 (2001), 6169-6179.
    [27] P. Magal, X.-Q. Zhao, Global attractors and steady states for uniformly persistent dynamical systems, SIAM J. Math. Anal., 37 (2005), 251-275.
    [28] H. R. Thieme, Convergence results and a Poincare-Bendixson trichotomy for asymptotically autonomous differential equations, J. Math. Biol., 30 (1992), 755-763.
  • This article has been cited by:

    1. Vladimir P. Zhdanov, Joshua A. Jackman, Analysis of the initiation of viral infection under flow conditions with applications to transmission in feed, 2020, 196, 03032647, 104184, 10.1016/j.biosystems.2020.104184
    2. Shawn A. Means, Md A. Ali, Harvey Ho, Illuminating HBV with multi-scale modeling, 2023, 3, 2674-0702, 10.3389/fsysb.2023.1045754
  • 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(3787) PDF downloads(424) Cited by(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog