We study a spatial susceptible-infected-susceptible(SIS) model in heterogeneous environments with vary advective rate. We establish the asymptotic stability of the unique disease-free equilibrium(DFE) when R0<1 and the existence of the endemic equilibrium when R0>1. Here R0 is the basic reproduction number. We also discuss the effect of diffusion on the stability of the DFE.
Citation: Xiaowei An, Xianfa Song. A spatial SIS model in heterogeneous environments with vary advective rate[J]. Mathematical Biosciences and Engineering, 2021, 18(5): 5449-5477. doi: 10.3934/mbe.2021276
[1] | 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 |
[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] | 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 |
[4] | 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 |
[5] | Pengfei Liu, Yantao Luo, Zhidong Teng . Role of media coverage in a SVEIR-I epidemic model with nonlinear incidence and spatial heterogeneous environment. Mathematical Biosciences and Engineering, 2023, 20(9): 15641-15671. doi: 10.3934/mbe.2023698 |
[6] | Danfeng Pang, Yanni Xiao . The SIS model with diffusion of virus in the environment. Mathematical Biosciences and Engineering, 2019, 16(4): 2852-2874. doi: 10.3934/mbe.2019141 |
[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] | 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 |
[9] | Maoxing Liu, Jie Zhang, Zhengguang Li, Yongzheng Sun . Modeling epidemic in metapopulation networks with heterogeneous diffusion rates. Mathematical Biosciences and Engineering, 2019, 16(6): 7085-7097. doi: 10.3934/mbe.2019355 |
[10] | Cruz Vargas-De-León, Alberto d'Onofrio . Global stability of infectious disease models with contact rate as a function of prevalence index. Mathematical Biosciences and Engineering, 2017, 14(4): 1019-1033. doi: 10.3934/mbe.2017053 |
We study a spatial susceptible-infected-susceptible(SIS) model in heterogeneous environments with vary advective rate. We establish the asymptotic stability of the unique disease-free equilibrium(DFE) when R0<1 and the existence of the endemic equilibrium when R0>1. Here R0 is the basic reproduction number. We also discuss the effect of diffusion on the stability of the DFE.
In this paper, we are concerned with the following susceptible-infected-susceptible(SIS) model
{St=(dSSx−a′(x)S)x−β(x)SIS+I+γ(x)I, 0<x<L, t>0,It=(dIIx−a′(x)I)x+β(x)SIS+I−γ(x)I, 0<x<L, t>0,dSSx−a′(x)S=dIIx−a′(x)I=0, x=0,L, t>0,S(x,0)=S0(x), I(x,0)=I0(x),0<x<L. | (1.1) |
Here S(x,t) and I(x,t) denote the density of susceptible and infected individuals in a given spatial interval (0,L), dS and dI are positive constants which stand for the diffusion coefficients for the susceptible and infected populations, a′(x) is a smooth nonnegative function which represents the advection speed rate, while β(x) and γ(x) represent the rates of disease transmission and recovery at location x, which are Hölder continuous functions on (0,L). In addition, S0(x) and I0(x) are continuous and satisfy
(A1)S0(x)≥0 and I0(x)≥0 for x∈(0,L),∫L0I0(x)dx>0. |
We would like to give the survey of some results on SIS model. In [1], Allen et al. investigated a discrete SIS model, in [2], they also proposed the SIS model with no advection in a given spatial region Ω, where they dealt with the existence, uniqueness and asymptotic behaviors of the endemic equilibrium as the diffusion rate of the susceptible individuals approaches to zero. Many authors also considered the SIS reaction–diffusion model, including the global stability of the endemic equilibrium, the effects of large and small diffusion rates of the susceptible and infected population on the persistence and extinction of the disease, discuss how the disease vanish or spreading in high-risk or low-risk domain, and so on. For the dynamics and asymptotic profiles of steady states of an epidemic model in advective environments, we can see[3]. For A SIS reaction-diffusion-advection model in a low-risk and high-risk domain, we can see [4]. For Dynamics of an SIS reaction-diffusion epidemic model for disease transmission, we can see[5], For Concentration profile of endemic equilibrium of a reaction- diffusion-advection SIS epidemic model, we can see [6]. For the varying total population enhances disease persistence, we can see [7]; For the asymptotic profiles of the positive steady state for an SIS epidemic reaction-diffusion model, we can see [8]. For the global stability of the steady states of an SIS epidemic reaction-diffusion model, we can see [9]. For the asymptotic profile of the positive steady state for an SIS epidemic reaction- diffusion model: effects of epidemic risk and population movement, we can see [10]; For reaction-diffusion SIS epidemic model in a time-periodic environment, we can see [11]. For the global dynamics and traveling waves for a periodic and diffusive chemostat model with two nutrients and one microorganism, we can see [12]. For more information about dynamical systems in population biology, we also can refer to see [13] and the references therein. Recently, Cui and Lou studied (1.1) when a′(x)≡q for x∈[0,L] in [14], that is, it is a constant advection. Besides establishing the asymptotic stability of the unique disease-free equilibrium(DFE) when R0<1 and the existence of the endemic equilibrium when R0>1, they found that the DFE changes its stability at most once as dI varies from zero to infinity, which is strong contrast with the case of no advection. Since (1.1) has vary advection, an natural and interesting question is whether we can establish the similar results on (1.1) to those in the case of no advection or not.
Since the functions a′(x), β(x), γ(x), S0(x) and I0(x) are continuous in (0,L), by the standard theory for a system of semilinear parabolic equations, (1.1) is locally wellposedness in (0,Tmax). Noticing (A1), by the maximum principle, S(x,t) and I(x,t) are positive and bounded for x∈[0,L] and t∈(0,Tmax). Hence, by the results in [15], Tmax=∞ and (1.1) posses a unique classical solution (S(x,t),I(x,t)) for all time.
It is easy to verify that
∫L0[S(x,t)+I(x,t)]dx=∫L0[S(x,0)+I(x,0)]dx:=N>0,t>0. | (1.2) |
Inspired by [2] and [14], we say that (0,L) is a low-risk domain if ∫L0β(x)dx<∫L0γ(x)dx and high-risk domain if ∫L0β(x)dx>∫L0γ(x)dx.
The corresponding equilibrium system of (1.1) is
{(dS˜Sx−a′(x)˜S)x−β(x)˜S˜I˜S+˜I+γ(x)˜I=0, 0<x<L,(dI˜Ix−a′(x)˜I)x+β(x)˜S˜I˜S+˜I−γ(x)˜I=0, 0<x<L,dS˜Sx−a′(x)˜S=dI˜Ix−a′(x)˜I=0, x=0,L. | (1.3) |
The half trivial solution (˜S(x),0) of (1.3) is called a disease-free equilibrium(DFE), while the solution (˜S(x),˜I(x)) of (1.3) is called endemic equilibrium(EE) if ˜I(x)>0 for some x∈(0,L).
We also introduce the following basic reproduction number as those in literatures [2] and [14]. We also can refer to [16] and see the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, refer to [17] and see reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, see basic reproduction numbers for reaction-diffusion epidemic models [18].
R0=supφ∈H1((0,L)),φ≠0{∫L0β(x)ea(x)dIφ2dxdI∫L0ea(x)dIφ2xdx+∫L0γ(x)ea(x)dIφ2dx}. | (1.4) |
Our first result is concerned with the qualitative properties for R0.
Theorem 1.1. Let ˆR0 be the basic reproduction number when a(x)≡0 which was introduced in [2]. Then the following conclusions hold.
(1) For any given a′(x)>0, R0→β(L)γ(L) as dI→0 and R0→∫L0β(x)dx∫L0γ(x)dx as dI→+∞;
(2) For any given dI>0, R0→ˆR0 as maxx∈[0,L]a′(x)→0 and R0→β(L)γ(L) as minx∈[0,L]a′(x)→+∞;
(3) If β(x)>(<)γ(x) on [0,L], then R0>(<1) for any given dI>0 and a′(x)>0.
Our second result deals with the stability of DFE, which will extend those of [2] and [14].
Theorem 1.2. The DFE is unstable if R0>1 while it is globally asymptotically stable if R0<1.
We will analyze (1.1) under the following assumptions on β(x) and γ(x):
(C1) β(0)−γ(0)<0<β(L)−γ(L), i.e., β(x)−γ(x) changes sign from negative to positive,
or
(C2) β(0)−γ(0)>0>β(L)−γ(L), i.e., β(x)−γ(x) changes sign from positive to negative.
In the point view of biological,
(C1) all lower-risk sites are located at the upstream and all high-risk sites are at the downstream,
or
(C2) all high-risk sites are distributed at the upstream and lower-risk sites are at the downstream.
To state other results, in convenience, let q=maxx∈[0,L]a′(x) and denote a(x)=q˜a(x) sometimes in the sequels.
We can get further properties of R0 when ∫L0β(x)dx>∫L0γ(x)dx.
Theorem 1.3. Assume that ∫L0β(x)dx>∫L0γ(x)dx. Denote R0=R0(dI,q).
(i) If (C1) holds, then the DFE is unstable for any q>minx∈[0,L]a′(x)>0 and dI>0;
(ii) If (C2) holds, then there exists a unique curve in dI–q plane
Γ1={(dI,ρ1(dI)):R0(dI,ρ1(dI))=1,dI∈(0,+∞)} |
with the function ρ1=ρ1(dI):(0,+∞)→(0,+∞) satisfying
limdI→0+ρ1(dI)=0,limdI→+∞ρ1(dI)dI=θ1, |
and such that for every dI>0, the DFE is unstable for 0<minx∈[0,L]a′(x)<q<ρ1(dI) and it is globally and asymptotically stable for q>minx∈[0,L]a′(x)>ρ1(dI).
Here θ1 is the unique positive solution of
∫L0[β(x)−γ(x)]eθ1˜a(x)dx=0. |
Similarly, we can get further properties of R0 when ∫L0β(x)dx<∫L0γ(x)dx.
Theorem 1.4. Assume that ∫L0β(x)dx<∫L0γ(x)dx. Let d∗I is the unique positive root of the equation ˆR0=1, where ˆR0 was introduced in [2].
(1) If (C1) holds, then the DFE is unstable for any q>minx∈[0,L]a′(x)>0 and dI∈(0,d∗I], while for dI∈(d∗I,+∞) there exists a unique curve in dI–q plane
Γ2={(dI,ρ2(dI)):R0(dI,ρ2(dI))=1,dI∈(d∗I,+∞)} |
with the monotone function ρ2=ρ2(dI):(d∗I,+∞)→(0,+∞) satisfying
limdI→d∗I+ρ2(dI)=0,limdI→+∞ρ2(dI)dI=θ2, |
and such that the DFE is unstable for 0<minx∈[0,L]a′(x)<q<ρ2(dI) and it is globally asymptotically stable for q>minx∈[0,L]a′(x)>ρ2(dI).
Here θ2 is the unique positive solution of
∫L0[β(x)−γ(x)]eθ2˜a(x)dx=0. |
(2) If (C2) holds, then for dI∈(0,d∗I), there exists a unique curve in dI−q plane
Γ3={(dI,ρ3(dI)):R0(dI,ρ3(dI))=1,dI∈(0,d∗I)} |
with the function ρ3=ρ3(dI):(0,d∗I)→(0,+∞) satisfying
limdI→0+ρ3(dI)=0,limdI→d∗I−ρ3(dI)=0, |
and such that the DFE is unstable for 0<minx∈[0,L]a′(x)<q<ρ3(dI) and it is globally and asymptotically stable for q>minx∈[0,L]a′(x)>ρ3(dI), while for dI∈(d∗I,+∞), the DFE is globally and asymptotically stable for any q>minx∈[0,L]a′(x)>0.
The following theorem deals with the existence of EE.
Theorem 1.5. Assume that β(x)−γ(x) changes sign once in (0,L). If R0>1, then problem (1.3) possesses at least one EE.
The last theorem will consider the results on (1.1) when β(x)−γ(x) changes sign twice in (0,L).
Theorem 1.6. Assume that β(x)−γ(x) changes sign twice in (0,L).
(1) If ∫L0β(x)dx>∫L0γ(x)dx and β(L)<γ(L), then there exists some positive constant Λ which is independent of dI and q such that for every dI>Λ, we can find a positive constant Q which depends on dI such that R0>1 when 0<minx∈[0,L]a′(x)<q<Q and R0<1 when q>Q.
(2) If ∫L0β(x)dx>∫L0γ(x)dx and β(L)>γ(L), then there exists some positive constant Λ which is independent of dI and q such that for every dI>Λ one of the following conclusions holds:
(i) R0>1 for any q>minx∈[0,L]a′(x)>0;
(ii) There exists a positive constant ˆQ which is independent of dI and satisfies that R0>1 for q≠ˆQ and R0=1 when q=ˆQ;
(iii) There exist two positive constants Q2>Q1 both depending on dI such that R0>1 when q∈(0,Q1)∪(Q2,+∞) while R0<1 when q∈(Q1,Q2).
(3) If ∫L0β(x)dx<∫L0γ(x)dx and β(L)>γ(L), then there exists some positive constant Λ>d∗I which is independent of dI and q such that for every dI>Λ, we can find a positive constant Q which depends on dI such that R0<1 when 0<minx∈[0,L]a′(x)<q<Q and R0>1 when q>Q.
(4) If ∫L0β(x)dx<∫L0γ(x)dx and β(L)<γ(L), then there exists some positive constant Λ>d∗I which is independent of dI and q such that for every dI>Λ one of the following conclusions holds:
(iv) R0<1 for any q>minx∈[0,L]a′(x)>0;
(v) There exists a positive constant ˆQ which is independent of dI and satisfies that R0<1 for q≠ˆQ and R0=1 when q=ˆQ;
(vi) There exist two positive constants Q2>Q1 both depending on dI and satisfy that R0<1 when q∈(0,Q1)∪(Q2,+∞) while R0>1 when q∈(Q1,Q2).
The rest of this paper is organized as follows. In Section 2, we give the proofs of Theorem 1.1 and Theorem 1.2. In Section 3, we will prove Theorem 1.3. In Section 4, we will prove Theorem 1.4. In Section 5, we will prove Theorem 1.5. In Section 6, we will prove Theorem 1.6.
In this section, we first give some qualitative properties of R0, then we deal with the stability of DFE, and we can finish the proofs of Theorem 1.1 and Theorem 1.2.
By the definition of R0, there exits some positive function Φ(x)∈C2([0,L]) such that
{−[dIΦx−a′(x)Φ]x+γ(x)Φ=1R0β(x)Φ,0<x<L,dIΦx(0)−a′(0)Φ(0)=0,dIΦx(L)−a′(L)Φ(L)=0. | (2.1) |
Letting φ(x)=e−a(x)dIΦ(x), we have
{−dIφxx−a′(x)φx+γ(x)φ=1R0β(x)φ,0<x<L,φx(0)=0,φx(L)=0. | (2.2) |
Linearizing (1.1) around (ˆS,0) and letting ˉξ(x,t)=S(x,t)−ˆS(x,t), ˉη(x,t)=I(x,t), we have
{ˉξt=(dSˉξx−a′(x)ˉξ)x−[β(x)−γ(x)]ˉη,0<x<L, t>0,ˉηt=(dIˉηx−a′(x)ˉη)x+[β(x)−γ(x)]ˉη,0<x<L, t>0. |
For the linear system, seeking for the solution which is separation of variables, i.e., ˉξ(x,t)=e−λtξ(x) and ˉη(x,t)=e−λtη(x), we have
{(dSξx−a′(x)ξ)x−[β(x)−γ(x)]η+λξ=0,0<x<L,(dIηx−a′(x)η)x+[β(x)−γ(x)]η+λη=0,0<x<L, | (2.3) |
subject to boundary conditions
{dSξx(0)−a′(0)ξ(0)=0,dSξx(L)−a′(L)ξ(L)=0,dSηx(0)−a′(0)η(0)=0,dSηx(L)−a′(L)η(L)=0. | (2.4) |
By the conservation of total population, we need to impose that
∫L0[ξ(x)+η(x)]dx=0. | (2.5) |
Noticing that the second equation of (2.3) is independent of ξ, letting ζ(x)=e−a(x)dIη(x), we only need to consider the following eigenvalue problem
{dIζxx+a′(x)ζx+[β(x)−γ(x)]ζ(x)+λζ(x)=0,0<x<L,ζx(0)=ζx(L)=0. | (2.6) |
By the results of [19], all the eigenvalues are real, the smallest eigenvalue λ1(dI,q) is simple, and its corresponding eigenfunction ϕ1 can be chosen positive.
We will show a fact below.
Lemma 2.1.1. For any dI and q>minx∈[0,L]a′(x)>0, λ1(dI,q)<0 if R0>1, λ1(dI,q)=0 if R0=1 and λ1(dI,q)>0 if R0<1.
Proof. Note that (λ1(dI,q),ϕ1) satisfies
{−dI(ϕ1)xx−a′(x)(ϕ1)x+[γ(x)−β(x)]ϕ1(x)=λ1(dI,q)ϕ1(x), 0<x<L,(ϕ1)x(0)=(ϕ1)x(L)=0. | (2.7) |
Multiplying (2.1) by ea(x)dIϕ1 and (2.7) by ea(x)dIΦ, integrating by parts in (0,L), and subtracting the resulting equations, we get
∫L0(1R0−1)β(x)Φ(x)ϕ1(x)dx=∫L0λ1(dI,q)Φ(x)ϕ1(x)dx. |
Using the mean value theorem of integrating, we have
(1R0−1)β(x1)Φ(x1)ϕ1(x1)=λ1(dI,q)Φ(x2)ϕ1(x2) |
for some 0≤x1≤L and 0≤x2≤L. Using β(x1)Φ(x1)ϕ1(x1)>0 and Φ(x2)ϕ1(x2)>0, we know that
(1R0−1)has the same sign ofλ1(dI,q), |
which implies the conclusions are true.
Lemma 2.1.2. If dIq→0 and dIq2→0, ˜a′(x)>δ>0 for some constant δ, then R0→β(L)γ(L).
Proof. Let w(x)=e−qdIA˜a(x)Φ(x), where Φ(x) is the solution of (2.1), A is a constant which will be chosen later. It is easy to verify that w satisfies
{ [q2A(A−1)dI(˜a′(x))2+q(A−1)˜a″(x)+1R0β(x)−γ(x)]w=−dIwxx+(1−2A)a′(x)wx,0<x<L, t>0,dIwx(0)=a′(0)(1−A)w(0),dIwx(L)=a′(L)(1−A)w(L). | (2.8) |
First we chose A=1+C1dIq2, where C1 is a positive constant to be chosen later. Then (2.8) becomes
{ [C1(1+C1dIq2)(˜a′(x))2+q(1+C1dIq2)˜a″(x)+1R0β(x)−γ(x)]w=−dIwxx−(1+2C1dIq2)a′(x)wx,0<x<L, t>0,dIwx(0)=−C1dIq˜a′(0)w(0),dIwx(L)=−C1dIq˜a′(L)w(L). |
Assume that w(x∗)=minx∈[0,L]w(x). We will show that x∗=L below. wx(0)<0 implies that x∗≠0. If x∗∈(0,L), then wxx(x∗)≥0 and wx(x∗)=0, (2.9) means that
[C1(1+C1dIq2)(˜a′(x∗))2+q(1+C1dIq2)˜a″(x∗)+1R0β(x∗)−γ(x∗)]≤0 |
Taking C1=Kq with K large enough, we can get a contradiction. Therefore, x∗=L and w(x)≥w(L) for x∈[0,L], which implies that
Φ(x)Φ(L)≥e−qdI(1+C1dIq2)[˜a(L)−˜a(x)]. | (2.9) |
Next, we chose A=1−C2dIq2, where C2 is a positive constant to be chosen later. Then (2.8) becomes
{ [C2(1−C2dIq2)(˜a′(x))2+q(1−C2dIq2)˜a″(x)+1R0β(x)−γ(x)]w=−dIwxx−(1−2C2dIq2)a′(x)wx,0<x<L, t>0,dIwx(0)=C2dIq˜a′(0)w(0),dIwx(L)=C2dIq˜a′(L)w(L). |
Assume that w(x∗)=maxx∈[0,L]w(x). We will show that x∗=L below. wx(0)>0 implies that x∗≠0. If x∗∈(0,L), then wxx(x∗)≥0 and wx(x∗)=0, (2.10) means that
[C2(1−C2dIq2)(˜a′(x∗))2+q(1−C2dIq2)˜a″(x∗)+1R0β(x∗)−γ(x∗)]≤0 |
Taking C2=K′q with K′ large enough, we can get a contradiction. Therefore, x∗=L and w(x)≤w(L) for x∈[0,L], which implies that
Φ(x)Φ(L)≤e−qdI(1−C2dIq2)[˜a(L)−˜a(x)]. | (2.10) |
Dividing (2.1) by Φ(L) and integrating the result in (0,L), we have
∫L0γ(x)Φ(x)Φ(L)dx=1R0∫L0β(x)Φ(x)Φ(L)dx. | (2.11) |
Letting y=q[˜a(L)−˜a(x)]dI, i.e., x=˜a−1[˜a(L)−dIyq], we have
e−(1+C1dIq)y≤Φ(˜a−1[˜a(L)−dIyq])Φ(L)≤e−(1−C2dIq)y | (2.12) |
and
∫q[˜a(L)−˜a(0)]dI0γ(˜a−1[˜a(L)−dIyq])Φ(˜a−1[˜a(L)−dIyq])˜a′(˜a−1[˜a(L)−dIyq])Φ(L)dy=1R0∫q[˜a(L)−˜a(0)]dI0β(˜a−1[˜a(L)−dIyq])Φ(˜a−1[˜a(L)−dIyq])˜a′(˜a−1[˜a(L)−dIyq])Φ(L)dy. | (2.13) |
Using (2.12), by Lebesgue dominant convergence theorem, then passing to the limit in (2.13), we get
limdI/q→0,dI/q2→0R0=limdI/q→0,dI/q2→0∫q[˜a(L)−˜a(0)]dI0β(˜a−1[˜a(L)−dIyq])Φ(˜a−1[˜a(L)−dIyq])˜a′(˜a−1[˜a(L)−dIyq])Φ(L)dy∫q[˜a(L)−˜a(0)]dI0γ(˜a−1[˜a(L)−dIyq])Φ(˜a−1[˜a(L)−dIyq])˜a′(˜a−1[˜a(L)−dIyq])Φ(L)dy=∫∞0β(L)˜a′(L)e−ydy∫∞0γ(L)˜a′(L)e−ydy=β(L)γ(L). | (2.14) |
We have the following corollary.
Corollary 2.1.1. The following statements hold.
(i) Given dI>0, R0→ˆR0 as q→0;
(ii) Given dI>0, R0→β(L)γ(L) as q→+∞;
(iii) Given q>0, R0→β(L)γ(L) as dI→0;
(iv) Given q>0, R0→∫L0β(x)dx∫L0γ(x)dx as dI→+∞.
Proof. (i) For any fixed φ∈H1((0,L)), φ≠0, we have
limq→0dI∫L0ea(x)dIφ2xdx+∫L0γ(x)ea(x)dIφ2dx∫L0β(x)ea(x)dIφ2dx=dI∫L0φ2xdx+∫L0γ(x)φ2dx∫L0β(x)φ2dx. |
Taking infφ∈H1((0,L)),φ≠0 both sides, we have 1R0→1ˆR0 as q→0.
(ii) and (iii) are the direct conclusions of Lemma 2.2.
(iv) By the definition of 1R0, for φ≡1, we have
1R0≤∫L0γ(x)ea(x)dIdx∫L0β(x)ea(x)dIdx≤maxx∈[0,L]γ(x)minx∈[0,L]β(x), |
which implies that 1R0 is uniformly bounded for dI>0, passing to a subsequence if necessary, it has a finite limit 1ˉR0 as dI→∞.
On the other hand, by the standard elliptic regularity and the Sobolev embedding theorem, Φ is uniformly bounded for all dI≥1. Dividing both sides of (2.1) by dI and letting dI→+∞, we have Φxx→0 for x∈(0,L) and Φx(0)→0, Φx(L)→0. Consequently, there exists a positive constant ˉΦ such that Φ(x)→ˉΦ as dI→+∞. Integrating (2.1) by parts over (0,L), we can get
qdI∫L0e−a(x)dI[dIΦx−a′(x)Φ(x)]dx+∫L0e−a(x)dIγ(x)Φ(x)dx=1R0∫L0e−a(x)dIβ(x)Φ(x)dx. |
Letting dI→+∞, we obtain ˉR0=∫L0β(x)dx∫L0γ(x)dx.
Lemma 2.1.3. The following statements hold.
(i) If β(x)>γ(x) on [0,L], then R0>1 for any dI>0 and q>minx∈[0,L]a′(x)>0;
(i) If β(x)<γ(x) on [0,L], then R0<1 for any dI>0 and q>minx∈[0,L]a′(x)>0.
Proof. (i) If β(x)>γ(x) on [0,L], by the definition of 1R0, for φ≡1, we have
1R0≤∫L0γ(x)ea(x)dIdx∫L0β(x)ea(x)dIdx<1, |
i.e., R0>1.
(ii) Subtracting both sides of (2.2) by β(x)φ, multiplying by ea(x)dIφ, we have
−dIφxxea(x)dIφ−a′(x)φxea(x)dIφ+[γ(x)−β(x)]ea(x)dIφ2=(1R0−1)β(x)ea(x)dIφ2. |
Integrating it by parts over (0,L), using φx(0)=φx(L)=0, we obtain
dI∫L0ea(x)dI(φx)2dx+∫L0[γ(x)−β(x)]ea(x)dIφ2dx=(1R0−1)∫L0β(x)ea(x)dIφ2dx. |
Since β(x)<γ(x) on [0,L], the left side of the above equality is positive, and
(1R0−1)∫L0β(x)ea(x)dIφ2dx>0, |
which implies that R0<1.
Proof. Theorem 1.1 is the direct results of Lemma 2.1.2, Corollary 2.1.1 and Lemma 2.1.3.
Next we will consider the stability of DFE.
Lemma 2.1.4. The DFE is stable if R0<1, while it is unstable if R0>1.
Proof. 1. Assume contradictorily the DFE is unstable if R0<1. Then we can find (λ,ξ,η) which is a solution of (2.3)–(2.4) subject to (2.5), with at least one of ξ and η is not identical zero, and ℜ(λ)≤0. Suppose that η≡0, then ξ≢0 on [0,L]. Using (2.3)–(2.4), we have
{−(dSξx−a′(x)ξ)x=λξ,0<x<L,dSξx(0)−a′(0)ξ(0)=0,dSξx(L)−a′(L)ξ(L)=0. | (2.15) |
It is easy to see that λ is real and nonnegative, and therefore λ=0. We find that ξ=ξ0eqdI˜a(x), where ξ0 is some constant to be determined later. By (1.2), we impose that ∫L0[ξ(x)+η(x)]dx=0, ξ0=0, i.e., ξ≡0 on [0,L]. This is a contradiction. Then we conclude that η≡0 on [0,L]. From (2.6), λ must be real and λ≤0. Since λ1(dI,q) is the principal eigenvalue, then λ1(dI,q)≤λ≤0. Lemma 2.1 implies that R0≥1, which is a contradiction. Then we conclude that if (λ,ξ,η) is a solution of (2.3)–(2.4), with at least one of ξ and η not identical zero on [0,L], then ℜ(λ)>0. This proves the linear stability of the DFE.
2. Suppose that R0>1. Since (λ1(dI,q),ϕ1) is the principal eigen-pair of (2.6), (λ1(dI,q),ea(x)dIϕ1) satisfies
{[dI(ϕ1)x−a′(x)ϕ1]x+[β(x)−γ(x)]ϕ1+λ1(dI,q)ϕ1=0,0<x<L,dI(ϕ1)x−a′(x)ϕ1=0,x=0, L. |
By the result of Lemma 2.1.1, λ1(dI,q)<0. On the other hand,
{(dSξx−a′(x)ξ)x+λξ=[β(x)−γ(x)]ea(x)dIϕ1,0<x<L,dSξx(0)−a′(0)ξ(0)=0,dSξx(L)−a′(L)ξ(L)=0. | (2.16) |
There exists a unique solution ξ1 of (2.16). And (2.5) becomes
∫L0[ξ1(x)+ea(x)dIϕ1(x)]dx=0, |
which implies that (2.3)–(2.4) has a solution (λ1(dI,q),ξ1,ea(x)dIϕ1(x)) satisfying λ1(dI,q)<0 and ea(x)dIϕ1(x)>0 in (0,L). Therefore, the DFE is linearly unstable.
Lemma 2.1.5. If R0<1, then (S,I)→(ˆS,0) in C([0,L]) as t→+∞.
Proof. If R0<1, letting u(x,t)=Me−λ1(dI,q)tea(x)dIϕ1(x), then we have
{ut=[dIux−a′(x)u]x+[β(x)−γ(x)]u,0<x<L,t>0,dIux(0,t)−a′(0)u(0,t)=0,dIux(L,t)−a′(L)u(L,t)=0, t>0. |
Here (λ1(dI,q),ϕ1) is the principal eigen-pair, λ1(dI,q)>0 and ϕ1(x)>0 on [0,L]. M is large enough such that I(x,0)≤u(x,0) for every x∈(0,L). Noticing that
{It=[dIIx−a′(x)I]x+[β(x)−γ(x)]I,0<x<L,t>0,dIux(0,t)−a′(0)u(0,t)=0,dIux(L,t)−a′(L)u(L,t)=0, t>0. |
By the comparison principle, we have I(x,t)≤u(x,t) for every x∈(0,L) and t≥0. Obviously, u(x,t)→0 for every x∈(0,L) as t→∞, which implies that I(x,t)→0 for every x∈(0,L) as t→∞.
Now we will show that S→ˆS as t→+∞. Since
St=(dSSx−a′(x)S)x−β(x)SIS+I+γ(x)I, 0<x<L, t>0, |
we have
|St−(dSSx−a′(x)S)x|≤(‖β‖∞+‖γ‖∞)I≤Ce−λ1(dI,q)t, |
for 0<x<L, t>0. Noticing that
limt→+∞e−λ1(dI,q)t→0 |
as t→+∞, we know that there exists a positive function ˜S(x) such that
limt→+∞S(x,t)=˜S(x),∫L0˜S(x)dx=N. |
Therefore, limt→+∞S(x,t)=˜S(x)=ˆS(x).
Proof. Theorem 1.2 is the direct results of Lemma 2.1.4 and Lemma 2.1.5.
In this section, we will study further properties of R0 in the case of β(x)−γ(x) changing sign once.
Lemma 2.2.1. Assume that ϕ1 is a positive eigenfunction corresponding to R0=1, β(x)−γ(x) changes sign once in (0,L). If assumption (C1)(or (C2)) holds, then (ϕ1)x>0(or (ϕ1)x<0) in (0,L).
Proof. If β(x)−γ(x) changes sign once in (0,L) and assumption (C1) holds, then there exists some x0∈(0,L) such that β(x)−γ(x)<0 in (0,x0), β(x0)=γ(x0) and β(x)−γ(x)>0 in (x0,L).
By the definition of ϕ1, we have
{−dI(ϕ1)xx−a′(x)(ϕ1)x=[β(x)−γ(x)]ϕ1,0<x<L,(ϕ1)x(0)=(ϕ1)x(L)=0. | (2.17) |
Multiplying (2.17) by ea(x)dI, we obtain
−dI(ea(x)dI(ϕ1)x)x=[β(x)−γ(x)]ea(x)dIϕ1. |
Under the assumptions on β(x) and γ(x), we can obtain (ea(x)dI(ϕ1)x)x>0 in (0,x0), (ea(x)dI(ϕ1)x)x=0 at x0 and (ea(x)dI(ϕ1)x)x<0 in (x0,L). That is, ea(x)dI(ϕ1)x is strictly increasing in (0,x0) and strictly decreasing in (x0,L). Noticing that (ϕ1)x(0)=(ϕ1)x(L)=0, we can get ea(x)dI(ϕ1)x>0 in (0,L). So (ϕ1)x>0 in (0,L).
Similarly, if β(x)−γ(x) changes sign once in (0,L) and assumption (C2) holds, (ϕ1)x<0 in (0,L). We omit the details here.
Now we prove two general lemmas below.
For any continuous function m(x) on [0,L], define
F(η)=∫L0˜a′(x)eη˜a(x)m(x)dx,0≤η<∞. |
Lemma 2.2.2. Assume that m(x)∈C1([0,L]) and m(L)>0(or m(L)<0). Then there exists some positive constant M such that F(η)>0(or F(η)<0) for any η>M.
Proof. Since m′(x) and ˜a′(x) is uniformly bounded independent of η, we have
limη→+∞ηe−η˜a(L)F(η)=limη→+∞∫L0η˜a′(x)e−η[˜a(x)−˜a(L)]m(x)dx=m(L)−limη→+∞(m(0)eη[˜a(0)−˜a(L)]+∫L0m′(x)eη[˜a(x)−˜a(L)]dx)=m(L)−limη→+∞(m(0)eη[˜a(0)−˜a(L)]+∫L0m′(x)e˜a′(ξ)[x−L]dx)=m(L)>0(<0). |
Therefore, there exists some positive constant M such that F(η)>0(<0) for η>M.
Lemma 2.2.3. Assume that m(x) changes sign once in (0,L). Then
(i) If m(L)>0 and ∫L0˜a′(x)m(x)dx>0, then F(η)>0 for any η>0;
(ii) If m(L)<0 and ∫L0˜a′(x)m(x)dx<0, then F(η)<0 for any η>0;
(iii) If m(L)>0 and ∫L0˜a′(x)m(x)dx<0, then there exists a unique η1∈(0,+∞) such that F(η1)=0 and F′(η1)>0;
(iv) If m(L)<0 and ∫L0˜a′(x)m(x)dx>0, then there exists a unique η1∈(0,+∞) such that F(η1)=0 and F′(η1)<0.
Proof. We only prove part (i) and part (iii). The proofs of part (ii) and part (iv) are similar.
(i) If m(L)>0 and m(x) changes sign once in (0,L), then there exists x1∈(0,L) such that m(x)<0 for x∈(0,x1) and m(x)>0 for x∈(x1,L). Since ˜a(x) is increasing, we have m(x)[˜a(x)−˜a(x1)]>0 for x∈(0,L) and x≠x1. And
[e−˜a(x1)ηF(η)]′=e−˜a(x1)η[F′(η)−˜a(x1)F(η)]=e−˜a(x1)η∫L0[˜a(x)−˜a(x1)]m(x)˜a′(x)eη˜a(x)dx>0, | (2.18) |
which implies that e−˜a(x1)ηF(η) is strictly increasing in η∈(0,∞), e−˜a(x1)ηF(η)>F(0)=∫L0˜a′(x)m(x)dx>0. Consequently, F(η)>0 for any η>0. Here the prime notation denotes differentiation by η. Part (i) is proved.
(iii) ∫L0˜a′(x)m(x)dx<0 means that F(0)<0, while, by the result of Lemma 2.2.2, m(L)>0 means that F(η)>0 for η>M with M large enough. By continuity, there at least exists a positive root for F(η)=0. But e−˜a(x1)ηF(η) is increasing in η∈(0,∞), so F(η)=0 only has a unique positive root η1. By (2.18), we have F′(η1)>a(x1)F(η1)=0. Part (iii) is proved.
In this section, we consider the stability of DFE. First we have
Lemma 2.3.1. Assume that β(x)−γ(x) changes sign once in (0,L) and ∫L0β(x)dx>∫L0γ(x)dx.
(i) If β(x) and γ(x) satisfy (C1), then R0>1 for dI>0 and q>minx∈[0,L]a′(x)>0;
(ii) If β(x) and γ(x) satisfy (C2), then for every dI>0, there exists a unique ˉq=ˉq(dI) such that R0>1 for 0<minx∈[0,L]a′(x)<q<ˉq, R0=1 for q=ˉq and R0<1 for q>ˉq.
Proof. (i) Subtracting both sides of (2.2) by β(x)φ, multiplying by ea(x)dIφ, we have
[−dIφxx−a′(x)φx]ea(x)dIφ+[γ(x)−β(x)]ea(x)dI=(1R0−1)β(x)ea(x)dI. |
Integrating it by parts over (0,L), using φx(0)=φx(L)=0, we obtain
dI∫L0ea(x)dI(φx)2φ2dx+∫L0[β(x)−γ(x)]ea(x)dIdx=(1−1R0)∫L0β(x)ea(x)dIdx. |
Using Lemma 2.2.3(i) with m(x)=[β(x)−γ(x)]˜a′(x), ∫L0[β(x)−γ(x)]ea(x)dIdx>0, and
(1−1R0)∫L0β(x)ea(x)dIφ2dx>0, |
which implies that R0>1.
(ii) Differentiating both sides of (2.2) with respect to q, denoting the differentiation with respect to q by the dot notation, we obtain
{−dI˙φxx−˜a′(x)φx−˜a′(x)˙φx+γ(x)˙φ=−˙R0R20β(x)φ+1R0β(x)˙φ,0<x<L,˙φx(0)=˙φx(L)=0. | (2.19) |
Multiplying (2.19) by ea(x)dIφ and integrating the resulting equation in (0,L), we have
dI∫L0ea(x)dI˙φxφxdx−∫L0ea(x)dIφxφ˜a′(x)dx+∫L0γ(x)ea(x)dI˙φφdx=−˙R0R20∫L0β(x)ea(x)dIφ2dx+1R0∫L0β(x)ea(x)dI˙φφdx. | (2.20) |
Multiplying (2.2) by ea(x)dI˙φ and integrating the resulting equation in (0,L), we get
dI∫L0ea(x)dI˙φxφxdx+∫L0γ(x)ea(x)dI˙φφdx=1R0∫L0β(x)ea(x)dI˙φφdx. | (2.21) |
Subtracting (2.20) and (2.21), we obtain
∂R0∂q=R20∫L0ea(x)dIφxφ˜a′(x)dx∫L0β(x)ea(x)dIφ2dx. | (2.22) |
By the result of Corollary 2.1.1, we know that
limq→∞R0=β(L)γ(L)<1. |
Meanwhile, we have
limq→0R0=ˆR0>1 |
for any dI. Then there must exist at least some ˉq such that R0(ˉq)=1. By Lemma 2.1.1, for any ˉq>0 satisfying R0(ˉq)=1, (ϕ1)x<0 in (0,L). Recalling (2.22), we have
∂R0∂ˉq=∫L0eˉqdI˜a(x)(ϕ1)xϕ1dx∫L0β(x)eˉqdI˜a(x)(ϕ1)2dx<0, |
which implies that ˉq is the unique point satisfying R0(ˉq)=1.
The following lemma will tell us that there exists a function q=ρ1(dI) such that R0(dI,ρ1(dI))=1 and give the asymptotic profile of ρ1(dI) if ∫L0β(x)dx>∫L0γ(x)dx.
Lemma 2.3.2. Assume that β(x)−γ(x) changes sign once in (0,L), ∫L0β(x)dx>∫L0γ(x)dx, and θ1 is the unique solution of
∫L0[β(x)−γ(x)]eθ1˜a(x)dx=0. |
Suppose that β(x) and γ(x) satisfy (C2). Then there exists a function ρ1:(0,∞)→(0,∞) such that R0(dI,ρ1(dI))=1. And ρ1 satisfies
limdI→0ρ1(dI)=0,limdI→∞ρ1(dI)dI=θ1. |
Proof. 1. Let's first consider the limit of ρ1(dI)dI as dI→∞. Assume that ρ1(dI)dI→∞ as dI→∞. Under the assumption (C2), by Lemma 2.1.4, we have
limρ1(dI)→∞,ρ1(dI)dI→∞R0(dI,ρ1(dI))=β(L)γ(L)<1, |
which is a contradiction to R0(dI,ρ1(dI))=1.
Next, we will prove that ρ1(dI)dI→θ1 as dI→∞. Here θ1 is the unique positive root of ∫L0[β(x)−γ(x)]eθ1˜a(x)dx=0. By the discussions above, we know that ρ1(dI)dI is bounded for large dI. Passing to a subsequence if necessary, we suppose that ρ1(dI)dI→θ∗ for some nonnegative number θ∗ as dI→∞. Let ˜φ be the unique normalized eigenfunction of the eigenvalue R0(dI,ρ1(dI))=1. Then
{−dI(eρ1(dI)dI˜a(x)˜φx)x+[γ(x)−β(x)]eρ1(dI)dI˜a(x)˜φ=0,0<x<L,˜φx(0)=˜φx(L)=0. | (2.23) |
Integrating (2.23) in (0,L), we get
∫L0[β(x)−γ(x)]eρ1(dI)dI˜a(x)˜φdx=0. | (2.24) |
Recalling that, up to a subsequence if necessary, ˜φ→1 in C([0,1]) as dI→∞. Letting dI→∞ in (2.24), we have
∫L0[β(x)−γ(x)]eθ∗˜a(x)dx=0. |
By Lemma 2.2.3 with m(x)=[β(x)−γ(x)]˜a′(x), F(η) has a unique positive root, i.e., θ∗=θ1.
2. Contradictorily, assume that q=ρ1(dI)→q∗>0 or q=ρ1(dI)→∞ as dI→0. By Lemma 2.1.4, we know that
limρ1(dI)→q∗,ρ1(dI)dI→∞R0(dI,ρ1(dI))=β(L)γ(L)<1 |
or
limρ1(dI)→∞,ρ1(dI)dI→∞R0(dI,ρ1(dI))=β(L)γ(L)<1, |
which is a contradiction to R0(dI,ρ1(dI))=1. Therefore, we have limdI→0ρ1(dI)=0.
To study the properties of R0 when ∫L0β(x)dx<∫L0γ(x)dx, we need the following results which were stated in [2]:
Proposition 2.3.1. Assume that β(x)−γ(x) changes sign in (0,L).
(i) ˆR0 is a monotone decreasing function of dI with ˆR0→max{β(x)/γ(x):x∈[0,L]} as dI→0 and ˆR0→∫L0β(x)dx/∫L0γ(x)dx as dI→+∞;
(ii) ˆR0>1 for all dI>0 if ∫L0β(x)dx≥∫L0γ(x)dx;
(iii) There exists a threshold value d∗I∈(0,+∞) such that ˆR0>1 for dI<d∗I and ˆR0<1 for dI>d∗I if ∫L0β(x)dx<∫L0γ(x)dx.
Lemma 2.3.3. Assume that β(x)−γ(x) changes sign once in (0,L) and ∫L0β(x)dx<∫L0γ(x)dx. Then there exists some constant d∗I>0 such that d∗I is the unique positive root of the equation ^R0(dI)=1 and the following statements hold.
1. If β(x) and γ(x) satisfy (C1), then
(i) for dI∈(0,d∗I], R0>1 for any q>minx∈[0,L]a′(x)>0;
(ii) for dI∈(d∗I,∞), there exists a unique ˉq=ˉq(dI) such that R0<1 for any 0<minx∈[0,L]a′(x)<q<ˉq and R0>1 for any q>ˉq.
2. If β(x) and γ(x) satisfy (C2), then
(iii) for dI∈(0,d∗I], there exists a unique ˉq=ˉq(dI) such that R0>1 for any 0<minx∈[0,L]a′(x)<q<ˉq and R0<1 for any q>ˉq;
(iv) for dI∈(d∗I,∞), R0<1 for any q>minx∈[0,L]a′(x)>0.
Proof. (i) Noticing that β(x) and γ(x) satisfy (C1), similar to the proof of (ii) in Lemma 2.1.4, we can prove that there exists a unique ˉq>0 satisfying R0(ˉq)=1 and R′0(ˉq)>0. Hence, the conclusion is true for dI∈(d∗I,+∞).
For dI∈(0,d∗I], by the results of Proposition 2.3.1, we have limq→0R0=ˆR0≥1. By the results of Corollary 2.1.1, limq→+∞R0=β(L)/γ(L)>1 under the condition (C1). Hence R0>1 for any q>0.
(ii) The proof of Lemma 2.3.3 under the condition (C2) is similar to that of Lemma 2.1.4, we omit the details here.
Lemma 2.3.4. Assume that β(x)−γ(x) changes sign once in (0,L) and ∫L0β(x)dx<∫L0γ(x)dx. Then there exists a constant d∗I>0 such that d∗I is the unique positive root of the equation ^R0(dI)=1 and the following statements hold.
1. If β(x) and γ(x) satisfy (C1), then there exists a function ρ2:(d∗I,∞)→(0,∞) such that ρ2 is a monotone increasing function of dI and R0(dI,ρ2(dI))=1. Let θ2 be the unique solution of
∫L0[β(x)−γ(x)]eθ2˜a(x)dx=0. |
Then
limdI→d∗I+ρ2(dI)=0,limdI→∞ρ2(dI)dI=θ2. |
2. If β(x) and γ(x) satisfy (C2), then there exists a function ρ3:(0,d∗I)→(0,∞) such that R0(dI,ρ3(dI))=1 and
limdI→0+ρ3(dI)=0,limdI→d∗I−ρ3(dI)dI=0. |
Proof. 1. If we can prove that ρ′2(dI)>0 for dI∈(d∗I,∞), then ρ2(dI) is a monotone increasing function of dI. Here the prime notation denotes differentiation by dI. Since R0(dI,ρ2(dI))=1, we can get
∂R0∂qρ′2(dI)+∂R0∂dI=0. | (2.25) |
By Lemma 2.3.1, ∂R0∂q>0 for R0(dI,ρ2(dI))=1. So we need to prove that ∂R0∂dI<0.
Differentiating both sides of (2.2) with respect to dI, denoting the differentiation with respect to dI by the dot notation, we obtain
{−φxx−dI˙φxx−a′(x)˙φx+γ(x)˙φ=−˙R0R20β(x)φ+1R0β(x)˙φ,0<x<L,˙φx(0)=˙φx(L)=0. | (2.26) |
Multiplying (2.26) by ea(x)dIφ and integrating the resulting equation in (0,L), we obtain
−∫L0ea(x)dIφxxφdx+dI∫L0ea(x)dI˙φxφxdx+∫L0γ(x)ea(x)dI˙φφdx=−˙R0R20∫L0β(x)ea(x)dIφ2dx+1R0∫L0β(x)ea(x)dI˙φφdx. | (2.27) |
Multiplying (2.2) by ea(x)dI˙φ and integrating the resulting equation in (0,L), we get
dI∫L0ea(x)dI˙φxφxdx+∫L0γ(x)ea(x)dI˙φφdx=1R0∫L0β(x)ea(x)dI˙φφdx. | (2.28) |
Subtracting (2.27) and (2.28), we have
∂R0∂dI=R20∫L0ea(x)dIφxxφdx∫L0β(x)ea(x)dIφ2dx=−R20∫L0ea(x)dI(φx)2dx∫L0β(x)ea(x)dIφ2dx−R20∫L0ea(x)dIφxφa′(x)dxdI∫L0β(x)ea(x)dIφ2dx. | (2.29) |
By Lemma 2.2.1, for any dI satisfying R0(dI,q)=1, (ϕ1)x>0, we can get
∂R0∂dI=−R20∫L0ea(x)dI[(ϕ1)x]2dx∫L0β(x)ea(x)dIϕ21dx−R20∫L0ea(x)dI(ϕ1)xϕ1a′(x)dxdI∫L0β(x)ea(x)dIϕ21dx<0. | (2.30) |
(2.25) and (2.30) imply that ρ′2(dI)>0 for dI∈(d∗I,∞).
The proof of limdI→∞ρ2(dI)dI=θ2(θ2 is the unique solution of ∫L0[β(x)−γ(x)]eθ2a(x)dx=0) is similar to the proof of Lemma 2.3.2, we omit the details here.
Now we will prove that limdI→d∗I+ρ2(dI)=0. Assume that there exists q∗ such that q=ρ2(dI)→q∗ as dI→d∗I+. Then there exists a positive function ϕ∗(x)∈C2([0,L]) such that
{−d∗Iϕ∗xx−q∗˜a′(x)ϕ∗x+γ(x)ϕ∗=β(x)ϕ∗,0<x<L,ϕ∗x(0)=ϕ∗x(L)=0. | (2.31) |
Noticing that d∗I is the unique positive root of ˆR0=1 and the definition of ˆR0 implies q=0, there exists a positive function ˆϕ(x)∈C2([0,L]) such that
{−d∗Iˆϕxx+γ(x)ˆϕ=β(x)ˆϕ,0<x<L,ˆϕx(0)=ˆϕx(L)=0. | (2.32) |
Multiplying (2.31) by ˆϕ, (2.32) by ϕ∗, subtracting the two resulting equations, then integrating by parts over (0,L), we get
q∗∫L0˜a′(x)ϕ∗xˆϕdx=0. |
Since ϕ∗x is positive(by Lemma 2.2.1), we have q∗=0. Therefore, limdI→d∗I+ρ2(dI)=0.
2. Using the arguments above, similar to the proof of Lemma 2.3.2, we can obtain the conclusions.
In this section, we will show that: If the disease-free equilibrium is unstable, then we can use the bifurcation analysis and degree theory to study the existence of endemic equilibrium.
Letting ˜S=ea(x)dSˉS, ˜I=ea(x)dIˉI, we have
{dSˉSxx+a′(x)ˉSx−β(x)ea(x)dIˉSˉIea(x)dSˉS+ea(x)dIˉI+γ(x)e(1dI−1dS)a(x)ˉI=0, 0<x<L,dIˉIxx+a′(x)ˉIx+β(x)ea(x)dSˉSˉIea(x)dSˉS+ea(x)dIˉI−γ(x)ˉI=0, 0<x<L,ˉSx(0)=ˉSx(L)=0,ˉIx(0)=ˉIx(L)=0, ∫L0[ea(x)dSˉS+ea(x)dIˉI]dx=N. | (2.33) |
Since the structure of the solution set of (2.33) is the same as that of (1.3), we study (2.33) instead of (1.3). Denote the unique disease-free equilibrium of (2.33) by (ˆˉS,0)=(N∫L0ea(x)dS,0). We will consider a branch of positive solutions of (2.33) bifurcating from the branch of semi-trivial solutions given by
ΓS:={(q,(ˆˉS,0)):0<minx∈[0,L]a′(x)<q<∞} |
through using the local and global bifurcation theorems. For fixed dS, dI>0, we take q as the bifurcation parameter. Let
X={u∈W2,p((0,L)):ux(0)=ux(L)=0},Y=Lp((0,L)) |
for p>1 and the set of positive solution of (2.33) to be
O={(q,(S,I))∈R+×X×X:q>minx∈[0,L]a′(x)>0,S>0,I>0,(q,(S,I)) satisfies (2.33)}. |
Lemma 2.4.1 Assume that dS, dI>0 and β(x)−γ(x) changes sign once in (0,L). Then
1. q∗>0 is a bifurcation point for the positive solutions of (2.33) from the semi-trivial branch ΓS if and only if q∗ satisfies R0(dI,q∗)=1. That is,
(I) If ∫L0β(x)dx>∫L0γ(x)dx, then such q∗ exists uniquely for any dI>0 if and only if assumption (C2) holds;
(II) If ∫L0β(x)dx<∫L0γ(x)dx, let d∗I be the unique positive root of ˆR0=1, then such q∗ exists uniquely for any dI>0 if and only if either β(x) and γ(x) satisfy condition (C1) and d>d∗I or they satisfy condition (C2) and 0<d<d∗I.
2. There exits some δ>0 such that all positive solutions of (2.33) near (q∗,(ˆˉS,0)))∈R×X×X can be parameterized as
Γ={(q(τ),(ˆˉS+ˉS1(τ),ˉI1(τ))):τ∈[0,δ)}, | (2.34) |
where (q(τ),(ˆˉS+ˉS1(τ),I1(τ))) is a smooth curve with respect to τ and satisfies q(0)=q∗, ˆS1(0)=I1(0)=0.
3. There exists a connected component Σ of ˉO satisfying Γ⊆Σ, and Σ possesses some properties as follows.
Case (I) Assume that ∫L0β(x)dx>∫L0γ(x)dx and (C2) holds. Then there exists some endemic equilibrium (ˆS∗,ˆI∗) of (2.33) when q=0 such that for Σ, the projection of Σ to the q-axis satisfies ProjqΣ=[0,q∗] and the connected component Σ connects to (0,(ˆS∗,ˆI∗)).
Case (II) Assume that ∫L0β(x)dx<∫L0γ(x)dx. Then
(i) If (C1) holds and dI>d∗I, then (2.33) has no positive solution for 0<minx∈[0,L]a′(x)<q<q∗ and for Σ, the projection of Σ to the q-axis satisfies ProjqΣ=[q∗,∞).
(ii) If (C2) holds and 0<dI<d∗I, then there exists some endemic equilibrium (ˆS∗,ˆI∗) of (2.33) when q=0 such that for Σ, the projection of Σ to the q-axis satisfies ProjqΣ=[0,q∗] and the connected component Σ connects to (0,(ˆS∗,ˆI∗)).
Proof. 1. Let F:R+×X×X→Y×Y×R be the mapping as follows.
F(q,(ˉS,ˉI))=(dSˉSxx+a′(x)ˉSx−β(x)ea(x)dIˉSˉIea(x)dSˉS+ea(x)dIˉI+γ(x)e(1dI−1dS)a(x)ˉIdIˉIxx+a′(x)ˉIx+β(x)ea(x)dSˉSˉIea(x)dSˉS+ea(x)dIˉI−γ(x)ˉI∫L0[ea(x)dSˉS+ea(x)dIˉI]dx−N). |
It is to verify that the pair (ˉS,ˉI) is a solution of (2.33) if only if F(q,(ˉS,ˉI))=0. Obviously, F(q,(ˆˉS,0))=0 for any q>minx∈[0,L]a′(x)>0. The Frˊechet derivatives of F at (ˆˉS,0) are given by
D(ˉS,ˉI)F(q,(ˆˉS,0))[ΦΨ]=(dSΦxx+˜a′(x)Φx+[γ(x)−β(x)]e(1dI−1dS)a(x)ΨdIΨxx+˜a′(x)Ψx+[β(x)−γ(x)]Ψ∫L0[ea(x)dSΦ+ea(x)dIΨ]dx), |
Dq,(ˉS,ˉI)F(q,(ˆˉS,0))[ΦΨ]=(˜a′(x)Φx+(a(x)dI−a(x)dS)[γ(x)−β(x)]e(1dI−1dS)a(x)Ψ˜a′(x)Ψx∫L0[a(x)dSea(x)dSΦ+a(x)dIea(x)dIΨ]dx), |
D(ˉS,ˉI),(ˉS,ˉI)F(q,(ˆˉS,0))[ΦΨ]2=(2ˆˉSβ(x)e2(qdI−qdS)˜a(x)Ψ2−2ˆˉSβ(x)e(1dI−1dS)a(x)Ψ20). |
If (Φ1,Ψ1) is a nontrivial solution of the following problem
{dSΦxx+˜a′(x)Φx+[γ(x)−β(x)]e(1dI−1dS)a(x)Ψ=0,0<x<L,dIΨxx+˜a′(x)Ψx+[β(x)−γ(x)]Ψ=0,0<x<L,Φx(0)=Φx(L)=Ψx(0)=Ψx(L)=0,∫L0[ea(x)dSΦ+ea(x)dIΨ]dx=0, | (2.35) |
then (q∗,(ˆˉS,0))) is degenerate solution of (2.33). The second equation of (2.33) has a positive solution Ψ1 only if q=q∗ satisfies R0(dI,q∗)=1. And Φ1 satisfies
{dS(Φ1)xx+˜a′(x)(Φ1)x+[γ(x)−β(x)]e(1dI−1dS)a(x)Ψ1=0,0<x<L,(Φ1)x(0)=(Φ1)x(L)=0,∫L0[ea(x)dSΦ1+ea(x)dIΨ1]dx=0, | (2.36) |
Obviously, Φ1 is uniquely determined by Ψ1 in (2.36). Therefore, q=q∗ is the only possible bifurcation point along ΓS where positive solutions of (2.33) bifurcates and such q∗ exists if and only if R0=1. We can obtain the necessary and sufficient conditions for the occurrence of bifurcation by Lemma 2.3.1 and Lemma 2.3.3.
2. At (q,(ˉS,ˉI))=(q∗,(ˆˉS,0)), the kernel
Ker(D(ˉS,ˉI)F(q∗,(ˆˉS,0)))=span{(Φ1,Ψ1)}, |
where (Φ1,Ψ1) is the solution of (2.35) with q=q∗. Up to a multiple of constant, (Φ1,Ψ1) is unique. And the range of D(ˉS,ˉI)F(q∗,(ˆˉS,0)) is given by
Range(D(ˉS,ˉI)F(q∗,(ˆˉS,0)))={(f,g,k)∈Y×Y×RN:∫L0gΨ1ea(x)dIdx=0}, |
and it is co-dimension one. By the result of Lemma 2.1.1, (Ψ1)x keeps one sign in (0,L) and ∫L0(Ψ1)xΨ1ea(x)dIdx≠0, which implies that
Dq,(ˉS,ˉI)F(q∗,(ˆˉS,0))[(Φ1,Ψ1)]∉Range(Dq,(ˉS,ˉI)F(q∗,(ˆˉS,0))). |
Therefore, using the local bifurcation theorem in [20] to F(q,(ˉS,ˉI)) at (q∗,(ˆˉS,0)), we know that the set of positive solutions of (2.33) is a smooth curve
Γ={(q(τ),(ˆˉS+ˉS1(τ),ˉI1(τ))):τ∈[0,δ)} |
satisfying q(0)=q∗, ˉS1(τ)=τˆˉS+o(|τ|) and I1(τ)=o(|τ|). Similar to the procedure in [21] and [22], (also see [23]), we can compute
q′=−<l,D(ˉS,ˉI),(ˉS,ˉI)F(q∗,(ˆˉS,0))[Φ1,Ψ1]2>2<l,Dq,(ˉS,ˉI)F(q∗,(ˆˉS,0))[(Φ1,Ψ1)]=∫L0β(x)e(1dI−1dS)a(x)ϕ31dxˆˉS∫L0ea(x)dIϕ1(ϕ1)xdx. |
Here l is the linear functional on Y×Y×R defined by <l,[f,g,k]>=∫L0gΨ1ea(x)dIdx.
3. By the global bifurcation theorem in [23] and [24], we can get the existence of the connected component Σ. Moreover, Σ is either unbounded, or connects to another (q,(ˆˉS,0)), or Σ connects to another point on the boundary of O.
Case (I) Assume that ∫L0β(x)dx>∫L0γ(x)dx and (C2) holds. By Lemma 2.2.1 and the proof of part 2, we see that there exits a unique q∗ such that the local bifurcation occurs at (q∗,(ˆˉS,0)) and q′(0)<0, which means that the bifurcation direction is subcritical. Therefore, there exists some small δ>0 such that (2.33) has a positive solution if q∗−δ<q<q∗. By Lemma 2.1.4, R0>1 if q∗−δ<q<q∗ for δ>0 small enough. By Lemma 2.1.5, (2.33) has no positive solution if R0<1, which implies that (2.33) has no positive solution if q>q∗. Consequently, the projection of Σ to the q-axis ProjqΣ⊂[0,q∗]. And Σ must be bounded in ˉO because the positive solutions are uniformly bounded in L∞ for 0≤q≤q∗. So the third option must happen here. Hence Σ must connect to (0,(ˉS∗,ˉI∗)), so 0∈ProjqΣ. Here (ˉS∗,ˉI∗) is the unique endemic equilibrium of (2.33) when q=0.
Case (II) Assume that ∫L0β(x)dx<∫L0γ(x)dx.
(i) If (C1) holds and dI>d∗I, by Lemma 2.2.1 and the bifurcation analysis above, there exists unique bifurcation point q∗ satisfying q′(0)>0, which means the bifurcation direction is supercritical. Then there exists some small δ>0 such that (2.33) has a positive solution if q∗<q<q∗+δ. By Lemma 2.3.3, R0>1 if q∗<q<q∗+δ for some δ>0 small enough. By Lemma 2.1.5, (2.33) has no positive solution if R0<1, which implies that (2.33) has no positive solution if 0<q<q∗. So the first option must happen here. If there exists some finite q∗>q∗ such that ProjqΣ=[q∗,q∗), then it contradicts to the fact that all positive solutions are uniformly bounded in L∞ for q=q∗. Consequently, the projection of Σ to the q-axis ProjqΣ=[q∗,∞).
(ii) If (C2) holds and 0<dI<d∗I, the proof is similar to that of Case (I), we omit the details here.
We will give the Leray-Schauder degree argument.
Lemma 2.4.2. For any ϵ>0, there exist two constants C_ and ˉC which depend on dI, ϵ, ‖β‖∞, ‖γ‖∞ and N such that if R0≠1, then for any positive solution of (2.33),
C_≤ˉS(x),ˉI(x)≤ˉCfor any x∈[0,L] | (2.37) |
for any ϵ≤dS≤1ϵ and 0≤q≤1ϵ.
Proof. ∫L0[ea(x)dSˉS+ea(x)dIˉI]dx=N means that ˉS(x) and ˉI(x) are bounded in L1 space. Using the standard theory of elliptic equation, it is easy to see that ˉS and ˉI have the upper bound ˉC depending on dI, ϵ, ‖β‖∞, ‖γ‖∞ and N.
Therefore, we just need to prove that ˉS and ˉI have lower bounds.
Suppose contradictorily that there exist a sequence of {(dS,i,qi)}∞i=1 satisfies ϵ≤dS,i≤1ϵ and 0≤qi≤1ϵ and R0≠1, and {(ˉSi(x),ˉIi(x))}∞i=1 are the corresponding positive solutions of (2.33) satisfying
maxx∈[0,L]Ii(x)→0,as i→∞, |
and (ˉSi(x),ˉIi(x)) satisfies
{dS,i(ˉSi)xx+qi˜a′(x)(ˉSi)x−β(x)eqidI˜a(x)ˉSiˉIieqidS,i˜a(x)ˉSi+eqidI˜a(x)ˉIi+γ(x)e(qidI−qidS,i)˜a(x)ˉIi=0, 0<x<L,dI(ˉIi)xx+qi˜a′(x)(ˉIi)x+β(x)eqidS,i˜a(x)ˉSiˉIieqidS,i˜a(x)ˉSi+eqidI˜a(x)ˉIi−γ(x)ˉIi=0, 0<x<L,(ˉSi)x(0)=(ˉSi)x(L)=0,(ˉIi)x(0)=(ˉIi)x(L)=0, ∫L0[eqidS,i˜a(x)ˉSi+eqidI˜a(x)ˉIi]dx=N. | (2.38) |
Up to a subsequence, we assume that dS,i→dS>0 and qi→q≥0. Note that ‖ˉIi‖∞ are uniformly bounded. Letting ˜ˉIi=ˉIi‖ˉIi‖∞, we have
{dI(˜ˉIi)xx+qi˜a′(x)(˜ˉIi)x+β(x)˜ˉIieqidS,i˜a(x)ˉSieqidS,i˜a(x)ˉSi+eqidI˜a(x)ˉIi−γ(x)˜ˉIi=0, 0<x<L,(˜ˉIi)x(0)=(˜ˉIi)x(L)=0. |
By standard regularity and Sobolev embedding theorem in [25], up to a subsequence, ˉIi→0 in C1([0,L]) and there exists I∗>0 such that ˜ˉIi→I∗ in C1([0,L]) and ‖I∗‖∞=1. Since ˉIi→0 in C1([0,L]) and ∫L0[eqidS,i˜a(x)ˉSi+eqidI˜a(x)ˉIi]dx=N implies that ˉSi is bounded in L1([0,L]), using the equation of ˉSi, we get ˉSi→ˆˉS>0 in C1([0,L]). Letting i→∞ in the equation of ˉIi, we have
{dII∗xx+a′(x)I∗x+[β(x)−γ(x)]I∗=0, 0<x<L,I∗x(0)=I∗x(L)=0. | (2.39) |
Since I∗>0, (2.39) means that 0 is the principle eigenvalue, which is a contradiction of the assumption of R0≠1 for any dI>0 and 0≤q≤1ϵ. Therefore, there must exist some positive constant C_ such that maxx∈[0,L]I(x)≥C_. Similar to the argument in [26], by Harnack inequality, we have
maxx∈[0,L]ˉI(x)≤C∗minx∈[0,L]ˉI(x) |
for some constant C∗ depending on dI, ϵ, ‖β‖∞, ‖γ‖∞ and N, which implies that ˉI(x) has uniformly positive lower bound.
Now we prove that S(x) has a uniform positive lower bound. Let S(x0)=minx∈[0,L]S(x). Using the minimum principle in [27], we have
β(x0)eqdI˜a(x0)ˉS(x0)eqdS˜a(x0)ˉS(x0)+eqdI˜a(x0)ˉI(x0)−γ(x0)e(qdI−qdS)˜a(x0)≥0. |
Consequently,
β(x0)ˉS(x0)ˉI(x0)≥β(x0)eqdI˜a(x0)ˉS(x0)eqdS˜a(x0)ˉS(x0)+eqdI˜a(x0)ˉI(x0)≥γ(x0)e(qdI−qdS)˜a(x0) |
and
ˉS(x0)≥γ(x0)e(qdI−qdS)˜a(x0)ˉI(x0)β(x0)ˉI(x0)≥Cminx∈[0,L]ˉI(x), |
which completes the proof.
Lemma 2.4.3. Assume that β(x)−γ(x) changes sign once in (0,L) and one of the following conditions holds:
(i) dI>0, q>minx∈[0,L]a′(x)>0, ∫L0β(x)dx>∫L0γ(x)dx and (C2) holds;
(ii) 0<dI<d∗I, q>minx∈[0,L]a′(x)>0, ∫L0β(x)dx<∫L0γ(x)dx and (C1) holds.
Then (2.33) has at least an endemic equilibrium.
Proof. Note that we can extend the ranges of f and g properly for any nonnegative pair (f,g)∈C([0,L])×C([0,L]) such that the function fgeτa(x)dSf+eτa(x)dIg is Lipschitz continuous for f,g∈R and τ∈[0,1]. Therefore we define the following compact operator family from C([0,L])×C([0,L]) to C([0,L])×C([0,L]):
{(τdS+(1−τ)dI)uxx+τa′(x)ux+γ(x)e(τdI−τdS)a(x)v=β(x)efgτa(x)dIfeτa(x)dS+geτa(x)dI,0<x<L,dIvxx+τa′(x)vx−γ(x)v=−β(x)fgeτa(x)dSfeτa(x)dS+geτa(x)dI,0<x<L,ux(0)=ux(L)=0,vx(0)=vx(L)=0,∫L0[eτa(x)τdS+(1−τ)dIu+eτa(x)dIv]dx=N. | (2.40) |
Since the operator dId2dx2+τa′(x)ddx−γ(x) is invertible, then for any τ∈[0,1] and (f,g)∈C([0,L])×C([0,L]), by the second equation of (2.40), v is uniquely determined. Substituting this v into the first and last equations of (2.40), u is also uniquely determined. Therefore, we can define Gτ(f,g):=(u,v).
Under conditions (i) and (ii), R0,τ>1 for any τ∈[0,1]. Here
R0,τ=supφ∈H1((0,L)),φ≠0{∫L0β(x)eτa(x)dIφ2dxdI∫L0β(x)eτa(x)dIφ2xdx+∫L0γ(x)eτa(x)dIφ2dx}. |
By the result of Lemma 2.4.2, for any τ∈[0,1], there exist two positive constant ˉC and C_ depending on dS, dI, q, ‖β‖∞, ‖γ‖∞ and N such that C_≤u,v≤ˉC for any solution of (2.40).
Let
D={(u,v)∈C([0,L])×C([0,L]):C_2≤u,v≤2ˉC}. |
Then (ˉS,ˉI)≠G(τ,(ˉS,ˉI)) for any τ∈[0,1] and (ˉS,ˉI)∈∂D, which implies that Leray-Schauder degree deg(I−G(τ,(⋅,⋅)),D,0) is well defined, and it is independent of τ. Here I is the identity map. Moreover, (ˉS,ˉI) is a solution of (2.33) if and only if (ˉS,ˉI) satisfies (ˉS,ˉI)=G(1,(ˉS,ˉI)). If (ˉS,ˉI)∈D and (I−G(0,(⋅,⋅)))(ˉS,ˉI)=0, then (ˉS,ˉI) is a positive solution of
{dIˉSxx−β(x)ˉSˉIˉS+ˉI+γ(x)ˉI=0,0<x<L,dIˉIxx+β(x)ˉSˉIˉS+ˉI−γ(x)ˉI=0,0<x<L,ˉSx(0)=ˉSx(L)=0,ˉIx(0)=ˉIx(L)=0,∫L0[ˉS+ˉI]dx=N. | (2.41) |
By the result of [2], (2.41) has a unique positive solution (S∗,I∗) satisfying S∗+I∗=NL if the basic reproduction number ˆR0>1. Linearizing (2.41) around (S∗,I∗), we get
{−dIΦxx+β(x)I2∗(S∗+I∗)2Φ+β(x)S2∗(S∗+I∗)2Ψ−γ(x)Ψ=μΦ,0<x<L,−dIΨxx−β(x)S2∗(S∗+I∗)2Ψ+γ(x)Ψ−β(x)I2∗(S∗+I∗)2Φ=μΨ,0<x<L,Φx(0)=Φx(L)=0,Ψx(0)=Ψx(L)=0,∫L0[Φ+Ψ]dx=N. | (2.42) |
Adding the first two equations of (2.42) and using the boundary condition Φx=Ψx=0, x=0,L, we get
−dI(Φxx+Ψxx)=μ(Φ+Ψ),x∈(0,L),(Φ+Ψ)x=0,x=0,L. |
Solving it, we have Φ=−Ψ. Substituting this relation into the first equation of (2.42), we obtain
−dIΦxx+(2Lβ(x)NI∗+γ(x)−β(x))Φ=μΦ. |
Since I∗ is a positive solution of (2.40), we know that −dId2dx2+2LNβ(x)I∗+γ(x)−β(x) is a positive operator, so μ>0. Hence the unique positive solution (S∗,I∗) is linearly stable. Using Leray-Schauder degree index (see Theorem 1.2.8.1 in [28]), we obtain
deg(I−G(0,(⋅,⋅)),D,0)=1. |
Consequently, using the homotopy invariance of Leray-Schauder degree, we have
deg(I−G(1,(⋅,⋅)),D,0)=deg(I−G(0,(⋅,⋅)),D,0)=1 |
for (dI,q)∈ΩUhh∪ΩU1lh. By the properties of degree, G(1,(⋅,⋅) has a fixed point in D if (dI,q)∈ΩUhh∪ΩU1lh, which implies that (2.33) has at least one positive solution.
In this section, we consider the properties of R0 when β(x)−γ(x) changes sign twice. We also need the results on the positive roots of F(η) which is defined as
F(η)=∫L0˜a′(x)m(x)eη˜a(x)dx,0≤η<∞, |
for any given continuous function m(x) on [0,L].
Lemma 2.5.1. Assume that there exists 0<x1<x2<L such that m(x1)=m(x2)=0, i.e., m(x) change sign twice for x∈[0,L]. Then
(i) If m(L)<0 and ∫L0˜a′(x)m(x)dx>0, then F(η) has a unique positive root η1 for η∈(0,+∞) satisfying F′(η1)<0;
(ii) If m(L)>0 and ∫L0˜a′(x)m(x)dx<0, then F(η) has a unique positive root η1 for η∈(0,+∞) satisfying F′(η1)>0;
(iii) If m(L)>0 and ∫L0˜a′(x)m(x)dx>0, then F(η) has at most two positive roots for η∈(0,+∞);
(iv) If m(L)<0 and ∫L0˜a′(x)m(x)dx<0, then F(η) has at most two positive roots for η∈(0,+∞).
Proof. We only prove part (i) and part (iii). The proofs of part (ii) and part (iv) are similar.
(i). Let G1(η):=e−˜a(x2)η[˜a(x1)F(η)−F′(η)] and the prime notation denote differentiation with respect to η. Since m(L)<0 and m(x) changes sign twice, it is easy to see that m(x)<0 for x∈(0,x1)∪(x2,L) and m(x)>0 for x∈(x1,x2). Note that ˜a(x) is increasing. We know that
m(x)[˜a(x)−˜a(x1)][˜a(x)−˜a(x2)]<0 |
for x∈(0,L) and x≠xi(i=1,2). As a result, for any η>0, we have
G′1(η)=−e−˜a(x2)η(F″(η)−[˜a(x1)+˜a(x2)]F′(η)+˜a(x1)˜a(x2)F(η))=−∫L0eη[˜a(x)−˜a(x2)]˜a′(x)m(x)[˜a(x)−˜a(x1)][˜a(x)−˜a(x2)]dx>0, |
which implies that G′1(η) is a strictly increasing function for η∈(0,∞). By Lemma 2.2.2 and m(L)<0, F(η)<0 for η>M if M is large enough. But F(0)=∫L0˜a′(x)m(x)dx>0, so there exits at least a positive root of F(η). Let η1 be the smallest positive one, then F′(η1)≤0.
If F′(η1)=0, since
F″(η)−[˜a(x1)+˜a(x2)]F′(η)+˜a(x1)˜a(x2)F(η)=∫L0eη[˜a(x)−˜a(x2)]˜a′(x)m(x)[˜a(x)−˜a(x1)][˜a(x)−˜a(x2)]dx<0, |
then
F″(η1)−[˜a(x1)+˜a(x2)]F′(η1)+˜a(x1)˜a(x2)F(η1)=F″(η1)<0. |
That is, η1 is a strict local maximum value point of F(η), which is a contradiction. So F′(η1)<0. Now we will prove that η1 is the unique positive root of F(η). Assume contradictorily that η2>η1 is the first number such that F(η2)=0. Since F(η1)=0 and F′(η1)<0, then F(η)<0 in (η1,η2), which implies that F′(η2)≥0. By the definition of G1(η), and noticing that F(η1)=F(η2)=0, we have G1(η1)=−˜a(x1)e˜a(x2)η1F′(η1)>0 and G1(η2)=−˜a(x1)e˜a(x2)η2F′(η2)≤0, which is a contradiction to the fact that G1(η) is strictly increasing.
(iii) By Lemma 2.2.2 and m(L)>0, we see that F(η)>0 for η>M if M is large enough. Then either F(η)>0 for any η>0 or F(η) has positive roots in (0,∞). Let G2(η)=e−˜a(x2)η[F′(η)−˜a(x1)F(η)] and η1 be the first positive root of F(η)=0. Similar to the proof of part (i), it is easy to prove that G2(η) is strictly monotone increasing in (0,+∞) and F′(η1)≤0. We discuss in two cases.
Case 1: F′(η1)=0. We will show that η1 is the unique positive root of F(η). Since
F″(η)−[˜a(x1)+˜a(x2)]F′(η)+˜a(x1)˜a(x2)F(η)=∫L0eη[˜a(x)−˜a(x2)]˜a′(x)m(x)[˜a(x)−˜a(x1)][˜a(x)−˜a(x2)]dx>0 |
then F″(η1)−[˜a(x1)+˜a(x2)]F′(η1)+˜a(x1)˜a(x2)F(η1)=F″(η1)>0. That is, F(η) attains a strict local minimum at η1. Now we will prove that η1 is the unique positive root of F(η). Assume contradictorily that η2>η1 is the first number such that F(η2)=0. Since η1 is a strict local minimum value point, we have F(η)>0 in (η1,η2), which implies that F′(η2)≤0. By the definition of G2(η), and noticing that F(η1)=F(η2)=0, we have G2(η1)=0 and G2(η2)=ea(x2)η2F′(η2)≤0, which is a contradiction to the fact that G2(η) is strictly increasing. So F(η) only has a unique positive root η1 in this case.
Case 2. F′(η1)<0. Since F(η1)=0, so F(η)<0 if η>η1 and η close to η1 enough. By Lemma 3.2 and m(L)>0, F(η)>0 for η>M if M is large enough. Therefore, there exists at least a root of F(η)=0 in (η1,∞). Assume that η2 is the first root of F(η)=0 in (η1,∞). Then F(η)<0 in (η1,η2) and F′(η2)≥0. If F′(η2)=0, then
F″(η2)=F″(η2)−[˜a(x1)+˜a(x2)]F′(η2)+˜a(x1)˜a(x2)F(η2)=∫L0eη2[˜a(x)−˜a(x2)]˜a′(x)m(x)[˜a(x)−˜a(x1)][˜a(x)−˜a(x2)]dx>0. |
And F(η) attains a strict local minimum at η2, which is a contradiction. Hence F′(η2)>0.
We need to show that there is no positive root of F(η)=) for η>η2. Assume contradictorily that there exists η3>η2 such that F(η3)=0 and F(η)>0 in (η2,η3). Then F′(η3)<0. And G2(η2)=e˜a(x2)η2F′(η2)>0 and G2(η3)=ea(x2)η3F′(η3)<0, which contradicts the fact that G2(η) is strictly increasing. Therefore we have proved that there exists a unique η2>η1 such that F(η2)=0 and F′(η2)>0.
Now we give the proof of Theorem 1.6 below.
Proof. We only prove part(i) and (iii). The proofs of (ii) and (iv) are similar.
Part (i): Similar to the proofs of Lemma 2.3.2 and 2.3.3, it is easy to prove that there exists some positive constant Λ which is independent of dI and q and for each dI>Λ, there exists some ˜q=˜q(dI) which satisfies R0(dI,˜q)=1 and ˜qdI→η0 as dI→∞. Here η0 is the unique positive root of F(η)=0.
Next, we will prove that
∂R0∂q(dI,˜q)<0 |
for any ˜q satisfying R0(dI,˜q)=1 if dI is large enough.
Let ˜φ be the unique normalized eigenfunction of the eigenvalue R0(dI,˜q)=1, i.e., max[0,L]˜φ=1 and
{−dI(e˜qdI˜a(x)˜φx)x+[γ(x)−β(x)]e˜qdI˜a(x)˜φ=0,0<x<L,˜φx(0)=˜φx(L)=0. | (2.43) |
By (2.22), we have
∂R0∂q(dI,˜q)=R20∫L0e˜qdI˜a(x)˜φx˜φ˜a′(x)dx∫L0β(x)e˜qdI˜a(x)˜φ2dx. | (2.44) |
Multiplying (2.43) by ∫x0˜φ(s)ds and integrating it over (0,L), we get
dI∫L0e˜qdI˜a(x)˜φx˜φ˜a′(x)dx+∫L0[γ(x)−β(x)]e˜qdI˜a(x)˜φ(∫x0˜φ(s)ds)dx=0. |
Substitute it into (2.44), we obtain
dI∂R0∂q(dI,˜q)=∫L0[β(x)−γ(x)]e˜qdI˜a(x)˜φ(∫x0˜φ(s)ds)dx∫L0β(x)e˜qdI˜a(x)˜φ2dx. |
As dI→∞, ˜qdI→η0 and ˜φ→1, we have
limdI→∞dI∂R0∂q(dI,˜q)=∫L0x[β(x)−γ(x)]eη0˜a(x)dx∫L0β(x)eη0˜a(x)dx. |
By Lemma 2.5.1(i),
∫L0x[β(x)−γ(x)]eη0˜a(x)dx=F′(η0)<0. |
Hence, there exists some constant Q>0(dependent on dI) such that R0>1 for 0<q<Q and R0<1 for q>Q.
Part (iii). According to the results of Lemma 2.5.1(iii), we divide into three cases to prove it.
Case 1. F(η)>0 for any η>0. It is easy to show that there exists some positive constant Λ independent of dI and q such that R0>1 for every dI>Λ and any q>0.
Case 2. F(η) has a unique positive root η1 for η∈(0,+∞) and F′(η1)=0. Similar to the proof of part (i), we can prove that there exists some positive constant Λ independent of dI and q such that for every dI>Λ, there exists some ˜q=˜q(dI) such that R0(dI,˜q)=1 and ˜qdI→η0 as dI→∞, where η0 is the unique positive root of F(η)=0. Moreover, ∂R0∂q(dI,˜q)=0. Therefore there exists some positive constant Λ which is independent of dI and q such that for every dI>Λ, there exists a constant Q>0 dependent on dI satisfying R0=1 for q=Q and R0>1 for q∈(0,Q)∪(Q,∞).
Case 3. F(η) has two positive roots η1 and η2(η1<η2) for η∈(0,+∞) and F′(η1)<0, F′(η2)>0. Similar to the discussion of part (i), for each dI>0, there exist ˜q1=˜q1(dI) and ˜q2=˜q2(dI) such that R0(dI,˜qi)=1(i=1,2) and ˜q1dI→η1, ˜q2dI→η2 as dI→∞. And
∂R0∂q(dI,˜q1)<0,∂R0∂q(dI,˜q2)>0. |
Consequently, there exist two constants Q2>Q1>0 which depend on dI and satisfy that R0>1 for q∈(0,Q1)∪(Q2,∞), R0<1 for q∈(Q1,Q2).
In this section, we will summarize the main results of this paper.
Theorem 1.1 gives some properties for the basic reproduction number R0 and Theorem 1.2 says that R0=1 is the watershed for judging whether the DFE is stable or not. Theorem 1.3 and Theorem 1.4 deal with the stable and unstable regions of the DFE. Theorem 1.5 establishes the existence of EE. Theorem 1.6 considers the results on (1.1) when β(x)−γ(x) changes sign twice in (0,L).
We only establish the results on (1.1) under the assumption of a′(x)>0 in this paper. However, it is much more difficult to obtain the results on (1.1) if there exists some x0∈(0,L) satisfying a′(x0)=0.
Biologically, the influence of advection is from the upstream to the downstream, small diffusion or large advection tends to force the individuals to concentrate at the downstream end. Therefore, the disease persists for arbitrary advection rate if the habitat is a high-risk domain and the downstream end is a high-risk site. While the advection transports the individuals to a favorable location and thus it can help eliminate the disease if the downstream end is a low-risk site. In conclusion, when advection is strong or the diffusion is small, the disease will be eliminated if the downstream end is a low–risk site, while the disease will persist if the downstream end is a high–risk site.
The authors thank the anonymous referees for their helpful suggestions.
Xiaowei An was supported by Natural Science Foundation of China People's Police University(No.ZKJJPY201723).
All authors declare no conflicts of interest in this paper.
[1] |
L. J. S. Allen, B. M. Bolker, Y. Lou, A. L. Nevai, Asymptotic profiles of the steady states for an SIS epidemic patch model, SIAM J. Appl. Math., 67 (2007), 1283–1309. doi: 10.1137/060672522
![]() |
[2] |
L. J. S. Allen, B. M. Bolker, Y. Lou, A. L. Nevai, Asymptotic profiles of the steady states for an SIS epidemic reaction–diffusion model, Discrete Contin. Dyn. Syst., 21 (2008), 1–20. doi: 10.3934/dcds.2008.21.1
![]() |
[3] |
R. H. Cui, K. Y. Lam, Y. Lou, Dynamics and asymptotic profiles of steady states of an epidemic model in advective environments, J. Differ. Equations, 263 (2017), 2343–2373. doi: 10.1016/j.jde.2017.03.045
![]() |
[4] |
J. Ge, K. I. Kim, Z. G. Lin, H. P. Zhu, A SIS reaction–diffusion–advection model in a low–risk and high–risk domain, J. Differ. Equations, 259 (2015), 5486–5509. doi: 10.1016/j.jde.2015.06.035
![]() |
[5] |
W. Z. Huang, M. A. Han, K. Y. Liu, Dynamics of an SIS reaction–diffusion epidemic model for disease transmission, Math. Biosci. Eng., 7 (2010), 51–66. doi: 10.3934/mbe.2010.7.51
![]() |
[6] | K. Kuto, H. Matsuzawa, R. Peng, Concentration profile of endemic equilibrium of a reaction–diffusion–advection SIS epidemic model, Calc. Var. Partial Differential Equations, 56 (2017), Paper No. 112, 28 pp. |
[7] |
H. C. Li, R. Peng, F. B. Wang, Varying total population enhances disease persistence: qualitative analysis on a diffusive SIS epidemic model, J. Differ. Equations, 262 (2017), 885–913. doi: 10.1016/j.jde.2016.09.044
![]() |
[8] |
R. Peng, Asymptotic profiles of the positive steady state for an SIS epidemic reaction–diffusion model. I, J. Differ. Equations, 247 (2009), 1096–1119. doi: 10.1016/j.jde.2009.05.002
![]() |
[9] |
R. Peng, S. Q. Liu, Global stability of the steady states of an SIS epidemic reaction–diffusion model, Nonlinear Anal., 71 (2009), 239–247. doi: 10.1016/j.na.2008.10.043
![]() |
[10] |
R. Peng, F. Q. 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. doi: 10.1016/j.physd.2013.05.006
![]() |
[11] |
R. Peng, X. Q. Zhao, A reaction–diffusion SIS epidemic model in a time–periodic environment, Nonlinearity, 25 (2012), 1451–1471. doi: 10.1088/0951-7715/25/5/1451
![]() |
[12] |
W. Wang, W. Ma, Z. Feng, Global dynamics and traveling waves for a periodic and diffusive chemostat model with two nutrients and one microorganism, Nonlinearity, 33 (2020), 4338–4380. doi: 10.1088/1361-6544/ab86ca
![]() |
[13] | X. Q. Zhao, Dynamical systems in population biology 2nd edition. CMS Books in Mathematics/Ouvrages de Mathmatiques de la SMC. Springer, Cham, 2017. xv+413 pp. ISBN: 978-3-319-56432-6; 978-3-319-56433-3. |
[14] |
R. H. Cui, Y. Lou, A spatial SIS model in advective heterogeneous environments, J. Differ. Equations, 261 (2016), 3305–3343. doi: 10.1016/j.jde.2016.05.025
![]() |
[15] | D. Henry, Geometric Theory of Semilinear Parabolic Equations, Lecture Notes in Mathematics, vol.840, Springer-Verlag, Berlin-New York, 1981. |
[16] | O. Diekmann, J. A. P. Heesterbeek, J. A. J. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28 (1990), 365–382. |
[17] |
P. van den Driessche, J. Watmough, Reproduction numbers and sub–threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. doi: 10.1016/S0025-5564(02)00108-6
![]() |
[18] |
W. D. Wang, X. Q. Zhao, Basic reproduction numbers for reaction–diffusion epidemic models, SIAM J. Appl. Dyn. Syst., 11 (2012), 1652–1673. doi: 10.1137/120872942
![]() |
[19] | H. L. Smith, Monotone Dynamical Systems, Mathematical Surveys and Monographs, vol.41, American Mathemat-ical Society, Providence, RI, 1995. |
[20] |
M. G. Crandall, P. H. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321–340. doi: 10.1016/0022-1236(71)90015-2
![]() |
[21] |
Y. H. Du, J. P. Shi, Allee effect and bistability in a spatially heterogeneous predator-prey model, Trans. Amer. Math. Soc., 359 (2007), 4557–4593. doi: 10.1090/S0002-9947-07-04262-6
![]() |
[22] |
J. P. Shi, Persistence and bifurcation of degenerate solutions, J. Funct. Anal., 169 (1999), 494–531. doi: 10.1006/jfan.1999.3483
![]() |
[23] |
J. P. Shi, X. F. Wang, On global bifurcation for quasilinear elliptic systems on bounded domains, J. Differ. Equations, 246 (2009), 2788–2812. doi: 10.1016/j.jde.2008.09.009
![]() |
[24] |
P. H. Rabinowitz, Some global results for nonlinear eigenvalue problems, J. Funct. Anal., 7 (1971), 487–513. doi: 10.1016/0022-1236(71)90030-9
![]() |
[25] | D. Gilbarg, N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, 2nd edition, Springer-Verlag, Berlin, 1983. |
[26] |
C. S. Lin, W. M. Ni, I. Takagi, Large amplitude stationary solutions to a chemotaxis system, J. Differ. Equations, 72 (1988), 1–27. doi: 10.1016/0022-0396(88)90147-7
![]() |
[27] |
Y. Lou, W. M. Ni, Diffusion, self–diffusion and cross–diffusion, J. Differ. Equations, 131 (1996), 79–131. doi: 10.1006/jdeq.1996.0157
![]() |
[28] | L. Nirenberg, Topics in Nonlinear Functional Analysis, New York University, Courant Institute of Mathematical Sciences/American Mathematical Society, New York/Providence, RI, 2001. |
1. | Mengxin Chen, Ranchao Wu, Qianqian Zheng, QUALITATIVE ANALYSIS OF A DIFFUSIVE COVID-19 MODEL WITH NON-MONOTONE INCIDENCE RATE, 2023, 13, 2156-907X, 2229, 10.11948/20220450 | |
2. | Yantao Luo, Yunqiang Yuan, Zhidong Teng, ANALYSIS OF A DEGENERATED DIFFUSION SVEQIRV EPIDEMIC MODEL WITH GENERAL INCIDENCE IN A SPACE HETEROGENEOUS ENVIRONMENT, 2024, 14, 2156-907X, 2704, 10.11948/20230346 | |
3. | Yun Li, Shigui Ruan, Zhi-Cheng Wang, Asymptotic behavior of endemic equilibria for a SIS epidemic model in convective environments, 2025, 426, 00220396, 606, 10.1016/j.jde.2025.01.055 |