Processing math: 94%
Research article Special Issues

Analysis on a diffusive SIS epidemic system with linear source and frequency-dependent incidence function in a heterogeneous environment

  • In this paper, we consider a diffusive SIS epidemic reaction-diffusion model with linear source in a heterogeneous environment in which the frequency-dependent incidence function is SI/(c + S + I) with c a positive constant. We first derive the uniform bounds of solutions, and the uniform persistence property if the basic reproduction number R0>1. Then, in some cases we prove that the global attractivity of the disease-free equilibrium and the endemic equilibrium. Lastly, we investigate the asymptotic profile of the endemic equilibrium (when it exists) as the diffusion rate of the susceptible or infected population is small. Compared to the previous results [1, 2] in the case of c = 0, some new dynamical behaviors appear in the model studied here; in particular, R0 is a decreasing function in c∈[0, ∞) and the disease dies out once c is properly large. In addition, our results indicate that the linear source term can enhance the disease persistence.

    Citation: Jinzhe Suo, Bo Li. Analysis on a diffusive SIS epidemic system with linear source and frequency-dependent incidence function in a heterogeneous environment[J]. Mathematical Biosciences and Engineering, 2020, 17(1): 418-441. doi: 10.3934/mbe.2020023

    Related Papers:

    [1] Roberto A. Saenz, Herbert W. Hethcote . Competing species models with an infectious disease. Mathematical Biosciences and Engineering, 2006, 3(1): 219-235. doi: 10.3934/mbe.2006.3.219
    [2] Wenzhang Huang, Maoan Han, Kaiyu Liu . Dynamics of an SIS reaction-diffusion epidemic model for disease transmission. Mathematical Biosciences and Engineering, 2010, 7(1): 51-66. doi: 10.3934/mbe.2010.7.51
    [3] Hongbin Guo, Michael Yi Li . Global dynamics of a staged progression model for infectious diseases. Mathematical Biosciences and Engineering, 2006, 3(3): 513-525. doi: 10.3934/mbe.2006.3.513
    [4] C. Connell McCluskey . Global stability of an SIR epidemic model with delay and general nonlinear incidence. Mathematical Biosciences and Engineering, 2010, 7(4): 837-850. doi: 10.3934/mbe.2010.7.837
    [5] Yan’e Wang , Zhiguo Wang, Chengxia Lei . Asymptotic profile of endemic equilibrium to a diffusive epidemic model with saturated incidence rate. Mathematical Biosciences and Engineering, 2019, 16(5): 3885-3913. doi: 10.3934/mbe.2019192
    [6] Xiaowei An, Xianfa Song . A spatial SIS model in heterogeneous environments with vary advective rate. Mathematical Biosciences and Engineering, 2021, 18(5): 5449-5477. doi: 10.3934/mbe.2021276
    [7] Zhen Jin, Guiquan Sun, Huaiping Zhu . Epidemic models for complex networks with demographics. Mathematical Biosciences and Engineering, 2014, 11(6): 1295-1317. doi: 10.3934/mbe.2014.11.1295
    [8] Baoxiang Zhang, Yongli Cai, Bingxian Wang, Weiming Wang . Dynamics and asymptotic profiles of steady states of an SIRS epidemic model in spatially heterogenous environment. Mathematical Biosciences and Engineering, 2020, 17(1): 893-909. doi: 10.3934/mbe.2020047
    [9] Yicang Zhou, Zhien Ma . Global stability of a class of discrete age-structured SIS models with immigration. Mathematical Biosciences and Engineering, 2009, 6(2): 409-425. doi: 10.3934/mbe.2009.6.409
    [10] Shouying Huang, Jifa Jiang . Global stability of a network-based SIS epidemic model with a general nonlinear incidence rate. Mathematical Biosciences and Engineering, 2016, 13(4): 723-739. doi: 10.3934/mbe.2016016
  • In this paper, we consider a diffusive SIS epidemic reaction-diffusion model with linear source in a heterogeneous environment in which the frequency-dependent incidence function is SI/(c + S + I) with c a positive constant. We first derive the uniform bounds of solutions, and the uniform persistence property if the basic reproduction number R0>1. Then, in some cases we prove that the global attractivity of the disease-free equilibrium and the endemic equilibrium. Lastly, we investigate the asymptotic profile of the endemic equilibrium (when it exists) as the diffusion rate of the susceptible or infected population is small. Compared to the previous results [1, 2] in the case of c = 0, some new dynamical behaviors appear in the model studied here; in particular, R0 is a decreasing function in c∈[0, ∞) and the disease dies out once c is properly large. In addition, our results indicate that the linear source term can enhance the disease persistence.


    In the study of transmission of infectious disease, people have realised that environmental heterogeneity and individual motility are significant factors that should be incorporated into the mathematical models. In recent decades, more and more research works have been devoted to the investigation of the dynamics of infectious disease modelled by reaction-diffusion systems in which the migration of population and environmental heterogeneity are taken into account. One may refer to [1,3,4,5,6,7,8,9] and the references therein.

    In a recent paper [1], Allen et al. investigated a frequency-dependent SIS (susceptible-infected-susceptible) epidemic reaction-diffusion model, which reads as

    {StdSΔS=β(x)SIS+I+γ(x)I,xΩ,t>0,ItdIΔI=β(x)SIS+Iγ(x)I,xΩ,t>0,Sν=Iν=0,xΩ,t>0,S(x,0)=S0(x)0,I(x,0)=I0(x),0. (1.1)

    Here, S and I stand for the density of susceptible and infected population at location x and time t respectively; the positive constants dS and dI represent the motility of susceptible and infected individuals, respectively; the function β(x) is the rate of disease transmission, and γ(x) is the recovery rate of infected individuals, all of which are positive Hölder continuous functions on ¯Ω. The habitat ΩRN(N1) is a bounded domain with smooth boundary Ω, and the Neumann boundary conditions mean that no population flux crosses the boundary Ω. For realistic implication, the initial data S0 and I0 are assumed to be nonnegative continuous functions on ¯Ω, and there is a positive number of infected individuals, i.e., ΩI0(x)dx>0.

    It is easily seen from (1.1) that

    Ω(S(x,t)+I(x,t))dx=Ω(S0(x)+I0(x))dx,  t>0,

    which means that the total number of population is conserved.

    Clearly, model (1.1) dose not account into account the birth rate of the susceptible population and the death rate induced by disease. Indeed, these factors are important in the evolution of disease transmission; see [7,10,11]. With such a consideration, in the paper [12], the authors studied a varying total population model in which the linear external source term Λ(x)S was introduced. That is, the model (1.1) becomes the following:

    {StdSΔS=Λ(x)Sβ(x)SIS+I+θγ(x)I,xΩ, t>0,ItdIΔI=β(x)SIS+Iγ(x)I,xΩ, t>0,Sν=Iν=0,xΩ, t>0,S(x,0)=S0(x)0,I(x,0)=I0(x),0,xΩ. (1.2)

    where dS,dI,β,γ,S and I have the same epidemiological interpretation as in (1.1). The parameter θ[0,1] represent the number of infected population becomes susceptible. The positive functions β,γ,Λ also be assumed to be Hölder continuous over ¯Ω. In the linear external source term, Λ(x) and S, respectively, account for the birth rate of the susceptible population and the disease-induced death rate. It is worth mentioning that in some cases, people ignore the effect of external source on the infected population; one may see, for instance, [3,9,13] for related discussion.

    Different from model (1.1), a new feature in (1.2) is that the total population of susceptible and infected individuals are varying with respect to time t>0. On the other hand, the works in [14,15,16,17] have shown that, in certain circumstances, the frequency-dependent incidence function SIS+I used in models (1.1) and (1.2) may not be appropriate to describe the transmission process of disease; instead an alternate incidence function should be SIc+S+I, where c is a positive constant. Based on model (1.2), in this paper we are led to study the following SIS epidemic model:

    {StdSΔS=a(x)μ(x)Sβ(x)SIc+S+I+γ(x)I,xΩ, t>0,ItdIΔI=β(x)SIc+S+I[γ(x)+μ(x)]I,xΩ, t>0,Sν=Iν=0,xΩ, t>0, S(x,0)=S0(x)0,I(x,0)=I0(x),0,xΩ. (1.3)

    In model (1.3), it should be noted that we also assume that the infected population allows the same natural death rate as for the susceptible population, which is represented by the function μ(x). With such a consideration, the system (1.3) is more realistic to describe the disease transmission in some cases as suggested in [15,16,17]. From now on, we always assume that c is a nonnegative constant, the positive function a(x) stands for the recruitment rate of the susceptible corresponding to births and immigration; a(x) and μ(x) are also positive Hölder continuous functions on ¯Ω. All the other parameters have the same assumptions as in (1.1) and (1.2).

    Since c>0, the term SIc+S+I is a smooth function of S and I in the first quadrant. By the standard theory for parabolic equations, combined with our assumption on the initial data, it is well known that (1.3) admits a unique classical solution (S,I) (namely, S,IC2,1(¯Ω×(0,)). Moreover, it follows from the strong maximum principle and the Hopf boundary lemma for parabolic equations that both S(x,t) and I(x,t) are positive for x¯Ω and t(0,).

    The aim of the present paper is to provide the theoretical analysis of solution to (1.3) and its steady-state (i.e., equilibrium) problem. The extinction or persistence behavior of the infectious disease in the long run is one of our main focuses. Once the disease can persist, what we are particularly interested in is the spatial distribution of the disease when the diffusion (migration) rate (represented by dS or dI in our current context) is controlled to be small. Such information will be useful for decision-makers to understand and predict the pattern of disease occurrence and then to take more effective actions/measures to eradicate diseases; one may refer to Section 5 for further discussion.

    We would like to mention that many research works have been devoted to the study of the related epidemic systems; one may see, for instance, [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34].

    The rest of our paper is organized as follows. In section 2, we derive the uniform bounds of the solution of (1.3) and discuss the uniform persistence in terms of the basic reproduction number. In section 3, we study the global attractivity of the disease-free equilibrium and the endemic equilibrium in some special cases. In section 4, we analyze the asymptotic profile of the endemic equilibrium (if it exists) as the diffusion coefficient dS or dI goes to zero. Section 5 ends the paper with some discussion of the epidemiological implications of the theoretical results obtained in this paper.

    In this section, we will establish the uniform boundedness of solutions to (1.3), and then study the uniform persistence property. In the following, we use DFE and EE to represent the disease-free equilibrium and the endemic equilibrium, respectively.

    From now on, for notational simplicity, we denote

    F=maxxˉΩF(x)andF=minxˉΩF(x)

    for any given function FC(¯Ω).

    Lemma 2.1. Assume that dS=dI. For any solution (S,I) of (1.3), there holds

     S(x,t)+I(x,t)max{am, s0+(1+ε1)I0},  x¯Ω, t0. (2.1)

    Here, ε1 is any given positive constant so that 1ε1β>0 and

    m1=min{με1β, (1+ε1)μ+ε1γ1+ε1}.

    Proof. Let dS=dI=d. For any given positive constant ε1 so that 1ε1β>0, we set

    V1(x,t)=S(x,t)+(1+ε1)I(x,t).

    In view of (1.3), we then have

    V1tdΔV1=a(x)μ(x)S+ε1β(x)SIC+S+Iμ(x)Iε1[γ(x)+μ(x)]Ia(x)(με1β)S(1+ε1)μ+ε1γ1+ε1(1+ε1)Ia(x)m1V1.

    It is easily seen that V1 and max{am1,S0+(1+ε1)I0} is a pair of upper and lower solutions to the initial-boundary value problem:

    {vtdΔv=a(x)m1v,xΩ,t>0,vν=0,xΩ,t>0,v(x,0)=S0(x)+(1+ε1)I0(x),xΩ.

    By the comparison principle for parabolic equations, we obtain

    S(x,t)+(1+ε1)I(x,t)V1(x,t)max{am1,S0+(1+ε1I0)},   x¯Ω ,t0.

    For the general case, we also derive the uniform boundedness of solution to (1.3). Indeed, we can state the following result.

    Lemma 2.2. For any solution (S,I) of (1.3), there exists a positive constant M1 depending on initial data such that

    S(,t)L(Ω)+I(,t)L(Ω)M1, t0. (2.2)

    Moreover, there exists some positive constant M2 independent of initial data such that

    S(,t)L(Ω)+I(,t)L(Ω)M2, tT, (2.3)

    for some large T>0.

    Proof. For any fixed constant ε2>0 such that με2β>0, we denote

    m2=min{uε2β, μ+ε2(γ+μ)1+ε2}.

    Then we set

    V2(t)=Ω[S(x,t)+(1+ε2)I(x,t)]dx,  t0.

    In view of (1.3), it can be easily shown that

    dV2dt=Ωa(x)dxΩμ(x)Sdx+ε2Ωβ(x)SIc+S+Idx            Ω[μ(x)+ε2(γ(x)+μ(x)]Idx       Ωa(x)dxμΩSdx+ε2βΩSdx[μ+ε2(γ+μ)]ΩIdx       =Ωa(x)dx(με2β)ΩSdxμ+ε2(γ+μ)1+ε2(1+ε2)ΩIdx       Ωa(x)dxm2V2,

    from which it follows that

    dV2dt+m2V2Ωa(x)dx|Ω|a.

    Hence, we have

     V2(t)V2(0)em2t+|Ω|am2(1em2t),  t0. (2.4)

    Note that S and I are nonnegative. Thus together with (2.4), one can use [6,Lemma 2.1] (or [35]) with σ=p0=1 to derive (2.2). Moreover, the inequality (2.4) infers that

    lim suptV2|Ω|am,

    which is independent of initial data. Making use of [6,Lemma 2.1] again, we can derive (2.3).

    In what follows, we discuss the uniform persistence of solution. To this aim, we first consider the elliptic problem:

     dSΔS=a(x)μ(x)S, xΩ;   Sν=0, xΩ. (2.5)

    Obviously, (2.5) admits a uniform solution ˆS>0, and (ˆS,0) is a unique disease-free equilibrium of (1.3), which we call as DFE.

    We then define the basic reproduce number R0:

    R0=supφH1(Ω),φ0{ΩβˆSc+ˆSφ2dxΩdI|φ|2+(γ+μ)φ2dx}. (2.6)

    Indeed, one can follow the idea of next generation operators in [35] to introduce the basic reproduction number, which coincides with the value R0.

    It should be noticed that the basic reproduction number R0 defined by (2.6) implicitly depends on the diffusion rate dS of the susceptible population; this qualitatively differs from the basic reproduction number R0 defined in [1] and [12].

    Let λ0 be the principal eigenvalue of the following eigenvalue problem

    {dIΔψ(γ+μβˆSc+ˆS)ψ+λψ=0,  xΩ,ψν=0,   xΩ. (2.7)

    Then, we have the following proposition; the proof is the same as [1,Lemma 2.3] and is omitted here.

    Proposition 2.3. The following statements hold.

    (a) R0 is a monotone decreasing function of dI with R0maxx¯ΩβˆS(c+ˆS)(γ+μ) as dI0 and R0ΩβˆS(c+ˆS)/Ω(γ+μ) as dI.

    (b) If ΩβˆS(x)c+ˆS(x)dx<Ω(γ(x)+μ(x))dx, and βˆSc+ˆS(γ+μ) changes sign, then there exists a threshold value dI(0,) so that R0<1 for dI>dI and R0>1 for dI<dI.

    (c) If ΩβˆS(x)c+ˆS(x)dxΩ(γ(x)+μ(x))dx, then R0>1 for all dI.

    (d) R0>1 when λ<0, R0=1 when λ=0, and R0<1 when λ>0.

    (e) R0 is a monotone decreasing function of c, and R0<1 if c>c for some c0.

    Proposition 2.4. If R0>1 then the DFE (ˆS,0) is unstable, and if R0<1, it is stable.

    The proof of Proposition 2 is similar to the proof of [1,Lemma 2.4] and hence the details are omitted.

    In view of (2.3) of Lemma 2.2, we can establish the uniform persistence of (1.3) when R0>1. In fact, according to the theory developed by Magal and Zhao (see [36,Theorem 4.5] or [37]), we can conclude the following theorem.

    Theorem 2.5. If R0>1, then there exists some real number η>0 independent of the initial data, such that any solution (S,I) of (1.3) satisfies

    liminftS(x,t)ηandliminftI(x,t)ηuniformlyforxˉΩ,

    and hence, the disease persists uniformly. Furthermore, (1.3) admits at least one EE when R0>1.

    This section is devoted to the study of the global attractivity of the DFE and EE of (1.3). In the first subsection, we obtain the global attractivity of the DFE in the the nonhomogeneous environment. In the second subsection, we will derive that the global attractivity of the EE in the homogeneous environment.

    For later purpose, we need a useful lemma; see [38,Lemma 2.5.1].

    Lemma 3.1. Let a1 and a2>0 be any constants. Assume that φ,ψC1([a1,)),ψ(t)0 in [a1,) and φ is bounded from below. If φ(t)a2ψ(t) and ψ(t)K in [a1,) for some constant K, then limtψ(t)=0.

    Our main result of this subsection reads as follows.

    Theorem 3.2. The DFE (ˆS,0) is globally attractive if one of the following conditions holds:

    (ⅰ) β(x)γ(x)+μ(x), x¯Ω;

    (ⅱ) c>c for some constant c>0.

    Proof. We first handle case (ⅰ). To verify our result, we construct the following Lyapunov function

    V(t)=12ΩI2(x,t)dx,  t0.

    Hereafter (S,I) is the solution of (1.3).

    Then some elementary calculation yields

    V(t)=ΩI Itdx   =ΩI[dIΔI+β(x)SIc+S+I(γ(x)+μ(x))I]dx   =Ω|I|2dx+Ωβ(x)Sc+S+II2dxΩ(γ(x)+μ(x))I2dx   Ω|I|2dx+Ωβ(x)I2dxΩ(γ(x)+μ(x))I2dx   (γ(x)+μ(x)β(x))ΩI2dx0.

    This motivates us to define

    ψ(t)=(γ(x)+μ(x)β(x))ΩI2dx0,  t0.

    Due to the Lemma 2.2, we know that both S(,t)L(Ω) and I(,t)L(Ω) are bounded. Thus, by [39,Theorem A2], we have

    S(,t)C2+α(¯Ω)+I(,t)C2+α(¯Ω)C0,  t1, (3.1)

    for some constant C0>0. In addition, using the second equation of (1.3), we can see that ψ(t) is bounded from above for t[1,). In view of Lemma 3 (by taking φ(t)=V(t)), we can conclude that

    I(,t)0  in L2(Ω),  as t. (3.2)

    Moreover, in light of (3.1), it is clear that the set {I(,t): t1} is compact in C2(¯Ω). Combining this fact with (3.2), we assert that

    I(,t)0  in C2(¯Ω),   as t.

    Thus, for any small ϵ>0, it follows that

    I(x,t)ϵ,  x¯Ω, tT,

    for some large T.

    Using the above fact, it is easy to find that S is a lower solution of the following parabolic problem:

    {¯wtdSΔ¯w=a(x)μ(x)¯w+εγ,xΩ, t>T,¯wν=0,xΩ, t>T,¯w(x,T)=S(x,T),xΩ. (3.3)

    Let w1 be the solution of (3.3). Then by the comparison principle, we have

    S(x,t)w1(x,t),    x¯Ω,tT.

    Similarly, we can find that S is an upper solution to

    {w_tdSΔw_=a(x)μ(x)w_εβ,xΩ, t>T,w_ν=0,xΩ, t>T,w(x,T)=S(x,T),xΩ. (3.4)

    Thus, by letting w2 be the solution of (3.4), we have

    S(x,t)w2(x,t),  x¯Ω, tT.

    It is standard to show that problem (3.3) and problem (3.4) exist a unique positive steady state, denoted by ˆS(ε,x) and ˆS+(ε,x), respectively. Moreover, we have

    w1(x,t)ˆS(ε,x)andw2(x,t)ˆS+(ε,x)uniformlyonˉΩ,ast,

    On the other hand, it is easy to check that

    ˆS(ε,x),ˆS+(ε,x)ˆS(x)uniformlyforxˉΩ,asε0.

    According to the arbitrariness of ε, we conclude that

    S(x,t)ˆS(x)  uniformly on ¯Ω, as t.

    This proves our assertion in case (ⅰ).

    We next consider case (ⅱ). First of all, one can check the proof of Lemma 2.2 and claim that

    S(x,t)C0,   x¯Ω, t0, (3.5)

    for some positive constant C0, depending on the initial data but independent of c0. Hence, by (3.5), it is easily noticed that I is a lower solution of the following parabolic problem:

    {¯wtdIΔ¯w=β(x)C0c+C0¯w[γ(x)+μ(x)]¯w,xΩ, t>0,¯wν=0,xΩ, t>0,¯w(x,0)=I0(x),xΩ. (3.6)

    Let w3 be the solution of (3.6). Since C0 does not depend on c, one can take c to be large so that

    β(x)C0c+C0[γ(x)+μ(x)]<0,   x¯Ω.

    Then, a simple analysis, together with the parabolic comparison principle, shows that

    I(x,t)w3(x,t)0  uniformly for x¯Ω,  as t.

    Now, in view of the above assertion, a similar argument as in case (ⅰ) allows us to conclude that

    S(x,t)ˆS(x)  uniformly on ¯Ω, as t.

    This proves our assertion in case (ⅱ). The proof is complete.

    In this subsection, we will consider the global attractivity of the EE by assuming that all of the parameters a,β,μ and γ are positive constant (that is, the environment is spatially homogeneous).

    In this situation, in view of (3.1), we can see that the unique DFE is given by (ˆS,0)=(aμ,0). And the unique EE (˜S,˜I) exists if and only if R0=βaμ(γ+μ)(c+aμ)>1, where

    ˜S=(γ+μ)(c+aμ)β=1R0aμ,   ˜I=aμ1β(μ+γ)(c+aμ)=(γ+μ)(c+aμ)β(R01).

    Our result is stated as follows.

    Theorem 3.3. Assume that the parameters a,β,μ and γ are positive constant and dS=dI. If R0>1, then the EE (˜S,˜I) is globally attractive.

    Proof. By setting dS=dI=d, we construct the following Lyapunov functional

    W(t)=ΩM(S(x,t),I(x,t))dx,  t>0,

    where

    M(S,I)=(S˜S)+(I˜I)(c+˜S+˜I)lnc+S+Ic+˜S+˜I              +2μ(c+˜S+˜I)β(c+˜I)(I˜I˜IlnI˜I).

    For simplicity, we note

    g1(S,I)=aμSβSIc+S+I+γI,   g2(S,I)=βSIc+S+I(γ+μ)I.

    Then simple calculation gives

    W(t)=Ω[MS(S,I)St+MI(S,I)It]dx=dΩ[MS(S,I)ΔS+MI(S,I)ΔI]dx +Ω[MS(S,I)g1(S,I)+MI(S,I)g2(S,I)]dx.

    Moreover, integrating by parts, we have

    ΩMS(S,I)ΔSdx=Ω[MSS(S,I)|S|2+MSI(S,I)SI]dx,
    ΩMI(S,I)ΔIdx=Ω[MIS(S,I)SI+MII(S,I)|I|2]dx.

    It is easy to see that

    MSS=MSI=MIS=c+˜S+˜I(c+S+I)2,   MII=c+˜S+˜I(c+S+I)2+2μ(c+˜S+˜I)˜Iβ(c+˜I)I2.

    Thus

    dΩ[MS(S,I)ΔS+MI(S,I)ΔI]dx=dΩ{c+˜S+˜I(c+S+I)2|S|2+2c+˜S+˜I(c+S+I)2SI  +[c+˜S+˜I(c+S+I)2+2μ(c+˜S+˜I)˜Iβ(c+˜I)I2]|I|2}dx=dΩ{c+˜S+˜I(c+S+I)2|(S+I)|2+2μ(c+˜S+˜I)˜Iβ(c+˜I)I2|I|2}dx0. (3.7)

    In addition, by direct computations, we have

    MS(S,I)g1(S,I)+MI(S,I)g2(S,I)=(1c+˜S+˜Ic+S+I)g1+[1c+˜S+˜Ic+S+I+2μ(c+˜S+˜I)β(c+˜I)(1˜II)]g2=(1c+˜S+˜Ic+S+I)(g1+g2)+2μ(c+˜S+˜I)β(c+˜I)(1˜II)g2=(1c+˜S+˜Ic+S+I)(aμSμI)+2μ(c+˜S+˜I)β(c+˜I)(1˜II)[βSIc+S+I(γ+μ)I]=(S˜S)+(I+˜I)c+S+I[μ(S˜S)μ(I˜I)]+2μ(c+˜S+˜I)β(c+˜I)(I˜I)(βSc+S+Iβ˜Sc+˜S+˜I)=μ[(S˜S)+(I+˜I)]2c+S+I+2μ(c+˜S+˜I)c+˜I(I˜I)(Sc+S+I˜Sc+˜S+˜I)=μ(S˜S)2c+S+I2μ(S˜S)(I˜I)c+S+Iμ(I˜I)2c+S+I+2μ(c+˜S+˜I)c+˜I(I˜I)(c+˜I)(S˜S)˜S(I˜I)(c+S+I)(c+˜S+˜I)=μ(S˜S)2c+S+I2μ(S˜S)(I˜I)c+S+Iμ(I˜I)2c+S+I+2μ(S˜S)(I˜I)c+S+I2μ˜S(I˜I)2(c+˜I)(c+S+I)=μ(S˜S)2c+S+Iμ(I˜I)2c+S+I2μ˜S(I˜I)2(c+˜I)(c+S+I)0.

    Here, we have used the fact:

    a=μ(˜S+˜I),   γ+μ=β˜Sc+˜S+˜I.

    According to (3.7), we thus derive that W(t)0, t>0 along all trajectories. By a standard argument, we can obtain

    (S(,t),I(,t))(˜S,˜I)  in[L2(Ω)]2,  as t.

    Due to the Lemma 2.2, we know that both S(,t)L(Ω) and I(,t)L(Ω) are bounded. As a consequence, by [39,Theorem A2], we have

    S(,t)C2+α(¯Ω)+I(,t)C2+α(¯Ω)C0,  t1,

    Thus, {(S(,t),I(,t)): t1} is compact in C2(¯Ω)×C2(¯Ω). Then, combining this with the above L2-convergence, we obtain that

    (S(,t),I(,t))(˜S,˜I)  in[C2(¯Ω)]2,  as t.

    Thus, the EE (˜S,˜I) is globally attractive.

    In this section, we consider the asymptotic behavior of the positive solution of the following elliptic system:

    {dSΔS=a(x)μ(x)Sβ(x)SIc+S+I+γ(x)I,xΩ,dIΔI=β(x)SIc+S+I[γ(x)+μ(x)]I,xΩ,Sν=Iν=0,xΩ, (4.1)

    when dS or dI goes to zero.

    Via a singular perturbational argument, it is easily seen that (2.5) has a unique positive solution ˆS which converges uniformly to ˆS1(x)=a(x)μ(x) as dS0. We also recall that λ0 is the principal eigenvalue of problem (2.7). Then, it follows that

    λ0˜λ,  as  dS0,

    where ˜λ is the principle eigenvalue of the following eigenvalue problem

    {dIΔψ(γ+μβˆS1c+ˆS1)ψ+λψ=0,  xΩ,ψν=0,   xΩ. (4.2)

    In order to ensure that the elliptic system (4.1) admits a positive solution, we always assume ˜λ<0 in this subsection. Then, we will investigate the asymptotic behavior of positive solution of (4.1) as dS0 while dI>0 is fixed.

    Theorem 4.1. Fix dI>0 and assume that ˜λ<0. Let dS0, then every positive solution (SdS,IdS) of (4.1) satisfies (up to a subsequence of dS0)

    (SdS,IdS)(WS,WI)  uniformly on ¯Ω,

    where

    WS(x)=J(x,WI(x)):=12{a+γWIμ(c+WI)βWIμ+[a+γWIμ(c+WI)βWIμ]2+4(a+γWI)(c+WI)μ}

    and WI is a positive solution of

    {dIΔWI=β(x)J(x,WI)WIc+J(x,WI)+WI[γ(x)+μ(x)]WI, xΩ,WIν=0, xΩ. (4.3)

    Proof. Mentioned as before, (4.1) has at least one EE for all small dS when ˜λ<0. In what follows, we divide the proof into three steps to derive the conclusion.

    Step 1. A priori bounds for S,I. Integrating the first and the second equation of (4.1) over Ω, respectively, we have

    Ω[μ(x)S+β(x)SIc+S+I]dx=Ωa(x)dx+Ωγ(x)Idx (4.4)

    and

    Ωβ(x)SIc+S+Idx=Ω[γ(x)+μ(x)]Idx. (4.5)

    Inserting (4.5) into (4.4) gives

    Ωμ(x)Idx+Ωμ(x)Sdx=Ωa(x)dx. (4.6)

    From (4.6), we get

    ΩIdx+ΩSdx|Ω|aμ. (4.7)

    It is obvious that the L1-bounds of S and I are independent of both dS and dI.

    We write the I-equation as follows

    {ΔI+1dI[β(x)Sc+S+Iγ(x)μ(x)]I=0,xΩ,Iν=0,xΩ. (4.8)

    Then, we apply the Harnack-type inequality (refer to [40] or [41,Lemma 2.2]) to (4.8) to assert that

    max¯ΩICmin¯ΩI. (4.9)

    Hereafter, the positive constant C is independent of dS>0, and it may vary from place to place.

    By (4.7) and (4.9), we can derive

    I(x)max¯ΩICmin¯ΩIC|Ω|ΩIdxC,  x¯Ω. (4.10)

    Step 2. Convergence of I. From (4.10), we get

    1dI[β(x)Sc+S+Iγ(x)μ(x)]ILp(Ω)C,p>1.

    Applying the standard Lp-estimate for elliptic equations ([42]), it follows that

    IW2,p(Ω)Cforanygivenp>1.

    Then, taking p to be sufficiently large and using the embedding theorem [42], we can see that

    IC1+α(¯Ω)C  for some 0<α<1.

    Hence, there exist a subsequence of dS0, say di:=dS,i, satisfying di0 as i, and a corresponding positive solution (Si,Ii):=(SdS,i,IdS,i) of (4.1) with dS=di, such that

    IiWI  uniformly on ¯Ω,  as i, (4.11)

    where WIC1(¯Ω) and WI0. By (4.9), we know that

    either WI0  on ¯Ω  or  WI>0  on ¯Ω. (4.12)

    Suppose that WI0. That is,

    IiWI0  uniformly on ¯Ω,  as i.

    Then for sufficiently small ε with 0<ε<minx¯Ωa(x), we have 0Ii(x)ε,x¯Ω, for all large i. Combining this fact with the first equation of (4.1), for all large i, one sees that (Si,Ii) satisfies

    diΔSia(x)μ(x)Si+εγ, xΩ;  Siν=0, xΩ.

    This leads us to consider the following auxiliary system:

    diΔSi=a(x)μ(x)Si+εγ, xΩ;  Siν=0, xΩ. (4.13)

    It is clear that (4.13) admits a unique positive solution, denoted by ui. A simple upper and lower solution argument guarantees that

    Siui  on ¯Ω,  for all large i. (4.14)

    Similarly, for all large i, (Si,Ii) satisfies

    diΔSia(x)μ(x)Siεβ, xΩ;  Siν=0, xΩ.

    We also consider the following auxiliary system:

    diΔSi=a(x)μ(x)Siεβ, xΩ;  Siν=0, xΩ. (4.15)

    Let vi denote the unique positive solution of (4.15). Similarly as before, we have

    Sivi   on ¯Ω,  for all large i. (4.16)

    By a singular perturbation argument as in [43,Lemma 2.4], it is easy to show that

    uia(x)+εγμ(x),  via(x)εβμ(x)  uniformly on ¯Ω, as i.

    Hence, sending i, by (4.14) and (4.16), we find

    a(x)εβμ(x)lim infiSi(x)lim supiSi(x)a(x)+εγμ(x) on ¯Ω.

    Due to the arbitrariness of ε, we obtain that

    Sia(x)μ(x)  uniformly on ¯Ω,  as i. (4.17)

    We now consider the second equation of (4.1) and then know that Ii satisfies

    dIΔIi=β(x)SiIic+Si+Ii[γ(x)+μ(x)]Ii, xΩ;   Iiν=0, xΩ. (4.18)

    Define ˆIi:=IiIiL(Ω). Then ˆIiL(Ω)=1 for all i1, and ˆIi solves

    dIΔˆIi=[β(x)Sic+Si+Iiγ(x)+μ(x)]ˆIi, xΩ;   Iiν=0, xΩ. (4.19)

    As before, using the standard compactness argument for the elliptic equation, after passing to a further subsequence if necessary, we assume that

    ˆIiˆI  in C1(¯Ω),  as i,

    where ˆIC1(¯Ω) with ˆI0 on ¯Ω and ˆIL(Ω)=1.

    Together with the fact Ii0 uniformly on ¯Ω as i, from (4.17) and (4.19), one can show that ˆI fulfills

    dIΔˆI=[β(x)a(x)μ(x)c+a(x)μ(x)γ(x)+μ(x)]ˆI, xΩ;   ˆIν=0, xΩ. (4.20)

    Applying the Harnack-type inequality ([40] or [41,Lemma 2.2]) to (4.20), we obtain ˆI>0 on ¯Ω. This implies that ˜λ of the eigenvalue problem (4.2) must be zero. It contradicts our assumption that ˜λ<0. Hence, the latter always holds in (4.12). That is

    IiWI>0  uniformly on ¯Ω,  as i. (4.21)

    Step 3. Convergence of S. Observe that Si fulfills

    {diΔSi=a(x)μ(x)Siβ(x)SiIic+Si+Ii+γ(x)Ii,xΩ,Siν=0,xΩ.

    From (4.21), given any small ε>0, it holds

    0<WIεIiWI+ε,   xΩ, (4.22)

    for all large i. For simplicity, denote W±,I=WI±ε. Hence, we have

    aμSiβSiIic+Si+Ii+γIiaμSiβSiW,Ic+Si+W,I+γW+,I=μ(J1,ϵ+(x,WI(x))Si)(SiJ1,ϵ(x,WI(x)))c+Si+(WIϵ),

    where

    J1,ϵ±(x,WI(x))=12{a+γW+,Iμ(c+W,I)βW,Iμ±[a+γW+,Iμ(c+W,I)βW,Iμ]2+4(a+γW+,I)(c+W,I)μ},

    for all large i.

    Notice that J1,ϵ+>0 and J1,ϵ<0 on ¯Ω. Then, given large i, we consider the following auxiliary elliptic problem:

    {diΔz=μ(J1,ϵ+(x,WI(x))z)(zJ1,ϵ(x,WI(x)))c+z+W,I, xΩ,zν=0,xΩ. (4.23)

    It is easily observed that Si is a lower solution of (4.23). And any sufficiently large constant C>0 satisfies SiC is an upper solution of (4.23). Therefore, (4.23) admits at least one positive solution, denoted by Zi, which satisfies SiZiC on ¯Ω. By similar arguments as in the proof of [43,Lemma 2.4], we find that

    ZiJ1,ϵ+(x,WI(x))   uniformly on ¯Ω as i.

    Since Si is a lower solution of (4.23), we have

    lim supiSi(x)J1,ϵ+(x,WI(x))  uniformly on ¯Ω. (4.24)

    On the other hand, from (4.22), for all large i, we have

    aμSiβSiIic+Si+Ii+γIiaμSiβSiW+,Ic+Si+W+,I+γW,I=μ(J1,ϵ+(x,WI(x))Si)(SiJ1,ϵ(x,WI(x)))c+Si+(WI+ϵ),

    where

    J2,ϵ±(x,WI(x))=12{a+γW,Iμ(c+W+,I)βW+,Iμ±[a+γW,Iμ(c+W+,I)βW+,Iμ]2+4(a+γW,I)(c+W+,I)μ},

    with J2,ϵ+(x,WI(x))>0 and J2,ϵ(x,WI(x))<0 on ¯Ω.

    As before, for any given large i, we also consider the following auxiliary elliptic problem:

    {diΔz=μ(J2,ϵ+(x,WI(x))z)(zJ2,ϵ(x,WI(x)))c+z+W+,I, xΩ,zν=0,xΩ. (4.25)

    Observe that Si and 0 is a pair of upper and lower solution of (4.25). Hence, we can assert that (4.25) admits at least one positive solution. Using a similar argument as before, we further get

    lim infiSi(x)J2,ϵ+(x,WI(x))  uniformly on ¯Ω. (4.26)

    Notice that

    J1,0+(x,WI(x))=J2,0+(x,WI(x))=J(x,WI(x)).

    Due to the arbitrariness of ε, by (4.24) and (4.26), we derive that

    Si(x)J(x,WI(x))uniformlyforxˉΩ,asi.

    In addition, by (4.18), we can easily see that WI satisfies (4.3). The proof is complete.

    In this subsection, we will analyze the asymptotic behavior of positive solution of (4.1) as dI0 with dS>0 being fixed. According to Proposition 2.3(a) and Theorem 2.5, we need to assume that {β(x)ˆS(x)/c+ˆS(x)>γ(x)+μ(x): x¯Ω} is nonempty so that (4.1) has positive solution for all small di.

    As usual, we denote g+=max{g,0}. Our main result can be stated as follows.

    Theorem 4.2. Fix dS>0 and assume that {β(x)ˆS(x)/c+ˆS(x)>γ(x)+μ(x): x¯Ω} is nonempty. Let dI0, then every positive solution (SdI,IdI) of (4.1) fulfills

    (SdI,IdI)(WS,WI)   uniformly on ¯Ω,

    where WS is the unique positive solution of

    {dSΔWS=a(x)μ(x)WSβ(x)WSWIc+WS+WI+γ(x)WI, xΩ,WSν=0, xΩ, (4.27)

    and WI is a nonnegative function

    WI={[β(x)γ(x)μ(x)]WSc[γ(x)+μ(x)]γ(x)+μ(x)}+. (4.28)

    Proof. We divide the proof into four steps for clarity. In the following, let us denote m to be a positive constant which does not depend on dI>0 and may vary from line to line.

    Step 1. Lower bound of S. Pick x1¯Ω so that S(x1)=min¯ΩS(x). In light of the first equation of (4.1), it follows from [44,Proposition 2.2] that

    a(x1)μ(x1)S(x1)β(x1)S(x1)I(x1)c+S(x1)+I(x1)+γ(x1)I(x1)0. (4.29)

    By (4.29), we obtain

    a<a(x1)+γ(x1)I(x1)μ(x1)S(x1)+β(x1)S(x1)I(x1)c+S(x1)+I(x1)μS(x1)+βS(x1),

    from which we further have

    S(x)S(x1)aμ+β>0. (4.30)

    Step 2. W1,q-bound of S for some q1. We now write the S-equation as

    {dSΔS+[μ(x)β(x)Ic+S+I]S=a(x)+γ(x)I, xΩ,Sν=0, xΩ. (4.31)

    From (4.7), we derive Ω|a(x)+γ(x)I|dxm. Hence, by the L1-estimate theory for elliptic equations (see [45,Lemma 2.2] or [46]), we get

    SW1,q(Ω)m,q[1,N/(N1)),(or 1q< if N=1). (4.32)

    Step 3. Lp-bound of S and I. In view of (4.32), by the Sobolev embedding theorem, we can see that W1,q(Ω) is compactly embedded into Lp0(Ω), p0[1,N/(Nq)). This implies that

    SLp0(Ω)m,1<p0NqNq.

    As q is close to N/(N1), it is clear that

    SLp0(Ω)m,1<p0<NN2. (4.33)

    Notice that (4.33) holds for any 1<p0< when N=2.

    We now multiply the second equation of (4.1) by Ik for any fixed k>0 and then integrate by parts to obtain that

    0dIkΩIk1|I|2dx=Ωβ(x)SIk+1c+S+IdxΩ[γ(x)+μ(x)]Ik+1dx.

    By direct calculation, we get

    [γ+μ]ΩIk+1dxβΩSIk+1c+S+Idx. (4.34)

    Taking 1p0+1q0=1 (notice 1q01=p01) and k0=1q0, by (4.7), (4.33), (4.34) and Hölder inequality, we have

    [γ+μ]ΩIk0+1dxβΩSIk0+1c+S+Idxβ(ΩSp0dx)1/p0(ΩIdx)1/q0m.

    This means that

    ILk0+1(Ω)m. (4.35)

    Then, we take k1=(k0+1)/q0=1/q0+1/q20. As before, by (4.34) and Hölder inequality, together (4.33) and (4.35), we infer that

    [γ+μ]ΩIk1+1dxβΩSIk1+1c+S+Idxβ(ΩSp0dx)1/p0(ΩIdx)1/q0m,

    that is,

    ILk1+1(Ω)m.

    Repeating the iteration as above, we can easily see that

    ILk+1(Ω)m, (4.36)

    where

    k=1q0+1q20+1q30+=1q01=p01.

    Notice that (4.36) can be deduced through finitely many times of iterations with the help of (4.33). Thus, we deduce

    ILp0(Ω)m. (4.37)

    Combining (4.33) and (4.37), from the equation (4.31) by using the well-known Lp-theory, one can assert that

    SW2,p0(Ω)m.

    By the Sobolev embedding theorem again, W2,p0(Ω) is compactly embedded into Lp1(Ω), p1(1,Np0/(N2p0)). Observe that Np0N2p0NN4 as p0NN2 (see (4.33)). Thus, we have

    SLp1(Ω)m,1<p1<NN4  or 1<p1< if N4.

    By a similar argument as in deducing (4.37), one gets

    ILp1(Ω)m.

    Making use of (4.31), the Sobolev embedding theorem and the well-known Lp-theory repeatedly, we can eventually conclude that

    SLp(Ω),ILp(Ω)m,1p<. (4.38)

    Step 4. Convergence of S and I. According to (4.38), for the equation (4.31), it holds that

    SW2,p(Ω)C,1<p<.

    Then taking sufficiently large p, the standard embedding theorem enables us to conclude that, up to a sequence of dI0, denoted by dj:=dI,j, with dj0 as j, the corresponding positive solution sequence (Sj,Ij):=(SdI,j,IdI,j) of (4.1) with dI=dI,j fulfills

    SjWS  in C1(¯Ω), as j, (4.39)

    where WSC1(¯Ω) and WS>0 on ¯Ω due to (4.30). Observe that Ij fulfills

    {djΔIj=β(x)SjIjc+Sj+Ij[γ(x)+μ(x)]Ij, xΩ,Ijν=0, xΩ. (4.40)

    In light of (4.39) and (4.40), using a simple upper and lower solution similarly as in step 3 of Theorem 4.1, we have

    IjWIinC1(ˉΩ),asj,

    where WI is given by (4.28).

    It is clear that WS satisfies (4.27). Moreover, by the expression of WI, we can see that (4.27) admits a unique positive solution (refer to [47,Lemma A.1]). Thus, we can conclude that all the above limits hold without passing to a subsequence. This proof is complete.

    In this paper, we have studied the SIS reaction-diffusion model (1.3) in which we have taken into account the natural mortality of the susceptible and infected populations. First of all, we have established the uniform bounds of solution to (1.3); see Lemma 2.1 and Lemma 2.2. Then, we define the basic reproduction number R0 associated with (1.3):

    R0=supφH1(Ω),φ0{ΩβˆSc+ˆSφ2dxΩdI|φ|2+(γ+μ)φ2dx}.

    It is worth mentioning that \mathcal{R}_0 depends on the diffusion rates d_{S} and d_{I} when c > 0 , while \mathcal{R}_0 depends only on the diffusion rate d_{I} when c = 0 . Thus, compared with the model (1.2) (i.e., c = 0 ), the parameters d_S and c play vital roles in the dynamics of the infectious disease in (1.3). In particular, we have proved that \mathcal{R}_0 is decreasing with respect to c\in[0, \infty) , and when c is larger than a value, the basic reproduction number \mathcal{R}_0 < 1 so that the disease dies out in the long run; see Theorem 3.1(ⅱ).

    In another special case that \beta(x)\leq \gamma(x)+\mu(x), \ \forall x\in\overline{\Omega} , we have proved the global stability of the disease-free equilibrium via a Lyapunov function method; see Theorem 3.1. On the other hand, when the spatial environment is homogeneous, that is, all the parameters in (1.3) are positive constants, we have shown the global stability of the endemic equilibrium provided that the diffusion rates are equal and the basic reproduction number \mathcal{R}_0 > 1 ; see Theorem 3.3. This result means that the disease will persist all the time. In the general situation of spatially heterogeneous environment, once \mathcal{R}_0 > 1 , the uniform persistence property has been proved so that the disease exists eventually in the whole habitat; refer to Theorem 2.5. We suspect that the uniform persistence property holds if \mathcal{R}_0 > 1 whereas the disease extinction occurs if \mathcal{R}_0\leq1 ; this is a challenging problem and deserves future investigation.

    According to Theorem 2.5 as well as the above discussion, once \mathcal{R}_0 > 1 , (1.3) has an endemic equilibrium exists and the infectious disease will uniformly persist in space, and vice versa. Therefore, it becomes important to understand how the heterogeneity of spatial environment and the mobility of population dispersal (reflected by the change of the migration rates d_I and d_S ) prescribe the spatial profile of the endemic equilibrium, because this will help decision-makers to predict the pattern of disease occurrence and henceforth to conduct effective/optimal control strategies of disease eradication. This leads us to explore the asymptotic behavior of endemic equilibrium with respect to small diffusion rate d_S or d_I , which in turn will tell us the spatial distribution of susceptible and infected population

    Theorems 4.1 and 4.2 show that as the mobility of the susceptible or infected population goes to zero, the infectious disease will always exist in space at least in some region. Especially, it follows from Theorem 4.1 that the disease exists in the entire habitat even if the mobility of the susceptible is restricted to be small enough. This result is in sharp contrast with that of (1.1), as shown in [1] by Allen et al, where they proved that as the mobility of the susceptible goes to zero, the density of the infected population will vanish and so the disease dies out eventually. However, Peng in [12] showed that, for the model (1.2) which also includes the linear external source term \Lambda(x)-S , the density of the infected population will not vanish when the mobility of the susceptible or infected population goes to zero; that is, the infectious disease will always exist.

    The above results suggest that simply controlling the migration rate of the susceptible or infected population can not eliminate the disease modelled by (1.2) and (1.3). In other words, the presence of the linear external source term \Lambda(x)-S enhances the persistence of disease and the infectious disease will become more threatening and hard to control. As a consequence, in such a situation, more effective measures should be taken to eradicate diseases.

    We would like to express our sincere thanks to the referees for their careful reading and valuable suggestions, which improve the presentation of the manuscript.

    The authors declare there is no conflicts of interest.



    [1] L. J. S. Allen, B. M. Bolker, Y. Lou, et al., Asymptotic profiles of the steady states for an SIS epidemic reaction-diffusion model, Dis. Contin. Dyn. Syst. A, 21 (2008), 1-20.
    [2] R. Peng, Asymptotic profiles of the positive steady state for an SIS epidemic reaction-diffusion model. Part I, J. Differ. Equations, 247 (2009), 1096-1119.
    [3] R. M. Anderson and R. M. May, Populaition biology of infectious diseases, Nature 280 (1979), 361-367.
    [4] R. Cui and Y. Lou, A spatial SIS model in advective heterogeneous environments, J. Differ. Equations, 261 (2016), 3305-3343.
    [5] R. Cui, K.-Y. Lam and Y. Lou, Dynamics and asymptotic profiles of steady states of an epidemic model in advective environments, J. Differ. Equations, 263 (2017), 2343-2373.
    [6] Z. Du and R. Peng, A priori L-estimates for solutions of a class of reaction-diffusion systems, J. Math. Biol., 72 (2016), 1429-1439.
    [7] H. W. Hethcote, Epidemiology models with variable population size, Mathematical understanding of infectious disease dynamics, 63-89, Lect. Notes Ser. Inst. Math. Sci. Natl. Univ. Singap., 16, World Sci. Publ., Hackensack, NJ, 2009.
    [8] W. O. Kermack and A. G. McKendrick, Contribution to the mathematical theory of epidemics-I, Proc. Roy. Soc. London Ser. A, 115 (1927), 700-721.
    [9] M. Martcheva, An introduction to mathmatical epidemiology, Springer,New York, (2015).
    [10] H. W Hethcote, The mathematics of infectious diseases, SIAM Rev., 42 (2000), 599-653.
    [11] B. Li, H. Li and Y. Tong, Analysis on a diffusive SIS epidemic model with logistic source, Z. Angew. Math. Phys., 68 (2017), Art. 96, 25pp.
    [12] H. Li, R. Peng and F. B. Wang, Varying total population enhances disease persistence: qualitative analysis on a diffusive SIS epidemic model, J. Differ. Equations, 262 (2017), 885-913.
    [13] M. E. Alexander and S. M. Moghadas, Bifurcation Analysis of an SIRS Epidemic Model with Generalized Incidence, SIAM J. Appl. Math., 65 (2001), 1794-1816.
    [14] R. M. Anderson and R. M. May, Regulation and stability of host-parasite interactions.I. Regulatory processes, J. Anim. Ecol., 47 (1978), 219-247.
    [15] O. Diekmann and M. Kretzschmar, Patterns in the effects of infectious diseases on population growth, J. Math. Biol., 29 (1991), 539-570.
    [16] J. A. P. Heesterbeck and J. A. J. Metz, The saturating contact rate in marriage and epidemic models, J. Math. Biol., 31 (1993), 529-539.
    [17] M. G. Roberts, The dynamics of bovine tuberculosis in possum populations and its eradication or control by culling or vaccination, J. Anim. Ecol., 65 (1996), 451-464.
    [18] Y. Cai, K. Wang and W. Wang, Global transmission dynamics of a Zika virus model, Appl. Math. Lett., 92 (2019), 190-195.
    [19] L. Chen and J. Sun, Optimal vaccination and treatment of an epidemic network model, Physics Lett. A, 378 (2014), 3028-3036.
    [20] L. Chen and J. Sun, Global stability and optimal control of an SIRS epidemic model on heterogeneous networks, Physica A, 410 (2014), 196-204.
    [21] K. Deng and Y. Wu, Dynamics of a susceptible-infected-susceptible epidemic reaction-diffusion model, Proc. Roy. Soc. Edinburgh Sect. A, 146 (2016), 929-946.
    [22] X. Gao, Y. Cai, F. Rao, et al., Positive steady states in an epidemic model with nonlinear incidence rate, Comput. Math. Appl., 75 (2018), 424-443. doi: 10.1016/j.camwa.2017.09.029
    [23] J. Ge, C. Lei and Z. Lin, Reproduction numbers and the expanding fronts for a diffusion-advection SIS model in heterogeneous time-periodic environment, Nonlinear Anal. Real World Appl., 33 (2017), 100-120.
    [24] K. Kuto, H. Matsuzawa and R. Peng, Concentration profile of endemic equilibrium of a reactiondiffusion-advection SIS epidemic model, Calc. Var. Partial Dif., 56 (2017), Art. 112, 28 pp.
    [25] C. Lei, F. Li and J. Liu, Theoretical analysis on a diffusive SIR epidemic model with nonlinear incidence in a heterogeneous environment, Discrete Contin. Dyn. Syst. Ser. B, 23 (2018), 4499- 4517.
    [26] B. Li and Q. Bie, Long-time dynamics of an SIRS reaction-diffusion epidemic model, J. Math. Anal. Appl., 475 (2019), 1910-1926.
    [27] H. Li, R. Peng and Z. Wang, On a diffusive SIS epidemic model with mass action mechanism and birth-death effect: analysis, simulations and comparison with other mechanisms, SIAM J. Appl. Math., 78 (2018), 2129-2153.
    [28] H. Li, R. Peng and T. Xiang, Dynamics and asymptotic profiles of endemic equilibrium for two frequency-dependent SIS epidemic models with cross-diffusion, Eur. J. Appl. Math., 2019, https://doi.org/10.1017/S0956792518000463, in press.
    [29] Z. Lin, Y. Zhao and P. Zhou, The infected frontier in an SEIR epidemic model with infinite delay, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2355-2376.
    [30] R. Peng and S. Liu, Global stability of the steady states of an SIS epidemic reaction-diffusion model, Nonlinear Anal., 71 (2009), 239-247.
    [31] L. Pu and Z. Lin, A diffusive SIS epidemic model in a heterogeneous and periodically evolving environment, 16 (2019), 3094-3110.
    [32] X. Wen, J. Ji and B. Li, Asymptotic profiles of the endemic equilibrium to a diffusive SIS epidemic model with mass action infection mechanism, J. Math. Anal. Appl., 458 (2018), 715-729.
    [33] Y. Wu and X. Zou, Asymptotic profiles of steady states for a diffusive SIS epidemic model with mass action infection mechanism, J. Differ. Equations, 261 (2016), 4424-4447.
    [34] M. Zhu, X. Guo and Z. Lin, The risk index for an SIR epidemic model and spatial spreading of the infectious disease, Math. Biosci. Eng., 14 (2017), 1565-1583.
    [35] R. Peng and X. Zhao, A reaction-diffusion SIS epidemic model in a time-periodic environment, Nonlinearity, 25 (2012), 1451-1471.
    [36] P. Magal and X.-Q Zhao, Global attractive and steady states for uniformly persistent dynamical systems, SIAM. J. Math. Anal., 37 (2005), 251-275.
    [37] X.-Q. Zhao, Dynamical Systems in Populaition Biology, Springer-Verlag, New York,(2003)
    [38] M. Wang, Nonlinear Partial Differential Equations of Parabolic Type, Science Press, Beijing, 1993(in chinese).
    [39] K. J. Brown, P. C. Dunne and R. A. Gardner, A semilinear parabolic system arising in the theory of superconductivity, J. Differ. Equations, 40 (1981), 232-252.
    [40] G. M. Lieberman, Bounds for the steady-state Sel'kov model for arbitrary p in any number of dimensions, SIAM J. Math. Anal., 36 (2005), 1400-1406.
    [41] R. Peng, J. Shi and M. Wang, On stationary patterns of a reaction-diffusion model with autocatalysis and saturation law, Nonlinearity, 21 (2008), 1471-1488.
    [42] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equation of Second Order, Springer, (2001).
    [43] Y. Du, R. Peng and M. Wang, Effect of a protection zone in the diffusive Leslie predator-prey model, J. Differ. Equations, 246 (2009), 3932-3956.
    [44] Y. Lou and W.-M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Differ. Equations, 131 (1996), 79-131.
    [45] W.-M. Ni and I. Takagi, On the Neumann problem for some semilinear elliptic equations and eystems of activator-inhibitor type, Trans. Amer. Math. Soc., 297 (1986), 351-368.
    [46] H. Brezis and W. A. Strauss, Semi-linear second-order elliptic equations in L1, J. Math. Soc. Japan, 25 (1973), 565-590.
    [47] R. Peng and F. Yi, Asymptotic profile of the positive steady state for an SIS epidemic reactiondiffusion model: Effects of epidemic risk and population movement, Phys. D, 259 (2013), 8-25.
  • This article has been cited by:

    1. Shuyu Han, Chengxia Lei, Xiaoyan Zhang, Qualitative analysis on a diffusive SIRS epidemic model with standard incidence infection mechanism, 2020, 71, 0044-2275, 10.1007/s00033-020-01418-1
    2. M. A. Aziz-Alaoui, Parimita Roy, Deciphering role of inter and intracity human dispersal on epidemic spread via coupled reaction-diffusion models, 2020, 1598-5865, 10.1007/s12190-020-01450-4
    3. Lingmin Dong, Bo Li, Guanghui Zhang, Analysis on a Diffusive SI Epidemic Model with Logistic Source and Saturation Infection Mechanism, 2022, 45, 0126-6705, 1111, 10.1007/s40840-022-01255-7
    4. Chengxia Lei, Yi Shen, Guanghui Zhang, Yuxiang Zhang, Analysis on a diffusive SEI epidemic model with/without immigration of infected hosts, 2021, 14, 1937-1632, 4259, 10.3934/dcdss.2021131
    5. Lingmin Dong, Shuai Hou, Chengxia Lei, Global attractivity of the equilibria of the diffusive SIR and SEIR epidemic models with multiple parallel infectious stages and nonlinear incidence mechanism, 2022, 134, 08939659, 108352, 10.1016/j.aml.2022.108352
    6. Wenjie Li, Weiran Zhao, Jinde Cao, Lihong Huang, Dynamics of a linear source epidemic system with diffusion and media impact, 2024, 75, 0044-2275, 10.1007/s00033-024-02271-2
    7. Fang Wang, Yuting Liu, Yuxue Chen, Global stabilization and boundary control of coupled Fisher–Stream equation and application to SIS–Stream model, 2024, 1598-5865, 10.1007/s12190-024-02226-w
    8. Rui Peng, Zhi-An Wang, Guanghui Zhang, Maolin Zhou, Novel spatial profiles of some diffusive SIS epidemic models, 2023, 87, 0303-6812, 10.1007/s00285-023-02013-4
    9. Daozhou Gao, Chengxia Lei, Rui Peng, Benben Zhang, A diffusive SIS epidemic model with saturated incidence function in a heterogeneous environment * , 2024, 37, 0951-7715, 025002, 10.1088/1361-6544/ad1495
    10. Yan-Xia Feng, Wan-Tong Li, Fei-Ying Yang, Asymptotic profiles of a nonlocal dispersal SIS epidemic model with saturated incidence, 2024, 0308-2105, 1, 10.1017/prm.2024.62
  • 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(4580) PDF downloads(519) Cited by(10)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog