Processing math: 95%
Review Topical Sections

Molecular modeling studies of peptoid polymers

  • Peptoids, or poly-N-substituted glycines, are synthetic polymers composed of a protein backbone with side chains attached to the nitrogen atoms rather than the α-carbons. Peptoids are biomimetic and protease resistant and have been explored for a variety of applications including pharmaceuticals and coatings. They are also foldamer-type materials that can adopt diverse structures based on the sequences of their side chains. Design of new peptoid sequences may lead to the creation of many interesting materials. Given the large number of possible peptoid side chains, computer models predicting peptoid structure-side chain relationships are desirable. In this paper, we provide a survey of computational efforts to understand and predict peptoid structures. We describe simulations at several levels of theory, along with their assumptions and results. We also discuss some challenges for future peptoid computational research.

    Citation: Laura J. Weiser, Erik E. Santiso. Molecular modeling studies of peptoid polymers[J]. AIMS Materials Science, 2017, 4(5): 1029-1051. doi: 10.3934/matersci.2017.5.1029

    Related Papers:

    [1] Cameron Browne . Immune response in virus model structured by cell infection-age. Mathematical Biosciences and Engineering, 2016, 13(5): 887-909. doi: 10.3934/mbe.2016022
    [2] Shaoli Wang, Jianhong Wu, Libin Rong . A note on the global properties of an age-structured viral dynamic model with multiple target cell populations. Mathematical Biosciences and Engineering, 2017, 14(3): 805-820. doi: 10.3934/mbe.2017044
    [3] Ran Zhang, Shengqiang Liu . Global dynamics of an age-structured within-host viral infection model with cell-to-cell transmission and general humoral immunity response. Mathematical Biosciences and Engineering, 2020, 17(2): 1450-1478. doi: 10.3934/mbe.2020075
    [4] Sukhitha W. Vidurupola, Linda J. S. Allen . Basic stochastic models for viral infection within a host. Mathematical Biosciences and Engineering, 2012, 9(4): 915-935. doi: 10.3934/mbe.2012.9.915
    [5] Jinliang Wang, Xiu Dong . Analysis of an HIV infection model incorporating latency age and infection age. Mathematical Biosciences and Engineering, 2018, 15(3): 569-594. doi: 10.3934/mbe.2018026
    [6] Churni Gupta, Necibe Tuncer, Maia Martcheva . Immuno-epidemiological co-affection model of HIV infection and opioid addiction. Mathematical Biosciences and Engineering, 2022, 19(4): 3636-3672. doi: 10.3934/mbe.2022168
    [7] Yuyi Xue, Yanni Xiao . Analysis of a multiscale HIV-1 model coupling within-host viral dynamics and between-host transmission dynamics. Mathematical Biosciences and Engineering, 2020, 17(6): 6720-6736. doi: 10.3934/mbe.2020350
    [8] Ali Moussaoui, Vitaly Volpert . The impact of immune cell interactions on virus quasi-species formation. Mathematical Biosciences and Engineering, 2024, 21(11): 7530-7553. doi: 10.3934/mbe.2024331
    [9] Andrei Korobeinikov, Conor Dempsey . A continuous phenotype space model of RNA virus evolution within a host. Mathematical Biosciences and Engineering, 2014, 11(4): 919-927. doi: 10.3934/mbe.2014.11.919
    [10] Conrad Ratchford, Jin Wang . Multi-scale modeling of cholera dynamics in a spatially heterogeneous environment. Mathematical Biosciences and Engineering, 2020, 17(2): 948-974. doi: 10.3934/mbe.2020051
  • Peptoids, or poly-N-substituted glycines, are synthetic polymers composed of a protein backbone with side chains attached to the nitrogen atoms rather than the α-carbons. Peptoids are biomimetic and protease resistant and have been explored for a variety of applications including pharmaceuticals and coatings. They are also foldamer-type materials that can adopt diverse structures based on the sequences of their side chains. Design of new peptoid sequences may lead to the creation of many interesting materials. Given the large number of possible peptoid side chains, computer models predicting peptoid structure-side chain relationships are desirable. In this paper, we provide a survey of computational efforts to understand and predict peptoid structures. We describe simulations at several levels of theory, along with their assumptions and results. We also discuss some challenges for future peptoid computational research.


    Several mathematical models in the field of viral dynamics simply regard the host of an infection as a well-mixed compartment [1,2,3] and aim to theoretically determine the threshold, from an overall perspective for determining whether an infection is extinct or persistent. Under the assumption of spatial homogeneity, various functions of different tissues and organs may be overlooked. Given the complexity of the human body, it is essential to regard the host body as composed of multiple compartments and to explore the effect of connection between compartments on the infection [4,5,6].

    Viral propagation in human immunodeficiency virus (HIV) infection occurs mainly in lymphoid tissues [5,7], which are composed of primary and secondary lymphoid organs. The primary lymphoid organs include the thymus and bone marrow, which are where lymphocytes are generated; the lymph nodes and the spleen constitute the secondary lymphoid organs, which maintain mature naive lymphocytes and initiate adaptive immune response [8]. Between these tissues, connections are constructed through afferent or efferent lymphatics (or both) [9,10,11], by which cluster or differentiation 4 (CD4)-positive or CD8-positive T lymphocytes and other cells can migrate to execute their functions. This gives rise to the need to explore how the connection of various tissues influences the viral dynamics and curative effect of treatment.

    To analyze the spatiotemporal patterns of viral progression, the authors in [12] proposed a reaction-diffusion model by fitting the data to a radially symmetric propagation and numerically identified the initial propagation of the infection front followed by a state of stagnation due to an antiviral response. Infected CD4 T cells can actively migrate over different lymph nodes [13]. Accordingly, another scenario for studying the spatial heterogeneity of viral dynamics is using multi-compartment models to describe the occurrence of infection in various organs. In [14], a heterogeneous environment with various parameters for several grids was constructed, wherein the virus randomly spreading between the grids was modeled as local diffusion. The main result showed that the local dispersion of a virus can reduce the amplitude of viral oscillation. In a recent study [5], the spread of HIV infection throughout the lymphoid tissue was considered from a perspective of network structure. That study numerically determined that HIV infection may persist even under drug administration because a portion of lymphoid tissues fails to receive a sufficient amount of the dose. These studies clarify that considering the spatial heterogeneity is essential in studying viral dynamics.

    Most within-host virus models assume transmission solely through virus-to-cell infection, although it is now appreciated that there are multiple infection routes resulting in heterogenous kinetics. The cell-to-cell transmission process of HIV infection has been explored in a compartment such as the lymph nodes and the brain, because CD4+ T lymphocytes are densely agammaegated and frequently interact in lymphoid tissues, and the virus can disseminate through direct transmission from infected cells to uninfected cells through the HIV virological synapse [15,16]. Among the studies that have considered the delayed intracellular reaction but without distinction on infection age, HIV infection was explored in [1] by applying ordinary and delayed differential models, and sustained oscillations of infection were revealed for certain culture parameters. In [17], the infection-age-dependent infectivity of individuals was considered and the dynamics of HIV transmission were studied in a population group. The age-dependent production rate of viral particles and death rate of productively infected cells were considered in [18], in which an in vivo age-structured model of HIV infection was constructed. In order to explore distinct intracellular and extracellular infection kinetics, it is necessary to extend previous virus models by incorporating both age-since-infection and a multi-compartment environment.

    Global convergence in viral dynamics models is crucial for predicting the evolution of viral infection, and age-structure of infected cells tends to complicate matters. Our single-compartment model considered in this paper is related to that in [19], wherein a susceptible-infective (SI) epidemic model was incorporated with infection-age and an incidence function derived from the law of mass action. In that study, the system was addressed using an integrated semigroup to determine the persistence of disease; the results demonstrated global convergence to either disease-free or endemic equilibrium through the application of Lyapunov functionals. Notably, the probability of disease transmission was found to depend on the age of infection [19]. Another difficulty in the field arises because of the multigroup structure of the models. In the recent study [20], the authors investigated a multigroup (susceptible-infective-recovered) SIR model with age structure and elaborated on the convergence dynamics when the morality and removal rate were age-independent; because of the constant total host population, the model involved a constant boundary condition. Furthermore, in [21], an aged model in a network environment was considered, and both the age-dependent transmission rate and degree of the network connectivity were found to affect the spread of disease; on the basis of the constant total population, the authors demonstrated that global convergence to the endemic equilibrium occurs when it exists. In [22], a two-group model was adopted to investigate the globally asymptotic behavior of a SI epidemic model incorporating the age of infection. The Lyapunov functional employed in that study was skillfully constructed for the two-group structure. The age-dependent infectivity of infected cells renders it impossible to rewrite the viral infection model as the classic (without aged-structure) SIS or SIR epidemiological models or variant predator-prey models, and it increases the difficulty of exploring the global convergence dynamics, especially in deriving convergence to the endemic equilibrium state.

    In this paper, we construct and analyze a virus model consisting of multiple compartments with distinct cell infection-age structured infectivity kernels describing infection spread through various organs. In section 2, we detail the model formulation, along with establishing boundedness and compactness of the generated semiflow. In sections 3 and 4, we derive the basic reproduction number, and show that it is a global threshold determining the viral persistence or extinction. Section 5 concerns extending analysis to non-strongly connected networks through a sequence of threshold values controlling the infection pattern in the whole system. Finally in Section 6, we numerically investigate the viral dynamics for different connection topologies, along with examples of distinct cell infectivity kernels based on data for cell-free and cell-to-cell infection modes.

    An infected host with n compartments is described by

    dTk(t)dt=λkdkTk(t)αkTk(t)0pk(a)ik(t,a)dajN0mjkTk(t)+jN0mkjTj(t),ik(t,a)t+ik(t,a)a=δk(a)ik(t,a)jN0mjkik(t,a)+jN0mkjij(t,a), (2.1)

    for kN0:={1,2,,n}, with the boundary condition

    ik(t,0)=αkTk(t)0pk(a)ik(t,a)da, (2.2)

    and the initial condition

    Tk(0)=Tk0>0,ik(0,a)=ik0(a)L1+((0,+),R), (2.3)

    where L1+((0,+),R) is the nonnegative cone of L1((0,+),R). In the kth compartment, at time t, Tk(t) denotes the concentration of healthy cells, and ik(t,a) is that of infected cells with age a since infection. The parameters λk and dk respectively represent the production and death rates of healthy cells, and αk accounts for the rate of both virus-to-cell and cell-to-cell transmission. Here, we neglect the virus compartment due to its fast kinetics. The function pk(a) accounts for the net infectivity of infected cells from both infection routes, variant in the infection age, and δk(a) represents the death rate of infected cells. All these parameters describe the viral environment in the kth compartment. They may be different from compartment to compartment because of individual function. For instance, the infectivity kernel pk(a) may change according to primary transmission mode (cell-to-cell versus viral), or infection induced death rate, δk(a), may be different in distinct compartments based on predominant target cell type (e.g. T-cell in blood vs. macrophage in brain). The parameter mjk, jk, denotes the migration rate of the cells from the kth compartment to the jth compartment. Assuming no loss in the process of migration, it satisfies that nj=1mjk=0 and then mkk=njk,j=1mjk. As mentioned in Section 1, the migration may be asymmetric between any two compartments. In addition, there is no isolated compartments to be considered, however this assumption is relaxed later in Section 5.

    Here, we assume that the migration matrix M=(mkj)n×n is irreducible and, for kN0,

    (A1) λk,dk,αk>0;

    (A2) δk()C([0,+),R+), dkδk(a)δk,max<, for a0;

    (A3) pk()C([0,+),R+), 0<pk(a)pk,max<, aI0, for some finite interval I0[0,+), and pk(a)=0 for aI0.

    Denoting T(t)=(T1(t),,Tn(t))τ and i(t,a)=(i1(t,a),,in(t,a))τ, where ()τ means the transpose of a vector, the system (2.1) is equivalent to the vector form:

    dT(t)dt=ΛDT(t)Γdiag(T(t))0P(a)i(t,a)da+MT(t),i(t,a)t+i(t,a)a=˜D(a)i(t,a)+Mi(t,a), (2.4)

    with the boundary condition

    i(t,0)=Γdiag(T(t))0P(a)i(t,a)da

    and the initial condition T(0)=T0>0,i(0,)=i0()L1+((0,+),Rn), where Λ=(λ1,,λn)τ, D=diag(d1,,dn), Γ=diag(α1,,αn), diag(T(t))=diag(T1(t),,Tn(t)), ˜D(a)=diag(δ1(a),,δn(a)), P(a)=diag(p1(a),,pn(a)) and

    M=(jN0,j1mj1m12m1nm21jN0,j2mj2m2nmn1mn2jN0,jnmjn). (2.5)

    Following the approach proposed in [23], we reformulate (2.4) as a semilinear Cauchy problem. To this end, we first consider the extended state spaces

    X=Rn×Rn×L,L=L1((0,+),Rn),X0=Rn×{0Rn}×L,X+=Rn+×Rn+×L+,L+=L1((0,+),Rn+),X0+=X0X+,

    and equip the space X with the norm

    u=kN0(|uk|+|vk|+0|wk(a)|da), (2.6)

    for u=(u1,,un,v1,,vn,w1(),,wn())τX. Define the linear operator A:Dom(A)XX, Dom(A)=Rn×{0Rn}×W1,1((0,+),Rn), by

    A(T0Rni)=((D+M)Ti(0)i(˜D()M)i),

    where W1,1 is a Sobolev space, and the operator F:X0X by

    F(T0Rni)=(ΛΓdiag(T(t))0P(a)i(t,a)daΓdiag(T(t))0P(a)i(t,a)da0L).

    By denoting u(t)=(T(t),0Rn,i(t))τ, we regard system (2.4) as the abstract Cauchy problem:

    du(t)dt=Au(t)+F(u(t)),fort0,u(0)X0+, (2.7)

    and assert, by [23,Theorems 2 and 3], that there exists a unique solution semiflow Ψ(t):X0+X0 related to system (2.7).

    Next, by employing the boundary condition and the initial condition, we write (2.4) into the Volterra-type equation,

    dT(t)dt=ΛDT(t)Γdiag(T(t))0P(a)i(t,a)da+MT(t),i(t,a)={exp(aat(˜D(s)M)ds)i0(at),ifa>t,exp(a0(˜D(s)M)ds)Q(ta),ifat,

    where Q(t)=i(t,0) and satisfies

    Q(t)=Γdiag(T(t))t0P(a)exp(a0(˜D(s)M)ds)Q(ta)da+Γdiag(T(t))tP(a)exp(a0(˜D(s)M)ds)i0(at)da. (2.8)

    Denote

    Ω(a0,a)=(Ωkj(a0,a))n×n=exp(aa0(˜D(s)M)ds).

    First, we show a priority about the matrix Ω(a0,a).

    Lemma 2.1. For fixed a00, there exists a constant γ>0 such that

    0Ωkj(a0,a)γed_(aa0) (2.9)

    for all k,jN0 and aa0, where d_=minkN0{dk/2}>0.

    Proof. Denote ˜d=maxkN0{δk,max}+maxkN0{mkk} and I as the n×n identity matrix. Then, for aa0, the matrix aa0˜D(s)ds+˜d(aa0)I+(aa0)M is nonnegative and

    Ω(a0,a)=exp(˜d(aa0)Iaa0˜D(s)ds+˜d(aa0)I+(aa0)M)=exp(˜d(aa0)I)exp(aa0˜D(s)ds+˜d(aa0)I+(aa0)M)0.

    In addition,

    Ω(a0,a)=exp(d_(aa0)Iaa0(˜D(s)d_I)ds+(aa0)M)=exp(d_(aa0)I)exp(aa0(˜D(s)d_I)ds+(aa0)M)exp(d_(aa0)I)exp((aa0)(d_I+M)),

    where the last inequality holds since the fact that the quasi-positive matrices A=(akj)B=(bkj) with 0akj=bkj for all kj implies exp(A)exp(B). Since the matrix d_I+M is strictly diagonally dominant and each diagonal entry is negative, all its eigenvalues, say μj, j=1,2,,m, mn, have negative real parts [24,Theorem 6.1.10]. We write d_I+M in its Jordan canonical form, J, with d_I+M=PJP1, J=diag(J1,,Jm) and each Jj is a standard Jordan block related to the eigenvalue μj. Obviously,

    exp((aa0)(d_I+M))=Pexp((aa0)J)P1,exp((aa0)J)=diag(exp((aa0)J1),,exp((aa0)Jm)).

    Since the real part of μj is negative,

    exp((aa0)Jj)=(e(aa0)μj000000e(aa0)μj)exp(0aa000aa0000)0,

    as aa0+. Thus, there exists a positive constant γ such that exp((aa0)(d_I+M))γI for all aa0, where I is the matrix with a value of 1 for all entries, and the assertion is true.

    Denote

    Y=Rn×L,Y+=Rn+×L+,Y0={0Rn}×L,

    and equip the space Y with the norm

    (T,i())=kN0(|Tk|+0|ik(a)|da).

    Then the semiflow generated by (2.1) is point dissipative, as demonstrated in the following lemma.

    Lemma 2.2. The solution semiflow generated by (2.1) is point dissipative. Explicitly, the subset of phase space

    Ξ={(T(t),i(t,a))Y+|kN0Tk(t)+kN00ik(t,a)dakN0λk/dmin},

    is positively invariant and attracts all nonnegative solutions, where dmin=minkN0{dk}.

    Proof. It is easy to show that all solutions of (2.1) are nonnegative. Denote the total infected cells by I(t)=0i(t,a)da and the total cells by N(t)=T(t)+I(t). Integrating the i-equation in (2.4) with respect to a leads to

    dI(t)dt=Γdiag(T(t))0P(a)i(t,a)da0˜D(a)i(t,a)da+MI(t).

    Obviously,

    dN(t)dt=ΛDT(t)0˜D(a)i(t,a)da+MT(t)+MI(t),

    and then

    dkN0Nk(t)dt=kN0λkkN0dkTk(t)kN00δk(a)ik(t,a)dakN0λkdminkN0Nk(t).

    By a comparison principle, we derive that

    lim supt+kN0Nk(t)kN0λk/dmin.

    Since each solution of (2.1) remains nonnegative, as previously mentioned, the assertion holds true.

    Deriving the uniform persistence and global properties of the solution dynamics is essential in studying viral infection and it is necessary to show that the semiflow Φ(t) generated by (2.1) is asymptotically smooth. Denote

    ˜Pk(t)=0pk(a)ik(t,a)da,kN0.

    By integrating along the characteristic lines and incorporating the boundary condition and the initial condition of the model, we write equation (2.1) in the following equivalent Volterra integral equation,

    dTk(t)dt=λkdkTk(t)αkTk(t)0pk(a)ik(t,a)dajN0mjkTk(t)+jN0mkjTj(t),ik(t,a)={jN0Ωkj(at,a)ij0(at),ifa>t,jN0Ωkj(0,a)αj˜Pj(ta)Tj(ta),ifat. (2.10)

    Few studies have investigated asymptotically smooth semiflows in a system incorporating both the age structure and the migration of a population. Here, we show the property in the semiflow Φ(t) by using Lemma 1 and a result reported in [25,Lemma 3.2.3].

    Lemma 2.3. The semiflow Φ(t) generated by (2.10) ((2.1)) is asymptotically smooth.

    Proof. It is easy to check that each function ˜Pk() is bounded, say with the bound ˉPk, and Lipschitz continuous on R+, say with the Lipschitz constant lk (see [26]). Define Φ(t)=Λ1(t)+Λ2(t), where

    (Λ1(t)x)(a)={(0Rn,0L),t>a,(0Rn,i(t,a)),at;(Λ2(t)x)(a)={(T(t),i(t,a)),t>a,(T(t),0L),at,

    for xY. From (2.10) and Lemma 2.1,

    Λ1(t)x=kN0t|ik(t,a)|da=kN0tjN0Ωkj(at,a)ij0(at)dakN0tjN0γed_tij0(at)danγed_tx.

    Define Δ(r,t)=nrγed_t. Then Δ(r,t)0 as t+ and Λ1(t)xΔ(r,t) for xr.

    Let BY be a bounded subset such that Φ(t)BB. Choose r0>0 such that xr0 for all xB. Note, from Lemma 2.2, that (T0,i0())B{T(t)} is bounded in Rn and then is precompact in Rn. Hence, in order to show that Λ2(t)B is precompact, it suffices to verify that the set ˜Λ2(t)B is precompact for

    (˜Λ2(t)x)(a)={i(t,a),t>a,0L,at. (2.11)

    From [27], it is sufficient to verify the following conditions:

    (ⅰ) limh0kN00{|ik(t,a)ik(t,a+h)|}da=0 uniformly for i(t,a)˜Λ2(t)B.

    (ⅱ) limh+kN0h{|ik(t,a)|}da=0 uniformly for i(t,a)˜Λ2(t)B.

    From (2.11), kN0h{|ik(t,a)|}da=0 for ht, and then the criterion (ⅱ) holds for the set ˜Λ2(t)B. To show the criterion (ⅰ), we directly calculate for kN0 and ht

    0|ik(t,a)ik(t,a+h)|da=th0|jN0Ωkj(0,a)αjTj(ta)˜Pj(ta)jN0Ωkj(0,a+h)αjTj(tah)˜Pj(tah)|da+tth|jN0Ωkj(0,a)αjTj(ta)˜Pj(ta)|dajN0th0γed_a|αjTj(ta)˜Pj(ta)αjTj(ta)˜Pj(ta)Ωkj(0,a+h)Ωkj(0,a)|da+jN0th0γed_aΩkj(0,a+h)Ωkj(0,a)|αjTj(ta)˜Pj(ta)αjTj(tah)˜Pj(tah)|da+jN0tth|Ωkj(0,a)αjTj(ta)˜Pj(ta)|da=:C(Φ(t)x,h)+D(Φ(t)x,h)+E(Φ(t)x,h). (2.12)

    Note that

    C(Φ(t)x,h)=jN0th0γed_aαjTj(ta)˜Pj(ta)|1Ωkj(0,a+h)Ωkj(0,a)|dajN00γed_aαjr0ˉPj|1Ωkj(0,a+h)Ωkj(0,a)|da, (2.13)

    and

    E(Φ(t)x,h)jN0γαjr0ˉPjh. (2.14)

    According to the assumption (A2), there exists a positive constant h0 such that Ωkj(0,a+h)Ωkj(0,a)<2 for |h|<h0 since that Ωkj(0,a) is continuous in a. Hence,

    D(Φ(t)x,h)jN002γed_adasups[h,t][αjTj(s)|˜Pj(s)˜Pj(sh)|+αj˜Pj(sh)|Tj(s)Tj(sh)|] (2.15)

    Solve the equation (2.10) to obtain, for jN0,

    Tj(t)=Tj(0)+t0[λjdjTj(ξ)αjTj(ξ)˜Pj(ξ)+lN0mjlTj(ξ)]dξ.

    Accordingly, we have that

    |Tj(s)Tj(sh)|ssh[|λjdjTj(ξ)|+αjTj(ξ)˜Pj(ξ)+lN0|mjl|Tj(ξ)]dξmaxη[0,r0](|λjdjη|+αjηˉPj+lN0|mjl|η)h=:Cjh. (2.16)

    From (2.15), (2.16) and the Lipschitz constant, lk, of ˜Pk, it holds that

    D(Φ(t)x,h)02γed_adajN0(αjr0lj+αjˉPjCj)h. (2.17)

    By (2.12)-(2.14) and (2.17), it concludes that the criterion (ⅱ) holds for the set ˜Λ2(t)B, and then ˜Λ2(t)B is compact. From [25,Lemma 3.2.3], the assertion is true.

    From Lemma 2.2, the semiflow Φ(t) is point dissipative and then there exists a positively invariant absorbing set under the semiflow Φ(t), and then Φ(t) maps any bounded subset of Y+ to a precompact set in Y+ and is then compact for any t>0. From Lemma 2.3, the semiflow is asymptotically smooth. Combining these properties and the result on the existence of global attractors in [25,Theorem 3.4.6] (or see [28,Theorem 2.6]), the following lemma is implied.

    Lemma 2.4. The solution semiflow Φ(t) generated by system (2.1) in Y+ admits a compact global attractor AY+.

    First, we demonstrate the existence of infection-free equilibrium by examining the convergence dynamics of the no-infection model.

    Lemma 3.1. ([29,Sec. 2]) Consider the no-infection model

    d˜Tk(t)dt=λkdk˜Tk(t)jN0mjk˜Tk(t)+jN0mkj˜Tj(t),˜Tk(0)=Tk0. (3.1)

    There exists a positive equilibrium (ˉT1,,ˉTn) that is globally asymptotically stable with respect to Rn+.

    Hence, the model (2.1) admits an infection-free equilibrium ˉE=(ˉT,0Rn)τ=(ˉT1,,ˉTn,0,,0)τ. The linearization of (2.4) at the equilibrium ˉE is the system

    dT(t)dt=Γdiag(ˉT)0P(a)i(t,a)daDT(t)+MT(t),i(t,a)t+i(t,a)a=(˜D(a)M)i(t,a),i(t,0)=Γdiag(ˉT)0P(a)i(t,a)da. (3.2)

    To determine the basic reproduction number of (2.1), we consider the decoupled i-equation in (3.2):

    i(t,a)t+i(t,a)a=(˜D(a)M)i(t,a),i(t,0)=Γdiag(ˉT)0P(a)i(t,a)da,i(0,)=i0()L+((0,+),Rn). (3.3)

    Define the operators ˆA:L+L+ by

    ˆA(a)i(a)=(˜D(a)+M)i(a),

    and

    A(0f)=(f(0)f+ˆA()f),B(0f)=(Γdiag(ˉT)0P(a)f(a)da0). (3.4)

    Let ˜Φ=A+B and ˜Φ0 be the restriction of ˜Φ in Y0; that is, Dom(˜Φ0)={(0,f)τ:˜ΦfY0}. It is easy to see that (0,f)τDom(˜Φ0) if it holds that f(0)=0P(a)f(a)da. From the argument in [30,Sec. 6], the basic reproduction number of system (2.1) is

    R0=ρ(BA1),

    where ρ() denotes the spectral radius of an operator (also for that of a matrix). In fact, for (A1)(x,g)τ=(0,f)τ, it holds that

    f(a)=exp(a0(˜D(s)M)ds)x+a0exp(aν(˜D(s)M)ds)g(ν)dν,

    and then

    BA1(x,g)=(Γdiag(ˉT)0P(a)[exp(a0(˜D(s)M)ds)x+a0exp(aν(˜D(s)M)ds)g(ν)dν]da0).

    Note that the operator BA1 has the same spectral radius on Y and Rn×{0L}. Hence,

    R0=ρ(Θ),whereΘ=Γdiag(ˉT)0P(a)exp(a0(˜D(s)M)ds)da. (3.5)

    By the result in [30] (Theorems 3.16, 3.17), R0 plays the threshold value to determine local stability of the infection-free equilibrium, as stated in the following theorem.

    Theorem 3.1. The infection-free equilibrium ˉE is locally asymptotically stable (unstable) in (2.1) if R0<1 (R0>1).

    Now, as a concrete example we consider a two-compartment system, n=2, with δk()=δ() for k=1,2. Then Θ=Θ2, where

    Θ2=(α1ˉT100α2ˉT2)0(p1(a)00p2(a))exp(a0(δ(s)+m1m2m1δ(s)+m2)ds)da.

    Through a direct calculation, we derive that

    Θ2=(α1ˉT100α2ˉT2)×0(p1(a)exp(a0δ(s)ds)(1η1(a))p1(a)exp(a0δ(s)ds)η2(a)p2(a)exp(a0δ(s)ds)η1(a)p2(a)exp(a0δ(s)ds)(1η2(a)))da,

    where

    ηk(a)=mkm1+m2(1e(m1+m2)a),k=1,2.

    Since the matrix Θ2 is nonnegative and irreducible, ρ(Θ2) is real and positive [24]. Denote ρ(Θ2)=eξ for ξR. Then

    det(IeξΘ2)=0. (3.6)

    To explore the effect of the migration rate on the dynamics of (2.1), we choose pk(a)=epa, δk(a)=δ, mk=m, k=1,2, for positive p, δ and nonnegative m. Then (3.6) is equivalent to

    Gm(ξ):=[12α1ˉT1eξ(1p+δ+1p+δ+2m)1]×[12α2ˉT2eξ(1p+δ+1p+δ+2m)1]14α1α2ˉT1ˉT2e2ξ(1p+δ1p+δ+2m)2=0. (3.7)

    Referring to [31,32,33], we choose λ1=λ2=1, d1=d2=1, α1=0.00065, α2=0.00035, p=0.15, δ=0.4 and adjust m to determine the root of the function Gm. Figure 1 shows that (3.7) admits a positive root, which indicates that ρ(Θ2)>1, when m=0, and there is no nonnegative root to (3.7), which indicates that ρ(Θ2)<1, when m=2. Hence, different values of migration rate m may affect the extinction or persistence of virus in (2.1). Explicitly, in this case, strong circulation facilitates clearing the virus. In Section 6, we shall discuss the effect of the migration rate on the viral dynamics in more cases.

    Figure 1.  The graph of function Gm, m = 0, 2.

    When the reproduction number R0<1, we will demonstrate that the virus becomes extinct. First, the following lemma claims that infection can not occur under non-viable initial infection.

    Lemma 4.1. Consider the virus equation in an environment with a constant vector of healthy cells, T0,

    ik(t,a)t+ik(t,a)a=(˜D(a)M)i(t,a),i(t,0)=Γdiag(T0)0P(a)i(t,a)da, (4.1)

    with i0()L+ and 0P(a)i0(a)da=0. Then i(t,)=0() for all t0.

    Proof. The well know result in [34] of linear age-structured models reveals that

    i(t,a)={exp(aat(˜D(s)M)ds)i0(at),ifa>t,exp(a0(˜D(s)M)ds)˜Q(ta),ifat,

    where ˜Q satisfies

    ˜Q(t)=Γdiag(T0)t0P(a)exp(a0(˜D(s)M)ds)˜Q(ta)da+Γdiag(T0)tP(a)exp(a0(˜D(s)M)ds)i0(at)da.

    The initial condition, with 0P(a)i0(a)da=0, leads to

    ˜Q(t)=Γdiag(T0)t0P(a)exp(a0(˜D(s)M)ds)˜Q(ta)da.

    Since ˜Q(0)=0, we have that ˜Q(t)=0 for all t0, and then i(t,)=0() for all t0.

    With both the multi-compartmental structure and infection-age, it is a challenge to demonstrate the global dynamics in (2.1) by constructing a Lyapunov function [20]. However, the following theorem shows that the virus goes to extinction in system (2.1) when the basic reproduction number is less than or equal to unity.

    Theorem 4.1. When R01, the infection-free equilibrium ˉE is globally asymptotically stable in (2.1).

    Proof. We first show the assertion for the case with R0<1. By the standard comparison theorem, the solution of (2.1) satisfies that

    Tk(t)˜Tk(t), (4.2)

    where ˜Tk(t) is the solution to (3.1). From the assumption R0<1, there exists a sufficiently small ε>0 such that ρ(Θε)<1, where

    Θε=Γ(ˉT+εI)0P(a)exp(a0(˜D(s)M)ds)da.

    From Lemma 3.1 and (4.2), there is a t1>0 such that Tk(t)ˉTk+ε, kN0, for tt1. Since the matrix Θε is nonnegative and irreducible, ρ(Θε) is a positive eigenvalue which corresponds to a positive eigenvector, say v=(v1,,vn), with vτΘε=ρ(Θε)vτ. Define

    W(t)=vτ0P(a)a0exp(as(˜D(r)M)dr)i(t,s)dsda.

    We calculate the derivative of W(t) along the solution of (2.4) and derive, for tt1,

    dW(t)dt=vτ0P(a)a0exp(as(˜D(r)M)dr)ti(t,s)dsda=vτ0P(a)a0exp(as(˜D(r)M)dr)(si(t,s)(˜D(s)M)i(t,s))dsda=vτ0P(a)[exp(as(˜D(r)M)dr)i(t,s)]s=as=0da=vτ{0P(a)exp(a0(˜D(r)M)dr)i(t,0)da0P(a)i(t,a)da}=vτ{Γdiag(T(t))0P(a)exp(a0(˜D(r)M)dr)daI}0P(a)i(t,a)da=vτ[Γdiag(T(t))diag(ˉT+εI)1Γ1ΘεI]0P(a)i(t,a)da(ρ(Θε)vτvτ)0P(a)i(t,a)da0,

    and the equality holds if and only if 0P(a)i(t,a)da=0 for tt1. We claim that the largest positively invariant subset M0 of {dW(t)dt=0} is {(T,0())|TRn}. For ˜X(t)=(˜T(t),˜i(t,a))M0, a solution of (2.4), there is a ˜t1>0 such that 0P(a)˜i(t,a)da=0 for t˜t1. Denote ˜q0()=˜i(˜t1,) and consider the time-rescaled solution ˜Y(t)=˜X(t+˜t1), but remain the symbol ˜Y(t)=(˜T(t),˜i(t,a)). Then

    ˜i(t,a)ˆi(t,a),fort0, (4.3)

    where ˆi(t,a) satisfies (4.1) with T0=ˉT+εI and ˆi0()=˜q0(). Note that 0P(a)˜q0()da=0. From Lemma 4.1, ˆi(t,)=0() for all t0. From (4.3), M0 is in fact the set{(T,0())|TRn}. In addition, by using the theory of asymptotically autonomous systems [35,Theorem 2.3] and [36], we show that ˉE is globally asymptotically stable in system (2.1), when R0<1.

    When R0=1, we show it by contradiction. Suppose there is a k0N0 such that lim supt+0ik0(t,a)da>0, then by Lemma 4.1 we have

    ϵ0:=lim supt+˜Pk0(t)(=lim supt+0pk0(a)ik0(t,a)da)>0.

    From T-equation in (2.1),

    dTk0(t)dt=λk0dk0Tk0(t)αk0˜Pk0(t)Tk0(t)jN0mjk0Tk0(t)+jN0mk0jTj(t).

    We claim that lim supt+Tk0<ˉTk0. Denoting ˜P(t)=diag(˜P1(t),,˜Pn(t)), we rewrite the T-equation in system (2.1) into the vector form

    dT(t)dt=Λ+(D+MΓ˜P(t))T(t). (4.4)

    From the variation of constant formula, the solution to (4.4) satisfies

    T(t)=et(DM)T(0)+t0e(ts)(DM)Λdst0e(ts)(DM)Γ˜P(s)T(s)ds. (4.5)

    From Lemma 3.1, we have limt+(et(DM)T(0)+t0e(ts)(DM)Λds)=ˉT. We next estimate the third term in (4.5). Since the assumption ϵ0>0, there exist sequences of positive numbers {t(k0)l} and {τl} with limlt(k0)l=+ such that ˜Pk0(t)ϵ02 for t[t(k0)lτl,t(k0)l+τl].

    We claim that ˜Pk0(t) is a bounded function. From the assumption (A3), it holds true whenever 0ik0(t,a)tda is a bounded function of t. In fact, (2.1) and (2.2) imply that

    0ik0(t,a)tda=ik0(t,0)0δk0(a)ik0(t,a)dajN0mjk00ik0(t,a)da+jN0mk0j0ij(t,a)da=αk0Tk0(t)0pk0(a)ik0(t,a)da0δk0(a)ik0(t,a)dajN0mjk00ik0(t,a)da+jN0mk0j0ij(t,a)da.

    It shows that 0ik0(t,a)tda is a bounded function of t due to the assumptions (A2) and (A3) and Lemma 2.2. Hence ˜Pk0(t) is a bounded function and then the sequence {τl} can be chosen such that limlτl>0. It is easy to show that T(t) has a positive lower bound. In addition, since the matrix e(DM) is positive and irreducible, the k0-th component of t0e(ts)(DM)Γ˜P(s)T(s)ds has a positive lower bound. Therefore, there exist ε1>0 and t1>0 such that Tk0(t)<ˉTk0ε1 for t>t1.

    From the monotonicity of the spectral of nonnegative matrices ([37,Corollary (1.5) in Ch2]), and the assumption R0=1, we have ρ(Θε1)<1, where

    Θε1=Γdiag(ˉT1,,ˉTk01,ˉTk0ε1,ˉTk0+1,,ˉTn)×0P(a)exp(a0(˜D(s)M)ds)da.

    In addition, since the spectral of a matrix is continuous on each component, there exists a sufficiently small ε2>0 such that ρ(Θε1,ε2)<1, where

    Θε1,ε2=Γdiag(ˉT1+ε2,,ˉTk01+ε2,ˉTk0ε1,ˉTk0+1+ε2,,ˉTn+ε2)×0P(a)exp(a0(˜D(s)M)ds)da.

    By using the method in case of R0<1, we can show that ik(t,)0() as t+ for each kN0. This contradicts to the assumption on ik0(t,). Therefore, we conclude that ˉE is globally asymptotically stable in system (2.1), when R0=1.

    In this subsection, we study the uniform persistence for the model (2.4) when R0>1, by using the persistence theory developed in [38] (or see [19]). Recall that the semiflow Ψ(t) is generated by the Cauchy problem (2.7) which is equivalent to (2.4). Define

    U=Rn×{0Rn}׈U,ˆU={(w1(),,wn())τL+|kN00wk(a)da>0},U=X0+U,ˆU=L+ˆU.

    Lemma 4.2. The subsets U and U are positively invariant under the semiflow Ψ(t). Furthermore, limt+Ψ(t)u=ˉu=(ˉT,0Rn,0L)τ for each uU.

    Proof. First, we show the positive invariance of the set U. Let u0=(T0,0,i0())τU. Denote

    Ψ(t)(u0)=(T(t),0,i(t))τ

    and define

    Υ(t)=kN00ik(t,a)da.

    Since u0U, Υ(0)>0. Through a direct calculation, we have

    Υ(t)=kN00(ik(t,a)aδk(a)ik(t,a)kN0mjkik(t,a)+kN0mkjij(t,a))da=kN0(ik(t,0)0δ(a)ik(t,a)da)δmaxkN00ik(t,a)da=δmaxΥ(t),

    where δmax=max1kn{δk,max}. Hence, we obtain that

    Υ(t)eδmaxtnk=10ik0(a)da>0

    for t0, and then Ψ(t)UU.

    Next, in order to show the positive invariance of the set U, we consider u0U. It is clear from Lemma 2.2 that T(t)ˆT for some ˆT>0. Then the comparison theorem implies that

    i(t,)ˆi(t,), (4.6)

    where ˆi(t,a) is the solution of the following system

    ˆi(t,a)t+ˆi(t,a)a=(˜D(a)U)ˆi(t,a),ˆi(t,0)=Γdiag(ˆT)0P(a)ˆi(t,a)da,ˆi(0,)=i0(). (4.7)

    Since 0i0(a)da=0 and the assumption (A3), it holds that 0P(a)ˆi(0,a)da=0. From Lemma 4.1, we obtain that ˆi(t,a)=0() for all t0. The comparison in (4.6) implies that i(t,a)=0() for all t0 and then U is positively invariant under the semifolw Ψ(t). In addition, from Lemma 3.1, it is clear that for the solution remaining in U we have T(t)ˉT. Hence, limt+Ψ(t)u=ˉu for each uU.

    Next, we demonstrate uniform persistence of system (2.7).

    Theorem 4.2. When R0>1, the semiflow Ψ(t) generated by (2.7) is uniformly persistent with respect to (U,U); that is, there exists a constant ς>0 such that for each uU,

    lim inft+d(Ψ(t)u,U)ς,

    where d(,) is the distance associated to the norm in (2.6). Furthermore, there exists a compact subset A0U which is a global attractor for {Ψ(t)}t0 in U.

    Proof. In the following, we will prove that Ws({ˉu})U=,whereWs({ˉu})={uX0+|limt+Ψ(t)u=ˉu}. Recall from Lemma 4.2 that the infection-free equilibrium ˉu is GAS in U. It is sufficient to show that there exists σ>0 satisfying for each u{vU|vˉuσ} there exists ˜t00 such that Ψ(˜t0)uˉu>σ. Suppose by contradiction that for each integer m0 there is a um{vU|vˉu1m+1} such that Ψ(t)umˉu1m+1 for t0. Denote Ψ(t)um=(Tm(t),0Rn,im(t,a))τ=(Tm1(t),,Tmn(t),0Rn,im1(t,a),,imn(t))τ then |Tmj(t)ˉTj|1m+1,jN0, for all t0. Consider

    im(t,a)t+im(t,a)a=˜D(a)im(t,a)+Mim(t,a),im(t,0)=Γdiag(Tm(t))0P(a)im(t,a)da,Tm(0)=Tm0,im(0,)=im0(),

    with Tm0=(Tm10,,Tmn0)τ0, im0()=(im10(),,imn0())τ0() and kN00imk0(a)da>0. From the comparison theorem,

    im(t,)˜im(t,), (4.8)

    for t0, where ˜im(t,) is the solution of the following system

    ˜im(t,a)t+˜im(t,a)a=˜D(a)˜im(t,a)+M˜im(t,a),˜im(t,0)=Γdiag(ˉT1m+1I)0P(a)˜im(t,a)da,˜im(0,)=im0(),

    or equivalently

    dxdt=(ˆA+ˆLm)x(t), (4.9)

    for t0, x(0)¯ˆD(ˆA), the closure of ˆD(ˆA)={0Rn}×(W1,1)n, where x(t)=(0Rn,˜im(t,))τ and the operators ˆA,ˆLm are defined as

    ˆA(0˜im(a))=(˜im(0)d˜im(a)da+(˜D(a)+M)˜im(a)),ˆLm(0˜im(a))=(Γdiag(ˉT1m+1I)0P(a)˜im(a)da0).

    Since R0>1, there is a m0>0 large enough such that, for mm0,

    Rm0:=ρ(Γdiag(ˉT1m+1I)0P(a)exp(a0(˜D(s)M)ds)da)>1.

    The dominant eigenvalue λm of system (4.9) satisfies the characteristic equation Δm(λ)=0, where

    Δm(λ):=IΓdiag(ˉT1m+1I)0P(a)exp(a0(λI+˜D(s)M)ds)da.

    Note that Rm0>1 implies the existence of λm>0 such that Δm(λm)=0. Denote the solution semiflow of (4.9) by {ˆΨm(t)}t0 and let ˆΠm(ˆx):ˆXˆX, ˆX:=Rn×L, be the projection of given ˆxˆX on the eigenspace associated to the dominant eigenvalue λm. By the result in [19,Section 2.3], we see an explicit expression for the following projection

    ˆΠm(ˆΨm(t)ˆx0)=eλmtˆΠm(ˆx0)=eλmtlimλλm(λλm)(λIˆAˆLm)1ˆx0, (4.10)

    for ˆx0=(0Rn,im0(),)τ and

    (λIˆAˆLm)1=(λIˆA)1(IˆLm(λIˆA)1)1. (4.11)

    By using (4.11) and directly calculating (λIˆA)1, (IˆLm(λIˆA)1)1 respectively, we obtain that

    (λIˆAˆLm)1(0Rn,ϱ(a))τ=(0Rn,ϑ(a))τ

    if and only if

    ϑ(a)=ρ1exp(a0(λI+˜D(s)M)ds)+a0exp(a0(λI+˜D(r)M)dr)ρ2(s)ds,

    and

    ρ1=Δ(λ)1Γdiag(ˉT1m+1I)0P(a)a0exp(as(λI+˜D(r)M)dr)ϱ(s)dsdaρ2()=ϱ().

    Since Δm(λm)=0, we have

    limλλmλλmΔm(λ)=(dΔ(λm)dλ)1. (4.12)

    By using Reλm>0, (4.10), (4.11) and (4.12), it deduces that limt+˜im(t)=+. The comparison (4.8) implies limt+vm(t)=+, which is a contradiction to the boundedness of the solution. Hence, Ws({ˉu})U=. Accordingly, we derive from [38,Theorem 4.1,Theorem 4.2] that the semiflow {Ψ(t)}t0 is uniform persistence with respect to (U,U). In addition, the result in [28,Theorem 3.7] implies that there exists a compact global attractor A0U for {Ψ(t)}t0 in U.

    Remark 1. According to Theorem 4.2, the uniform persistence of (2.4) indicates that there exists a constant ζ>0 such that for each solution in (2.4) satisfies

    lim inftkN00ik(t,a)daζ. (4.13)

    We further note that there exists a constant ζ0>0 such that

    lim inft0ik(t,a)daζ0, (4.14)

    for each kN0. Otherwise, there is a k0N0 such that limt0ik0(t,a)da=0, which contradicts to the ik0-equation in (2.1) and the fact (4.13).

    The dichotomy of viral persistence or extinction within a host is based on the assumption of irreducible migration matrix. However, the migration matrix may be reducible due to functions of afferent or efferent lymphatics. In this case, the theory of direct graphs developed in [39] provides a method to describe directional connections between all within-host compartments (see also [40]). By building n vertices and assigning a directed edge from vertex j to vertex k when a flow from the j-th compartment to k-th compartment is available, we can associate the matrix M with a direct graph. Then the graph of M reflects the connection structure within a host, and the matrix M is irreducible if and only if for any pair of two compartments there is a path from one compartment to the other. In such a case, we say that the viral environment assumes a strongly connected within-host structure, whereas the case with a reducible matrix M is referred to non-strongly connected structure.

    When the within-host structure in non-strongly connected, that is, the matrix M is reducible, one can use a permutation operator to reach a triangular block form, still denoted by matrix M,

    M=(M1100M21M220Mp1Mpp), (5.1)

    where pn and each Mll, 1lp, is a rl×rl irreducible square matrix. We denote P0={1,2,,p} and write N0=lP0N(l)0, where N(l)0={Σl1s=1rs+1,Σl1s=1rs+2,,Σls=1rs}. By this way, we divide the whole within-host environment into p parts of which each part consists of strongly connected compartments. We indicate, by block-l, the part consisting of k-th compartment for kN(l)0. Explicitly, given lP0 with rl2, the compartments in block-l are strongly connected, whereas rl=1 means that the block-l consists of a single compartment. For later convenience, we denote N(l)0={ξl,,ξl}, that is, ξl=Σl1s=1rs+1 and ξl=Σls=1rs. Notably, under the form of (5.1), there is a connection from the block-l1 into the block-l2 only for 1l1<l2p. Denote P_(l)0={˜l|1˜l<l,Ml˜l0} for 1<lp and ˉP(l)0={˜l|l<˜lp,M˜ll0} for 1l<p. P_(l)0 is the set of indexes of blocks where each one is connected to block-l through a directional pathway, whereas ˉP(l)0 is that of blocks where each one is connected by a directional pathway from block-l.

    Now, for lP0, we say that the block-l is infection free if

    limtik(t,)=0(),forkN(l)0,

    for all solutions, and is infected if the infected population is uniformly persistent, that is, from Remark 1, there exists a constant ζ0>0 such that

    lim inft0ik(t,a)da>ζ0,forkN(l)0.

    We set the order of the blocks according to (5.1) and employ mathematical induction to analyze the virus dynamics in (2.1). We first introductorily define a sequence of basic reproduction numbers. In the system with only block-1, the basic reproduction number can be formulated as in (3.5) via replacing the matrix M by M11. However, when there are more than one block, it is necessary to reformulate the threshold value to determine the viral dynamics. Explicitly, for kN(1)0, we have

    dTk(t)dt=λkdkTk(t)αkTk(t)0pk(a)ik(t,a)dajN(1)0mjkTk(t)j˜lˉP(1)0N(˜l)0mjkTk(t)+jN(1)0mkjTj(t),ik(t,a)t+ik(t,a)a=δk(a)ik(t,a)jN(1)0mjkik(t,a)j˜lˉP(1)0N(˜l)0mjkik(t,a)+jN(1)0mkjij(t,a). (5.2)

    Denote X(t)=(T1(t),,Tξ1(t))τ, y(t,a)=(i1(t,a),,iξ1(t,a))τ, L(1)=(λ1,,λξ1)τ, D(1)=diag(d1,,dξ1), G(1)=diag(α1,,αξ1), diag(X(t))=diag(T1(t),,Tξ1(t)), ˜D(1)(a)=diag(δ1(a),,δξ1(a)), P(1)(a)=diag(p1(a),,pξ1(a)) and

    M(1)=(ˉm(1)11m12m1ξ1m21ˉm(1)22m2ξ1mξ11mξ12ˉm(1)ξ1ξ1) (5.3)

    with

    ˉm(1)kk=jN0,jkmjkj˜lˉP(1)0N(˜l)0mjk.

    We rewrite (5.2) into

    dX(t)dt=L(1)D(1)X(t)G(1)diag(X(t))0P(1)(a)y(t,a)da+M(1)X(t),y(t,a)t+y(t,a)a=˜D(1)(a)y(t,a)+M(1)y(t,a), (5.4)

    with

    y(t,0)=G(1)diag(X(t))0P(1)(a)y(t,a)da.

    In order to determine the threshold value to conclude viral persistence or extinction in (5.2), we denote

    R(1)0=ρ(Θ(1)),whereΘ(1)=G(1)diag(ˉT(1))0P(1)(a)exp(a0(˜D(1)(s)M(1))ds)da, (5.5)

    and ˉT(1) satisfies L(1)D(1)ˉT(1)+M(1)ˉT(1)=0. Note that R(1)0 is not the basic reproductive number for isolated block-1 except that ˉP(1)0=.

    Next, we start the process to determine the viral dynamics from the first block and the following is a straightforward result from Theorem 4.1 and Theorem 4.2.

    Proposition 5.1. The first block is infection free when R(1)01, whereas it is infected when R(1)0>1.

    For given lP0, we suppose that from block-1 to block-(l1) are all determined to be infected or infection free. In an infection free block, it satisfies limtTk(t)=ˉTk for constants ˉTk>0 derived as in Lemma 3.1 and the virus goes extinction. Next, we check on the connection from these l1 blocks to the block-l. Explicitly, we say that the block-l is susceptible from an infected block if there exists ˜l{1,2,,l1} such that the block-˜l is infected and Ml˜l0.

    Here, we denote by an index to determine whether the block-l is susceptible from an infected block. First, we denote

    S(Ml˜l)={0,ifellementsofMl˜lareall0s,1,ifoneellementofMl˜lisnonzero.

    Then there is a (no) connection from the block-˜l to the block-l when S(Ml˜l)=1 (S(Ml˜l)=0). We further denote

    χ(l)=max1˜l<l{sign(max{R(˜l)01,0})×S(Ml˜l)}.

    Then the block-l is susceptible from an infected ˜l block with ˜l<l if χ(l)=1, whereas it is not susceptible from any infected block-˜l with ˜l<l if χ(l)=0.

    Suppose that the block-l is not susceptible from any infected block-˜l with ˜l<l, i.e. χ(l)=0. Then the subsystem with kN(l)0 has the limit system

    dTk(t)dt=λkdkTk(t)αkTk(t)0pk(a)ik(t,a)dajN(l)0mjkTk(t)j˜lˉP(l)0N(˜l)0mjkTk(t)+j˜lP_(l)0N(˜l)0mkjˉTj+jN(l)0mkjTj(t),ik(t,a)t+ik(t,a)a=δk(a)ik(t,a)jN(l)0mjkik(t,a)j˜lˉP(l)0N(˜l)0mjkik(t,a)+jN(l)0mkjij(t,a). (5.6)

    Denote X(t)=(Tξl(t),,Tξl(t))τ, y(t,a)=(iξl(t,a),,iξl(t,a))τ, L(l)=(ˉλξl,,ˉλξl)τ,

    ˉλk=λk+j˜lP_(l)0N(˜l)0mkjˉTj, (5.7)

    D(l)=diag(dξl,,dξl), G(l)=diag(αξl,,αξl), diag(X(t))=diag(Tξl(t),,Tξl(t)), ˜D(l)(a)=diag(δξl(a),,δξl(a)), P(l)(a)=diag(pξl(a),,pξl(a)) and

    M(l)=(ˉm(l)ξlξlmξlξl+1mξlξlmξl+1ξlˉm(l)ξl+1ξl+1mξl+1ξlmξlξlmξlξl+1ˉm(l)ξlξl)

    with

    ˉm(l)kk=jN0,jkmjkj˜lˉP(l)0N(˜l)0mjk.

    We rewrite (5.6) into

    dX(t)dt=L(l)D(l)X(t)G(l)diag(X(t))0P(l)(a)y(t,a)da+M(l)X(t),y(t,a)t+y(t,a)a=˜D(l)(a)y(t,a)+M(l)y(t,a), (5.8)

    with

    y(t,0)=G(l)diag(X(t))0P(l)(a)y(t,a)da.

    In order to determine the threshold value to conclude viral persistence or extinction in (5.8), we denote

    R(l)0=ρ(Θ(l)),whereΘ(l)=G(l)diag(ˉT(l))0P(l)(a)exp(a0(˜D(l)(s)M(l))ds)da, (5.9)

    and ˉT(l) satisfies L(l)D(l)ˉT(l)+M(l)ˉT(l)=0. Note that R(l)0 is not the basic reproductive number for isolated block-l except that P_(l)0ˉP(l)0=.

    Accordingly, the virus dynamics in the block-l can be determined as the following.

    Theorem 5.2. For 1<lp, if χ(l)=1, that is, the block-l is susceptible from an infected block-˜l with ˜l<l, then the block-l is infected; if χ(l)=0, that is, the block-l is not susceptible from any infected block-˜l with ˜l<l, then the block-l is infection free when R(l)01, whereas it is infected when R(l)0>1.

    Proof. Consider the block-l with X(l)=1. It holds that, for kN(l)0,

    ik(t,a)t+ik(t,a)aδk,maxik(t,a)jN0mjkik(t,a)+jN0mkjij(t,a).

    Also note that there exists t0 and ˇTk such that Tk(t)ˇTk for all kN(l)0 for tt0. Hence, we consider an auxiliary system

    ˜ik(t,a)t+˜ik(t,a)a=δk,max˜ik(t,a)jN0mjk˜ik(t,a)+jN0mkj˜ij(t,a),kN(l)0,

    with ˜ik(t,0)=αkˇTk0pk(a)˜ik(t,a)da and ˜ik(0,a)=ik0(a). By a comparison theory, it reveals that ik(t,a)˜ik(t,a) for a0 and tt0. Denote ˜Ik(t)=0˜ik(t,a)da for kN(l)0, then it satisfies that

    d˜Ik(t)dt=˜ik(t,0)δk,max˜Ik(t)jN0mjk˜Ik(t)+jN0mkj˜Ij(t)δk,max˜Ik(t)jN0mjk˜Ik(t)+jN(l)0mkj˜ij(t,a)+jl1˜l=1N(˜l)0mkj˜Ij(t)δk,max˜Ik(t)jN0mjk˜Ik(t)+jN(l)0mkj˜Ij(t)+˜ζk,

    where ˜ζk:=jl1˜l=1N(˜l)0mkjζ(˜l)0 and ζ(˜l)0 is the lower bound in (4.14). Note that the assumption X(l)=1 implies ˜ζk>0. We further consider the auxiliary system

    dˇIk(t)dt=δk,maxˇIk(t)jN0mjkˇIk(t)+jN(l)0mkjˇIj(t)+˜ζk,kN(l)0,

    Since the matrix Mll is irreducible, there exists a positive equilibrium which is globally asymptotically stable as in Lemma 3.1. Thus we reveal that

    lim inft˜Ik(t)lim inftˇIk(t)ˇζk,kN(l)0,

    for some positive constants ˇζk. In addition, since ik(t,a)˜ik(t,a) for a0 and tt0, we see that 0ik(t,a)da0˜ik(t,a)da and then

    lim inft0ik(t,a)dalim inft0˜ik(t,a)daˇζk,kN(l)0,

    which concludes that the virus population uniformly persists in block-l.

    Next, we suppose that the block-l is not susceptible from any infected block-˜l with ˜l<l, that is X(l)=0. Then in the block-l we obtain a limit system as in (5.4) for kN(l)0. Therefore, the value of R(l)0 will determine the viral dynamics and it completes the proof.

    Remark 2. Note that, for kN(l)0, the value of ˉλk in (5.7) may depend on the connection from block-˜l with ˜l<l. From Theorem 5.2, we see that even if the block-l is not susceptible from any infected block-˜l with ˜l<l, the involved viral dynamics may be affected by the connection from an infection free block.

    Considering each compartment with or without afferent and efferent lymphatics, there are several different connection structures to build the whole system. In this section, we investigate the viral extinction and persistence in the model (2.1) in cases of strongly and non-strongly connected structures, along with compartments containing distinct infection characteristics.

    Deep lymph nodes of the head and neck are arranged in a vertical chain along the internal jugular vein and a circular chain consisting of occipital nodes, submental nodes, submandibular nodes, buccal or facial nodes and parotid nodes [41]. Hence, it is necessary to consider the group as a circular chain and to study the influence of removing one connection between lymphoid tissues on the viral infection. In this subsection, we consider three-compartment models with two types of connection matrices; one with a complete connection and the other without the path from compartment 3 to compartment 2, see Figure 2. To explore the effect of the connection topology on the viral infection, we assume identical compartments, λk=λ, dk=d, αk=α, pk()=p(), and δk()=δ(), k=1,2,3, and an identical migration rate, mkj=m for kj. In the case with the complete connection, it yields that

    Θ=λαd0P(a)exp(a0(δ(s)+2mmmmδ(s)+2mmmmδ(s)+2m)ds)da.
    Figure 2.  The diagrams of connected three compartments, (a) with a complete connection, (b) lacking the path from compartment 3 to compartment 2.

    A direct calculation gives

    exp(a0(δ(s)+2mmmmδ(s)+2mmmmδ(s)+2m)ds)=(ϕ(a)(13+23e3ma)ϕ(a)(1313e3ma)ϕ(a)(1313e3ma)ϕ(a)(1313e3ma)ϕ(a)(13+23e3ma)ϕ(a)(1313e3ma)ϕ(a)(1313e3ma)ϕ(a)(1313e3ma)ϕ(a)(13+23e3ma)),

    where ϕ(a)=exp(a0δ(s)ds). Denoting Q1(m,a)=13+23e3ma and Q2(m,a)=1313e3ma, we derive that

    Θ=(λαd0p(a)ϕ(a)Q1(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q1(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q2(m,a)daλαd0p(a)ϕ(a)Q1(m,a)da).

    Note that R0>1 if and only if there exists a positive constant ξ such that det(IeξΘ)=0, which is equivalent to that the determinant

    |1S1(ξ,m)S2(ξ,m)S2(ξ,m)S2(ξ,m)1S1(ξ,m)S2(ξ,m)S2(ξ,m)S2(ξ,m)1S1(ξ,m)|=0, (6.1)

    where

    S1(ξ,m)=eξλαd0p(a)ϕ(a)Q1(m,a)da,S2(ξ,m)=eξλαd0p(a)ϕ(a)Q2(m,a)da.

    Equation (6.1) holds if and only if

    eξλαd0p(a)ϕ(a)e3mada=1oreξλαd0p(a)ϕ(a)da=1.

    Hence, there exists a positive ξ satisfying det(IeξΘ)=0 if and only if

    λαd0p(a)ϕ(a)e3mada>1orλαd0p(a)ϕ(a)da>1, (6.2)

    that is

    λαd0p(a)ϕ(a)da>1. (6.3)

    Thus, the migration of cells between three nodes has no influence on the viral dynamics.

    However, in the case without the path from compartment 3 to compartment 2,

    Θ=αdiag(ˉT1,ˉT2,ˉT3)0P(a)exp(a0(δ(s)+2mmmmδ(s)+2m0mmδ(s)+m)ds)da,

    where the infection-free equilibrium (ˉT1,0,ˉT2,0,ˉT3,0) is obtained as in Lemma 3.1. A direct calculation reveals that

    exp(a0(δ(s)+2mmmmδ(s)+2m0mmδ(s)+m)ds)=(ϕ(a)(13+23e3ma)ϕ(a)(1313e3ma)ϕ(a)(1313e3ma)ϕ(a)(1623e3ma+12e2ma)ϕ(a)(16+13e3ma+12e2ma)ϕ(a)(16+13e3ma12e2ma)ϕ(a)(1212e2ma)ϕ(a)(1212e2ma)ϕ(a)(1212e2ma)),

    and then

    Θ=(αˉT10p(a)ϕ(a)˜Q11(m,a)daαˉT10p(a)ϕ(a)˜Q12(m,a)daαˉT10p(a)ϕ(a)˜Q13(m,a)daαˉT20p(a)ϕ(a)˜Q21(m,a)daαˉT20p(a)ϕ(a)˜Q22(m,a)daαˉT20p(a)ϕ(a)˜Q23(m,a)daαˉT30p(a)ϕ(a)˜Q31(m,a)daαˉT30p(a)ϕ(a)˜Q32(m,a)daαˉT30p(a)ϕ(a)˜Q33(m,a)da),

    where

    ˜Q11(m,a)=13+23e3ma,˜Q12(m,a)=1313e3ma=˜Q13(m,a),˜Q21(m,a)=1623e3ma+12e2ma,˜Q22(m,a)=16+13e3ma+12e2ma,˜Q23(m,a)=16+13e3ma12e2ma,˜Q31(m,a)=1212e2ma=˜Q32(m,a),˜Q33(m,a)=1212e2ma.

    In this circular chain, we choose parameters d=0.09, λ=10, p(a)=e0.15a, δ=0.4 [42] and change the value of α to see possible effect of migration rate, m, on the viral persistence. When α=0.0046, there is a m>0 such that R0<1 for 0m<m and R0>1 for m>m. Hence sufficiently slow migration of cells will drives the virus to extinction, whereas faster migration of cells will lead to viral persistence, which is different from the non-influence of migration in the completely connected circular chain. Moreover, when α=0.0042, there exist positive constants m<m such that R0<1 for 0m<m or m>m and R0>1 for m<m<m. It means that the influence of migration on the viral dynamics is complicated, from eradicating virus in the whole system with sufficiently small migration rate to initiating viral infection by medium value of migration rate, and eradicating virus again with sufficiently large migration rate.

    Consider a host consisting of four compartments as in Figure 4, where the 1st and 2nd compartments (involved in block-1) are strongly connected and also for the 3rd and 4th compartments (involved in block-2). The only connection between the two blocks is from the 2nd compartment to the 3rd compartment, that is, the block-2 may be susceptible from the block-1 when the latter one is infected. Denote m=m32. Hence, R(1)0=ρ(Θ(1)), where

    Θ(1)=G(1)diag(ˉT(1))0P(1)(a)exp(a0(˜D(1)(s)M(1))ds)da,
    Figure 3.  Influence of the migration rate m on the value of R0. When α=0.0046, there is a m>0 such that R0<1 for 0m<m and R0>1 for m>m. When α=0.0042, there exist positive constants m<m such that R0<1 for 0m<m or m>m and R0>1 for m<m<m.
    Figure 4.  Two blocks that are non-strongly connected. The 1st and 2nd compartments (involved in block-1) are strongly connected and also for the 3rd and 4th compartments (involved in block-2), but the only connection between the two blocks is from the 2nd compartment to the 3rd compartment.

    D(1)=diag(d1,d2), G(1)=diag(α1,α2), ˜D(1)(a)=diag(δ1(a),δ2(a)), P(1)(a)=diag(p1(a),p2(a)),

    L(1)=(λ1λ2),M(1)=(m21m12m21m12m)

    and ˉT(1)=(ˉT(1)1,ˉT(1)2)τ=(D(1)M(1))1L(1). By direct calculation, it yields that

    ˉT(1)2(m)=λ1m21+λ2(d1+m21)(d1+m21)(d2+m12+m)m12m21.

    The inner integral in Θ(1) is a positive matrix because of the irreducibility of matrix M(1), and then Θ(1) is positive for all m0. We see how the value of m affects that of R(1)0 in the following. All proofs in this subsection will be postponed to Appendix.

    Proposition 6.1. The value of R(1)0 is decreasing in m.

    On the other hand, R(2)0=ρ(Θ(2)), where

    Θ(2)=G(2)diag(ˉT(2))0P(2)(a)exp(a0(˜D(2)(s)M(2))ds)da,

    D(2)=diag(d3,d4), G(2)=diag(α3,α4), ˜D(2)(a)=diag(δ3(a),δ4(a)), P(2)(a)=diag(p3(a),p4(a)),

    L(2)=(λ3+mˉT(1)2(m)λ4),M(2)=(m43m34m43m34),

    and ˉT(2)=(ˉT(2)1,ˉT(2)2)τ=(D(2)M(2))1L(2). Accordingly, we can also see how the value of m affects that of R(2)0.

    Proposition 6.2. The value of R(2)0 is increasing in m.

    Biologically, a unidirectional connection between two blocks benefits the one with exportation in viral eradication but damages the other one with importation.

    In these two non-strongly connected blocks, we fix the viral environment in each block, including migration rates within the blocks, and change the migration rate between two blocks to explore its influence on the viral dynamics. Explicitly, we choose parameters relevant to block-1 by α1=α2=0.0085, d1=d2=0.09, λ1=λ2=10, m21=m12=1, δ=0.4, p(a)=e0.15a and those relevant to block-2 by α3=α4=0.00275, d3=d4=0.09, λ3=λ4=10, m43=m34=1, δ=0.4, p(a)=e0.15a. As depicted in Figure 5, the value of R(1)0 is decreasing in the migration rate, m, whereas the value of R(2)0 is increasing in m. There exist positive constants m<m with R(1)0(m)=1 and R(2)0(m)=1. Hence, according to Theorem 5.1, we observe the change of viral dynamics from viral persistence in the whole system with small value of m (0m<m) to virus extinction in the whole system with medium value of m (mmm), and to virus persistence only in the block-2 when the value of m is sufficiently large (m>m).

    Figure 5.  Decreasing R(1)0 and increasing R(2)0 related to the migration rate, m, from the 2nd compartment to the 3rd compartment.

    Recent studies show that for HIV the viral production and cell infection kernels are gamma distributed [43,44] after an intracellular delay for integration of viral genome. The predominant infected cell type or infection mode may determine the shape characteristics of the kernels. In particular, cell-to-cell transmission results in faster infection kinetics [44], and certain tissue compartments may have cells arranged closer together facilitating this infection model [45]. In order to model, compartments with distinct infection kinetics, we consider infected cell death rates δk(a) and infectivity kernels pk(a) of the following piecewise form:

    δk(a)={dk0a<τkνkτk<a,pk(a)={00a<τkgk(aτk)τk<a, (6.4)

    where τk is the intracellular delay, νk is the cell death rate after integration of viral genome, and gk(s) is the gamma-distributed (gk Gamma(κk,θk)) infectivity kernel, for compartment kN(l)0. Note that the mean of the gamma distribution gk(κk,θk) is μk=κk/θk, representing the mean time of (secondary) cell infection for an infected cell in the absence of death. In the special case that each compartment has identical delays τk=τ and gamma p.d.f.s, gk=g(κ,θ), with shape parameter κ=n as a positive integer, then the linear chain trick can yield the following equivalent system of delay differential equations:

    dT(t)dt=ΛDT(t)Γdiag(T(t))Yn(t)+MT(t),dY1(t)dt=Γexp((DM)τ)(T(tτ)Yn(tτ))(˜DM)Y1(t)θY1(t),dYm(t)dt=θ(Ym1(t)Ym(t))(˜DM)Ym(t),m=2,n, (6.5)

    where Ym(t):=0θmam1(m1)!eθai(t,a+τ)da,m=1,n, ˜D=diag(νk), and T(tτ)Yn(tτ) refers to entry-wise product of the vectors. Note that previous examples fall under this scenario with n=1 and τ=0. Thus for the previous results on how migration rate affects R0 in Figure 3 and Figure 5, we can verify extinction versus persistence by transforming to the ODE to obtain numerical solutions. For the general case of patches with distinct intracellular delays and/or gamma distribution parameters in (6.4), however the system will be infinite-dimensional as our original PDE model (2.1) and we utilize a finite-difference method for example simulations below.

    We explore the dynamics in this general setting by considering parameters representative of the study [44], where distinct modes of infection produce different infectivity kernels. Here we consider the case of three compartments connected in a circular chain lacking path from compartment-3 to compartment-2, as in Figure 2(b), with infectivity kernels as (6.4). In Figure 6, we display time-dependent solutions for 3 distinct scenarios; all compartments utilize cell-free infection, all compartments utilize cell-to-cell infection, and finally a mix where the first two compartments display cell-free while compartment-3 utilizes cell-to-cell infection. We also investigate how varying intracellular delay, τ3, and gamma distribution mean, μ3, of compartment-3 affect the overall R0. The fixed parameters utilized are d=0.09, λ=100, ν=0.4, α=0.0011, mij=m=1, and in Figure 6(a)-(c), intracellular delay τ=2/3 and gamma distribution mean μ1,μ2,μ3 either equal to 4/5 (cell-free infection) or 1/2 (cell-to-cell infection) with shape parameter κ fixed at 4. Shorter values of τk and μk result in larger R0, along with faster infection dynamics and larger number of infected cells. Thus, compartments which facilitate cell-to-cell infection may be important for allowing the virus to persist and establish reservoirs, which has been found in studies of HIV under drug treatment [46] or during initial bottleneck of host infection [16].

    Figure 6.  Three compartment circular chain lacking path from compartment-3 to compartment-2 in the case of example (6.4) with intracellular delay and gamma distribution infectivity kernel. Time-dependent solutions of Ik(t)=0ik(t,a)da are shown for intracellular delay τk=τ=2/3 day and gamma distribution mean (a) μk=μ=4/5, (b)μk=μ=1/2, and (c) μ1=μ2=4/5,μ3=1/2, consistent with cell-free infection, cell-cell infection, and mixed compartment infection mode, respectively. For the mixed compartment case in (c), the intracellular delay τ3 and gamma distribution mean μ3 are varied and the contour plot of R0 is plotted in (d).

    The assumption of well-mixed viral environment may ignore diversity of viral dynamics, for example, the existence of viral sanctuary sites and variant viral clearances, and then underestimate the infection within a host. We propose a multi-compartment model incorporating age-since-infection to imitate the migration of cells through afferent and efferent vessels between diverse organs. A threshold, depending on the rate of cell migration, the structure of organic connection and infection kinetics, was found to determine the extinction or persistence of infection. Accordingly, in this paper, we theoretically discussed the effect of the connection structure on the extinction or persistence of infection. Comparing systems with complete connection (with bi-directional connection between any two compartments of the system) and only lacking one connection, we see the influence of the migration rate in latter case. In both strongly connected systems, the influence may non-monotonously depend on the migration rate. As for a non-strongly connected system, we constructed a sequence of threshold values to conclude the infection pattern within a host and revealed that increasing the migration rate from a strongly connected block to another strongly connected block does not always help to eradicate virus in the whole system. Furthermore our inclusion of cell infection-age allows for organ-specific infectivity kinetics based on predominance of cell-free versus cell-to-cell transmission or target cell type. In particular, we demonstrated that individual compartments dominated by cell-to-cell infection may precipitate viral persistence in a strongly connected system with distinct gamma-distributed infectivity kernels representative of experimental studies. These findings complement and extend research on models consisting of a well-mixed compartment.

    In this study, the global convergence to the infection-free equilibrium was demonstrated in the case when the basic reproduction number is less than or equal to unity. On the other hand, the global convergence to the unique infection state is a challenging problem. In [20] for studying a multi-group SIR epidemic model with an age structure, an analogous problem was partially solved by using the method of Lyapunov functional and graph theory. However, the convergence dynamics to an infection state in this study remains as an unsolved problem since the model incorporates both the age structure and the migration behavior of cells. Moreover, the connection topology between a group significantly affects the evolution dynamics, and hence attracts researchers to contribute toward either theoretical analysis or application-oriented studies. A portion of a host body can be imitated by a strongly connected structure in which the viral dynamics holds clearance or uniformly persists in the complete part. However, biological evidences show that a functional system within a host, for example, the lymphatic system or the blood circulation system, consists of several such portions which are not strongly connected. Motivated by the study in [40] for infectious diseases via transportation networks, we systematically constructed a sequence of threshold values to find out the infection pattern in the whole system.

    We explicitly explored the effect of connection topology or cell infection-age kinetics on the viral dynamics in two basic structures: (a) a strongly connected circular chain with or without a certain directional connection, and (b) a system with one pair of two-compartment blocks connected via a unidirectional pathway. In the first example, either with or without the indicated directional connection, it remains the type of strongly connected. Nevertheless, in the latter case, it may show multiple switches between viral eradication and viral infection in the whole system when increasing the identical migration rate of cells. The second example was devoted to study the connection topology between two strongly-connected blocks. In particular, we explored the effect of the unidirectional migration between two compartments, each is located in a strongly-connected block. An explicit formula shows that increasing the migration rate between two blocks may first change the viral persistence only in the first block to viral eradication in the whole system, and then, for even larger migration rate, trigger the infection in only the flow-in block. Biologically, this unidirectional connection between two blocks benefits the one with exportation in viral eradication but damages the other one with importation. Although these example structures may be limited in number of compartments, they reveal the necessity in distinguishing the connection topology within a host. Finally, we investigated how incorporating intracellular delays and distinct gamma distributed infectivity kernels based on cell-free and cell-to-cell infection data [44] can affect viral dynamics in the strongly connected circular chain. We find that the faster kinetics associated with cell-to-cell infection promote viral persistence and spread, even when only a single compartment in the chain has this transmission mode.

    Variable time-dependent drug efficacy is another issue that requires consideration. For drug treatments, such as those administrated through targeted therapy, the treatment focuses on the targeted organ (the target compartment), and the drug efficacy is constant in each compartment. However, under several traditional therapies, drugs may spread throughout the entire body of a host via the circulatory system, in which the drug efficacy would no longer be constant in each compartment. Therefore, it is necessary to formulate the drug circulation among neighboring compartments along afferent and efferent vessels. These considerations will be explored in our next research.

    C. J. Browne was partially supported by a U.S. National Science Foundation grant (DMS-1815095); C.-Y. Cheng was partially supported by the Ministry of Science and Technology of Taiwan, R.O.C. (MOST 108-2115-M-153-003).

    All authors declare no conflicts of interest in this paper.

    To show the dependence of R(1)0 and R(2)0 on the value of m, we begin first from a property of the following exponential matrix S0(β).

    Lemma A.1.

    Given c,d>0, the value of each component of the matrix

    S0(β)=:exp(cdcdβ)

    is decreasing in β, for β>0.

    Proof. The following matrix can be written into with its diagonal form

    (cdcdβ)=PDP1, (7.1)

    where

    D=(((c+d+β)24cβcdβ)/200((c+d+β)24cβcdβ)/2),P=(ψ1(β)+ψ2(β)(ψ1(β)+ψ2(β))2+4c2ψ1(β)ψ2(β)(ψ1(β)+ψ2(β))2+4c22c(ψ1(β)+ψ2(β))2+4c22c(ψ1(β)+ψ2(β))2+4c2),

    with

    ψ1(β)=(c+d+β)24cβ,ψ2(β)=cdβ.

    Note that ψ1(β)>0, (ψ1(β))2(ψ2(β))2=4cd>0, and then ψ1(β)ψ2(β)>0 and ψ1(β)+ψ2(β)>0. In addition, direct calculations give

    ψ1(β)=ψ1(β)ψ2(β), ψ2(β)=1. (7.2)

    Hence, from (7.1), we have

    S0(β)=(S11S12S21S22),

    where

    S11(β)=ec2ψ1(β)(e(ψ2(β)ψ1(β))/2(ψ1(β)+ψ2(β))+e(ψ1(β)+ψ2(β))/2(ψ1(β)ψ2(β))),S12(β)=decψ1(β)(e(ψ1(β)+ψ2(β))/2e(ψ2(β)ψ1(β))/2),S21(β)=cecψ1(β)(e(ψ1(β)+ψ2(β))/2e(ψ2(β)ψ1(β))/2),S22(β)=ec2ψ1(β)(e(ψ2(β)ψ1(β))/2(ψ1(β)ψ2(β))+e(ψ1(β)+ψ2(β))/2(ψ1(β)+ψ2(β))).

    By direct calculations and using (7.2), it yields that

    S11(β)=cdψ1(β)3eψ2(β)/2(eψ1(β)/2(ψ1(β)+2)+eψ1(β)/2(ψ1(β)2))S12(β)=dec2ψ1(β)3eψ2(β)/2(eψ1(β)/2(2ψ2(β)ψ1(β)ψ2(β)ψ1(β)2)+eψ1(β)/2(2ψ2(β)ψ1(β)ψ2(β)+ψ1(β)2))S21(β)=cdS12(β),S22(β)=ec4ψ1(β)2(e(ψ2(β)ψ1(β))/2(ψ1(β)ψ2(β))2e(ψ1(β)+ψ2(β))/2(ψ1(β)+ψ2(β))28cdψ1(β)eψ2(β)/2(eψ1(β)/2eψ1(β)/2)).

    Define

    k(ψ1(β))=eψ1(β)/2(ψ1(β)+2)+eψ1(β)/2(ψ1(β)2),h(ψ1(β),ψ2(β))=eψ1(β)/2(2ψ2(β)ψ1(β)ψ2(β)ψ1(β)2)+eψ1(β)/2(2ψ2(β)ψ1(β)ψ2(β)+ψ1(β)2).

    It is easy to show that k(x)>0 for all x>0 and h(x, y) < 0 for all x>0. Finally, the fact \psi_1(\beta)>0 implies S'_{ij}(\beta) < 0 for i, j = 1, 2, and it completes the assertion.

    Based on this property, we see the dependence of \mathcal{R}_0^{(1)} on m.

    Proof of Proposition 6.1:

    Proof. Since the spectral of a positive matrix positively depends on each component, it suffices to show that, for a fixed a>0, the value of each component of the matrix

    S(m): = \exp \left(-\int^a_0(\mathcal{\tilde{D}}^{(1)}(s)-\mathcal{M}^{(1)})ds \right)

    is decreasing in m. For given a>0, if \int^a_0\delta_2(s)da\geq \int^a_0\delta_1(s)da, we rewrite S(m) into

    \begin{eqnarray}\label{eq-d2} S(m) & = & \exp\left(\left(\begin{array}{cc} -\int^a_0\delta_1(s)ds & 0\\ 0 & -\int^a_0\delta_1(s)ds \end{array} \right) + \left(\begin{array}{cc} -m_1a & m_2a \\ m_1a & -m_2a-\beta(m) \end{array} \right) \right) \nonumber \\ & = & \exp\left(\begin{array}{cc} -\int^a_0\delta_1(s)ds & 0\\ 0 & -\int^a_0\delta_1(s)ds \end{array} \right) \exp\left(\begin{array}{cc} -m_1a & m_2a \\ m_1a & -m_2a-\beta(m) \end{array} \right), \end{eqnarray} (7.3)

    where

    \beta(m) = ma + \int^a_0(\delta_2(s)-\delta_1(s))ds.

    Note that \beta(m)>0 and \beta'(m)>0 for all m>0. From Lemma A.1, the value of each component of the matrix \exp(-\int^a_0(\mathcal{\tilde{D}}^{(1)}(s)-\mathcal{M}^{(1)})ds) is decreasing in m. A similar argument can be proceeded by exchanging \delta_1(\cdot) and \delta_2(\cdot) in (7.3) when \int^a_0\delta_1(s)da\geq \int^a_0\delta_2(s)da. In addition, since the spectral of a positive matrix positively depends on each component, the value of \mathcal{R}_0^{(1)} is decreasing in m.

    Proof of Proposition 6.2:

    Proof. In \Theta^{(2)}, \bar{T}^{(2)} is the only term tdepending on the value of m and, since the inner integral is a positive matrix, the spectral of \Theta^{(2)} depends increasingly on \bar{T}^{(2)}_2. It is easy to see that the spectral bound of the matrix -(\mathcal{D}^{(2)} - \mathcal{M}^{(2)}) is negative and then by the Perron-Frobenious Theorem, the matrix (\mathcal{D}^{(2)} - \mathcal{M}^{(2)})^{-1} is positive. Hence, the value of \bar{T}^{(2)}_2 positively depends on the value of F(m): = m\bar{T}^{(1)}_2(m). Moreover, since

    F'(m) = \frac{ (\lambda_1m_{21} + \lambda_2(d_1+m_{21})) ((d_1+m_{21}) (d_2+m_{12}) -m_{12}m_{21}) }{((d_1+m_{21}) (d_2+m_{12}+m) -m_{12}m_{21})^2} \gt 0,

    we conclude that the value of \mathcal{R}_0^{(2)} is increasing in m.

    [1] Sun J, Zuckermann RN (2013) Peptoid Polymers: A Highly Designable Bioinspired Material. ACS Nano 7: 4715–4732. doi: 10.1021/nn4015714
    [2] Seo J, Lee BC, Zuckermann RN (2011) Peptoids: Synthesis, Characterization, and Nanostructures. Compr Biomater 2: 53–76.
    [3] Chongsiriwatana NP, Patch JA, Czyzewski AM, et al. (2008) Peptoids that mimic the structure, function, and mechanism of helical antimicrobial peptides. Proc Natl Acad Sci USA 105: 2794–2799. doi: 10.1073/pnas.0708254105
    [4] Vollrath SBL, Fürniss D, Schepers U, et al. (2013) Amphiphilic peptoid transporters-synthesis and evaluation. Org Biomol Chem 11: 8197–8201. doi: 10.1039/c3ob41139g
    [5] Li N, Zhu F, Gao F, et al. (2010) Blockade of CD28 by a synthetical peptoid inhibits T-cell proliferation and attenuates graft-versus-host disease. Cell Mol Immunol 7: 133–142. doi: 10.1038/cmi.2009.120
    [6] Dohm MT, Kapoor R, Barron AE (2011) Peptoids: Bio-Inspired Polymers as Potential Pharmaceuticals. Curr Pharm Design 17: 2732–2747. doi: 10.2174/138161211797416066
    [7] Statz AR, Park JP, Chongsiriwatana NP, et al. (2008) Surface-immobilised antimicrobial peptoids. Biofouling 24: 439–448. doi: 10.1080/08927010802331829
    [8] Seurynck SL, Patch JA, Barron AE (2005) Simple, helical peptoid analogs of lung surfactant protein B. Chem Biol 12: 77–88. doi: 10.1016/j.chembiol.2004.10.014
    [9] Maayan G, Ward MD, Kirshenbaum K (2009) Folded biomimetic oligomers for enantioselective catalysis. Proc Natl Acad Sci USA 106: 13679–13684. doi: 10.1073/pnas.0903187106
    [10] Gellman SH (1998) Foldamers:  A Manifesto. Accounts Chem Res 31: 173–180.
    [11] Armand P, Kirshenbaum K, Falicov A, et al. (1997) Chiral N-substituted glycines can form stable helical conformations. Fold Design 2: 369–375. doi: 10.1016/S1359-0278(97)00051-5
    [12] Shah NH, Butterfoss GL, Nguyen K (2008) Oligo(N-aryl glycines): A New Twist on Structured Peptoids. J Am Chem Soc 130: 16622–16632. doi: 10.1021/ja804580n
    [13] Huang K, Wu CW, Sanborn TJ, et al. (2006) A threaded loop conformation adopted by a family of peptoid nonamers. J Am Chem Soc 128: 1733–1738. doi: 10.1021/ja0574318
    [14] Crapster JA, Guzei IA, Blackwell HE (2013) A Peptoid Ribbon Secondary Structure. Angew Chem Int Ed 52: 5079–5084. doi: 10.1002/anie.201208630
    [15] Mannige RV, Haxton TK, Proulx C, et al. (2015) Peptoid nanosheets exhibit a new secondary-structure motif. Nature 526: 415–420. doi: 10.1038/nature15363
    [16] Hebert ML, Shah DS, Blake P, et al. (2013) Tunable peptoid microspheres: effects of side chain chemistry and sequence. Org Biomol Chem 11: 4459–4464. doi: 10.1039/c3ob40561c
    [17] Murnen HK, Rosales AM, Jaworski JN, et al. (2010) Hierarchical Self-Assembly of a Biomimetic Diblock Copolypeptoid into Homochiral Superhelices. J Am Chem Soc 132: 16112–16119. doi: 10.1021/ja106340f
    [18] Sanii B, Kudirka R, Cho A, et al. (2011) Shaken, Not Stirred: Collapsing a Peptoid Monolayer To Produce Free-Floating, Stable Nanosheets. J Am Chem Soc 133: 20808–20815. doi: 10.1021/ja206199d
    [19] Dill KA, MacCallum JL (2012) The Protein-Folding Problem, 50 Years On. Science 338: 1042–1046. doi: 10.1126/science.1219021
    [20] Gorske BC, Blackwell HE (2006) Tuning peptoid secondary structure with pentafluoroaromatic functionality: A new design paradigm for the construction of discretely folded peptoid structures. J Am Chem Soc 128: 14378–14387. doi: 10.1021/ja065248o
    [21] Stringer JR, Crapster JA, Guzei IA, et al. (2010) Construction of Peptoids with All Trans-Amide Backbones and Peptoid Reverse Turns via the Tactical Incorporation of N-Aryl Side Chains Capable of Hydrogen Bonding. J Org Chem 75: 6068–6078. doi: 10.1021/jo101075a
    [22] Gorske BC, Nelson RC, Bowden ZS, et al. (2013) "Bridged" n→π* Interactions Can Stabilize Peptoid Helices. J Org Chem 78: 11172–11183. doi: 10.1021/jo4014113
    [23] Wu CW, Kirshenbaum K, Sanborn TJ, et al. (2003) Structural and spectroscopic studies of peptoid oligomers with alpha-chiral aliphatic side chains. J Am Chem Soc 125: 13525–13530. doi: 10.1021/ja037540r
    [24] Kirshenbaum K, Barron AE, Goldsmith RA, et al. (1998) Sequence-specific polypeptoids: A diverse family of heteropolymers with stable secondary structure. Proc Natl Acad Sci USA 95: 4303–4308. doi: 10.1073/pnas.95.8.4303
    [25] Armand P, Kirshenbaum K, Goldsmith RA, et al. (1998) NMR determination of the major solution conformation of a peptoid pentamer with chiral side chains. Proc Natl Acad Sci USA 95: 4309–4314. doi: 10.1073/pnas.95.8.4309
    [26] Dill KA (1990) Dominant forces in protein folding. Biochemistry 29: 31.
    [27] Dill KA, Ozkan SB, Shell MS, et al. (2008) The protein folding problem. Annu Rev Biophys 37: 289–316. doi: 10.1146/annurev.biophys.37.092707.153558
    [28] Sali A, Blundell TL (1993) Comparative protein modeling by satisfaction of spatial restraints. J Mol Biol 234: 779–815. doi: 10.1006/jmbi.1993.1626
    [29] Shen MY, Sali A (2006) Statistical potential for assessment and prediction of protein structures. Protein Sci 15: 2507–2524. doi: 10.1110/ps.062416606
    [30] Rohl CA, Strauss CEM, Misura KMS, et al. (2004) Protein structure prediction using Rosetta. Method Enzymol 383: 66–93. doi: 10.1016/S0076-6879(04)83004-0
    [31] Leach AR (2001) Molecular modelling : principles and applications, England: Pearson/Prentice Hall.
    [32] Shell MS (2016) Coarse-Graining with the Relative Entropy, In: Rice SA, Dinner AR, Advances in Chemical Physics, Malden: Wiley-Blackwell, 395–441.
    [33] Mackerell AD, Feig M, Brooks CL (2004) Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J Comput Chem 25: 1400–1415. doi: 10.1002/jcc.20065
    [34] Feigel M (1983) Rotation barriers of amides in the gas phase. J Phys Chem 87: 3054–3058. doi: 10.1021/j100239a019
    [35] Sui Q, Borchardt D, Rabenstein DL (2007) Kinetics and equilibria of cis/trans isomerization of backbone amide bonds in peptoids. J Am Chem Soc 129: 12042–12048. doi: 10.1021/ja0740925
    [36] Duffy EM, Severance DL, Jorgensen WL (1992) Solvent effects on the barrier to isomerization for a tertiary amide from ab initio and Monte Carlo calculations. J Am Chem Soc 114: 7535–7542. doi: 10.1021/ja00045a029
    [37] Torrie GM, Valleau JP (1977) Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J Comput Phys 23: 187–199. doi: 10.1016/0021-9991(77)90121-8
    [38] Sugita Y, Okamoto Y (1999) Replica-exchange molecular dynamics method for protein folding. Chem Phys Lett 314: 141–151. doi: 10.1016/S0009-2614(99)01123-9
    [39] Stringer JR, Crapster JA, Guzei IA, et al. (2011) Extraordinarily Robust Polyproline Type I Peptoid Helices Generated via the Incorporation of alpha-Chiral Aromatic N-1-Naphthylethyl Side Chains. J Am Chem Soc 133: 15559–15567. doi: 10.1021/ja204755p
    [40] Ramachandran GN, Ramakrishnan C, Sasisekharan V (1963) Stereochemistry of polypeptide chain configurations. J Mol Biol 7: 95–99. doi: 10.1016/S0022-2836(63)80023-6
    [41] Butterfoss GL, Renfrew PD, Kuhlman B, et al. (2009) A Preliminary Survey of the Peptoid Folding Landscape. J Am Chem Soc 131: 16798–16807. doi: 10.1021/ja905267k
    [42] Mohle K, Hofmann HJ (1996) Peptides and peptoids—A systematic structure comparison. J Mol Model 2: 307–311. doi: 10.1007/s0089460020307
    [43] Miertuš S, Scrocco E, Tomasi J (1981) Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects. Chem Phys 55: 117–129.
    [44] Pascual-Ahuir JL, Silla E, Tomasi J, et al. (1987) Electrostatic interaction of a solute with a continuum. Improved description of the cavity and of the surface cavity bound charge distribution. J Comput Chem 8: 778–787.
    [45] Parker BF, Knight AS, Vukovic S, et al. (2016) A Peptoid-Based Combinatorial and Computational Approach to Developing Ligands for Uranyl Sequestration from Seawater. Ind Eng Chem Res 55: 4187–4194. doi: 10.1021/acs.iecr.5b03500
    [46] Cancès E, Mennucci B, Tomasi J (1997) A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics. J Chem Phys 107: 3032–3041. doi: 10.1063/1.474659
    [47] Cornell W, Cieplek P, Bayly CI, et al. (1995) A Second Generation Force-Field for the Simulation of Proteins, Nucleic-Acids, and Organic-Molecules. J Am Chem Soc 117: 5179–5197. doi: 10.1021/ja00124a002
    [48] Hawkins GD, Cramer CJ, Truhlar DG (1998) Universal Quantum Mechanical Model for Solvation Free Energies Based on Gas-Phase Geometries. J Phys Chem B 102: 3257–3271. doi: 10.1021/jp973306+
    [49] Bradley EK, Kerr JM, Richter LS, et al. (1997) NMR Structural Characterization of Oligo-N-Substituted Glycine Lead Compounds from a Combinatorial Library. Mol Divers 3: 1–15. doi: 10.1023/A:1009698309407
    [50] Mann G, Yun RH, Nyland L, et al. (2002) The Sigma MD Program and a Generic Interface Applicable to Multi-Functional Programs with Complex, Hierarchical Command Structure, In: Schlick T, Gan HH, Computational Methods for Macromolecules: Challenges and Applications, Springer, Berlin, Heidelberg, 129–145.
    [51] Hermans J, Berendsen HJC, Van Gunsteren WF, et al. (1984) A consistent empirical potential for water–protein interactions. Biopolymers 23: 1513–1518. doi: 10.1002/bip.360230807
    [52] Butterfoss GL, Yoo B, Jaworski JN (2012) De novo structure prediction and experimental characterization of folded peptoid oligomers. Proc Natl Acad Sci USA 109: 14320–14325. doi: 10.1073/pnas.1209945109
    [53] Wang JM, Wolf RM, Caldwell JW, et al. (2004) Development and testing of a general amber force field. J Comput Chem 25: 1157–1174. doi: 10.1002/jcc.20035
    [54] Onufriev A, Bashford D, Case DA (2004) Exploring protein native states and large-scale conformational changes with a modified generalized born model. Proteins 55: 383–394. doi: 10.1002/prot.20033
    [55] MacKerell AD, Bashford D, Bellott M, et al. (1998) All-atom empirical potential for molecular modeling and dynamics studies of proteins. J Phys Chem B 102: 3586–3616. doi: 10.1021/jp973084f
    [56] Case DA, Cheatham TE, Darden T, et al. (2005) The Amber biomolecular simulation programs. J Comput Chem 26: 1668–1688. doi: 10.1002/jcc.20290
    [57] Jorgensen WL, Maxwell DS, TiradoRives J (1996) Development and testing of the OPLS all-atom force field on conformational energetics and properties of organic liquids. J Am Chem Soc 118: 11225–11236. doi: 10.1021/ja9621760
    [58] Moehle K, Hofmann HJ (1996) Peptides and peptoids—A quantum chemical structure comparison. Biopolymers 38: 781–790. doi: 10.1002/(SICI)1097-0282(199606)38:6<781::AID-BIP9>3.0.CO;2-N
    [59] Jorgensen W, Chandrasekhar J, Madura J, et al. (1983) Comparison of Simple Potential Functions for Simulating Liquid Water. J Chem Phys 79: 926–935. doi: 10.1063/1.445869
    [60] Tobias DJ, Brooks CL (1988) Molecular dynamics with internal coordinate constraints. J Chem Phys 89: 5115–5127. doi: 10.1063/1.455654
    [61] Wang J, Wang W, Kollman PA, et al. (2006) Automatic atom type and bond type perception in molecular mechanical calculations. J Mol Graph Model 25: 247–260. doi: 10.1016/j.jmgm.2005.12.005
    [62] Jakalian A, Jack DB, Bayly CI (2002) Fast, efficient generation of high-quality atomic charges. AM1-BCC model: II. Parameterization and validation. J Comput Chem 23: 1623–1641.
    [63] Mukherjee S, Zhou G, Michel C, et al. (2015) Insights into Peptoid Helix Folding Cooperativity from an Improved Backbone Potential. J Phys Chem B 119: 15407–15417. doi: 10.1021/acs.jpcb.5b09625
    [64] Lifson S, Roig A (1961) On the Theory of Helix-Coil Transition in Polypeptides. J Chem Phys 34: 1963–1974. doi: 10.1063/1.1731802
    [65] Mirijanian DT, Mannige RV, Zuckermann RN, et al. (2014) Development and use of an atomistic CHARMM-based forcefield for peptoid simulation. J Comput Chem 35: 360–370. doi: 10.1002/jcc.23478
    [66] Jordan PA, Bishwajit P, Butterfoss GL, et al. (2011) Oligo(N-alkoxy glycines): trans substantiating peptoid conformations. J Pept Sci 96: 617–626. doi: 10.1002/bip.21675
    [67] Nam KT, Shelby SA, Cho PH, et al. (2010) Free-floating ultrathin two-dimensional crystals from sequence-specific peptoid polymers. Nat Mater 9: 454–460. doi: 10.1038/nmat2742
    [68] Mannige RV, Kundu J, Whitelam S (2016) The Ramachandran Number: An Order Parameter for Protein Geometry. PLoS One 11: e0160023. doi: 10.1371/journal.pone.0160023
    [69] Reith D, Putz M, Muller-Plathe F (2003) Deriving effective mesoscale potentials from atomistic simulations. J Comput Chem 24: 1624–1636. doi: 10.1002/jcc.10307
    [70] Izvekov S, Voth GA (2005) Multiscale coarse graining of liquid-state systems. J Chem Phys 123: 134105. doi: 10.1063/1.2038787
    [71] Shell MS (2008) The relative entropy is fundamental to multiscale and inverse thermodynamic problems. J Chem Phys 129: 144108. doi: 10.1063/1.2992060
    [72] Haxton TK, Mannige RV, Zuckermann RN, et al. (2015) Modeling Sequence-Specific Polymers Using Anisotropic Coarse-Grained Sites Allows Quantitative Comparison with Experiment. J Chem Theory Comput 11: 303–315. doi: 10.1021/ct5010559
    [73] Sanii B, Haxton TK, Olivier GK, et al. (2014) Structure-Determining Step in the Hierarchical Assembly of Peptoid Nanosheets. ACS Nano 8: 11674–11684. doi: 10.1021/nn505007u
    [74] Haxton TK, Zuckermann RN, Whitelam S (2016) Implicit-Solvent Coarse-Grained Simulation with a Fluctuating Interface Reveals a Molecular Mechanism for Peptoid Monolayer Buckling. J Chem Theory Comput 12: 345–352. doi: 10.1021/acs.jctc.5b00910
    [75] Drew K, Renfrew PD, Butterfoss GL (2013) Adding Diverse Noncanonical Backbones to Rosetta: Enabling Peptidomimetic Design. PLoS One 8: e67051.
    [76] Kaufmann KW, Lemmon GH, DeLuca SL, et al. (2010) Practically Useful: What the Rosetta Protein Modeling Suite Can Do for You. Biochemistry 49: 2987–2998. doi: 10.1021/bi902153g
    [77] Renfrew PD, Craven TW, Butterfoss GL, et al. (2013) A Rotamer Library to Enable Modeling and Design of Peptoid Foldamers. J Am Chem Soc 136: 8772–8782.
    [78] Laio A, Gervasio FL (2008) Metadynamics: a method to simulate rare events and reconstruct the free energy in biophysics, chemistry and material science. Rep Prog Phys 71: 126601. doi: 10.1088/0034-4885/71/12/126601
  • This article has been cited by:

    1. Feng-Bin Wang, Chang-Yuan Cheng, A diffusive virus model with a fixed intracellular delay and combined drug treatments, 2021, 83, 0303-6812, 10.1007/s00285-021-01646-7
    2. Joshua C. Macdonald, Hayriye Gulbudak, Brianna Beechler, Erin E. Gorsich, Simon Gubbins, Eva Pérez-Martin, Anna E. Jolles, Within-Host Viral Growth and Immune Response Rates Predict Foot-and-Mouth Disease Virus Transmission Dynamics for African Buffalo, 2024, 204, 0003-0147, 133, 10.1086/730703
  • Reader Comments
  • © 2017 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(7054) PDF downloads(1328) Cited by(22)

Figures and Tables

Figures(11)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog