Processing math: 74%
Research article Special Issues

Asymptotic profile of endemic equilibrium to a diffusive epidemic model with saturated incidence rate

  • Received: 24 December 2018 Accepted: 02 April 2019 Published: 30 April 2019
  • We study the existence and asymptotic profile of endemic equilibrium (EE) of a diffusive SIS epidemic model with saturated incidence rate. By introducing the basic reproduction number R0, the existence of EE is established when R0>1. The effects of diffusion rates and the saturated coefficient on asymptotic profile of EE are investigated. Our results indicate that when the diffusion rate of susceptible individuals is small and the total population N is below a certain level, or the saturated coefficient is large, the infected population dies out, while the two populations persist if at least one of the diffusion rates of the susceptible and infected individuals is large. Finally, we illustrate the influences of the population diffusion and the saturation coefficient on this model numerically.

    Citation: Yan’e Wang , Zhiguo Wang, Chengxia Lei. Asymptotic profile of endemic equilibrium to a diffusive epidemic model with saturated incidence rate[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 3885-3913. doi: 10.3934/mbe.2019192

    Related Papers:

    [1] Yoichi Enatsu, Yukihiko Nakata . Stability and bifurcation analysis of epidemic models with saturated incidence rates: An application to a nonmonotone incidence rate. Mathematical Biosciences and Engineering, 2014, 11(4): 785-805. doi: 10.3934/mbe.2014.11.785
    [2] Jinzhe Suo, Bo Li . Analysis on a diffusive SIS epidemic system with linear source and frequency-dependent incidence function in a heterogeneous environment. Mathematical Biosciences and Engineering, 2020, 17(1): 418-441. doi: 10.3934/mbe.2020023
    [3] 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
    [4] 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
    [5] Yanyan Du, Ting Kang, Qimin Zhang . Asymptotic behavior of a stochastic delayed avian influenza model with saturated incidence rate. Mathematical Biosciences and Engineering, 2020, 17(5): 5341-5368. doi: 10.3934/mbe.2020289
    [6] James M. Hyman, Jia Li . Epidemic models with differential susceptibility and staged progression and their dynamics. Mathematical Biosciences and Engineering, 2009, 6(2): 321-332. doi: 10.3934/mbe.2009.6.321
    [7] Ruixia Zhang, Shuping Li . Analysis of a two-patch SIS model with saturating contact rate and one- directing population dispersal. Mathematical Biosciences and Engineering, 2022, 19(11): 11217-11231. doi: 10.3934/mbe.2022523
    [8] Abdennasser Chekroun, Mohammed Nor Frioui, Toshikazu Kuniya, Tarik Mohammed Touaoula . Global stability of an age-structured epidemic model with general Lyapunov functional. Mathematical Biosciences and Engineering, 2019, 16(3): 1525-1553. doi: 10.3934/mbe.2019073
    [9] Fang Wang, Juping Zhang, Maoxing Liu . Dynamical analysis of a network-based SIR model with saturated incidence rate and nonlinear recovery rate: an edge-compartmental approach. Mathematical Biosciences and Engineering, 2024, 21(4): 5430-5445. doi: 10.3934/mbe.2024239
    [10] Peter Rashkov, Ezio Venturino, Maira Aguiar, Nico Stollenwerk, Bob W. Kooi . On the role of vector modeling in a minimalistic epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 4314-4338. doi: 10.3934/mbe.2019215
  • We study the existence and asymptotic profile of endemic equilibrium (EE) of a diffusive SIS epidemic model with saturated incidence rate. By introducing the basic reproduction number R0, the existence of EE is established when R0>1. The effects of diffusion rates and the saturated coefficient on asymptotic profile of EE are investigated. Our results indicate that when the diffusion rate of susceptible individuals is small and the total population N is below a certain level, or the saturated coefficient is large, the infected population dies out, while the two populations persist if at least one of the diffusion rates of the susceptible and infected individuals is large. Finally, we illustrate the influences of the population diffusion and the saturation coefficient on this model numerically.


    To understand the dynamics of disease transmission in a spatially heterogeneous environment, an SIS epidemic reaction-diffusion model was proposed in [1], satisfying the parabolic system

    {ˉSt=dSΔˉSβ(x)ˉSˉIˉS+ˉI+γ(x)ˉI,xΩ, t>0,ˉIt=dIΔˉI+β(x)ˉSˉIˉS+ˉIγ(x)ˉI,xΩ, t>0,ˉSν=ˉIν=0,xΩ, t>0, (1.1)

    where ˉS(x,t) and ˉI(x,t) denote the densities of susceptible and infected individuals at position x and time t, respectively; the positive constants dS and dI are the diffusion rates of the susceptible and infected individuals; the habitat Ω is assumed to be a bounded domain in Rn(n1) with smooth boundary Ω; the positive Hölder-continuous functions β(x) and γ(x) on ˉΩ represent the rates of disease transmission and disease recovery at x, respectively; the homogeneous Neumann boundary condition means that there is no flux across the boundary Ω, and /ν is the outward normal derivative to Ω.

    In [1], under the assumption that the total population keep constant, the existence and uniqueness of the endemic equilibrium (EE) were achieved in terms of the basic reproduction number R0. Furthermore, the asymptotic profile of EE was obtained for small diffusion rate of susceptible individuals. To further understand the impact of large and small diffusion rates on the persistence and extinction of the disease, the global stability and asymptotic behavior of EE for system (1.1) were investigated in [20,21,22]. In [23], Peng and Zhao considered the diffusive SIS model with spatially heterogeneous and temporally periodic disease transmission and recovery rates. The authors in [4,5,11] studied the effects of diffusion and advection for a spatial SIS model in heterogeneous environments. Their results suggest that advection can help speed up the elimination of disease. Other related works on (1.1) can be found in [8,9,13,14,15].

    The aforementioned studies adopt the standard incidence rate βˉSˉI/(ˉS+ˉI). Another most frequently used incidence rate is the bilinear incidence rate βˉSˉI (see [2,10]), which gives rise to the dependence of the basic reproduction number on the total population. For the diffusive SIS epidemic model with the bilinear incidence rate, Deng and Wu discussed the existence and the global attractivity of the EE in [7]. In the continued work [25], Wu and Zou explored the asymptotic profile of EE for large and small diffusion rates of the susceptible and infected individuals. They observed some new interesting profiles for such model. In contrast, Capasso and Serio in [6] pointed out that the number of effective contacts between infective individuals and susceptible individuals cannot always increase linearly with I; the bilinear incidence rate might be true for a small number of infectives, but unrealistic for large I. They introduced a saturated incidence rate g(ˉI)ˉS into epidemic models based on the study of the cholera epidemic spread in Bari of Italy, where

    g(ˉI)=βˉI1+mˉI.

    Such an incident rate seems more realistic in certain situations because the number of effective contacts between infective individuals and susceptible individuals may saturate at high infective levels due to crowding of the infective individuals or due to the protection measures by the susceptible individuals. Here βˉI measures the infection force of the disease, 1/(1+mˉI) measures the inhibition effect from the behavioral change of the susceptible individuals when their number increases or from the crowding effect of the infective individuals, m>0 is the saturation coefficient. This type of incidence rate has been adopted by many authors [12,19,26,27].

    However, to our best knowledge, little work has been devoted to the study of the diffusive epidemic model with saturated incidence rate. Inspired by the above research, we here consider an SIS epidemic reaction-diffusion model with saturated incidence rate. We are interested in the existence of the EE and particularly the effects of the diffusion rates and the saturated coefficient on asymptotic profile of EE. In contrast to [1] and [25], for some special case, such as the rate of disease transmission β being a constant and Ω being a high-risk domain, our results indicate that it is not enough to just restrict the movement of the susceptible individuals to completely eradicate the disease in the whole habitat; however, if the inhibition effect is large, the infectious disease will extinct eventually (see Theorem 1.3 and Corollary 1.5). In general, we conclude that the infective individuals cannot persist if the saturated coefficient is large with fixed diffusion rates of the susceptible and infected individuals (see Theorem 1.9).

    In this paper, we are concerned with the following SIS epidemic reaction-diffusion model with saturated incidence rate:

    {ˉSt=dSΔˉSβ(x)ˉSˉI1+mˉI+γ(x)ˉI,xΩ, t>0,ˉIt=dIΔˉI+β(x)ˉSˉI1+mˉIγ(x)ˉI,xΩ, t>0,ˉSν=ˉIν=0,xΩ, t>0, (1.2)

    where the parameters are described as before. We assume that the initial data satisfies the following hypothesis.

    (H) ˉS(x,0), ˉI(x,0)0 are nonnegative continuous functions in ˉΩ, and the number of initial infectious individuals in the region is positive, i.e, ΩˉI(x,0)dx>0.

    By a similar argument as in [1], it is easy to show that system (1.2) admits a unique global classical solution (ˉS(x,t),ˉI(x,t)). Let

    N:=Ω(ˉS(x,0)+ˉI(x,0))dx>0

    be the total number of individuals in Ω at t=0. Adding the two equations in (1.2) and integrating over the domain Ω, we get

    tΩ(ˉS+ˉI)dx=0, t>0.

    Hence, the total population size is a constant, i.e.,

    Ω(ˉS(x,t)+ˉI(x,t))dx=N, t0. (1.3)

    In the current paper, we mainly focus on the nonnegative equilibrium of problem (1.2), which is the nonnegative solution of the following semilinear elliptic system:

    {dSΔSβ(x)SI1+mI+γ(x)I=0,xΩ,dIΔI+β(x)SI1+mIγ(x)I=0,xΩ,Sν=Iν=0,xΩ. (1.4)

    Here S(x) and I(x) denote the densities of susceptible and infected individuals at xˉΩ, respectively. In view of (1.3), we have to impose the additional hypothesis:

    Ω(S(x)+I(x))dx=N. (1.5)

    Obviously, system (1.4)–(1.5) always has a solution E0=(N/|Ω|,0), which is the uniquedisease-free equilibrium (DFE). On the other hand, a nonnegative solution E1=(S,I) of (1.4)–(1.5) with I(x)0,0 is called an endemic equilibrium (EE) of (1.4)–(1.5).

    Similar to [1,7], let us define the basic reproduction number R0 for (1.2) as follows:

    R0:=sup{NΩβφ2dx|Ω|Ω(dI|φ|2+γφ2)dx:φH1(Ω) and φ0},

    where H1(Ω)={u:uL2(Ω),DuL2(Ω)}. Denote the high-risk set and low-risk set respectively by

    Ω+:={xΩ:N|Ω|β(x)>γ(x)}

    and

    Ω:={xΩ:N|Ω|β(x)<γ(x)}.

    We say the domain Ω is a high-risk domain if N|Ω|Ωβ(x)dxΩγ(x)dx and it is a low-risk domain if N|Ω|Ωβ(x)dx<Ωγ(x)dx.

    We begin with some properties of R0 which is similar to Lemmas 2.2 and 2.3 in [1].

    Proposition 1.1. The basic reproduction number R0 has the following properties.

    (ⅰ) R0 is positive decreasing function of dI>0;

    (ⅱ) R0N|Ω|maxxˉΩβ(x)γ(x) as dI0+, and R0N|Ω|Ωβ(x)dxΩγ(x)dx as dI;

    (ⅲ) if Ω is a high-risk domain, then R0>1 for dI>0;

    (ⅳ) if Ω is a low-risk domain with nonempty Ω+, then there exists dI>0 such that R0=1 when dI=dI, R0>1 when dI<dI, and R0<1 when dI>dI;

    (ⅴ) R0>1 implies N|Ω|>minxˉΩγ(x)β(x).

    The first goal of this paper is to establish the existence of EE.

    Theorem 1.2. The following statements hold.

    (ⅰ) If dSdI, there exists a unique EE when R0>1 and EE does not exist when R01;

    (ⅱ) if dS<dI, there exists an EE when R0>1 and EE does not exist when R0dS/dI.

    Theorem 1.2(ⅰ) indicates that R0=1 is the critical value for the existence of EE when dSdI. However, if dS<dI, we do not know whether an EE exists or not in the case of R0(dS/dI,1).

    A combination of Proposition 1.1 and Theorem 1.2 implies that the EE always exists when Ω is a high-risk domain (see Figure 1(a)) or Ω is a low-risk domain with nonempty Ω+ and 0<dI<dI (see Figure 1(b)), where dI>0 is uniquely determined in Proposition 1.1(ⅳ).

    Figure 1.  The existence of the EE in dIdS plane. (a) High-risk domain; (b) low-risk domain with nonempty Ω+.

    The second goal of this paper is to investigate the effects of diffusion rates and saturation coefficient on asymptotic profiles of the EE when it exists. Here we consider the following three cases: (ⅰ) small diffusion, (ⅱ) large diffusion, (ⅲ) large saturation.

    The following theorem presents the asymptotic profile of EE when dS is sufficiently small or large.

    Theorem 1.3. Let dI and m be fixed. Assume R0>1 and (S(x),I(x)) is an EE of (1.2). Then the following statements hold.

    (ⅰ) (S,I)(S,I) in C2(ˉΩ) when dS, where I is the unique positive solution of the following problem

    {dIΔI=I[β1+mI(N|Ω|1|Ω|ΩIdx)γ],xΩ,Iν=0,xΩ (1.6)

    and

    S=N|Ω|1|Ω|ΩIdx.

    (ⅱ) (S,I)(S,I) in C(ˉΩ) when dS0+, where S and I satisfy

    (S,I)=(γβ|Ω|+mN|Ω|+mΩγ/βdx,NΩγ/βdx|Ω|+mΩγ/βdx), (1.7)

    or I=0 and

    S=N|Ω|+dI(1|Ω|ΩˇIdxˇI), (1.8)

    where ˇI>0 satisfies the following problem

    {dIΔI+I[β(N|Ω|+dI|Ω|ΩIdxdII)γ]=0,xΩ,Iν=0,xΩ. (1.9)

    Corollary 1.4. Suppose that R0>1 and N<Ωγ/βdx. Then, for any fixed dI>0 and m>0, the EE (S,I)(S,0) in C(ˉΩ) as dS0+, where S satisfies (1.8).

    Corollary 1.5. Fixed dI>0 and m>0. Suppose R0>1 and N>Ωγ/βdx. Then the EE (S,I)(S,I) in C(ˉΩ) as dS0+, where (S,I) satisfies (1.7) if one of the following conditions holds:

    (ⅰ) β is a positive constant;

    (ⅱ) N|Ω|>maxxˉΩγ(x)β(x);

    (ⅲ) γ(x)β(x)=r for any xˉΩ, where r is some positive constant.

    Corollary 1.5 implies that the infectious disease may persist even if the movement of the susceptible population is controlled to be very small, which is in sharp contrast to the epidemic model in [1].

    Next, we are going to explore the asymptotic profile of EE as dI0+ and dI/dSd>0.

    Theorem 1.6. Let m be fixed. Assume that Ω+ is nonempty. Then the following statements hold:

    (ⅰ) If dI0+ and dI/dSd(0,), then (S,I)(S,I) in C(ˉΩ), where I is the unique positive solution of

    (dβ(x)+mγ(x))I=[β(x)(N|Ω|1d|Ω|ΩIdx)γ(x)]+ (1.10)

    and S is given by

    S=N|Ω|1d|Ω|ΩIdxdI.

    (ⅱ) If d(0,1), then {xΩ:I>0}Ω+; if d(1,), then {xΩ:I>0}Ω+ and this inclusion is strict if Ω is nonempty.

    Remark 1.7. If d=1 in Theorem 1.6, then

    S=N|Ω|I, I=1β+mγ(βN|Ω|γ)+,

    which implies that {xΩ:I>0}=Ω+. It follows from Theorem 1.6(ii) that, in this situation, the ratio dI/dS plays a critical role in determining the existing region of the infected population. If d=1, the infected individuals survive exactly in the high-risk set; if d(0,1), the habitat of infected individual is confined within some subset of the high-risk set; if d>1, the infected individuals only die out at part of the low-risk sites.

    We now establish the asymptotic profile when the diffusion rate dI is large.

    Theorem 1.8. Let m be fixed. Suppose that Ω is a high-risk domain. Then the following statements hold.

    (ⅰ) If dI and dS, then (S,I)(S,I) in C2(ˉΩ), where S and I are positive constants satisfying

    S=ΩγdxΩβdx(1+mNΩβdx|Ω|Ωγdx|Ω|(Ωβdx+mΩγdx)), I=NΩβdx|Ω|Ωγdx|Ω|(Ωβdx+mΩγdx).

    (ⅱ) If dS is fixed, then there exists a sequence {dIn} with dIn as n such that the corresponding EE (Sn,In)(S,I) in C2(ˉΩ), where I is a positive constant and S is the positive solution of the following problem

    {dSΔS=βI1+mIS+γI,xΩ,Sν=0,xΩ,ΩSdx=N|Ω|I. (1.11)

    Furthermore, if dS0+ in (1.11), then (S,I)(ˆS,ˆI) in C1(ˉΩ), where ˆS and ˆI satisfy (1.7) or ˆI=0 and ˆS satisfies

    {ΔS=1K1(βS+γ),xΩ,Sν=0,xΩ,ΩSdx=N (1.12)

    with K1 being a positive constant.

    Finally, we describe the asymptotic profile when the saturated coefficient m is large.

    Theorem 1.9. Suppose that R0>1. Then for any fixed dI>0 and dS>0, the corresponding EE (S,I) of (1.2) satisfies (S,I)(N/|Ω|,0) in C(ˉΩ) as m. Furthermore, either mI or mIw as m, where w is the unique positive solution of the following problem

    {dIΔw=w(βN|Ω|(1+w)γ),xΩ,wν=0,xΩ. (1.13)

    Theorem 1.9 implies that large saturated coefficient can help to eliminate the disease. That is, if the susceptible individuals change the behavior when their number increases or the infective individuals produce crowding effect, the infectious disease may extinct eventually.

    The rest of this paper is arranged as follows. In Section 2, we focus on the existence, uniqueness and nonexistence of the EE, and give the proof of Theorem 1.2. In Section 3, the impacts of diffusion rates and saturated coefficient on the persistence and extinction of the infectious disease are studied. Then, we illustrate the influences of the population diffusion and the saturation coefficient on system (1.2) numerically in Section 4. In Section 5, we conclude the paper with some discussion of the epidemiological implication of our theoretical results. Finally, some well-known facts, which are frequently used in the proofs of our main results, are collected in the appendix.

    Since the existence of the EE is related to the stability of the DFE, we first investigate the stability of the DFE. To this end, we linearize (1.2) around the DFE to obtain

    {ηt=dSΔη(N|Ω|βγ)ξ,xΩ,t>0,ξt=dIΔξ+(N|Ω|βγ)ξ,xΩ,t>0.

    Here η(x,t)=¯S(x,t)N/|Ω| and ξ(x,t)=¯I(x,t). Let (η(x,t),ξ(x,t))=(eλtϕ(x),eλtψ(x)) be the solution of the linear system. Then, we derive an eigenvalue problem

    {dSΔϕ(N|Ω|βγ)ψ+λϕ=0,xΩ,dIΔψ+(N|Ω|βγ)ψ+λψ=0,xΩ (2.1)

    with boundary conditions

    ϕν=ψν=0, xΩ. (2.2)

    In view of (1.3), we impose an additional condition

    Ω(ϕ+ψ)dx=0. (2.3)

    Indeed, it suffices to consider the eigenvalue problem

    {dIΔψ+(N|Ω|βγ)ψ+λψ=0,xΩ,ψν=0,xΩ. (2.4)

    It is well known that all eigenvalues of (2.4) are real, and the principal eigenvalue, denoted by λ, is simple, and its corresponding eigenfunction ψ can be chosen positive on Ω. Furthermore, the eigenvalue λ is given by the variational characterization

    λ=inf{Ω[dI|φ|2+(γN|Ω|β)φ2]dx:φH1(Ω) and Ωφ2dx=1}.

    It has been shown in [7] that the basic reproduction number R0 and the principal eigenvalue λ has the following relationship.

    Lemma 2.1. 1R0 and λ have the same sign.

    As discussed in Lemma 2.4 of [1], the stability of the DFE depends on the value of R0.

    Lemma 2.2. The DFE is linearly stable if R0<1 and unstable if R0>1.

    To study the existence of the EE, we first convert problem (1.4)-(1.5) to an equivalent but more accessible problem.

    Lemma 2.3. The pair (S,I) is a nonnegative solution of problem (1.4)-(1.5) if and only if I is a nonnegative solution of the following problem

    {dIΔI+I[β1+mI(N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI)γ]=0,xΩ,Iν=0,xΩ (2.5)

    and

    S=N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI, xΩ. (2.6)

    Proof. By standard calculations, one can easily check that (S,I) is a nonnegative solution of problem (1.4)-(1.5) if and only if it solves the following problem:

    dSS+dII=κ,  xΩ, (2.7)
    dIΔI+β(x)SI1+mIγ(x)I=0,  xΩ, (2.8)
    Sν=Iν=0,  xΩ, (2.9)
    Ω(S+I)dx=N, (2.10)

    where κ is some positive constant independent of xΩ.

    Now we show the equivalence between problems (2.7)-(2.10) and (2.5)-(2.6).

    Suppose that (S,I) is a nonnegative solution of (2.7)-(2.10). By (2.7), we have dS(S+I)=κ+(dSdI)I. Integrating it over Ω and using (2.10), we get dSN=κ|Ω|+(dSdI)ΩIdx. Substituting (2.7) into the equation gives

    S=N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI, xΩ.

    That is, (2.6) holds. Substituting such S into (2.8), we get (2.5).

    Suppose that (S,I) is a nonnegative solution of problem (2.5)-(2.6). Substituting (2.6) into (2.5) yields (2.8). Clearly, S/ν=0, i.e., (2.9) holds. Integrating both sides of (2.6) over Ω, we get (2.10). Applying the Laplace operator to both sides of (2.6), we find that dSΔS=dIΔI which means Δ(dSS+dII)=0. Since ν(dSS+dII)=0, the maximum principle implies that dSS+dII is a constant. In view of (2.10), this constant must be positive, which yields (2.7).

    The nonlocal elliptic problem (2.5) has the following estimate.

    Lemma 2.4. If IC2(Ω)C1(ˉΩ) is a nonnegative solution of the nonlocal elliptic problem (2.5), then we have

    (1dIdS)1|Ω|ΩIdx+dIdSIN|Ω| for all xˉΩ. (2.11)

    Proof. It is easy to see that (2.11) holds if I0 on ˉΩ. Suppose I0,0. Then there exists some x0ˉΩ such that I(x0)=maxxˉΩI(x)>0. By Lemma A.3, we have

    β(x0)1+mI(x0)[N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI(x0)]γ(x0)0,

    which implies that

    N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI(x0)+γ(x0)β(x0)dIdSI. (2.12)

    The conclusion holds.

    Set

    S:=N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI, xˉΩ.

    If follows from Lemma 2.4 that S is nonnegative. Hence the pair (S,I) solves problem (2.5)–(2.6) as well as (1.4)–(1.5). Next, we focus on the existence of positive solution to the nonlocal elliptic problem (2.5) that only involves I.

    Let Γ={τ[0,):N(1dI/dS)τ0} and Y={zC2+α(ˉΩ):z/ν=0  on Ω}. Define a mapping F:Γ×YCα(ˉΩ) by

    F(τ,I)=dIΔI+If(τ,I)

    with

    f(τ,I)=β|Ω|(1+mI)[N(1dIdS)τdI|Ω|dSI]γ.

    Then I is a nonnegative solution of (2.5) if and only if F(τ,I)=0 and τ=ΩIdx.

    Now we consider the following problem:

    {dIΔI+If(τ,I)=0,xΩ,Iν=0,xΩ. (2.13)

    It is easy to check that (2.13) meets all the requirements of Lemma A.2. Thus the existence of positive solution of (2.13) is tightly related to an eigenvalue λ1(dI,f(τ,0)), which is defined by (8.1). For simplicity, we denote λτ=λ1(dI,f(τ,0)), and hence λ0=λ, where λ is the principal eigenvalue of (2.4).

    Lemma 2.5. Suppose that τ0.

    (ⅰ) If λτ0, the only nonnegative solution of (2.13) is I=0;

    (ⅱ) if λτ<0, there is a unique positive solution IY of (2.13).

    We are now ready to prove Theorem 1.2. To this end, we need to prove several lemmas as follows.

    Lemma 2.6. Suppose λ<0.

    (ⅰ) If dS>dI, then there exists a smooth curve (τ,Iτ(x)) in Γ×Y such that F(τ,Iτ)=0. Moreover, there is a Λ>0 such that IΛ=0 and Iτ(x)>0 for all xˉΩ, τ[0,Λ). Furthermore, Iτ is decreasing and continuously differentiable with respect to τ in (0,Λ);

    (ⅱ) if dS<dI, then there exists a smooth curve (τ,Iτ(x)) in [0,)×Y such that F(τ,Iτ)=0 with Iτ>0 for xˉΩ and τ[0,). Moreover, Iτ(x) is increasing and continuously differentiable in τ on (0,), and it satisfies the following estimate:

    ΩIτ(x)dxdSdIN+(1dSdI)τ. (2.14)

    Proof. (ⅰ) Suppose that (τ0,Iτ0(x))Γ×Y satisfies F(τ0,Iτ0)=0 and Iτ0(x)>0 on ˉΩ. The Frˊechet derivative of F is given by

    FI(τ0,Iτ0)w=dIΔw+[f(τ0,Iτ0)+fI(τ0,Iτ0)Iτ0]w

    for all wY, where

    fI(τ0,Iτ0)=β|Ω|(1+mIτ0)2{dI|Ω|dS+m[N(1dIdS)τ0]}<0.

    We claim that FI(τ0,Iτ0) is invertible. To this end, we need to show the unique solvability of the following problem for any hCα(ˉΩ),

    {dIΔw+[f(τ0,Iτ0)+fI(τ0,Iτ0)Iτ0]w=h,xΩ,wν=0,xΩ. (2.15)

    Since F(τ0,Iτ0)=0, i.e. dIΔIτ0+Iτ0f(τ0,Iτ0)=0, we see that λ1(dI,f(τ0,Iτ0))=0 and Iτ0 is a corresponding eigenvector by (8.1). It follows from Lemma A.1 and fI(τ0,Iτ0)<0 that λ1(dI,f(τ0,Iτ0)+fI(τ0,Iτ0)Iτ0)>λ1(dI,f(τ0,Iτ0))=0. So all eigenvalues of the problem

    {dIΔφ+[f(τ0,Iτ0)+fI(τ0,Iτ0)Iτ0]φ+λφ=0,xΩ,φν=0,xΩ (2.16)

    are positive. By the Fredholm alternative, (2.15) has a unique solution for every hCα(ˉΩ). The continuity of the unique solution follows from the classical Schauder estimate. Since λ0=λ<0, there exists a unique positive I0Y such that F(0,I0)=0 by Lemma 2.5. Hence, by the implicit function theorem, there is a unique IτY such that F(τ,Iτ)=0 for τ[0,ˆτ) with ˆτ>0, and Iτ is continuously differentiable with respect to τ.

    Now we show that Iτ is decreasing with respect to τ. Suppose that 0<τ1<τ2<ˆτ. Since dS>dI, we have that F(τ1,Iτ2)>F(τ2,Iτ2)=0. Hence, Iτ2 is a lower solution of the equation F(τ1,I)=0. On the other hand, we choose a sufficiently large number as an upper solution of the equation F(τ1,I)=0. Then, by the method of upper/lower solutions and the uniqueness of the positive solution of F(τ1,I)=0, we obtain that Iτ1>Iτ2.

    The curve (τ,Iτ) satisfying F(τ,Iτ)=0 will continue as long as Iτ>0, i.e., λτ<0, due to Lemma 2.5. By the variational formula, λτ is increasing with respect to τ and λτ>0 for large τ. It follows from Lemma 2.5 again that, there is no positive solution of F(τ,I)=0 if τ is large, i.e., Iτ=0 for large τ. Let [0,Λ) be the maximal interval of existence of τ such that Iτ>0. Then IΛ=0.

    (ⅱ) The existence and continuous differentiability of the curve (τ,Iτ) can be obtained by a similar argument as in the proof of (ⅰ). And one can check that Iτ is increasing with respect to τ since dS<dI. Thus the curve is continuous with respect to τ on [0,).

    To show (2.14), let Iτ(y0)=maxˉΩIτ(x). Applying Lemma A.3 to the first equation of (2.13), we obtain that

    f(τ,Iτ(y0))=β(y0)(1+mIτ(y0))[N|Ω|(1dIdS)τ|Ω|dIdSIτ(y0)]γ(y0)0,

    which implies

    β(y0)(1+mIτ(y0))[N|Ω|(1dIdS)τ|Ω|dIdSIτ(y0)]0,

    and hence

    Iτ(x)Iτ(y0)dSNdI|Ω|+(1dSdI)τ|Ω|

    for any xˉΩ. It follows that (2.14) holds by integrating the above inequality over Ω.

    Lemma 2.7. Suppose R0>1. Then there exists a unique EE if dSdI, and there exists at least one EE if dS<dI.

    Proof. If dS=dI, then λτ=λ<0 based on R0>1. The result follows directly from Lemma 2.5.

    For dS>dI, by Lemma 2.6 (ⅰ), there is a smooth curve (τ,Iτ) satisfying F(τ,Iτ)=0. By the definition of F, Iτ is a solution of problem (2.5) if τ=ΩIτdx. Let H(τ)=ΩIτdxτ. Then H(τ) is continuous and strictly decreasing with respect to τ in [0,Λ) because of the continuity and monotonicity of Iτ. Since ΩI0dx>0 and 0=ΩIΛdx<Λ, we have H(0)>0, H(Λ)<0. Then, there exists a unique τ0(0,Λ) such that H(τ0)=0, i.e. τ0=ΩIτ0dx. Hence problem (2.5) has a unique positive solution.

    For dS<dI, by Lemma 2.6 (ⅱ), there exists a smooth curve (τ,Iτ) satisfying F(τ,Iτ)=0. We also take H(τ)=ΩIτdxτ. Then it is continuous with respect to τ. The estimate (2.14) implies that H(τ)dSdI(Nτ). Since H(0)>0 and H(τ)<0 with τ>N, there exists a τ0>0 such that H(τ0)=0, i.e. τ0=ΩIτ0dx. Hence, problem (2.5) has at least one positive solution.

    Lemma 2.8. The EE does not exist if one of the following conditions holds:

    (i) dSdI and R01;

    (ii) dS<dI and R0dS/dI.

    Proof. (ⅰ) The case dS=dI follows directly from Lemma 2.5. We analyze the case dS>dI indirectly. Assume that an EE (S,I) exists if R01. Then there is a τ>0 such that τ=ΩIdx and F(τ,I)=0. By Lemma 2.5, we know λτ<0, which leads to λ=λ0λτ<0 since f(τ,0) is decreasing in τ when dS>dI. Then R0>1 by Lemma 2.1, which is a contradiction.

    (ⅱ) The case dS<dI. Assume to the contrary that an EE (S,I) exists when R0ds/dI. Let τ=ΩIdx. Then I is also the positive solution of F(τ,I)=0, and it satisfies (2.11) for all xˉΩ, i.e.

    (1dIdS)1|Ω|ΩIdx+dIdSIN|Ω|.

    Integrating this inequality over Ω, we get τ=ΩIdxN. Noting that λτ<0 by Lemma 2.5 and f(τ,0) is increasing in τ provided dS<dI, Lemma A.1 implies that λNλτ<0, where λN is the principal eigenvalue of the following problem

    {dIΔφ+(dINdS|Ω|βγ)φ=λφ,xΩ,φν=0,xΩ.

    Define

    R0:=sup{dINΩβφ2dxdS|Ω|Ω(dI|φ|2+γφ2)dx:φH1(Ω) and φ0}.

    Then R0>1 if and only if λN<0, which can be obtained as the properties of R0. Since R0=dIR0/dS and λN<0, we have R0>dS/dI, which is a contradiction.

    Proof of Theorem 1.2. Theorem 1.2 follows from Lemmas 2.6, 2.7 and 2.8.

    The goal of this section is to investigate the asymptotic profile of EE. To this end, we always assume R0>1, so that (1.2) has an EE. As a first step, we establish the priori estimates of any EE.

    Lemma 3.1. Assume that (S,I) is a nonnegative solution of (1.4)-(1.5). Then

    I(1+dSdI)N|Ω|, (3.1)
    minxˉΩ{γβ(1+mI)}SmaxxˉΩ{γβ(1+mI)}. (3.2)

    Proof. By (1.5), we have ΩIdxN. Applying the inequality (2.11), we get

    IdSdI(N|Ω|+dIdS1|Ω|ΩIdx)(1+dSdI)N|Ω|.

    Let S(x0)=maxxˉΩS(x), S(y0)=minxˉΩS(x). We apply Lemma A.3 to the first equation of (1.4) to obtain that

    β(x0)S(x0)1+mI(x0)+γ(x0)0,  β(y0)S(y0)1+mI(y0)+γ(y0)0,

    which imply S(x0)γ(x0)β(x0)(1+mI(x0)) and S(y0)γ(y0)β(y0)(1+mI(y0)). Hence (3.2) holds.

    Lemma 3.2. Assume that (S,I) is an EE of (1.4)-(1.5). Then I and S are uniformly bounded in L(ˉΩ) if dS/dI.

    Proof. Note that (S,I) satisfies (1.4)-(1.5) (or (2.5)-(2.6)). By (2.6), we have SN/|Ω| provided dS/dI. Then, we are going to derive a priori estimate of I when dS/dI by the Harnack inequality. Applying Lemma A.4 to the second equation of (1.4), we obtain that there is a positive constant C0 such that max¯ΩIC0min¯ΩI. In view of NΩIdx|Ω|min¯ΩI|Ω|max¯ΩI/C0, we conclude that IC0N/|Ω|.

    Now, we are ready to investigate the asymptotic profiles of the EE when dS is sufficiently small or large. To this end, we show the existence and uniqueness of the solution (1.6).

    Lemma 3.3. Suppose R0>1. Then (1.6) has a unique positive solution.

    Proof. Since R0>1 is equivalent to λ1(dI,Nβ/|Ω|γ)=λ<0 by Lemma 2.1. Taking

    f(τ,I)=β|Ω|(1+mI)(Nτ)γ

    and by similar arguments as in Lemma 2.7, we get that (1.6) has a unique positive solution.

    Proof of Theorem 1.3.. It follows from Theorem 1.2 that an EE (S,I) exists provided R0>1 for any dS>0.

    (ⅰ) We consider the asymptotic profile when dS. By Lemma 3.2, I and S are uniformly bounded in C(ˉΩ) for fixed dI>0 and dS. Then using the elliptic estimate and the Sobolev embedding theorem for (1.4), there exists a sequence {dSn} with dSn as n such that the corresponding EE (Sn,In)(S,I) in C2(ˉΩ). Letting n in (2.5), we get that I satisfies (1.6) which has a unique positive solution by Lemma 3.3. Thus, the strong maximum principle implies that there are two possibilities: I>0 or I0. By (2.5) and the positivity of In, we have

    λ1(dI,β1+mIn(N|Ω|(1dIdSn)1|Ω|ΩIndxdIdSnIn)γ)=0. (3.3)

    If I0, letting n in (3.3), we have λ1(dI,Nβ/|Ω|γ)=0, i.e. λ=0, which contradicts R0>1 by Lemma 1. Hence I is the positive solution of (1.6). By (1.4), S satisfies

    {ΔS=0,xΩ,Sν=0,xΩ. (3.4)

    The strong maximum principle implies that S is a constant. Furthermore, it follows from (1.5) that

    S=N|Ω|1|Ω|ΩIdx.

    (ⅱ) We analyze the asymptotic profile when dS0+. It follows from (3.1) that I is uniformly bounded for fixed dI and small dS. Hence there exists a sequence {dSn} with dSn0+ as n such that the corresponding EE (Sn,In) satisfies

    ΩIndxK for some K0.

    It then follows that

    Fn:=β(N|Ω|dSn(dSndI)1|Ω|ΩIndx)dIβK|Ω| as n.

    Hence, for any ϵ>0, there exists n1>0 such that for nn1,

    dIβ|Ω|(Kϵ)FndIβ|Ω|(K+ϵ) and 0<dSnmin{minx¯Ωβ(x)maxx¯Ωγ(x)ϵ,dI}. (3.5)

    We claim that

    InK|Ω| uniformly on ˉΩ as n. (3.6)

    Noting that In satisfies (2.5), we rewrite it as

    {dSndIΔIn+In(FndIβIn1+mIndSnγ(x))=0,xΩ,Inν=0,xΩ. (3.7)

    It follows from (3.5) that In is a lower solution of the problem

    {dSndIΔI+I[dIβ|Ω|(K+ϵ)dIβI]=0,xΩ,Iν=0,xΩ. (3.8)

    On the other hand, (3.1) and (3.5) imply that

    In(1+dSndI)N|Ω|2N|Ω| (3.9)

    for nn1. Meanwhile, In is an upper solution of the problem

    {dSndIΔI+I(dIβ(Kϵ)/|Ω|dIβI1+2mN/|Ω|βϵ)=0,xΩ,Iν=0,xΩ. (3.10)

    Observing that I=K+ϵ|Ω| is the unique positive solution of (3.8) and I=Kϵ|Ω|ϵdI(1+2mN|Ω|) is the unique positive solution of (3.10), we have

    Kϵ|Ω|ϵdI(1+2mN/|Ω|)InK+ϵ|Ω| for all nn1. (3.11)

    Since ϵ>0 is arbitrary, (3.11) indeed implies that InK|Ω| uniformly on ˉΩ as n.

    Now, we have two possibilities K>0 or K=0. First, we consider the case K>0. Obviously, Sn satisfies

    {dSnΔSn+(βSn1+mIn+γ)In=0,xΩ,Snν=0,xΩ. (3.12)

    By the fact that InK|Ω| and Lemma A.2, we can prove that

    Snγβ(1+mK/|Ω|) (3.13)

    uniformly on ˉΩ as n. Since (Sn,In) satisfies (1.5), letting n, we have

    (1+mK|Ω|)Ωγβdx+K=N,

    which implies that K=|Ω|(NΩγ/βdx)|Ω|+mΩγ/βdx when N>Ωγ/βdx. By (3.6) and (3.13), we know that (1.7) holds provided N>Ωγ/βdx.

    For the case K=0, we have In0 uniformly on ˉΩ as n. Passing to a subsequence if necessary, we then have either case (1) In/dSnC with C0, or case (2) In/dSn as n.

    If the case (1) occurs, then ΩIndx/dSnIn/dSnC. Let ˇIn:=In/dSn. Then ˇInC and ˇIn satisfies

    {dIΔˇIn=ˇIn[β1+mIn(N|Ω|(dSndI)1|Ω|ΩˇIndxdIˇIn)γ],xΩ,ˇInν=0,xΩ.

    Since the right hand of this equation is uniformly bounded in L(Ω), by standard elliptic regularity we know that {ˇIn} is precompact in C1(ˉΩ). Hence by passing to a subsequence we may assume that ˇInˇI0 in C(ˉΩ). Moreover, ˇI satisfies (1.9), i.e.

    {dIΔˇI=ˇI[β(N|Ω|+dI|Ω|ΩˇIdxdIˇI)γ],xΩ,ˇIν=0,xΩ. (3.14)

    We claim that ˇI>0. To this end, we assume ˇI0 on ˉΩ, i.e, ˇIn0 in C(ˉΩ). Let ˜In=ˇIn/ˇIn. Then ˜In=1 and ˜In satisfies

    {dIΔ˜In=˜In[β1+mIn(N|Ω|(dSndI)1|Ω|ΩˇIndxdIˇIn)γ],xΩ,˜Inν=0,xΩ.

    Similarly, by passing to a subsequence we may assume that ˜In˜I in C(ˉΩ). Furthermore, ˜In=1 and ˜I satisfies

    {dIΔ˜I=˜I(βN|Ω|γ),xΩ,˜Iν=0,xΩ. (3.15)

    It follows from the strong maximum principle that ˜I>0 on ˉΩ. Hence, the definition of R0 and (3.15) implies that R0=1, which is a contradiction. Therefore, ˇI0,0. By the strong maximum principle again, we get ˇI>0 on ˉΩ. Hence, ΩIn/dSndxΩˇIdx>0 in C(ˉΩ) as n. However, if In/dSn0, then ΩIn/dSndx0. It's a contradiction. Therefore, it remains In/dSnC1 with some C1>0. In this case, In0 in C(ˉΩ) and by passing to a subsequence

    Sn=N|Ω|(dSdI)1|Ω|ΩˇIndxdIˇInN|Ω|+dI|Ω|ΩˇIdxdIˇIinC1

    as n, i.e. (1.8) holds.

    If case (2) occurs, i.e., In/dSn. Recalling In0 uniformly on ˉΩ as n, we have In0. By Lemma 3.1, we know that Sn is uniformly bounded. Let ˜In=In/In. Then ˜In=1 and ˜In satisfies

    {dIΔ˜In=˜In(βSn1+mInγ),xΩ,˜Inν=0,xΩ. (3.16)

    By the standard elliptic estimates, ˜In is uniformly bounded in C1(ˉΩ) for fixed dI>0. So passing to a subsequence if necessary, we have ˜In˜I in C(ˉΩ) with ˜I=1. By the Harnack inequality, there is a positive constant K independent of n such that

    1=maxxˉΩ˜In(x)KminxˉΩ˜In(x).

    Hence minxˉΩ˜I1/K>0, i.e. ˜I is strictly positive. We now turn to consider the equation (3.12) for Sn. Dividing (3.12) by In, we have

    {dSnInΔSn=(βSn1+mIn+γ)InIn,xΩ,Snν=0,xΩ.

    Since dSn/In0+, In/In˜I and In0, it follows from Lemma A.2 that

    Snγβ uniformly on ˉΩ as n.

    Moreover by (1.5) and In0, we get

    Ωγβdx=N,

    which is a contradiction if Ωγ/βdxN. But if Ωγ/βdx=N, (1.7) holds.

    Proof of Corollary 1.4. From the proof of Theorem 1.3, we know that if R0>1 and N<Ωγ/βdx, then In0 uniformly on ˉΩ as n and In/dSnC with C0. In this case, S satisfies (1.8).

    Proof of Corollary 1.5. From the proof of Theorem 1.3, we only need to rule out I=0.

    (ⅰ) Suppose that β is a positive constant. Then N>Ωγ/βdx implies that N|Ω|Ωβdx>Ωγdx, i.e., Ω is a high-risk domain. By Proposition 1.1(ⅲ) and Theorem 1.2, we know that the EE (S,I) exists for all dI>0 and m>0. Multiplying both sides of the first equation of (2.5) by (1+mI)/I and integrating it over Ω, we get

    dIΩ|I|2I2dx+Ω{β[N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI]γ(1+mI)}dx=0,

    which implies that

    Ω{β[N|Ω|(1dIdS)1|Ω|ΩIdxdIdSI]γ(1+mI)}dx0.

    Since β is a constant, it follows from the above inequality that

    NΩIdxΩγβ(1+mI)dx0. (3.17)

    On the other hand, by Theorem 1.3, we know that (S,I)(S,I) in C(ˉΩ) as dS0, where I is a nonnegative constant. Letting dS0+, the inequality (3.17) implies that

    (|Ω|+mΩγβdx)INΩγβdx>0.

    Hence I>0, which indicates that (S,I) satisfies (1.7).

    (ⅱ)-(ⅲ). From (3.6), we know I=K/|Ω|. We need to prove K>0.

    If K=0, there are three cases to consider: (1) In/dSnC1>0, (2) In/dSn0, and (3) In/dSn. If N>Ωγ/βdx, from the proof of Theorem 1.3, the case (2) and (3) are excluded directly. It remains to consider case (1). In this case, In/dSnˇI>0, where ˇI satisfies (1.9). Next, we show that (1.9) has no positive solution under the conditions (ⅱ) or (ⅲ), which will deduce a contradiction. Take

    f1(τ,I)=β|Ω|(N+dIτdI|Ω|I)γ,  F1(τ,I)=dIΔI+If1(τ,I),

    and consider the problem

    {dIΔI+If1(τ,I)=0,xΩ,Iν=0,xΩ. (3.18)

    Repeating the arguments as in the proof of Theorem 1.2, we can prove that there exists a smooth curve (τ,˜Iτ(x)) in [0,)×Y such that F1(τ,˜Iτ)=0 with ˜Iτ>0 for all xˉΩ and τ[0,) if R0>1. Moreover, ˜Iτ(x) is strictly increasing and continuously differentiable with respect to τ in (0,). It is easy to see that ˜Iτ is a positive solution of (1.9) if and only if τ=Ω˜Iτdx.

    Let xτ,yτˉΩ satisfy ˜Iτ(xτ)=minxˉΩ˜Iτ(x) and ˜Iτ(yτ)=maxxˉΩ˜Iτ(x). Using Lemma A.3 to (3.18), it is easy to check that, for every τ[0,)

    1dI[N|Ω|γ(xτ)β(xτ)]+τ|Ω|˜Iτ1dI[N|Ω|γ(yτ)β(yτ)]+τ|Ω|.

    Then for every τ[0,), we have

    N|Ω|γ(xτ)β(xτ)dI|Ω|(Ω˜Iτdxτ)N|Ω|γ(yτ)β(yτ), (3.19)

    which implies that

    N|Ω|maxxˉΩγ(x)β(x)dI|Ω|(Ω˜Iτdxτ)N|Ω|minxˉΩγ(x)β(x). (3.20)

    Note that R0>1 implies N/|Ω|>minˉΩγ(x)/β(x) by Proposition 1.1(v). It follows from (3.20) that τΩIτdx for every τ[0,) provided N/|Ω|>maxˉΩγ(x)/β(x) or γ(x)/β(x)=r. Hence (1.9) has no positive solution under the conditions (ⅱ) or (ⅲ). K=0 is impossible and we complete the proof of Corollary 1.5.

    Next, we will prove Theorem 1.6. To this end, we first give the following result.

    Lemma 3.4. Assume that Ω+ is nonempty and d is a positive constant. Then the following equation

    (dβ(x)+mγ(x))I=[β(x)(N|Ω|1d|Ω|ΩIdx)γ(x)]+ (3.21)

    has a unique nonnegative solution.

    Proof. It is easy to see that (3.21) has a unique nonnegative solution I=[β(x)N/|Ω|γ(x)]+β+mγ if d=1. Hence, we only consider d(0,1)(1,+) below. Let

    Gτ:=[β(x)(N|Ω|1d|Ω|τ)γ(x)]+/(dβ(x)+mγ(x)).

    If d(0,1), then ΩGτdx is non-increasing in τ for τ0 and ΩGτdx=0 for sufficiently large τ. Define

    τ0:=min{τ0:ΩGτdx=0}.

    Since Ω+ is nonempty, we conclude that ΩG0dx=Ω[β(x)N/|Ω|γ(x)]+dβ(x)+mγ(x)dx>0 and ΩGτdx is decreasing with respect to τ[0,τ0]. Hence, there exists a unique τ(0,τ0), such that ΩGτdx=τ, i.e., Gτ is the unique nonnegative solution of (3.21).

    If d>1, then ΩGτdx is non-decreasing in τ for τ0 and ΩGτdx as τ. We notice that

    ΩGτdx1dΩ(N|Ω|γβ)+dx+(11d)τ.

    Since the right hand side of the above inequality is linear in τ with slope less than 1, there exists τ>0 such that ΩGτdx=τ, which implies that Gτ is a solution of (3.21). On the other hand, since ΩGτdx is concave up in τ, τ is the unique solution of ΩGτdx=τ. Hence, (3.21) has a unique nonnegative solution. The proof is complete.

    Proof of Theorem 1.6. Since Ω+ is nonempty, the EE (S,I) exists if dI is small by Proposition 1.1 and Theorem 1.2.

    First, we prove the conclusion for the case d<1. We claim that ΩIdxΩIdx as dI0+ and dI/dSd, where I is the unique solution of (3.21). Since ΩIdxN, there exist two sequences {dIn} and {dIn/dSn} with dIn0+ and dIn/dSnd as n such that the corresponding EE (Sn,In) satisfies ΩIndxK0[0,N]. Then, for any ϵ>0, there exists n>0 such that K0ϵ<ΩIndx<K0+ϵ and dϵ<dIn/dSn<d+ϵ when n>n. Therefore, In is a lower solution of the problem

    {dInΔI+I{β1+mI[N|Ω|1|Ω|(1dϵ)(K0ϵ)(dϵ)I]γ}=0,xΩ,Iν=0,xΩ, (3.22)

    and is an upper solution of the problem

    {dInΔI+I{β1+mI[N|Ω|1|Ω|(1d+ϵ)(K0+ϵ)(d+ϵ)I]γ}=0,xΩ,Iν=0,xΩ (3.23)

    for n>n. Denote the unique positive solutions of (3.22) and (3.23) by In,ϵ and In,+ϵ respectively if they exist; otherwise, let In,ϵ=0 (In,+ϵ=0). Then, it follows from an upper-lower solution argument that In,+ϵInIn,ϵ for n>n. Furthermore, by Lemma A.2, we have

    limnIn,±ϵ=I±ϵ

    in C(ˉΩ), where

    I±ϵ=[1β(d±ϵ)+mγ(βN|Ω|β(1d±ϵ)K0±ϵ|Ω|γ)]+. (3.24)

    Since ϵ>0 is arbitrary, by Lemma 3.4 we obtain

    K0=limnΩIndx=Ω[1βd+mγ(βN|Ω|β(1d)K0|Ω|γ)]+dx=ΩIdx.

    Now, we show II as dI0+ and dI/dSd. Substituting K0=ΩIdx into (3.24), we get that

    I[1βd+mγ(βN|Ω|β(1d)|Ω|ΩIdxγ)]+=I,

    and hence

    S=N|Ω|(1dIdS)1|Ω|ΩIdxdIdSIN|Ω|1d|Ω|ΩIdxdI=S

    as dI0+ and dI/dSd.

    The proof of the case d1 is similar, so we omit it here. The conclusion in (ⅱ) is easily obtained from equation (3.21).

    Proof of Theorem 1.8. Since Ω is a high-risk domain, the EE (S,I) always exists for any dS>0 and dI>0. By Lemmas 3.1 and 3.2, we know that I and S are uniformly bounded in L(Ω) for any positive dI and dS.

    (ⅰ) We consider the asymptotic profile of the EE when dI and dS. Note that βSI/(1+mI)γI is uniformly bounded in L(Ω). Applying the standard elliptic estimate arguments, we obtain that S and I are uniformly bounded in C2+α(ˉΩ) for all α(0,1), dS,dI1. It then follows from the compactness of the embedding C2+α(ˉΩ)C2(ˉΩ) that there exist sequences {dSn},{dIn} with dSn,dIn as n such that the corresponding EE (Sn,In)(S,I) in C2(ˉΩ), where (S,I) satisfies

    {ΔS=ΔI=0,xΩ,Sν=Iν=0,xΩ.

    By the strong maximum principle, S and I are constants. Let ˜In=In/In. By (1.4), we have

    {dInΔ˜In=(βSn1+mInγ)˜In,xΩ,˜Inν=0,xΩ. (3.25)

    Since ˜In=1 and In,Sn are uniformly bounded, it follows from the elliptic estimate and the Sobolev embedding theorem that ˜In is uniformly bounded in C2+α(ˉΩ). Passing to a subsequence if necessary, we have ˜In˜I in C2(ˉΩ), where ˜I satisfies

    {Δ˜I=0,xΩ,˜Iν=0,xΩ.

    Using the strong maximum principle again, ˜I is a constant. And then ˜I1 since ˜In=1. Integrating both sides of the first equation in (3.25) over Ω, we find

    Ω(βSn1+mInγ)˜Indx=0.

    Letting n, we have

    Ω(βS1+mIγ)dx=0.

    Noticing that Ω(S+I)dx=N and S,I,m are constants, we have

    I=NΩβdx|Ω|Ωγdx|Ω|(Ωβdx+mΩγdx), S=ΩγdxΩβdx[1+mNΩβdx|Ω|Ωγdx|Ω|(Ωβdx+mΩγdx)].

    (ⅱ) Now we consider the asymptotic profile of the EE when dI. The proof is similar to (ⅰ) and Theorem 1.3, so we sketch it in the following.

    Suppose dS is fixed. By the elliptic estimate, the Sobolev embedding theorem and the maximum principle, there exists a sequence {dIn} with dIn as n such that the corresponding EE (Sn,In)(S,I) in C2(ˉΩ), where I0 is a constant. If I=0, then S satisfies (3.4), which indicates that S is also a constant. As in the proof of (ⅰ), we also introduce ˜In=In/In. Then we can prove ˜In1 in C2(ˉΩ) as n, which leads to

    Ω(βSγ)dx=0,

    and so S=Ωγdx/Ωβdx. On the other hand, one can see from (1.5) that S=N/|Ω| when I=0. Hence, N|Ω|Ωβdx=Ωγdx, which contradicts the assumption that Ω is a high-risk domain. Therefore I is a positive constant, and S>0 satisfies (1.11).

    Furthermore, let dS0+. There exists a sequence {dSn} with dSn0+ as n such that the corresponding solution (Sn,In) of (1.11) satisfies dSn/In0, or dSn/In, or dSn/InK1 for some positive constant K1. If dSn/In0, we have (Sn,In)(˜S,˜I) in C(ˉΩ). Rewrite the equation of {Sn} as

    {dSnInΔSn+(βSn1+mIn+γ)=0,xΩ,Snν=0,xΩ.

    Letting n, we obtain

    ˜S=γβ(1+m˜I), ˜I=N|Ω|1|Ω|Ω˜Sdx,

    which implies that (˜S,˜I) satisfies (1.7).

    If dSn/InK1, by passing to a subsequence, we have (S_n^*, I_n^*)\rightarrow(\tilde{S}^*, 0) in C(\bar{\Omega}) as d_{S_n}\rightarrow 0^+ . Rewrite the equation of S_n^* as follow

    \begin{equation} \begin{cases} -\Delta S_n^* = \left(-\frac{ \beta S_n^*}{1+mI^*_n}+\gamma\right)\frac{ I^*_n}{d_{S_n}}, &x\in\Omega, \\ \frac{\partial S_n^*}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.26)

    Letting n\rightarrow\infty , we obtain \tilde{S}^* satisfies (1.12).

    If d_{S_n}/I^*_n\rightarrow\infty as d_{S_n}\rightarrow0^+ , by passing to a subsequence, we have (S_n^*, I_n^*)\rightarrow(\tilde{S}^*, 0) in C(\bar{\Omega}) and S_n^* satisfies (3.26). Letting n\rightarrow\infty , we know that \tilde{S}^* is a constant. By the last equation of (1.11), we actually have \tilde{S}^* = \frac{ N}{|\Omega|} . Integrating both sides of the first equation in (1.11), we get \int_\Omega\left[-\beta S_n^*/(1+mI_n^*)+\gamma\right]dx = 0 . Letting n\rightarrow\infty , we find \frac{N}{|\Omega|}\int_\Omega\beta dx = \int_\Omega\gamma dx , which contradicts the assumption that \Omega is a high-risk domain. Hence d_{S_n}/I^*_n\rightarrow\infty is impossible. The proof is complete.

    Proof of Theorem 1.9. Since \mathcal {R}_0 is independent of m , the EE (S, I) exists for any m > 0 provided \mathcal {R}_0 > 1 by Theorem 1.2. It follows from (3.1) that I is uniformly bounded for fixed d_I > 0, d_S > 0 . Hence there exists a sequence \{m_n\} with m_n\rightarrow\infty as n \rightarrow\infty such that the corresponding EE (S_n, I_n) satisfies

    \begin{equation} \begin{cases} -d_I \Delta I_n = I_n\left\{\frac{\beta} {1+m_nI_n}\left[\frac{ N}{|\Omega|}-\left(1-\frac{d_I}{d_S}\right)\frac{ 1}{|\Omega|}\int_\Omega I_ndx-\frac{d_I}{d_S}I_n\right]-\gamma\right\}, & x\in\Omega, \\ \frac{\partial I_n}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.27)

    Since the right hand of the above equation is uniformly bounded in L^\infty(\Omega) , by standard elliptic regularity we know that \{I_n\} is precompact in C^1(\Omega) . Hence by passing to a subsequence we may assume that I_n \rightarrow I_*\geq 0 in C(\bar\Omega) . There are two possibilities (a) \|m_nI_n\|_\infty\rightarrow\infty as n\rightarrow\infty or (b) \|m_nI_n\|_\infty\leq C , where C is a positive constant.

    If (a) occurs, then I_* satisfies

    \begin{equation} \begin{cases} -d_I \Delta I+\gamma I = 0, & x\in\Omega, \\ \frac{\partial I}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.28)

    The Fredholm alternative implies that I_* = 0 .

    If (b) occurs, then I_n\rightarrow 0 uniformly on \bar{\Omega} as n\rightarrow\infty . Denote m_nI_n = w_n . Then w_n is uniformly bounded in L^\infty(\Omega) and w_n\rightarrow w_*\geq 0 in C(\bar\Omega) , where w_* satisfies

    \begin{equation} \begin{cases} d_I \Delta w_*+w_*\left(\frac{\beta N}{|\Omega|(1+w_*)}-\gamma\right) = 0, & x\in\Omega, \\ \frac{\partial w_*}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.29)

    If w_*\equiv 0 , set \hat{w}_n = \frac{w_n}{\|w_n\|_\infty} . Then \hat{w}_n satisfies

    \begin{equation} \begin{cases} -d_I \Delta \hat{w}_n = \hat{w}_n\left\{\frac{\beta} {1+w_n}\left[\frac{ N}{|\Omega|}-\left(1-\frac{d_I}{d_S}\right)\frac{ 1}{|\Omega|}\int_\Omega I_ndx-\frac{d_I}{d_S}I_n\right]-\gamma\right\}, & x\in\Omega, \\ \frac{\partial \hat{w}_n}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.30)

    By standard elliptic regularity we may assume that \hat{w}_n\rightarrow \hat{w} in C(\bar\Omega) . Then \hat{w}\geq0, \not\equiv 0 satisfies

    \begin{equation} \begin{cases} d_I \Delta \hat{w}+\hat{w}\left(\frac{\beta N}{|\Omega|}-\gamma\right) = 0, & x\in\Omega, \\ \frac{\partial \hat{w}}{\partial \nu} = 0, &x\in\partial\Omega. \end{cases} \end{equation} (3.31)

    It follows from the strong maximum principle that \hat{w} > 0 on \bar{\Omega} . So \lambda_1(d_I, \beta N/|\Omega|-\gamma) = \lambda^* = 0 , which implies \mathcal {R}_0 = 1 . This is a contradiction to the assumption \mathcal {R}_0 > 1 . Hence w_*\geq0, \not\equiv 0 in \bar{\Omega} . Using the strong maximum principle again, we know that w_* > 0 on \bar{\Omega} . It follows from Lemma A.2 that (3.29) has a unique positive solution w_* as \lambda_1(d_I, \beta N/|\Omega|-\gamma) = \lambda^* < 0 .

    To illustrate the influences of the population diffusion and the saturation coefficient on system (1.2), we suppose the spatial domain \Omega = [0, 1] . Let \beta(x) = \sin(2\pi x)+a, \gamma(x) = \cos(2\pi x)+b , where a > 1, b > 1. Take S(x, 0) = \sin(\pi x)+1, I(x, 0) = \sin(\pi x)+1 , a = 1.5 , b = 4.5 . Then N = \int_\Omega(S(x, 0)+I(x, 0))dx = 3.2731 satisfies

    \begin{equation} 3 = \frac{\int_\Omega \gamma(x)dx}{\int_\Omega \beta(x)dx} \lt \frac{N}{|\Omega|} \lt \frac{1}{|\Omega|}\int_\Omega \frac{\gamma(x)}{\beta(x)}dx \lt \max\limits_\Omega \frac{\gamma(x)}{\beta(x)} = 9.1094, \end{equation} (4.1)

    where \frac{1}{|\Omega|}\int_\Omega \frac{\gamma(x)}{\beta(x)}dx = 4.0249 . Hence \Omega^+ , \Omega^- are nonempty, and \Omega is a high-risk domain which implies that \mathcal {R}_0 > 1 by Proposition 1.1 (ⅲ).

    (ⅰ) Firstly, we explore the influence of d_S on system (1.2). Fix d_I = 0.5 and m = 2 . A numerical positive equilibrium solution (S(x), I(x)) to (1.2) was computed in Figure 2(1) with d_S = 0.1 . When d_S is large, we find that S(x) tends to a constant and I(x) closely approximates the unique positive solution of (1.6) (see Figure 2(2) with d_S = 10^3 ), which is consistent with Theorem 1.3 (ⅰ). As d_S decreases, the value of I(x) decreases too. In Figure 2(3), I(x) closely approximates zero with d_S = 10^{-6} , which coincides with the results of Theorem 1.3(ⅱ) and Corollary 1.4.

    Figure 2.  The influence of d_S on the EE (S(x), I(x)) with d_I = 0.5 and m = 2 . Left column: the profile of S(x) ; Right column: the profile of I(x) .

    If we take a = 1.5 and b = 1.2 , then 3.2731 = \frac{N}{|\Omega|} > \max\limits_{x\in\bar\Omega}\frac{\gamma(x)}{\beta(x)} = 2.7514 , which implies that \mathcal {R}_0 > 1 by Proposition 1.1(ⅰ) and (ⅱ). By taking d_I = 0.5 and d_S = 10^{-6} , we find that I(x) closely approximates a positive constant (see Figure 2(4)), which coincides with the results of Theorem 1.3 (ⅱ) and Corollary 1.5.

    (ⅱ) Secondly, we analyze the asymptotic profile of the equilibrium solution (S(x), I(x)) to system (1.2) when d_S and d_I are small. Fix a = 1.5 , b = 4.5 and m = 2 . Then \Omega^+ and \Omega^- are nonempty from (4.1). By taking d_S = 10^{-4} and d_I = 0.2\times 10^{-4} , we find that (0.0470, 0.5590)\approx\{x\in \Omega: I^* > 0\}\subsetneqq\Omega^+ = (0.0281, 0.5663), which coincides with the result of Theorem 1.6 with d < 1 (see Figure 3(1)(2)). By taking {d_S} = 10^{-4} and d_I = 2\times 10^{-4} , we find that [0, 0.6954) = \{x\in \Omega: I^* > 0\}\supseteqq\Omega^+ = (0.0281, 0.5663), which coincides with the result of Theorem 1.6 with d > 1 (see Figure 3(1), (3)).

    Figure 3.  a = 1.5 , b = 4.5 and m = 2 . (1): Graph of \frac{N}{|\Omega|}\beta(x)-\gamma(x) ; \Omega^+ = (0.0281, 0.5663). (2): The profile of I(x) with d_S = 10^{-4} and d_I = 0.2\times 10^{-4} ; \{x\in \Omega: I^* > 0\}\approx (0.0470, 0.5590) . (3): The profile of I(x) with d_S = 10^{-4} and d_I = 2\times 10^{-4} ; \{x\in \Omega: I^* > 0\}\approx [0, 0.6954) . Here, I^* is the solution of (1.10).

    (ⅲ) Thirdly, we explore the asymptotic profile of the equilibrium solution (S(x), I(x)) to system (1.2) when d_I is large. Fix a = 1.5 , b = 4.5 and m = 2 . Then \Omega is a high-risk domain from (4.1). A numerical positive equilibrium solution (S(x), I(x)) to (1.2) was computed (see Figure 4). In Figure 4(1), by taking d_S = d_I = 10^4 , we find that (S(x), I(x)) closely approximates a constant equilibrium solution, which coincides with the result of Theorem 1.8(ⅰ). In Figure 4(2), by taking d_S = 0.1 and d_I = 10^{4} , we find that I(x) closely approximates a positive constant, which coincides with the result of Theorem 1.8(ⅱ).

    Figure 4.  The profile of the EE (S(x), I(x)) when d_I is large with a = 1.5 , b = 4.5 and m = 2 . Left column: S(x) ; Right column: I(x) .

    (ⅳ) Finally, we show the influence of saturation coefficient m on system (1.2). Fix a = 1.5 , b = 4.5 , d_S = 0.1 and d_I = 0.5 . Then \mathcal {R}_0 > 1 from (4.1). A numerical positive equilibrium solution (S(x), I(x)) to (1.2) was computed (see Figure 5). In Figure 5(1), by taking m = 2 , there is a positive equilibrium solution (S(x), I(x)) to (1.2). In Figure 5(2), by taking m = 10^3 , we find that (S(x), I(x)) closely approximates the DFE (N/|\Omega|, 0) , which coincides with the result of Theorem 1.9.

    Figure 5.  The influence of m on the EE (S(x), I(x)) with a = 1.5 , b = 4.5 , d_S = 0.1 and d_I = 0.5 . Left column: the profile of S(x) ; Right column: the profile of I(x) .

    In this paper, we investigate an SIS reaction-diffusion population model (1.2) with saturated incidence rate \beta\bar{I}\bar{S}/(1+m\bar{I}) . We focus on the existence of EE and particularly the effects of the diffusion rates and the saturated coefficient on asymptotic profiles of EE. Similar questions are addressed for the model with the standard incidence rate \beta \bar{S}\bar{I}/(\bar{S}+\bar{I}) in [1,20] or the bilinear rate \beta\bar{S}\bar{I} in [7,16,25].

    Firstly, by introducing the basic reproduction number \mathcal {R}_0 as in [1,7], we obtain for (1.2) that there is at least one EE if \mathcal {R}_0 > 1 , especially, EE exists uniquely when d_S\geq d_I . In contrast to [1], the definition of \mathcal {R}_0 , \Omega^+ and \Omega^- for our model (1.2) depends on the average population density, i.e. N/|\Omega| . From the epidemiology point of view, the more crowded the population is, the easier endemic a disease becomes.

    Secondly, we analyse the asymptotic profile of EE when it exists. Regarding this, it is worthwhile to mention the following three results.

    (ⅰ) The diffusion rate of the susceptible individuals d_S is small. Theorem 4 in [1] indicates that the asymptotic profile is some spatially inhomogeneous DFE. However, our results for (1.2) show that EE (if it exists) may approach a coexistence limiting equilibrium where the susceptible individuals spatially heterogeneously exist and the infected population is a spatially homogeneous state (see Theorem 1.3(ⅱ) and Corollary 1.5). Furthermore, we observe that the coexistence limiting equilibrium tends to a spatially inhomogeneous DFE when the saturated coefficient m is large. From a disease control point of view, if a disease is governed by system (1.2), it is not enough to just restrict the movement of the susceptible individuals to completely eradicate the disease in the whole habitat in certain situations, especially if the rate of disease transmission \beta is a constant and \Omega is a high-risk domain; see Corollary 1.5 (ⅰ).

    (ⅱ) d_I\rightarrow0 and d_I/d_S\rightarrow d\in(0, \infty). Theorem 1.1(2) and Corollary 1.1(ⅰ) of [20] show that the existence habitat of the infective individuals is exactly the high-risk set. However, in our results, the ratio d plays a key role (see Theorem 1.6 and Remark 1.7). If d = 1 , the infected individuals survive exactly in the high-risk set; if d\in(0, 1) , the habitat of infected individual is confined within some subset of the high-risk set; if d > 1 , the infected individuals only die out at part of the low-risk sites. This information suggests that reducing the radio d to less than 1 will help to control disease; in other words, the more isolated the patients become, the better disease control is.

    (ⅲ) The saturated coefficient m is large. For model (1.2), Theorem 1.9 indicates that the EE tends to the DFE, i.e., the infective individuals cannot persist. This result seems to coincide with the realistic intuition: the more inhibition effect from the behavioral change of the susceptible individuals when their number increases or from the crowding effect of the infective individuals, the better for disease control.

    Finally, we would like to mention some open problems left for future study: (1) the existence of EE when d_S < d_I and \mathcal {R}_0\in (d_S/d_I, 1) ; (2) the asymptotic profile of EE when \mathcal {R}_0 > 1 and \frac{1}{|\Omega|}\int_\Omega \frac{\gamma(x)}{\beta(x)} dx < \frac{N}{|\Omega|}\leq\max\limits_{x\in\bar\Omega}\frac{\gamma(x)}{\beta(x)} as d_S\rightarrow0^+ ; (3) the uniqueness of EE if it exists when d_S < d_I.

    The work of Y. Wang and Z. Wang was supported by the Natural Science Foundation of China (11671243, 61672021, 11771262, 11571274, 11501496), the Fundamental Research Funds for the Central Universities (GK201903088), the Natural Science Basic Research Plan in Shaanxi Province of China (2018JM1020, 2018JQ1021). The work of C. Lei was supported by the Natural Science Foundation of China (11801232), the Priority Academic Program Development of Jiangsu Higher Education Institution, the Natural Science Foundation of the Jiangsu Province(BK20180999), the Foundation of Jiangsu Normal University (17XLR008).

    The authors declare there is no conflict of interest.

    We recall the following well-known facts without proof.

    Let \lambda_1(d, g) be the principal eigenvalue of

    \begin{equation} \begin{cases} d \Delta \varphi+g(x)\varphi+\lambda\varphi = 0, &x\in\Omega, \\ \frac{\partial \varphi}{\partial \nu} = 0, &x\in\partial\Omega, \end{cases} \end{equation} (A.1)

    where g(x)\in L^\infty (\Omega) and d > 0 . It is folklore that \lambda_1(d, g) is given by

    \begin{equation} \lambda_1(d, g) = \inf\left\{\int_\Omega \left(d |\nabla\varphi|^2-g(x) \varphi^2\right)dx: \varphi\in H^1(\Omega) \ \text{and}\ \int_\Omega\varphi^2dx = 1\right\}. \end{equation} (A.2)

    The properties of \lambda_1(d, g) are stated as follows; they can be found in [1,3].

    Lemma A.1.

    (ⅰ) If g_1(x)\leq g_2(x) in \Omega with g_i\in L^\infty (\Omega) for i = 1, 2, then \lambda_1(d, g_1)\geq\lambda_1(d, g_2) with equality holds if and only if g_1 = g_2 a.e. in \Omega ;

    (ⅱ) if g\in L^\infty (\Omega) is non-constant, then \lambda_1(d_1, g) < \lambda_1(d_2, g) when d_1 < d_2 ;

    (ⅲ) \lambda_1(d, g) depends continuously on g and d , and it satisfies

    \begin{equation} \lim\limits_{d\rightarrow0^+}\lambda_1(d, g) = \min\limits_{x\in\bar{\Omega}}\{-g(x)\} \ \text{ and }\ \lim\limits_{d\rightarrow \infty}\lambda_1(d, g) = -\frac{1}{|\Omega|}\int_\Omega g(x)dx. \end{equation} (A.3)

    The following lemma is about the existence of an elliptic problem and its asymptotic profile (as d\rightarrow0^+ ), which can be found in [3] or can be directly proved by an upper/lower solution argument. Let us denote h^+(x): = \max\{h(x), 0\} for any function h defined on \bar{\Omega} .

    Lemma A.2. Suppose that positive functions a(x), b(x), d(x)\in C^\alpha(\bar{\Omega}) . Then the following statements hold for the problem

    \begin{equation} \begin{cases} d\Delta u+ \left(\frac{a(x)}{1+b(x)u}-c(x)\right)u = 0, &x\in\Omega, \\ \frac{\partial u}{\partial \nu} = 0, & x\in\partial\Omega.\\ \end{cases} \end{equation} (A.4)

    (ⅰ) If \lambda_1(d, a-c)\geq 0 , then u = 0 is the only non-negative solution of (A.4);

    (ⅱ) if \lambda_1(d, a-c) < 0 , then (A.4) has a unique positive solution u\in C^{2+\alpha}(\bar{\Omega}) . Furthermore, u\rightarrow \left(\frac{a-c}{bc}\right)^+ as d\rightarrow0^+ provided a(x_0)-c(x_0) > 0 for some x_0\in \bar{\Omega} .

    To obtain the priori estimates for solutions, the following maximum principle (due to Lou and Ni [18]) and the Harnack's inequality (see, e.g., [17]) are useful.

    Lemma A.3. Suppose that g\in C(\bar \Omega\times\mathbb{R}) .

    (ⅰ) Assume that \omega\in C^2(\Omega)\cap C^1(\bar\Omega) and satisfies

    \begin{equation*} \begin{cases} \Delta\omega(x)+g(x, \omega(x))\geq0, &x\in\Omega, \\ \frac{\partial\omega}{\partial \nu} = 0, & x\in\partial\Omega. \end{cases} \end{equation*}

    If \omega(x_0) = \max\limits_{x\in\bar\Omega}\omega(x) , then g(x_0, \omega(x_0))\geq0.

    (ⅱ) Assume that \omega\in C^2(\Omega)\cap C^1(\bar\Omega) and satisfies

    \begin{equation*} \begin{cases} \Delta\omega(x)+g(x, \omega(x))\leq 0, &x\in\Omega, \\ \frac{\partial\omega}{\partial \nu} = 0, & x\in\partial\Omega. \end{cases} \end{equation*}

    If \omega(x_0) = \min\limits_{x\in\bar\Omega}\omega(x) , then g(x_0, \omega(x_0))\leq0.

    Lemma A.4. Let w\in C^2(\Omega)\cap C^1(\bar{\Omega}) be a positive solution of

    \begin{equation*} \begin{cases} \Delta\omega(x)+c(x)\omega(x) = 0, &x\in\Omega, \\ \frac{\partial\omega}{\partial \nu} = 0, & x\in\partial\Omega, \end{cases} \end{equation*}

    where c(x)\in C(\bar{\Omega}) . Then there exists a positive constant C = C(n, \Omega, \|c\|_\infty) such that

    \max\limits_{\overline{\Omega}} w\leq C\min\limits_{\overline{\Omega}} w.


    [1] L. J. S. Allen, B. M. Bolker, Y. Lou, et al., A generalization of the Kermack–McKendrick deterministic epidemic model, Discrete Contin. Dyn. Syst., 21 (2008), 1–20.
    [2] F. Brauer and C. Castillo-Ch'avez, Mathematical models in population biology and epidemiology,Springer, 2001.
    [3] R. S. Cantrell and C. Cosner, Spatial Ecology via Reaction–Diffusion Equations, John Wiley and Sons Ltd. Chichester, UK, 2003.
    [4] 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.
    [5] R. Cui and Y. Lou, A spatial SIS model in advective heterogeneous environments, J. Differ. Equations, 261 (2016), 3305–3343.
    [6] V. Capasso and G. Serio, A generalization of the Kermack–McKendrick deterministic epidemic model, Math. Biosci., 42 (1978), 41–61.
    [7] K. Deng and Y. Wu, Dynamics of an SIS epidemic reaction–diffusion model, Proc. Roy. Soc. Edinburgh Sect. A, 146 (2016), 929–946.
    [8] J.Ge, K.I.Kim, Z.Lin, etal., ASISreaction-diffusion-advectionmodelinalow-riskandhigh-risk domain, J. Differ. Equations, 259 (2015), 5486–5509.
    [9] J. Ge, L. Lin and L. Zhang, A diffusive SIS epidemic model incorporating the media coverage impact in the heterogeneous environment, Discrete Contin. Dyn. Syst. Ser B, 22 (2017), 2763–2776.
    [10] W. O. Kermack and A. G. McKendrick, A contribution to the mathematical theory of epidemics, Proc. R. Soc. Lond. Ser. A, 115 (1927), 700–721.
    [11] K. Kousuke, H. Matsuzawa and R. Peng, Concentration profile of endemic equilibrium of a reaction-diffusion-advection SIS epidemic model, Calc. Var. Part. D. E., 56 (2017), 112.
    [12] A. Lahrouz and L. Omari, Extinction and stationary distribution of a stochastic SIRS epidemic model with non-linear incidence, Stat. Probabil. Lett., 83 (2013), 960–968.
    [13] C. Lei, Z. Lin and Q. Zhang, The spreading front of invasive species in favorable habitat or unfavorable habitat, J. Differ. Equations, 257 (2014), 145–166.
    [14] B. Li, H. Li and Y. Tong, Analysis on a diffusive SIS epidemic model with logistic source, Z. Angew. Math. Phys., 68 (2017), 96.
    [15] 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.
    [16] H. Li, R. Peng and F.-B. Wang, On a diffusive susceptible-infected-susceptible 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.
    [17] C.-S. Lin, W.-W. Ni and I. Takagi, Large amplitude stationary solutions to a chemotaxis system, J. Differ. Equations, 72 (1988), 1–27.
    [18] Y. Lou and W.-M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Differ. Equations, 131 (1996), 79–131.
    [19] X. Meng, S. Zhao, T. Feng, et al., Dynamics of a novel nonlinear stochastic SIS epidemic model with double epidemic hypothesis, J. Math. Anal. Appl., 433 (2016), 227–242.
    [20] R. Peng, Asymptotic profiles of the positive steady state for an SIS epidemic reaction–diffusion model. Part I, J. Differ. Equations, 247 (4) (2009), 1096–1119.
    [21] R. Peng and S. Liu, Global stability of the steady states of an SIS epidemic reaction–diffusion model, Nonlinear Anal., 71 (4) (2009), 239–247.
    [22] R. Peng and F. Yi, Asymptotic profile of the positive steady state for an SIS epidemic reaction–diffusion model: effects of epidemic risk and population movement, Phys. D, 259 (2013), 8–25.
    [23] R. Peng and X.-Q. Zhao, A reaction–diffusion SIS epidemic model in a time–periodic environment, Nonlinearity, 25 (2012), 1451–1471.
    [24] 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.
    [25] 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.
    [26] R. Xu and Z. Ma, Global stability of a SIR epidemic model with nonlinear incidence rate and time delay, Nonlinear Anal., 10 (2009), 3175–3189.
    [27] F. Zhang, Z. Jin and G. Sun, Bifurcation analysis of a delayed epidemic model, Appl. Math. Comput., 216 (2010), 753–767.
  • This article has been cited by:

    1. Linhe Zhu, Xiaoyuan Huang, Ying Liu, Zhengdi Zhang, Spatiotemporal dynamics analysis and optimal control method for an SI reaction-diffusion propagation model, 2021, 493, 0022247X, 124539, 10.1016/j.jmaa.2020.124539
    2. 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
    3. Jialiang Zhang, Renhao Cui, Asymptotic behavior of an SIS reaction–diffusion–advection model with saturation and spontaneous infection mechanism, 2020, 71, 0044-2275, 10.1007/s00033-020-01375-9
    4. Xin Huo, Renhao Cui, A reaction–diffusion SIS epidemic model with saturated incidence rate and logistic source, 2020, 0003-6811, 1, 10.1080/00036811.2020.1859495
    5. Xueying Sun, Renhao Cui, Analysis on a diffusive SIS epidemic model with saturated incidence rate and linear source in a heterogeneous environment, 2020, 490, 0022247X, 124212, 10.1016/j.jmaa.2020.124212
    6. Jialiang Zhang, Renhao Cui, Asymptotic profiles of the endemic equilibrium of a diffusive SIS epidemic system with saturated incidence rate and spontaneous infection, 2021, 44, 0170-4214, 517, 10.1002/mma.6754
    7. Renhao Cui, Asymptotic profiles of the endemic equilibrium of a reaction-diffusion-advection SIS epidemic model with saturated incidence rate, 2021, 26, 1553-524X, 2997, 10.3934/dcdsb.2020217
    8. Oyoon Abdul Razzaq, Najeeb Alam Khan, Muhammad Faizan, Asmat Ara, Saif Ullah, Behavioral response of population on transmissibility and saturation incidence of deadly pandemic through fractional order dynamical system, 2021, 26, 22113797, 104438, 10.1016/j.rinp.2021.104438
    9. Xueying Sun, Renhao Cui, Existence and asymptotic profiles of the steady state for a diffusive epidemic model with saturated incidence and spontaneous infection mechanism, 2021, 14, 1937-1632, 4503, 10.3934/dcdss.2021120
    10. Qi Cao, Yuying Liu, Wensheng Yang, Global dynamics of a diffusive SIR epidemic model with saturated incidence rate and discontinuous treatments, 2022, 10, 2195-268X, 1770, 10.1007/s40435-022-00935-3
    11. 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
    12. Yutong Guo, Jinliang Wang, Desheng Ji, Asymptotic profiles of a diffusive SIS epidemic model with vector-mediated infection and logistic source, 2022, 73, 0044-2275, 10.1007/s00033-022-01888-5
    13. Xuan Tian, Shangjiang Guo, Zhisu Liu, Qualitative analysis of a diffusive SEIR epidemic model with linear external source and asymptomatic infection in heterogeneous environment, 2022, 27, 1531-3492, 3053, 10.3934/dcdsb.2021173
    14. Chuanxin Liu, Renhao Cui, Qualitative analysis on an SIRS reaction–diffusion epidemic model with saturation infection mechanism, 2021, 62, 14681218, 103364, 10.1016/j.nonrwa.2021.103364
    15. Huicong Li, Tian Xiang, On an SIS epidemic model with power‐like nonlinear incidence and with/without cross‐diffusion, 2024, 153, 0022-2526, 10.1111/sapm.12683
    16. 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
    17. Anqi Qu, Jinfeng Wang, Asymptotic profiles of positive steady states in a reaction–diffusion benthic–drift model, 2024, 153, 0022-2526, 10.1111/sapm.12752
    18. Soufiane Bentout, Salih Djilali, Asymptotic profiles of a generalized reaction-diffusion SIS epidemic model with spatial heterogeneity, 2024, 75, 0044-2275, 10.1007/s00033-024-02373-x
    19. Han Lu, Yan’e Wang, Jianhua Wu, An Efficient Competitive Control Mechanism for Negative Information Spread in Online Social Networks, 2024, 34, 0218-1274, 10.1142/S0218127424501724
    20. Renhao Cui, Huicong Li, Zikun Wang, Saturated incidence tends to ease disease persistence: Analysis of an SIS epidemic patch model, 2025, 0, 1078-0947, 0, 10.3934/dcds.2025010
    21. Hongmin Zhang, Jian Zhang, Xin Huo, Theoretical Analysis on a Diffusive SIS Epidemic Model with Logistic Source, Saturated Incidence Rate and Spontaneous Infection Mechanism, 2025, 13, 2227-7390, 1244, 10.3390/math13081244
  • Reader Comments
  • © 2019 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(5142) PDF downloads(788) Cited by(21)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog