This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: The authors were supported by Natural Science Foundation of China (11871179; 11771374).

Traveling wave solutions for a class of discrete diffusive SIR epidemic model

Ran Zhang School of Mathematics
Harbin Institute of Technology
Harbin 150001 Heilongjiang
People’s Republic of China
[email protected]
   Jinliang Wang School of Mathematical Sciences
Heilongjiang University
Harbin 150080 Heilongjiang
People’s Republic of China
[email protected]
   Shengqiang Liu School of Mathematical Sciences
Tiangong University
Tianjin 300387 Tianjin
People’s Republic of China
[email protected]
(Date: January 1, 2004)
Abstract.

This paper is concerned with the conditions of existence and nonexistence of traveling wave solutions (TWS) for a class of discrete diffusive epidemic models. We find that the existence of TWS is determined by the so-called basic reproduction number and the critical wave speed: When the basic reproduction number 0>1\Re_{0}>1, there exists a critical wave speed c>0c^{*}>0, such that for each ccc\geq c^{*} the system admits a nontrivial TWS and for c<cc<c^{*} there exists no nontrivial TWS for the system. In addition, the boundary asymptotic behaviour of TWS is obtained by constructing a suitable Lyapunov functional and employing Lebesgue dominated convergence theorem. Finally, we apply our results to two discrete diffusive epidemic models to verify the existence and nonexistence of TWS.

Key words and phrases:
Lattice dynamical system, Schauder’s fixed point theorem, Traveling wave solutions, Diffusive Epidemic model, Lyapunov functional
1991 Mathematics Subject Classification:
35C07, 35K57, 92D30

1. Introduction

In a pioneering work, the classical Susceptible-Infectious-Recovered (SIR) epidemic model was introduced by Kermack and McKendrick [23] in 1927. Since then, epidemic modeling has became one of the most important tools to study spread of the disease, we refer readers to a good survey [20] on this topic. In order to understand the geographic spread of infectious disease, the spatial effect would give insights into disease spread and control. Due to this fact, epidemic models with spatial diffusion have been studied for decades. Considering spatial effects, Hosono and Ilyas [21] proposed and studied the following SIR epidemic model with diffusion:

{S(x,t)t=d1ΔS(t,x)βS(x,t)I(x,t),x,t>0,I(x,t)t=d2ΔI(t,x)+βS(x,t)I(x,t)γI(x,t),x,t>0,\left\{\begin{array}[]{ll}\vspace{2mm}\displaystyle\frac{\partial S(x,t)}{\partial t}=d_{1}\Delta S(t,x)-\beta S(x,t)I(x,t),&\ x\in\mathbb{R},\ t>0,\\ \displaystyle\frac{\partial I(x,t)}{\partial t}=d_{2}\Delta I(t,x)+\beta S(x,t)I(x,t)-\gamma I(x,t),&\ x\in\mathbb{R},\ t>0,\\ \end{array}\right. (1.1)

with initial conditions

S(x,0)=S0(x),I(x,0)=I0(x)>0,S(x,0)=S_{0}(x),\ \ I(x,0)=I_{0}(x)>0,

where S(x,t)S(x,t) and I(x,t)I(x,t) denote the densities of susceptible and infected individuals at position xx and time tt, respectively; di(i=1,2)d_{i}(i=1,2) are the diffusion rates of each compartments; β\beta denotes the transmission rate between susceptible and infected individuals; γ\gamma is the remove rate. All parameters in system (1.1) are assumed to be positive. The authors proved the existence of traveling wave solutions of system (1.1) with a constant speed when d2=1d_{2}=1. In recent years, many researchers have paid attention to study the traveling wave solutions for diffusive epidemic models (see, for example, [2, 11, 15, 18, 26, 28, 27, 36, 37, 51, 49] and references therein).

However, there are relatively few works on epidemic models with discrete spatial structure. In contrast to continuous media, lattice dynamical systems is more realistic in describing the discrete diffusion (for example, patch environment [33]). Lattice dynamical systems are systems of ordinary differential equations with a discrete spatial structure. Such systems arise from practical backgrounds, such as biology [38, 13, 41, 47, 17], chemical reaction [22, 12] and material science [3, 5]. In a recent paper [14], Fu et al. studied the existence of traveling wave solutions for a lattice dynamical system arising in a discrete diffusive epidemic model:

{dSn(t)dt=[Sn+1(t)+Sn1(t)2Sn(t)]βSn(t)In(t),n,dIn(t)dt=d[In+1(t)+In1(t)2In(t)]+βSn(t)In(t)γIn(t),n,\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]-\beta S_{n}(t)I_{n}(t),\ \ n\in\mathbb{Z},\\ \displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\beta S_{n}(t)I_{n}(t)-\gamma I_{n}(t),\ \ n\in\mathbb{Z},\\ \end{array}\right. (1.2)

where Sn(t)S_{n}(t) and In(t)I_{n}(t) denote the populations densities of susceptible and infectious individuals at niche nn and time tt, respectively; 11 and dd denote the random migration coefficients for susceptible and infectious individuals, respectively; β\beta is the transmission coefficient between susceptible and infectious individuals; γ\gamma is the recovery rate of infectious individuals. Note that system (1.2) is a spatially discrete version of system (1.1). It was proved in [14] that the conditions of existence and nonexistence of traveling wave solution for system (1.2) are determined by a threshold number and the critical wave speed cc^{*}. If the threshold number is greater than 1, then there exists a traveling wave solution for any c>cc>c^{*} and there is no traveling wave solutions for c<cc<c^{*}. Also, the non-existence of traveling wave solutions for the threshold number is less than 1 was derived. Furthermore, Wu [40] studied the existence of traveling wave solutions with critical speed c=cc=c^{*} of system (1.2). In [48] and [52], two models with saturated incidence rate are considered, and they also investigated the existence and nonexistence of traveling wave solutions. By introducing the constant recruitment, Chen et al. [9] studied the traveling wave solutions for the following discrete diffusive epidemic model:

{dSn(t)dt=[Sn+1(t)+Sn1(t)2Sn(t)]+μβSn(t)In(t)μSn(t),n,dIn(t)dt=d[In+1(t)+In1(t)2In(t)]+βSn(t)In(t)(γ+μ)In(t),n,\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]+\mu-\beta S_{n}(t)I_{n}(t)-\mu S_{n}(t),\ \ n\in\mathbb{Z},\\ \displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\beta S_{n}(t)I_{n}(t)-(\gamma+\mu)I_{n}(t),\ \ n\in\mathbb{Z},\\ \end{array}\right. (1.3)

where μ\mu is the input rate of the susceptible population, meanwhile, the death rates of susceptible and infectious individuals are also assumed to be μ\mu. In [9], the authors showed that the existence of traveling wave solutions connecting the disease-free equilibrium to the endemic equilibrium, but it still remains open whether the traveling wave solutions converge to the endemic equilibrium at ++\infty. As explained in [9], the main difficulties come from the fact that (1.3) is a system and is non-local. In fact, the traveling wave solutions of (1.2) and (1.3) are totally different: The disease will always die out for the system like (1.2) without constant recruitment, that is, II tends to 0 as ξ±\xi\rightarrow\pm\infty, where ξ=n+ct\xi=n+ct is the wave profile to be introduced in the next section; However, for the diffusive model with positive constant recruitment, it is more likely to get that I(ξ)0I(\xi)\rightarrow 0 as ξ\xi\rightarrow-\infty and I(ξ)II(\xi)\rightarrow I^{*} as ξ+\xi\rightarrow+\infty, where II^{*} is the positive endemic equilibrium (see [28] for nonlocal diffusive epidemic model; [15] for random diffusive epidemic model). Therefore, it naturally raises a question: For discrete diffusive systems, does the traveling wave solutions converge to the endemic equilibrium as ξ+\xi\rightarrow+\infty? This constitutes our first motivation of the present paper.

Our second motivation is the nonlinear incidence rate which plays a critical role in the epidemic modeling [1], for the discrete diffusive systems with nonlinear incidence rate, will the traveling wave solutions still converge to the endemic equilibrium as ξ+\xi\rightarrow+\infty? Traditionally, the incidence rate of an infectious disease in most of the literature is assumed to be of mass action form βSI\beta SI [1]. Yet the disease transmission process is generally unknown [24], some nonlinear incidence rates have been introduced and studied, for example, the saturated incidence rate with f(I)=I1+αIf(I)=\frac{I}{1+\alpha I} by [6], the saturated nonlinear incidence rate with f(I)=I1+αIp(0<p<1)f(I)=\frac{I}{1+\alpha I^{p}}(0<p<1) by [29], and so on. For more general cases, Capasso et al. [6] considered a more general incidence rate with the form Sf(I)Sf(I). It is seems that the general nonlinear incidence rate could bring nontrivial challenges in analysis. Therefore, it is of great significance to study the convergence property of traveling wave solutions of the system with nonlinear incidence rate.

In this paper, we consider a discrete diffusive SIR epidemic model with general nonlinear incidence rate. The main model of this paper is formulated as the following system:

{dSn(t)dt=d1[Sn+1(t)+Sn1(t)2Sn(t)]+ΛβSn(t)f(In(t))μ1Sn(t),n,dIn(t)dt=d2[In+1(t)+In1(t)2In(t)]+βSn(t)f(In(t))γIn(t)μ2In(t),n,dRn(t)dt=d3[Rn+1(t)+Rn1(t)2Rn(t)]+γIn(t)μ1Rn(t),n,\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=d_{1}[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]+\Lambda-\beta S_{n}(t)f(I_{n}(t))-\mu_{1}S_{n}(t),\ \ n\in\mathbb{Z},\\ \vspace{2mm}\displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d_{2}[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\beta S_{n}(t)f(I_{n}(t))-\gamma I_{n}(t)-\mu_{2}I_{n}(t),\ \ n\in\mathbb{Z},\\ \displaystyle\frac{{\rm d}R_{n}(t)}{{\rm d}t}=d_{3}[R_{n+1}(t)+R_{n-1}(t)-2R_{n}(t)]+\gamma I_{n}(t)-\mu_{1}R_{n}(t),\ \ n\in\mathbb{Z},\end{array}\right. (1.4)

where Sn(t)S_{n}(t), In(t)I_{n}(t) and Rn(t)R_{n}(t) denote the densities of susceptible, infectious and removed individuals at niche nn and time tt, respectively; di(i=1,2,3)d_{i}(i=1,2,3) is the random migration coefficients for each compartments; Λ\Lambda is the input rate of susceptible individuals. The biological meaning of other parameters are the same as in model (1.3). This paper aims to study the existence and convergence property of traveling wave solutions of model (1.4), and one of our results will answer the open problem proposed in [9].

Since Rn(t)R_{n}(t) is decoupled from other equations and denote μ2=γ+μ1\mu_{2}=\gamma+\mu_{1}, then we only need to study the following system:

{dSn(t)dt=d1[Sn+1(t)+Sn1(t)2Sn(t)]+ΛβSn(t)f(In(t))μ1Sn(t),n,dIn(t)dt=d2[In+1(t)+In1(t)2In(t)]+βSn(t)f(In(t))μ2In(t),n.\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=d_{1}[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]+\Lambda-\beta S_{n}(t)f(I_{n}(t))-\mu_{1}S_{n}(t),\ \ n\in\mathbb{Z},\\ \displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d_{2}[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\beta S_{n}(t)f(I_{n}(t))-\mu_{2}I_{n}(t),\ \ n\in\mathbb{Z}.\end{array}\right. (1.5)

We make the following assumptions on function ff.

Assumption 1.1.

Assume the function ff satisfying

(A1):

f(I)0f(I)\geq 0 and f(I)>0f^{\prime}(I)>0 for all I0I\geq 0, f(I)=0f(I)=0 if and only if I=0I=0.

(A2):

f(I)I\frac{f(I)}{I} is continuous and monotonously non-increasing for all I0I\geq 0 and limI0+f(I)I\lim\limits_{I\rightarrow 0^{+}}\frac{f(I)}{I} exists.

The conditions of Assumption 1.1 are satisfied in all the following specific incidence rates:

  1. (i)

    the bilinear incidence rate with f(I)=If(I)=I (see [1]);

  2. (ii)

    the saturated incidence rate with f(I)=I1+αIf(I)=\frac{I}{1+\alpha I} (see [6]);

  3. (iii)

    the saturated nonlinear incidence rate with f(I)=I1+αIpf(I)=\frac{I}{1+\alpha I^{p}}, where α>0\alpha>0 and 0<p<10<p<1 (a special case in [29], see also [30]);

  4. (iv)

    the nonlinear incidence rate with f(I)=I1+kI+1+2kIf(I)=\frac{I}{1+kI+\sqrt{1+2kI}} (see [19, 35]);

  5. (v)

    the nonlinear incidence rate with f(I)=I(ϵα+Iα)γf(I)=\frac{I}{(\epsilon^{\alpha}+I^{\alpha})^{\gamma}}, where ϵ,α,γ>0\epsilon,\alpha,\gamma>0 and αγ<1\alpha\gamma<1 (see [35]);

  6. (vi)

    the nonlinear incidence rate for pathogen transmission in infection of insects with f(I)=kln(1+νIk)f(I)=k\ln\left(1+\frac{\nu I}{k}\right), which could be described by epidemic model (see [4]).

Hence, system (1.5) covers many models as special cases. Now, we introduce some results on the system (1.5) without migration, which takes the form as:

{dS(t)dt=ΛβS(t)f(I(t))μ1S(t),dI(t)dt=βS(t)f(I(t))μ2I(t).\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S(t)}{{\rm d}t}=\Lambda-\beta S(t)f(I(t))-\mu_{1}S(t),\\ \displaystyle\frac{{\rm d}I(t)}{{\rm d}t}=\beta S(t)f(I(t))-\mu_{2}I(t).\end{array}\right. (1.6)

It is well-known that the global dynamics of (1.6) is completely determined by the basic reproduction number 0=βS0f(0)μ2\Re_{0}=\frac{\beta S_{0}f^{\prime}(0)}{\mu_{2}} (see [25]): that is, if the number is less than unity, then the disease-free equilibrium E0=(S0,0)=(Λ/μ1,0)E_{0}=(S_{0},0)=(\Lambda/\mu_{1},0) is globally asymptotically stable, while if the number is greater than unity, then a positive endemic equilibrium E=(S,I)E^{*}=(S^{*},I^{*}) exists and it is globally asymptotically stable, where EE^{*} satisfy

{ΛβSf(I)μ1S=0,βSf(I)μ2I=0.\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\Lambda-\beta S^{*}f(I^{*})-\mu_{1}S^{*}=0,\\ \displaystyle\beta S^{*}f(I^{*})-\mu_{2}I^{*}=0.\end{array}\right. (1.7)

The organization of this paper is as follows. In Section 2, we apply Schauder’s fixed point theorem to construct a family of solutions of the truncated problem. In Section 3, we show the existence and boundedness of traveling wave solutions. Further, we use a Lyapunov functional to show that the convergence of traveling wave solutions at ++\infty. In Section 4, we investigate the nonexistence of traveling wave solutions by using two-sided Laplace transform. At last, there is an application for our general results and a brief discussion.

2. Preliminaries

In this section, since system (1.5) does not enjoy the comparison principle, we will construct a pair upper and lower solutions and apply Schauder’s fixed point theorem to investigate the existence of traveling wave solutions of system (1.5). Consider traveling wave solutions which can be expressed as bounded profiles of continuous variable n+ctn+ct such that

Sn(t)=S(n+ct)andIn(t)=I(n+ct).S_{n}(t)=S(n+ct)\ \ \textrm{and}\ \ I_{n}(t)=I(n+ct). (2.1)

where cc denotes the wave speed. Let ξ=n+ct\xi=n+ct, then we can rewrite system (1.5) as follows:

{cS(ξ)=d1J[S](ξ)+Λμ1S(ξ)βS(ξ)f(I(ξ)),cI(ξ)=d2J[I](ξ)+βS(ξ)f(I(ξ))μ2I(ξ)\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle cS^{\prime}(\xi)=d_{1}J[S](\xi)+\Lambda-\mu_{1}S(\xi)-\beta S(\xi)f(I(\xi)),\\ \displaystyle cI^{\prime}(\xi)=d_{2}J[I](\xi)+\beta S(\xi)f(I(\xi))-\mu_{2}I(\xi)\end{array}\right. (2.2)

for all ξ\xi\in\mathbb{R}, where J[ϕ](ξ):=ϕ(ξ+1)+ϕ(ξ1)2ϕ(ξ)J[\phi](\xi):=\phi(\xi+1)+\phi(\xi-1)-2\phi(\xi). We want to find traveling wave solutions with the following asymptotic boundary conditions:

limξ(S(ξ),I(ξ))=(S0,0),\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0), (2.3)

and

limξ+(S(ξ),I(ξ))=(S,I).\lim_{\xi\rightarrow+\infty}(S(\xi),I(\xi))=(S^{*},I^{*}). (2.4)

where (S0,0)(S_{0},0) is the disease-free equilibrium and (S,I)(S^{*},I^{*}) is the positive endemic equilibrium, which have been defined in Section 1. Linearizing the second equation of system (2.2) at disease-free equilibrium (S0,0)(S_{0},0), we have

cI(ξ)=d2J[I](ξ)μ2I(ξ)+βS0f(0)I(ξ).cI^{\prime}(\xi)=d_{2}J[I](\xi)-\mu_{2}I(\xi)+\beta S_{0}f^{\prime}(0)I(\xi). (2.5)

Letting I(ξ)=eλξI(\xi)=e^{\lambda\xi} and substituting it into (2.5), yields

d2[eλ+eλ2]cλ+βS0f(0)μ2=0.d_{2}[e^{\lambda}+e^{-\lambda}-2]-c\lambda+\beta S_{0}f^{\prime}(0)-\mu_{2}=0.

Denote

Δ(λ,c)=d2[eλ+eλ2]cλ+βS0f(0)μ2.\Delta(\lambda,c)=d_{2}[e^{\lambda}+e^{-\lambda}-2]-c\lambda+\beta S_{0}f^{\prime}(0)-\mu_{2}. (2.6)

By some calculations, we have

Δ(0,c)=βS0f(0)μ2,limc+Δ(λ,c)=,\displaystyle\Delta(0,c)=\beta S_{0}f^{\prime}(0)-\mu_{2},\ \ \ \lim_{c\rightarrow+\infty}\Delta(\lambda,c)=-\infty,
Δ(λ,c)λ=d2[eλeλ]c,Δ(λ,c)c=λ<0,\displaystyle\frac{\partial\Delta(\lambda,c)}{\partial\lambda}=d_{2}[e^{\lambda}-e^{-\lambda}]-c,\ \ \ \frac{\partial\Delta(\lambda,c)}{\partial c}=-\lambda<0,
2Δ(λ,c)λ2=d2[eλ+eλ]>0,Δ(λ,c)λ|(0,c)=c<0,\displaystyle\frac{\partial^{2}\Delta(\lambda,c)}{\partial\lambda^{2}}=d_{2}[e^{\lambda}+e^{-\lambda}]>0,\ \ \ \frac{\partial\Delta(\lambda,c)}{\partial\lambda}\bigg{|}_{(0,c)}=-c<0,

for λ>0\lambda>0 and c>0c>0. Therefore, we have the following lemma.

Lemma 2.1.

Let 0>1.\Re_{0}>1. There exist c>0c^{*}>0 and λ>0\lambda^{*}>0 such that

Δ(λ,c)λ|(λ,c)=0andΔ(λ,c)=0.\frac{\partial\Delta(\lambda,c)}{\partial\lambda}\bigg{|}_{(\lambda^{*},c^{*})}=0\ \ \textrm{and}\ \ \Delta(\lambda^{*},c^{*})=0.

Furthermore,

(i):

if c=c,c=c^{*}, then Δ(λ,c)=0\Delta(\lambda,c)=0 has only one positive real root λ;\lambda^{*};

(ii):

if 0<c<c,0<c<c^{*}, then Δ(λ,c)>0\Delta(\lambda,c)>0 for all λ(0,λM),\lambda\in(0,\lambda_{M}), where λM(0,+]\lambda_{M}\in(0,+\infty];

(iii):

if c>c,c>c^{*}, then Δ(λ,c)=0\Delta(\lambda,c)=0 has two positive real roots λ1,\lambda_{1}, λ2\lambda_{2} with λ1<λ<λ2\lambda_{1}<\lambda^{*}<\lambda_{2}.

From Lemma 2.1, we have

Δ(λ,c){>0forλ<λ1,<0forλ1<λ<λ2,>0forλ>λ2.\Delta(\lambda,c)\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle>0\ \ \ {\rm for}\ \ \ \lambda<\lambda_{1},\\ \vspace{2mm}\displaystyle<0\ \ \ {\rm for}\ \ \ \lambda_{1}<\lambda<\lambda_{2},\\ \displaystyle>0\ \ \ {\rm for}\ \ \ \lambda>\lambda_{2}.\end{array}\right. (2.7)

In the following of this section, we always fix c>cc>c^{*} and 0>1\Re_{0}>1.

2.1. Construction of upper and lower solutions

Definition 2.2.

(S+(ξ),I+(ξ))(S^{+}(\xi),I^{+}(\xi)) and (S(ξ),I(ξ))(S^{-}(\xi),I^{-}(\xi)) are called a pair upper and lower solutions of (2.2) if S+,I+,S,IS^{+},I^{+},S^{-},I^{-} satisfy

{d1J[S+](ξ)cS+(ξ)+Λμ1S+(ξ)βS+(ξ)f(I(ξ))0,d2J[I+](ξ)cI+(ξ)+βS+(ξ)f(I+(ξ))μ2I+(ξ)0,d1J[S](ξ)cS(ξ)+Λμ1S(ξ)βS(ξ)f(I+(ξ))0,d2J[I](ξ)cI(ξ)+βS(ξ)f(I(ξ))μ2I(ξ)0.\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle d_{1}J[S^{+}](\xi)-c{S^{+}}^{\prime}(\xi)+\Lambda-\mu_{1}S^{+}(\xi)-\beta S^{+}(\xi)f(I^{-}(\xi))\leq 0,\\ \vspace{2mm}\displaystyle d_{2}J[I^{+}](\xi)-c{I^{+}}^{\prime}(\xi)+\beta S^{+}(\xi)f(I^{+}(\xi))-\mu_{2}I^{+}(\xi)\leq 0,\\ \vspace{2mm}\displaystyle d_{1}J[S^{-}](\xi)-c{S^{-}}^{\prime}(\xi)+\Lambda-\mu_{1}S^{-}(\xi)-\beta S^{-}(\xi)f(I^{+}(\xi))\geq 0,\\ \displaystyle d_{2}J[I^{-}](\xi)-c{I^{-}}^{\prime}(\xi)+\beta S^{-}(\xi)f(I^{-}(\xi))-\mu_{2}I^{-}(\xi)\geq 0.\\ \end{array}\right.

Define the following functions:

{S+(ξ)=S0,I+(ξ)=eλ1ξ,{S(ξ)=max{S0(1M1eε1ξ),0},I(ξ)=max{eλ1ξ(1M2eε2ξ),0},\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle S^{+}(\xi)=S_{0},\\ \displaystyle I^{+}(\xi)=e^{\lambda_{1}\xi},\end{array}\right.\ \ \left\{\begin{array}[]{l}\vspace{2mm}\displaystyle S^{-}(\xi)=\max\{S_{0}(1-M_{1}e^{\varepsilon_{1}\xi}),0\},\\ \displaystyle I^{-}(\xi)=\max\{e^{\lambda_{1}\xi}(1-M_{2}e^{\varepsilon_{2}\xi}),0\},\end{array}\right. (2.8)

where MiM_{i} and εi(i=1,2)\varepsilon_{i}(i=1,2) are some positive constants to be determined in the following lemmas. Now we show that (2.8) are a pair upper and lower solutions of (2.2).

Lemma 2.3.

The function I+(ξ)=eλ1ξI^{+}(\xi)=e^{\lambda_{1}\xi} satisfies

cI+(ξ)=d2J[I+](ξ)μ2I+(ξ)+βS0f(0)I+(ξ).c{I^{+}}^{\prime}(\xi)=d_{2}J[I^{+}](\xi)-\mu_{2}I^{+}(\xi)+\beta S_{0}f^{\prime}(0)I^{+}(\xi). (2.9)
Lemma 2.4.

The function S+(ξ)=S0S^{+}(\xi)=S_{0} satisfies

cS+(ξ)d1J[S+](ξ)+Λμ1S+(ξ)βS+(ξ)f(I(ξ)).c{S^{+}}^{\prime}(\xi)\geq d_{1}J[S^{+}](\xi)+\Lambda-\mu_{1}S^{+}(\xi)-\beta S^{+}(\xi)f(I^{-}(\xi)). (2.10)

The proof of the above two lemmas are straightforward, so we omit the details.

Lemma 2.5.

For each 0<ε1<λ10<\varepsilon_{1}<\lambda_{1} sufficiently small and M1>0M_{1}>0 large enough, the function S(ξ)=max{S0(1M1eε1ξ),0}S^{-}(\xi)=\max\{S_{0}(1-M_{1}e^{\varepsilon_{1}\xi}),0\} satisfies

cS(ξ)d1J[S](ξ)+Λμ1S(ξ)βS(ξ)f(I+(ξ))c{S^{-}}^{\prime}(\xi)\leq d_{1}J[S^{-}](\xi)+\Lambda-\mu_{1}S^{-}(\xi)-\beta S^{-}(\xi)f(I^{+}(\xi)) (2.11)

with ξ1ε1ln1M1:=𝔛1\xi\neq\frac{1}{\varepsilon_{1}}\ln\frac{1}{M_{1}}:=\mathfrak{X}_{1}.

Proof.

If ξ>𝔛1\xi>\mathfrak{X}_{1}, then inequality (2.11) holds since S(ξ)=S(ξ+1)=0S^{-}(\xi)=S^{-}(\xi+1)=0 and S(ξ1)0S^{-}(\xi-1)\geq 0. If ξ<𝔛1\xi<\mathfrak{X}_{1}, then

S(ξ)=S0(1M1eε1ξ),S(ξ1)=S0(1M1eε1(ξ1))andS(ξ+1)S0(1M1eε1(ξ+1)).S^{-}(\xi)=S_{0}(1-M_{1}e^{\varepsilon_{1}\xi}),\ \ S^{-}(\xi-1)=S_{0}(1-M_{1}e^{\varepsilon_{1}(\xi-1)})\ \ \textrm{and}\ \ S^{-}(\xi+1)\geq S_{0}(1-M_{1}e^{\varepsilon_{1}(\xi+1)}).

From the concavity of function f(I)f(I), we have

f(I+(ξ))f(0)I+(ξ),f(I^{+}(\xi))\leq f^{\prime}(0)I^{+}(\xi),

thus

d1J[S](ξ)+ΛμS(ξ)βS(ξ)f(I+(ξ))cS(ξ)\displaystyle\ d_{1}J[S^{-}](\xi)+\Lambda-\mu S^{-}(\xi)-\beta S^{-}(\xi)f(I^{+}(\xi))-c{S^{-}}^{\prime}(\xi)
\displaystyle\geq eε1ξS0[M1(d1eε1+d1eε12d1μcε1)β1f(eλ1ξ)eε1ξ+β1M1ε1f(eλ1ξ)]\displaystyle\ e^{\varepsilon_{1}\xi}S_{0}\left[-M_{1}(d_{1}e^{\varepsilon_{1}}+d_{1}e^{-\varepsilon_{1}}-2d_{1}-\mu-c\varepsilon_{1})-\beta_{1}f\left(e^{\lambda_{1}\xi}\right)e^{-\varepsilon_{1}\xi}+\beta_{1}M_{1}\varepsilon_{1}f\left(e^{\lambda_{1}\xi}\right)\right]
\displaystyle\geq eε1ξS0[M1(d1eε1+d1eε12d1μcε1)β1f(0)eλ1ξeε1ξ].\displaystyle\ e^{\varepsilon_{1}\xi}S_{0}\left[-M_{1}(d_{1}e^{\varepsilon_{1}}+d_{1}e^{-\varepsilon_{1}}-2d_{1}-\mu-c\varepsilon_{1})-\beta_{1}f^{\prime}(0)e^{\lambda_{1}\xi}e^{-\varepsilon_{1}\xi}\right].

Select 0<ε1<λ10<\varepsilon_{1}<\lambda_{1} small enough such that d1(2eε1eε1)μcε1<0-d_{1}(2-e^{\varepsilon_{1}}-e^{-\varepsilon_{1}})-\mu-c\varepsilon_{1}<0 and note that e(λ1ε1)ξ1e^{(\lambda_{1}-\varepsilon_{1})\xi}\leq 1 since ξ<𝔛1<0\xi<\mathfrak{X}_{1}<0. Hence, we need to choose

M1β1f(0)d1(2eε1eε1)+μ+cε1M_{1}\geq\frac{\beta_{1}f^{\prime}(0)}{d_{1}(2-e^{\varepsilon_{1}}-e^{-\varepsilon_{1}})+\mu+c\varepsilon_{1}}

large enough to make sure inequality (2.11) holds. This completes the proof. ∎

Lemma 2.6.

For each ε2>0\varepsilon_{2}>0 sufficiently small and M2>ε2ε1M1M_{2}>\frac{\varepsilon_{2}}{\varepsilon_{1}}M_{1} large enough, the function I(ξ)=max{eλ1ξ(1M2eε2ξ),0}I^{-}(\xi)=\max\{e^{\lambda_{1}\xi}(1-M_{2}e^{\varepsilon_{2}\xi}),0\} satisfies

cI(ξ)d2J[I](ξ)+βS(ξ)f(I(ξ))μ2I(ξ)cI^{\prime}(\xi)\leq d_{2}J[I](\xi)+\beta S^{-}(\xi)f(I(\xi))-\mu_{2}I(\xi) (2.12)

with ξ1ε2ln1M2:=𝔛2\xi\neq\frac{1}{\varepsilon_{2}}\ln\frac{1}{M_{2}}:=\mathfrak{X}_{2}.

Proof.

If ξ>𝔛2\xi>\mathfrak{X}_{2}, then inequality (2.12) holds since I(ξ)=I(ξ1)=0I^{-}(\xi)=I^{-}(\xi-1)=0 and I(ξ+1)0I^{-}(\xi+1)\geq 0. If ξ<𝔛2\xi<\mathfrak{X}_{2}, then

I(ξ)=eλ1ξ(1M2eε2ξ),I(ξ1)=eλ1(ξ1)(1M2eε2(ξ1)),I(ξ+1)eλ1(ξ+1)(1M2eε2(ξ+1)),I^{-}(\xi)=e^{\lambda_{1}\xi}(1-M_{2}e^{\varepsilon_{2}\xi}),\ \ I^{-}(\xi-1)=e^{\lambda_{1}(\xi-1)}(1-M_{2}e^{\varepsilon_{2}(\xi-1)}),\ \ I^{-}(\xi+1)\geq e^{\lambda_{1}(\xi+1)}(1-M_{2}e^{\varepsilon_{2}(\xi+1)}),

and inequality (2.12) is equivalent to the following inequality:

βS0f(0)I(ξ)βS(ξ)f(I(ξ))\displaystyle\ \beta S_{0}f^{\prime}(0)I^{-}(\xi)-\beta S^{-}(\xi)f(I^{-}(\xi)) (2.13)
\displaystyle\leq d2J[I](ξ)μ2I(ξ)cI(ξ)+βS0f(0)I(ξ).\displaystyle\ d_{2}J[I^{-}](\xi)-\mu_{2}I^{-}(\xi)-c{I^{-}}^{\prime}(\xi)+\beta S_{0}f^{\prime}(0)I^{-}(\xi).

Note that f(I)I\frac{f(I)}{I} is non-increasing on (0,)(0,\infty) in Assumption 1.1. For any ϵ(0,f(0))\epsilon\in(0,f^{\prime}(0)), then there exists a small positive number δ>0\delta>0 such that

f(I)If(0)ϵ, 0<I<δ.\frac{f(I)}{I}\geq f^{\prime}(0)-\epsilon,\ \ 0<I<\delta.

For 0<I<δ0<I<\delta, we have

βS0f(0)I(ξ)βS(ξ)f(I(ξ))=\displaystyle\beta S_{0}f^{\prime}(0)I^{-}(\xi)-\beta S^{-}(\xi)f(I^{-}(\xi))= (βS0βS(ξ)f(I(ξ))I(ξ))I(ξ)\displaystyle\left(\beta S_{0}-\beta S^{-}(\xi)\frac{f(I^{-}(\xi))}{I^{-}(\xi)}\right)I^{-}(\xi)
\displaystyle\leq (βS0βS(ξ)f(I(ξ))I(ξ)+I(ξ)2)2\displaystyle\left(\frac{\beta S_{0}-\beta S^{-}(\xi)\frac{f(I^{-}(\xi))}{I^{-}(\xi)}+I^{-}(\xi)}{2}\right)^{2}
\displaystyle\leq (βS0βS(ξ)(f(0)ϵ)+I(ξ))2.\displaystyle\left(\beta S_{0}-\beta S^{-}(\xi)(f^{\prime}(0)-\epsilon)+I^{-}(\xi)\right)^{2}. (2.14)

Recall that ξ<𝔛2\xi<\mathfrak{X}_{2}, we can choose M2M_{2} large enough such that

0<I(ξ)<δandS(ξ)S0.0<I^{-}(\xi)<\delta\ \ {\rm and}\ \ S^{-}(\xi)\rightarrow S_{0}.

Since (2.1) is valid for any ϵ\epsilon, we have

βS0f(0)I(ξ)βS(ξ)f(I(ξ))[I(ξ)]2.\beta S_{0}f^{\prime}(0)I^{-}(\xi)-\beta S^{-}(\xi)f(I^{-}(\xi))\leq[I^{-}(\xi)]^{2}.

Furthermore, the right-hand of (2.13) is satisfy

d2J[I](ξ)μ2I(ξ)cI(ξ)+βS0f(0)I(ξ)eλ1ξΔ(λ1,c)M2eλ1+ε2Δ(λ1+ε2,c).d_{2}J[I^{-}](\xi)-\mu_{2}I^{-}(\xi)-c{I^{-}}^{\prime}(\xi)+\beta S_{0}f^{\prime}(0)I^{-}(\xi)\geq e^{\lambda_{1}\xi}\Delta(\lambda_{1},c)-M_{2}e^{\lambda_{1}+\varepsilon_{2}}\Delta(\lambda_{1}+\varepsilon_{2},c).

Using the definition of Δ(λ,c)\Delta(\lambda,c) and [I(ξ)]2e2λ1ξ[I^{-}(\xi)]^{2}\leq e^{2\lambda_{1}\xi}, noticing that Δ(λ1+ε2,c)<0\Delta(\lambda_{1}+\varepsilon_{2},c)<0 for small ε2>0\varepsilon_{2}>0 by (2.7), then it suffices to show that

e(λ1ε2)ξM2Δ(λ1+ε2,c).e^{(\lambda_{1}-\varepsilon_{2})\xi}\leq-M_{2}\Delta(\lambda_{1}+\varepsilon_{2},c).

The above inequality holds for M2M_{2} large enough, since the left-hand side vanishes and the right-hand side tends to infinity as M2+M_{2}\rightarrow+\infty. This ends the proof. ∎

Hence, functions (2.8) are a pair upper and lower solutions of (2.2) by the Definition 2.2.

2.2. Truncated problem

Let X>𝔛2>0X>-\mathfrak{X}_{2}>0. Define the following set

ΓX:={(ϕ,ψ)C([X,X],2)|S(ξ)ϕ(ξ)S+(ξ),I(ξ)ψ(ξ)I+(ξ)ξ[X,X],ϕ(X)=S(X),ψ(X)=I(X).}.\Gamma_{X}:=\left\{(\phi,\psi)\in C([-X,X],\mathbb{R}^{2})\left|\begin{array}[]{l}\vspace{2mm}\displaystyle S^{-}(\xi)\leq\phi(\xi)\leq S^{+}(\xi),\ I^{-}(\xi)\leq\psi(\xi)\leq I^{+}(\xi)\ \ \forall\xi\in[-X,X],\\ \displaystyle\phi(-X)=S^{-}(-X),\ \ \psi(-X)=I^{-}(-X).\end{array}\right.\right\}.

It is clear that ΓX\Gamma_{X} is a nonempty bounded closed convex set in C([X,X],2)C([-X,X],\mathbb{R}^{2}). For any (ϕ,ψ)C([X,X],2)(\phi,\psi)\in C([-X,X],\mathbb{R}^{2}), extend it as

ϕ^(ξ)={ϕ(X),for ξ>X,ϕ(ξ),for ξ[X,X],S(ξ),for ξ<X,ψ^(ξ)={ψ(X),for ξ>X,ψ(ξ),for ξ[X,X],I(ξ),for ξ<X.\hat{\phi}(\xi)=\left\{\begin{array}[]{ll}\displaystyle\phi(X),&\mbox{for $\xi>X$,}\\ \displaystyle\phi(\xi),&\mbox{for $\xi\in[-X,X]$,}\\ \displaystyle S^{-}(\xi),&\mbox{for $\xi<-X$,}\\ \end{array}\right.\ \ \ \hat{\psi}(\xi)=\left\{\begin{array}[]{ll}\displaystyle\psi(X),&\mbox{for $\xi>X$,}\\ \displaystyle\psi(\xi),&\mbox{for $\xi\in[-X,X]$,}\\ \displaystyle I^{-}(\xi),&\mbox{for $\xi<-X$.}\\ \end{array}\right.

Consider the following truncated initial problem:

{cS(ξ)+(2d1+μ1+α)S(ξ)=d1ϕ^(ξ+1)+d1ϕ^(ξ1)+Λ+αϕ(ξ)βϕ(ξ)f(ψ(ξ)):=H1(ϕ,ψ),cI(ξ)+(2d2+μ2)I(ξ)=d2ψ^(ξ+1)+d2ψ^(ξ1)+βϕ(ξ)f(ψ(ξ)):=H2(ϕ,ψ),(S,I)(X)=(S,I)(X),\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle cS^{\prime}(\xi)+(2d_{1}+\mu_{1}+\alpha)S(\xi)=d_{1}\hat{\phi}(\xi+1)+d_{1}\hat{\phi}(\xi-1)+\Lambda+\alpha\phi(\xi)-\beta\phi(\xi)f(\psi(\xi)):=H_{1}(\phi,\psi),\\ \vspace{2mm}\displaystyle cI^{\prime}(\xi)+(2d_{2}+\mu_{2})I(\xi)=d_{2}\hat{\psi}(\xi+1)+d_{2}\hat{\psi}(\xi-1)+\beta\phi(\xi)f(\psi(\xi)):=H_{2}(\phi,\psi),\\ \displaystyle(S,I)(-X)=(S^{-},I^{-})(-X),\end{array}\right. (2.15)

where (ϕ,ψ)ΓX(\phi,\psi)\in\Gamma_{X} and α\alpha is a constant large enough such that H1(ϕ,ψ)H_{1}(\phi,\psi) is non-decreasing on ϕ(ξ)\phi(\xi). By the ordinary differential equation theory, system (2.15) has a unique solution (SX(ξ),IX(ξ))(S_{X}(\xi),I_{X}(\xi)) satisfying (SX(ξ),IX(ξ))C1([X,X],2)(S_{X}(\xi),I_{X}(\xi))\in C^{1}([-X,X],\mathbb{R}^{2}). Then we define an operator

𝒜=(𝒜1,𝒜2):ΓXC1([X,X],2)\mathcal{A}=(\mathcal{A}_{1},\mathcal{A}_{2}):\Gamma_{X}\rightarrow C^{1}\left([-X,X],\mathbb{R}^{2}\right)

by

SX(ξ)=𝒜1(ϕ,ψ)(ξ)andIX(ξ)=𝒜2(ϕ,ψ)(ξ).S_{X}(\xi)=\mathcal{A}_{1}(\phi,\psi)(\xi)\ \ {\rm and}\ \ I_{X}(\xi)=\mathcal{A}_{2}(\phi,\psi)(\xi).

Next we show the operator 𝒜=(𝒜1,𝒜2)\mathcal{A}=(\mathcal{A}_{1},\mathcal{A}_{2}) has a fixed point in ΓX\Gamma_{X} by Schauder’s fixed point theorem (see [7, Corollary 2.3.10]).

Lemma 2.7.

The operator 𝒜=(𝒜1,𝒜2)\mathcal{A}=(\mathcal{A}_{1},\mathcal{A}_{2}) maps ΓX\Gamma_{X} into itself.

Proof.

Firstly, we show that S(ξ)SX(ξ)S^{-}(\xi)\leq S_{X}(\xi) for any ξ[X,X].\xi\in[-X,X]. If ξ(𝔛1,X),\xi\in(\mathfrak{X}_{1},X), it follows that S(ξ)=0S^{-}(\xi)=0 and is a lower solution of the first equation of (2.15). If ξ(X,𝔛1),\xi\in(-X,\mathfrak{X}_{1}), then S(ξ)=S0(1M1eε1ξ)S^{-}(\xi)=S_{0}(1-M_{1}e^{\varepsilon_{1}\xi}), from Lemma 2.5, we have

cS(ξ)+(2d1+μ1+α)S(ξ)d1ϕ^(ξ+1)d1ϕ^(ξ1)Λαϕ(ξ)+βϕ(ξ)f(ψ(ξ))\displaystyle c{S^{-}}^{\prime}(\xi)+(2d_{1}+\mu_{1}+\alpha)S^{-}(\xi)-d_{1}\hat{\phi}(\xi+1)-d_{1}\hat{\phi}(\xi-1)-\Lambda-\alpha\phi(\xi)+\beta\phi(\xi)f(\psi(\xi))
\displaystyle\leq cS(ξ)d1J[S](ξ)Λ+μ1S(ξ)+βS(ξ)f(I+(ξ))\displaystyle c{S^{-}}^{\prime}(\xi)-d_{1}J[S^{-}](\xi)-\Lambda+\mu_{1}S^{-}(\xi)+\beta S^{-}(\xi)f(I^{+}(\xi))
\displaystyle\leq 0,\displaystyle 0,

which implies that S(ξ)=S0(1M1eε1ξ)S^{-}(\xi)=S_{0}(1-M_{1}e^{\varepsilon_{1}\xi}) is a lower solution of the first equation of (2.15). Thus S(ξ)SX(ξ)S^{-}(\xi)\leq S_{X}(\xi) for any ξ[X,X].\xi\in[-X,X].

Secondly, we show that SX(ξ)S+(ξ)=S0S_{X}(\xi)\leq S^{+}(\xi)=S_{0} for any ξ[X,X].\xi\in[-X,X]. In fact,

cS+(ξ)+(2d1+μ1+α)S+(ξ)d1ϕ^(ξ+1)d1ϕ^(ξ1)Λαϕ(ξ)+βϕ(ξ)f(ψ(ξ))\displaystyle c{S^{+}}^{\prime}(\xi)+(2d_{1}+\mu_{1}+\alpha)S^{+}(\xi)-d_{1}\hat{\phi}(\xi+1)-d_{1}\hat{\phi}(\xi-1)-\Lambda-\alpha\phi(\xi)+\beta\phi(\xi)f(\psi(\xi))
\displaystyle\geq βS0f(I(ξ))\displaystyle\beta S_{0}f(I^{-}(\xi))
\displaystyle\geq 0,\displaystyle 0,

thus S+(ξ)=S0S^{+}(\xi)=S_{0} is an upper solution to the first equation of (2.15), which gives us SX(ξ)S0S_{X}(\xi)\leq S_{0} for any ξ[X,X].\xi\in[-X,X].

Similarly, we can show that I(ξ)IX(ξ)I+(ξ)I^{-}(\xi)\leq I_{X}(\xi)\leq I^{+}(\xi) for any ξ[X,X].\xi\in[-X,X]. This completes the proof. ∎

Lemma 2.8.

The operator 𝒜\mathcal{A} is completely continuous.

Proof.

Suppose (ϕi(ξ),ψi(ξ))ΓX,i=1,2.(\phi_{i}(\xi),\psi_{i}(\xi))\in\Gamma_{X},\ i=1,2. Denote

SX,i(ξ)=𝒜1(ϕi(ξ),ψi(ξ))andIX,i(ξ)=𝒜2(ϕi(ξ),ψi(ξ)),\displaystyle S_{X,i}(\xi)=\mathcal{A}_{1}(\phi_{i}(\xi),\psi_{i}(\xi))\ \ {\rm and}\ \ I_{X,i}(\xi)=\mathcal{A}_{2}(\phi_{i}(\xi),\psi_{i}(\xi)),

We show that the operator 𝒜\mathcal{A} is continuous. By direct calculation, we have

SX(ξ)=S(X)e2d1+μ1+αc(ξ+X)+1cXξe2d1+μ1+αc(τ+X)H1(ϕ,ψ)(τ)dτ,S_{X}(\xi)=S^{-}(-X)e^{-\frac{2d_{1}+\mu_{1}+\alpha}{c}(\xi+X)}+\frac{1}{c}\int_{-X}^{\xi}e^{\frac{2d_{1}+\mu_{1}+\alpha}{c}(\tau+X)}H_{1}(\phi,\psi)(\tau){\rm d}\tau,

and

IX(ξ)=I(X)e2d2+μ2c(ξ+X)+1cXξe2d2+μ2c(τ+X)H2(ϕ,ψ)(τ)dτ,I_{X}(\xi)=I^{-}(-X)e^{-\frac{2d_{2}+\mu_{2}}{c}(\xi+X)}+\frac{1}{c}\int_{-X}^{\xi}e^{\frac{2d_{2}+\mu_{2}}{c}(\tau+X)}H_{2}(\phi,\psi)(\tau){\rm d}\tau,

where Hi(ϕ,ψ)(i=1,2)H_{i}(\phi,\psi)(i=1,2) are defined in (2.15). For any (ϕi,ψi)ΓX(\phi_{i},\psi_{i})\in\Gamma_{X}, i=1,2i=1,2, we have

|ϕ1(ξ)f(ψ1(ξ))ϕ2(ξ)f(ψ2(ξ))|\displaystyle\ |\phi_{1}(\xi)f(\psi_{1}(\xi))-\phi_{2}(\xi)f(\psi_{2}(\xi))|
\displaystyle\leq |ϕ1(ξ)f(ψ1(ξ))ϕ1(ξ)f(ψ2(ξ))|+|ϕ1(ξ)f(ψ2(ξ))ϕ2(ξ)f(ψ2(ξ))|\displaystyle\ |\phi_{1}(\xi)f(\psi_{1}(\xi))-\phi_{1}(\xi)f(\psi_{2}(\xi))|+|\phi_{1}(\xi)f(\psi_{2}(\xi))-\phi_{2}(\xi)f(\psi_{2}(\xi))|
\displaystyle\leq S0f(0)maxξ[X,X]|ψ1(ξ)ψ2(ξ)|+f(0)eλ1Xmaxξ[X,X]|ϕ1(ξ)ϕ2(ξ)|.\displaystyle\ S_{0}f^{\prime}(0)\max_{\xi\in[-X,X]}|\psi_{1}(\xi)-\psi_{2}(\xi)|+f^{\prime}(0)e^{\lambda_{1}X}\max_{\xi\in[-X,X]}|\phi_{1}(\xi)-\phi_{2}(\xi)|.

Thus, it is easy to see that the operator 𝒜\mathcal{A} is continuous. Next, we show 𝒜\mathcal{A} is compact. Indeed, since SXS_{X} and IXI_{X} are class of C1([X,X])C^{1}([-X,X]), note that

|c(SX,1(ξ)SX,2(ξ))+(2d1+μ1)(SX,1(ξ)SX,2(ξ))|\displaystyle\ \left|c(S_{X,1}^{\prime}(\xi)-S_{X,2}^{\prime}(\xi))+(2d_{1}+\mu_{1})(S_{X,1}(\xi)-S_{X,2}(\xi))\right|
\displaystyle\leq d1|(ϕ^1(ξ+1)ϕ^2(ξ+1))|+d1|(ϕ^1(ξ1)ϕ^2(ξ1))|+β|ϕ1(ξ)f(ψ1(ξ))ϕ2(ξ)f(ψ2(ξ))|\displaystyle\ d_{1}|(\hat{\phi}_{1}(\xi+1)-\hat{\phi}_{2}(\xi+1))|+d_{1}|(\hat{\phi}_{1}(\xi-1)-\hat{\phi}_{2}(\xi-1))|+\beta|\phi_{1}(\xi)f(\psi_{1}(\xi))-\phi_{2}(\xi)f(\psi_{2}(\xi))|
\displaystyle\leq βS0f(0)maxξ[X,X]|ψ1(ξ)ψ2(ξ)|+(2d1+βf(0)eλ1X)maxξ[X,X]|ϕ1(ξ)ϕ2(ξ)|.\displaystyle\ \beta S_{0}f^{\prime}(0)\max_{\xi\in[-X,X]}|\psi_{1}(\xi)-\psi_{2}(\xi)|+\left(2d_{1}+\beta f^{\prime}(0)e^{\lambda_{1}X}\right)\max_{\xi\in[-X,X]}|\phi_{1}(\xi)-\phi_{2}(\xi)|.

Same arguments with IXI_{X}^{\prime}, give us that SXS_{X}^{\prime} and IXI_{X}^{\prime} are bounded. Then the operator 𝒜\mathcal{A} is compact and is completely continuous. This ends the proof. ∎

Applying Schauder’s fixed point theorem, we have the following lemma.

Lemma 2.9.

There exists (SX,IX)ΓX(S_{X},I_{X})\in\Gamma_{X} such that

(SX(ξ),IX(ξ))=𝒜(SX,IX)(ξ)(S_{X}(\xi),I_{X}(\xi))=\mathcal{A}(S_{X},I_{X})(\xi)

for ξ[X,X]\xi\in[-X,X].

In the following, we show some prior estimates for (SX,IX)(S_{X},I_{X}). Define

C1,1([X,X])={uC1([X,X])|u,uare Lipschitz continuous}C^{1,1}([-X,X])=\{u\in C^{1}([-X,X])\ |\ u,u^{\prime}\textrm{are Lipschitz continuous}\}

with the norm

uC1,1([X,X])=maxx[X,X]|u|+maxx[X,X]|u|+supx,y[X,X]xy|u(x)u(y)||xy|.\displaystyle\|u\|_{C^{1,1}([-X,X])}=\max_{x\in[-X,X]}|u|+\max_{x\in[-X,X]}|u^{\prime}|+\sup_{\begin{subarray}{c}x,y\in[-X,X]\\ x\neq y\end{subarray}}\frac{|u^{\prime}(x)-u^{\prime}(y)|}{|x-y|}.
Lemma 2.10.

There exists a constant C(Y)>0C(Y)>0 such that

SXC1,1([Y,Y])C(Y)andIXC1,1([Y,Y])C(Y)\|S_{X}\|_{C^{1,1}([-Y,Y])}\leq C(Y)\ \ {\rm and}\ \ \|I_{X}\|_{C^{1,1}([-Y,Y])}\leq C(Y)

for 0<Y<X0<Y<X and X>𝔛2X>-\mathfrak{X}_{2}.

Proof.

Recall that (SX,IX)(S_{X},I_{X}) is the fixed point of the operator 𝒜,\mathcal{A}, then

{cSX(ξ)=d1S^X(ξ+1)+d1S^X(ξ1)(2d1+μ1)SX(ξ)+ΛβSX(ξ)f(IX(ξ)),cIX(ξ)=d2I^X(ξ+1)+d2I^X(ξ1)(2d2+μ2)IX(ξ)+βSX(ξ)f(IX(ξ)),\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle cS_{X}^{\prime}(\xi)=d_{1}\hat{S}_{X}(\xi+1)+d_{1}\hat{S}_{X}(\xi-1)-(2d_{1}+\mu_{1})S_{X}(\xi)+\Lambda-\beta S_{X}(\xi)f(I_{X}(\xi)),\\ \displaystyle cI_{X}^{\prime}(\xi)=d_{2}\hat{I}_{X}(\xi+1)+d_{2}\hat{I}_{X}(\xi-1)-(2d_{2}+\mu_{2})I_{X}(\xi)+\beta S_{X}(\xi)f(I_{X}(\xi)),\end{array}\right. (2.16)

where

S^X(ξ)={SX(X),for ξ>X,SX(ξ),for ξ[X,X],S(ξ),for ξ<X,I^X(ξ)={IX(X),for ξ>X,IX(ξ),for ξ[X,X],I(ξ),for ξ<X.\hat{S}_{X}(\xi)=\left\{\begin{array}[]{ll}\displaystyle S_{X}(X),&\mbox{for $\xi>X$,}\\ \displaystyle S_{X}(\xi),&\mbox{for $\xi\in[-X,X]$,}\\ \displaystyle S^{-}(\xi),&\mbox{for $\xi<-X$,}\\ \end{array}\right.\ \ \ \hat{I}_{X}(\xi)=\left\{\begin{array}[]{ll}\displaystyle I_{X}(X),&\mbox{for $\xi>X$,}\\ \displaystyle I_{X}(\xi),&\mbox{for $\xi\in[-X,X]$,}\\ \displaystyle I^{-}(\xi),&\mbox{for $\xi<-X$.}\\ \end{array}\right.

Since 0SX(ξ)S00\leq S_{X}(\xi)\leq S_{0} and 0IX(ξ)eλ1Y0\leq I_{X}(\xi)\leq e^{\lambda_{1}Y} for all ξ[Y,Y]\xi\in[-Y,Y], from (2.16) we have

|SX(ξ)|4d1+μ1cS0+Λc+βS0f(0)ceλ1Y,|S_{X}^{\prime}(\xi)|\leq\frac{4d_{1}+\mu_{1}}{c}S_{0}+\frac{\Lambda}{c}+\frac{\beta S_{0}f^{\prime}(0)}{c}e^{\lambda_{1}Y},

and

|IX(ξ)|4d2+μ2+βS0f(0)ceλ1Y.|I_{X}^{\prime}(\xi)|\leq\frac{4d_{2}+\mu_{2}+\beta S_{0}f^{\prime}(0)}{c}e^{\lambda_{1}Y}.

Thus there exists some constant C1(Y)>0C_{1}(Y)>0 such that

SXC1([Y,Y])C1(Y)andIXC1([Y,Y])C1(Y).\|S_{X}\|_{C^{1}([-Y,Y])}\leq C_{1}(Y)\ \ {\rm and}\ \ \|I_{X}\|_{C^{1}([-Y,Y])}\leq C_{1}(Y).

For any ξ,η[Y,Y]\xi,\eta\in[-Y,Y], it follows from [48] that

|S^X(ξ+1)S^X(η+1)|\displaystyle|\hat{S}_{X}(\xi+1)-\hat{S}_{X}(\eta+1)|
=\displaystyle= {|SX(Y)SX(Y)|=0,for ξ+1,η+1>Y,|SX(ξ+1)SX(η+1)|C1(Y)|ξη|,for ξ+1,η+1<Y,|SX(ξ+1)SX(Y)|C1(Y)(Yξ1)C1(Y)|ξη|,for ξ+1<Y,η+1>Y,|SX(X)SX(η+1)|C1(Y)(Yη1)C1(Y)|ξη|,for ξ+1>Y,η+1<Y.\displaystyle\left\{\begin{array}[]{ll}\displaystyle|S_{X}(Y)-S_{X}(Y)|=0,&\mbox{for $\xi+1,\eta+1>Y$,}\\ \displaystyle|S_{X}(\xi+1)-S_{X}(\eta+1)|\leq C_{1}(Y)|\xi-\eta|,&\mbox{for $\xi+1,\eta+1<Y$,}\\ \displaystyle|S_{X}(\xi+1)-S_{X}(Y)|\leq C_{1}(Y)(Y-\xi-1)\leq C_{1}(Y)|\xi-\eta|,&\mbox{for $\xi+1<Y,\eta+1>Y$,}\\ \displaystyle|S_{X}(X)-S_{X}(\eta+1)|\leq C_{1}(Y)(Y-\eta-1)\leq C_{1}(Y)|\xi-\eta|,&\mbox{for $\xi+1>Y,\eta+1<Y$.}\end{array}\right.

Then |S^X(ξ+1)S^X(η+1)|C1(Y)|ξη||\hat{S}_{X}(\xi+1)-\hat{S}_{X}(\eta+1)|\leq C_{1}(Y)|\xi-\eta| for all ξ,η[Y,Y]\xi,\eta\in[-Y,Y]. Similarly, we have

|S^X(ξ1)S^X(η1)|C1(Y)|ξη||\hat{S}_{X}(\xi-1)-\hat{S}_{X}(\eta-1)|\leq C_{1}(Y)|\xi-\eta|

for all ξ,η[Y,Y]\xi,\eta\in[-Y,Y]. Furthermore

|βSX(ξ)f(IX(ξ))βSX(η)f(IX(η))|\displaystyle\ |\beta S_{X}(\xi)f(I_{X}(\xi))-\beta S_{X}(\eta)f(I_{X}(\eta))|
\displaystyle\leq |βSX(ξ)f(IX(ξ))βSX(ξ)f(IX(η))|+|βSX(ξ)f(IX(η))βSX(η)f(IX(η))|\displaystyle\ |\beta S_{X}(\xi)f(I_{X}(\xi))-\beta S_{X}(\xi)f(I_{X}(\eta))|+|\beta S_{X}(\xi)f(I_{X}(\eta))-\beta S_{X}(\eta)f(I_{X}(\eta))|
\displaystyle\leq βf(0)C1(Y)(|SX(ξ)SX(η)|+|IX(ξ)IX(η)|)\displaystyle\ \beta f^{\prime}(0)C_{1}(Y)\left(|S_{X}(\xi)-S_{X}(\eta)|+|I_{X}(\xi)-I_{X}(\eta)|\right)

for all ξ,η[Y,Y]\xi,\eta\in[-Y,Y]. Hence, there exist some constant C(Y)>0C(Y)>0 such that

SXC1,1([Y,Y])C(Y).\|S_{X}\|_{C^{1,1}([-Y,Y])}\leq C(Y).

Similarly,

IXC1,1([Y,Y])C(Y)\|I_{X}\|_{C^{1,1}([-Y,Y])}\leq C(Y)

for any Y<XY<X. This completes the proof. ∎

3. Existence of traveling wave solutions

We first state the main results of this section as follows.

Theorem 3.1.

For any wave speed c>cc>c^{*}, system (1.5) admits a nontrivial traveling wave solution (S(ξ),I(ξ))(S(\xi),I(\xi)) satisfying

SS(ξ)S+andII(ξ)I+in.S^{-}\leq S(\xi)\leq S^{+}\ \ {\rm and}\ \ I^{-}\leq I(\xi)\leq I^{+}\ \ {\rm in}\ \ \mathbb{R}.

Furthermore,

limξ(S(ξ),I(ξ))=(S0,0)andlimξ+(S(ξ),I(ξ))=(S,I).\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0)\ \ {\rm and}\ \ \lim_{\xi\rightarrow+\infty}(S(\xi),I(\xi))=(S^{*},I^{*}).

The proof of Theorem 3.1 is divided into the following serval steps.

Step 1. We show that system (1.5) admits a nontrivial traveling wave solution (S(ξ),I(ξ))(S(\xi),I(\xi)) in \mathbb{R} and satisfying limξ(S(ξ),I(ξ))=(S0,0)\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0).

Choose {Xn}n=1+\{X_{n}\}_{n=1}^{+\infty} be an increasing sequence such that Xn>𝔛2X_{n}>-\mathfrak{X}_{2}, Xn>YX_{n}>Y and Xn+X_{n}\rightarrow+\infty as n+n\rightarrow+\infty for all nn\in\mathbb{N}, where YY is from Lemma 2.10. Denote (Sn,In)ΓXn(S_{n},I_{n})\in\Gamma_{X_{n}} be the solution of system (2.15). For any NN\in\mathbb{N}, since the function I+(ξ)I^{+}(\xi) is bounded in [XN,XN][-X_{N},X_{N}], then the sequences

{Sn}nNand{In}nN\{S_{n}\}_{n\geq N}\ \ {\rm and}\ \ \{I_{n}\}_{n\geq N}

are uniformly bounded in [XN,XN][-X_{N},X_{N}]. Then by (2.15), we can obtain that

{Sn}nNand{In}nN\{S^{\prime}_{n}\}_{n\geq N}\ \ {\rm and}\ \ \{I^{\prime}_{n}\}_{n\geq N}

are also uniformly bounded in [XN,XN][-X_{N},X_{N}]. Again with (2.15), we can express Sn′′(ξ)S^{\prime\prime}_{n}(\xi) and In′′(ξ)I^{\prime\prime}_{n}(\xi) in terms of Sn(ξ)S_{n}(\xi), In(ξ)I_{n}(\xi), Sn(ξ±1)S_{n}(\xi\pm 1), In(ξ±1)I_{n}(\xi\pm 1), Sn(ξ±2)S_{n}(\xi\pm 2) and In(ξ±2)I_{n}(\xi\pm 2), which give us

{Sn′′}nNand{In′′}nN\{S^{\prime\prime}_{n}\}_{n\geq N}\ \ {\rm and}\ \ \{I^{\prime\prime}_{n}\}_{n\geq N}

are uniformly bounded in [XN+2,XN2][-X_{N}+2,X_{N}-2]. By the Arzela-Ascoli theorem (see [32, Theorem A5]), we can use a diagonal process to extract a subsequence, denote by {Snk}k\{S_{n_{k}}\}_{k\in\mathbb{N}} and {Ink}k\{I_{n_{k}}\}_{k\in\mathbb{N}} such that

SnkS,InkI,SnkSandInkIask+S_{n_{k}}\rightarrow S,\ I_{n_{k}}\rightarrow I,\ S^{\prime}_{n_{k}}\rightarrow S^{\prime}\ {\rm and}\ I^{\prime}_{n_{k}}\rightarrow I^{\prime}\ {\rm as}\ k\rightarrow+\infty

uniformly in any compact subinterval of \mathbb{R}, for some functions SS and II in C1()C^{1}(\mathbb{R}). Thus (S(ξ),I(ξ))(S(\xi),I(\xi)) is a solution of system (2.2) with

S(ξ)S(ξ)S+(ξ)andI(ξ)I(ξ)I+(ξ)in.S^{-}(\xi)\leq S(\xi)\leq S^{+}(\xi)\ \ {\rm and}\ \ I^{-}(\xi)\leq I(\xi)\leq I^{+}(\xi)\ \ {\rm in}\ \ \mathbb{R}.

Furthermore, by the definition of S(ξ)S^{-}(\xi) and I(ξ)I^{-}(\xi) from (2.8), it follows that

limξ(S(ξ),I(ξ))=(S0,0).\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0).

Step 2. We claim that the functions S(ξ)S(\xi) and I(ξ)I(\xi) satisfy 0<S(ξ)<S0andI(ξ)>0in0<S(\xi)<S_{0}\ \ {\rm and}\ \ I(\xi)>0\ \ {\rm in}\ \ \mathbb{R}.

We first show that S(ξ)>0S(\xi)>0 for all ξ\xi\in\mathbb{R}. Assume reversely, that is assume that if there exists some real number ξ0\xi_{0} such that S(ξ0)=0S(\xi_{0})=0, then S(ξ0)=0S^{\prime}(\xi_{0})=0 and J[S](ξ0)0J[S](\xi_{0})\geq 0. By the first equation of (2.2), we have

0=d1J[S](ξ0)+Λ>0,0=d_{1}J[S](\xi_{0})+\Lambda>0,

which is a contradiction. Thus S(ξ)>0S(\xi)>0 for all ξ\xi\in\mathbb{R}.

Next, we show that I(ξ)>0I(\xi)>0 for all ξ\xi\in\mathbb{R}. By way of contradiction, we assume that if there exists ξ1\xi_{1} such that I(ξ1)=0I(\xi_{1})=0 and I(ξ)>0I(\xi)>0 for all ξ<ξ1\xi<\xi_{1}. From the second equation of (2.2), we have

I(ξ1+1)+I(ξ11)=0.I(\xi_{1}+1)+I(\xi_{1}-1)=0.

Consequently, I(ξ1+1)=I(ξ11)=0I(\xi_{1}+1)=I(\xi_{1}-1)=0 since I(ξ)0I(\xi)\geq 0 in \mathbb{R}, which is a contradiction to the definition of ξ1\xi_{1}.

To show that S(ξ)<S0S(\xi)<S_{0} for all ξ\xi\in\mathbb{R}, we assume that if there exists ξ2\xi_{2} such that S(ξ2)=S0S(\xi_{2})=S_{0}, it is easy to obtain

0=d1J[S](ξ2)βS(ξ2)f(I(ξ2))<0.0=d_{1}J[S](\xi_{2})-\beta S(\xi_{2})f(I(\xi_{2}))<0.

This contradiction leads to S(ξ)<S0S(\xi)<S_{0} for all ξ\xi\in\mathbb{R}.

Step 3. Boundedness of traveling wave solutions S(ξ)S(\xi) and I(ξ)I(\xi) in \mathbb{R}.

We need to consider two cases of the nonlinear incidence function f(x)f(x). In fact, the function f(x)f(x) satisfying Assumption 1.1 has two possibilities: (i) limx+f(x)\lim\limits_{x\rightarrow+\infty}f(x) exists; (ii)limx+f(x)=+\lim\limits_{x\rightarrow+\infty}f(x)=+\infty. For example, the saturated incidence with f(x)=bx1+cxf(x)=\frac{bx}{1+cx} satisfy (i) since limx+bx1+cx=bc\lim\limits_{x\rightarrow+\infty}\frac{bx}{1+cx}=\frac{b}{c} and the bilinear incidence with f(x)=bxf(x)=bx satisfy (ii).

Case 1. limx+f(x)\lim\limits_{x\rightarrow+\infty}f(x) exists. Without losing generality, we assume that limx+f(x)=f¯<+\lim\limits_{x\rightarrow+\infty}f(x)=\bar{f}<+\infty, then it is easy to verify that Λμ1+βf¯\frac{\Lambda}{\mu_{1}+\beta\bar{f}} is a lower solution of S(ξ)S(\xi) and βS0f¯μ2\frac{\beta S_{0}\bar{f}}{\mu_{2}} is an upper solution of I(ξ)I(\xi). Then we obtain

Λμ1+βf¯S(ξ)<S0and 0<I(ξ)βS0f¯μ2forallξ.\frac{\Lambda}{\mu_{1}+\beta\bar{f}}\leq S(\xi)<S_{0}\ \ {\rm and}\ \ 0<I(\xi)\leq\frac{\beta S_{0}\bar{f}}{\mu_{2}}\ \ {\rm for\ \ all}\ \ \xi\in\mathbb{R}. (3.1)

Case 2. limx+f(x)=+\lim\limits_{x\rightarrow+\infty}f(x)=+\infty. In this case, we have the following lemmas.

Lemma 3.2.

The functions I(ξ±1)I(ξ)\frac{I(\xi\pm 1)}{I(\xi)} and I(ξ)I(ξ)\frac{I^{\prime}(\xi)}{I(\xi)} are bounded in \mathbb{R}.

See Appendix A for the details of proof.

Lemma 3.3.

Let {ck,Sk,Ik}\{c_{k},S_{k},I_{k}\} be a sequence of traveling wave solutions of (1.5) with speed {ck}\{c_{k}\} in a compact subinterval of (0,)(0,\infty). If there is a sequence {ξk}\{\xi_{k}\} such that I(ξk)+I(\xi_{k})\rightarrow+\infty as k+k\rightarrow+\infty, then S(ξk)0S(\xi_{k})\rightarrow 0 as k+k\rightarrow+\infty.

See Appendix B for the details of proof.

Lemma 3.4.

If lim supξ+I(ξ)=+\limsup\limits_{\xi\rightarrow+\infty}I(\xi)=+\infty, then limξ+I(ξ)=+\lim\limits_{\xi\rightarrow+\infty}I(\xi)=+\infty.

The proof of Lemma 3.4 is similar to that of [9, Lemma 3.4], so we omit the details.

Lemma 3.5.

The function I(ξ)I(\xi) is bounded in \mathbb{R}.

See Appendix C for the details of proof.

By the above lemmas, we know that f(I(ξ))f(I(\xi)) is bounded from above since I(ξ)I(\xi) is bounded, then Proposition (3.1) follows. Hence, we obtained that S(ξ)S(\xi) and I(ξ)I(\xi) are bounded from above and S(ξ)S(\xi) has a strictly positive lower bound in \mathbb{R}. In the following, we will show I(ξ)I(\xi) could not approach zero.

Lemma 3.6.

Let 0<c1c20<c_{1}\leq c_{2} be given and (S(ξ),I(ξ))(S(\xi),I(\xi)) be a solution of system (2.2) with speed c[c1,c2]c\in[c_{1},c_{2}] satisfying 0<S(ξ)<S0andI(ξ)>0in0<S(\xi)<S_{0}\ \ {\rm and}\ \ I(\xi)>0\ \ {\rm in}\ \ \mathbb{R}. Then there exists some small enough constant ε0>0\varepsilon_{0}>0, such that I(ξ)>0I^{\prime}(\xi)>0 provided that I(ξ)ε0I(\xi)\leq\varepsilon_{0} for all ξ\xi\in\mathbb{R}.

See Appendix D for the details of proof.

Step 4. Convergence of the traveling wave solutions as ξ+\xi\rightarrow+\infty. The key point is to construct a suitable Lyapunov functional.

Let g(x)=x1lnxg(x)=x-1-\ln x, it is easy to check g(x)0g(x)\geq 0 since g(x)g(x) has the global minimum value 0 only at x=1x=1. Define the following Lyapunov functional:

L(S,I)(ξ)=W1(S,I)(ξ)+d1SW2(S,I)(ξ)+d2IW3(S,I)(ξ),L(S,I)(\xi)=W_{1}(S,I)(\xi)+d_{1}S^{*}W_{2}(S,I)(\xi)+d_{2}I^{*}W_{3}(S,I)(\xi),

where

W1(S,I)(ξ)=cSg(S(ξ)S)+cIg(I(ξ)I),W_{1}(S,I)(\xi)=cS^{*}g\left(\frac{S(\xi)}{S^{*}}\right)+cI^{*}g\left(\frac{I(\xi)}{I^{*}}\right),
W2(S,I)(ξ)=01g(S(ξθ)S)dθ10g(S(ξθ)S)dθW_{2}(S,I)(\xi)=\int_{0}^{1}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta-\int_{-1}^{0}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta

and

W3(S,I)(ξ)=01g(I(ξθ)I)dθ10g(I(ξθ)I)dθ.W_{3}(S,I)(\xi)=\int_{0}^{1}g\left(\frac{I(\xi-\theta)}{I^{*}}\right){\rm d}\theta-\int_{-1}^{0}g\left(\frac{I(\xi-\theta)}{I^{*}}\right){\rm d}\theta.

Thanks to the boundedness of S(ξ)S(\xi) and I(ξ)I(\xi) (see Step 3), we have W1(S,I)(ξ)W_{1}(S,I)(\xi) and W2(S,I)(ξ)W_{2}(S,I)(\xi) are well defined and bounded from below. Since limξI(ξ)=0\lim_{\xi\rightarrow-\infty}I(\xi)=0, we need to consider the process of ξ\xi approaching negative infinity for W3(S,I)(ξ)W_{3}(S,I)(\xi). For the ε0\varepsilon_{0} in Lemma 3.6, define ξ=min{ξ|I(ξ)=ε0}\xi^{*}=\min\{\xi\in\mathbb{R}|I(\xi)=\varepsilon_{0}\}, then I(ξ)I(\xi) is increasing in (,ξ](-\infty,\xi^{*}]. By the properties of function gg, we have W3(S,I)(ξ)0W_{3}(S,I)(\xi)\geq 0 for ξ(,ξ]\xi\in(-\infty,\xi^{*}]. Thus the Lyapunov function L(S,I)(ξ)L(S,I)(\xi) is well defined and bounded from below.

Next we show that the map ξL(S,I)(ξ)\xi\mapsto L(S,I)(\xi) is non-increasing. The derivative of W1(S,I)(ξ)W_{1}(S,I)(\xi) along the solution of (2.2) is calculated as follows

dW1(S,I)(ξ)dξ|(2.2)=\displaystyle\frac{{\rm d}W_{1}(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}= (1SS(ξ))cdS(ξ)dξ+(1II(ξ))cdI(ξ)dξ\displaystyle\left(1-\frac{S^{*}}{S(\xi)}\right)c\frac{{\rm d}S(\xi)}{{\rm d}\xi}+\left(1-\frac{I^{*}}{I(\xi)}\right)c\frac{{\rm d}I(\xi)}{{\rm d}\xi}
=\displaystyle= (1SS(ξ))(d1J[S](ξ)+Λμ1S(ξ)βS(ξ)f(I(ξ)))\displaystyle\left(1-\frac{S^{*}}{S(\xi)}\right)\left(d_{1}J[S](\xi)+\Lambda-\mu_{1}S(\xi)-\beta S(\xi)f(I(\xi))\right)
+(1II(ξ))(d2J[I](ξ)+βS(ξ)f(I(ξ))μ2I(ξ))\displaystyle+\left(1-\frac{I^{*}}{I(\xi)}\right)\left(d_{2}J[I](\xi)+\beta S(\xi)f(I(\xi))-\mu_{2}I(\xi)\right)
=\displaystyle= (1SS(ξ))d1J[S](ξ)+(1II(ξ))d2J[I](ξ)+Θ(ξ),\displaystyle\left(1-\frac{S^{*}}{S(\xi)}\right)d_{1}J[S](\xi)+\left(1-\frac{I^{*}}{I(\xi)}\right)d_{2}J[I](\xi)+\Theta(\xi),

where

Θ(ξ)=(1SS(ξ))(Λμ1S(ξ)βS(ξ)f(I(ξ)))+(1II(ξ))(βS(ξ)f(I(ξ))μ2I(ξ)).\Theta(\xi)=\left(1-\frac{S^{*}}{S(\xi)}\right)\left(\Lambda-\mu_{1}S(\xi)-\beta S(\xi)f(I(\xi))\right)+\left(1-\frac{I^{*}}{I(\xi)}\right)\left(\beta S(\xi)f(I(\xi))-\mu_{2}I(\xi)\right).

Note that the endemic equilibrium (S,I)(S^{*},I^{*}) of system (1.5) satisfying (1.7) and μ1=μ+α\mu_{1}=\mu+\alpha. By some calculation, we obtain

Θ(ξ)=\displaystyle\Theta(\xi)= μ1S(2SS(ξ)S(ξ)S)βSf(I)[g(SS(ξ))+g(IS(ξ)f(I(ξ))I(ξ)Sf(I))]\displaystyle\ \mu_{1}S^{*}\left(2-\frac{S^{*}}{S(\xi)}-\frac{S(\xi)}{S^{*}}\right)-\beta S^{*}f(I^{*})\left[g\left(\frac{S^{*}}{S(\xi)}\right)+g\left(\frac{I^{*}S(\xi)f(I(\xi))}{I(\xi)S^{*}f(I^{*})}\right)\right]
βSf(I)[g(I(ξ)I)g(f(I(ξ))f(I))].\displaystyle-\beta S^{*}f(I^{*})\left[g\left(\frac{I(\xi)}{I^{*}}\right)-g\left(\frac{f(I(\xi))}{f(I^{*})}\right)\right].

For W2(S,I)(ξ)W_{2}(S,I)(\xi), one has that

dW2(S,I)(ξ)dξ|(2.2)=\displaystyle\frac{{\rm d}W_{2}(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}= ddξ[01g(S(ξθ)S)dθ10g(S(ξθ)S)dθ]\displaystyle\ \frac{{\rm d}}{{\rm d}\xi}\left[\int_{0}^{1}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta-\int_{-1}^{0}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta\right]
=\displaystyle= 01ddξg(S(ξθ)S)dθ10ddξg(S(ξθ)S)dθ\displaystyle\ \int_{0}^{1}\frac{{\rm d}}{{\rm d}\xi}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta-\int_{-1}^{0}\frac{{\rm d}}{{\rm d}\xi}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta
=\displaystyle= 01ddθg(S(ξθ)S)dθ+10ddθg(S(ξθ)S)dθ\displaystyle\ -\int_{0}^{1}\frac{{\rm d}}{{\rm d}\theta}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta+\int_{-1}^{0}\frac{{\rm d}}{{\rm d}\theta}g\left(\frac{S(\xi-\theta)}{S^{*}}\right){\rm d}\theta
=\displaystyle= 2g(S(ξ)S)g(S(ξ1)S)g(S(ξ+1)S).\displaystyle\ 2g\left(\frac{S(\xi)}{S^{*}}\right)-g\left(\frac{S(\xi-1)}{S^{*}}\right)-g\left(\frac{S(\xi+1)}{S^{*}}\right).

Similarly,

dW3(S,I)(ξ)dξ|(2.2)=2g(I(ξ)I)g(I(ξ1)I)g(I(ξ+1)I).\frac{{\rm d}W_{3}(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}=2g\left(\frac{I(\xi)}{I^{*}}\right)-g\left(\frac{I(\xi-1)}{I^{*}}\right)-g\left(\frac{I(\xi+1)}{I^{*}}\right).

It can be shown that

(1SS(ξ))d1J[S](ξ)+SdW2(S,I)(ξ)dξ|(2.2)=d1S[g(S(ξ1)S(ξ))+g(S(ξ+1)S(ξ))],\left(1-\frac{S^{*}}{S(\xi)}\right)d_{1}J[S](\xi)+S^{*}\frac{{\rm d}W_{2}(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}=-d_{1}S^{*}\left[g\left(\frac{S(\xi-1)}{S(\xi)}\right)+g\left(\frac{S(\xi+1)}{S(\xi)}\right)\right],

and

(1II(ξ))d2J[I](ξ)+IdW3(S,I)(ξ)dξ|(2.2)=d2I[g(I(ξ1)I(ξ))+g(I(ξ+1)I(ξ))].\left(1-\frac{I^{*}}{I(\xi)}\right)d_{2}J[I](\xi)+I^{*}\frac{{\rm d}W_{3}(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}=-d_{2}I^{*}\left[g\left(\frac{I(\xi-1)}{I(\xi)}\right)+g\left(\frac{I(\xi+1)}{I(\xi)}\right)\right].

Thus

dL(S,I)(ξ)dξ|(2.2)=\displaystyle\frac{{\rm d}L(S,I)(\xi)}{{\rm d}\xi}\bigg{|}_{\eqref{WaveEqu}}= d1S[g(S(ξ1)S(ξ))+g(S(ξ+1)S(ξ))]d2I[g(I(ξ1)I(ξ))+g(I(ξ+1)I(ξ))]\displaystyle\ -d_{1}S^{*}\left[g\left(\frac{S(\xi-1)}{S(\xi)}\right)+g\left(\frac{S(\xi+1)}{S(\xi)}\right)\right]-d_{2}I^{*}\left[g\left(\frac{I(\xi-1)}{I(\xi)}\right)+g\left(\frac{I(\xi+1)}{I(\xi)}\right)\right]
β1Sf(I)[g(SS(ξ))+g(S(ξ)f(I(ξ))ISf(I)I(ξ))+g(I(ξ)I)g(f(I(ξ))f(I))]\displaystyle\ -\beta_{1}S^{*}f(I^{*})\left[g\left(\frac{S^{*}}{S(\xi)}\right)+g\left(\frac{S(\xi)f(I(\xi))I^{*}}{S^{*}f(I^{*})I(\xi)}\right)+g\left(\frac{I(\xi)}{I^{*}}\right)-g\left(\frac{f(I(\xi))}{f(I^{*})}\right)\right]
+μS(2SS(ξ)S(ξ)S).\displaystyle\ +\mu S^{*}\left(2-\frac{S^{*}}{S(\xi)}-\frac{S(\xi)}{S^{*}}\right).

Since the arithmetic mean of non-negative real numbers is greater than or equal to the geometric mean of the same list, then we have

2SS(ξ)S(ξ)S0.2-\frac{S^{*}}{S(\xi)}-\frac{S(\xi)}{S^{*}}\leq 0.

From Assumption 1.1, we can conclude that

(1f(I)f(I))(f(I)f(I)II)0.\left(1-\frac{f(I^{*})}{f(I)}\right)\left(\frac{f(I)}{f(I^{*})}-\frac{I}{I^{*}}\right)\leq 0.

Then, we have

g(f(I(ξ))f(I))g(I(ξ)I)=\displaystyle g\left(\frac{f(I(\xi))}{f(I^{*})}\right)-g\left(\frac{I(\xi)}{I^{*}}\right)=\ f(I)f(I)II+ln(If(I)If(I))\displaystyle\frac{f(I)}{f(I^{*})}-\frac{I}{I^{*}}+\ln\left(\frac{If(I^{*})}{I^{*}f(I)}\right)
\displaystyle\leq\ f(I)f(I)II+If(I)If(I)1\displaystyle\frac{f(I)}{f(I^{*})}-\frac{I}{I^{*}}+\frac{If(I^{*})}{I^{*}f(I)}-1
=\displaystyle=\ (1f(I)f(I))(f(I)f(I)II)\displaystyle\left(1-\frac{f(I^{*})}{f(I)}\right)\left(\frac{f(I)}{f(I^{*})}-\frac{I}{I^{*}}\right)
\displaystyle\leq\ 0.\displaystyle 0.

Here we use If(I)If(I)1ln(If(I)If(I))0\frac{If(I^{*})}{I^{*}f(I)}-1-\ln\left(\frac{If(I^{*})}{I^{*}f(I)}\right)\geq 0. Hence, the map ξL(S,I)(ξ)\xi\mapsto L(S,I)(\xi) is non-increasing. Consider an increasing sequence {ξk}k0\{\xi_{k}\}_{k\geq 0} with ξk>0\xi_{k}>0 such that ξk+\xi_{k}\rightarrow+\infty when k+k\rightarrow+\infty and denote

{Sk(ξ)=S(ξ+ξk)}k0and{Ik(ξ)=I(ξ+ξk)}k0.\{S_{k}(\xi)=S(\xi+\xi_{k})\}_{k\geq 0}\ \ \textrm{and}\ \ \{I_{k}(\xi)=I(\xi+\xi_{k})\}_{k\geq 0}.

Since the functions SS and II are bounded, the system (2.2) give us that the functions SS and II have bounded derivatives. Then by Arzela-Ascoli theorem, the functions {Sk(ξ)}\{S_{k}(\xi)\} and {Ik(ξ)}\{I_{k}(\xi)\} converge in Cloc()C_{loc}^{\infty}(\mathbb{R}) as k+k\rightarrow+\infty, up to extraction of a subsequence, one may assume that the sequences of {Sk(ξ)}\{S_{k}(\xi)\} and {Ik(ξ)}\{I_{k}(\xi)\} convergence towards some nonnegative CC^{\infty} functions SS_{\infty} and I.I_{\infty}. Furthermore, since L(S,I)(ξ)L(S,I)(\xi) is non-increasing on ξ\xi and bounded from below, then there exists a constant C0C_{0} and large kk such that

C0L(Sk,Ik)(ξ)=L(S,I)(ξ+ξk)L(S,I)(ξ).C_{0}\leq L(S_{k},I_{k})(\xi)=L(S,I)(\xi+\xi_{k})\leq L(S,I)(\xi).

Therefore there exists some δ\delta\in\mathbb{R} such that limkL(Sk,Ik)(ξ)=δ\lim\limits_{k\rightarrow\infty}L(S_{k},I_{k})(\xi)=\delta for any ξ\xi\in\mathbb{R}. By Lebesgue’s dominated convergence theorem (see [31, Theorem 11.32]), we have

limk+L(Sk,Ik)(ξ)=L(S,I)(ξ),ξ.\lim_{k\rightarrow+\infty}L(S_{k},I_{k})(\xi)=L(S_{\infty},I_{\infty})(\xi),\ \xi\in\mathbb{R}.

Thus

L(S,I)(ξ)=δ.L(S_{\infty},I_{\infty})(\xi)=\delta.

Note that dLdξ=0\frac{{\rm d}L}{{\rm d}\xi}=0 if and only if S(ξ)SS(\xi)\equiv S^{*} and I(ξ)II(\xi)\equiv I^{*}, it follows that

(S,I)(S,I).(S_{\infty},I_{\infty})\equiv(S^{*},I^{*}).

Hence, we complete the proof of Theorem 3.1.

Theorem 3.7.

For the wave speed c=cc=c^{*}, system (1.5) admits a nontrivial traveling wave solution (S(ξ),I(ξ))(S(\xi),I(\xi)) satisfying

SS(ξ)S+andII(ξ)I+in.S^{-}\leq S(\xi)\leq S^{+}\ \ {\rm and}\ \ I^{-}\leq I(\xi)\leq I^{+}\ \ {\rm in}\ \ \mathbb{R}.

Furthermore,

limξ(S(ξ),I(ξ))=(S0,0)andlimξ+(S(ξ),I(ξ))=(S,I).\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0)\ \ {\rm and}\ \ \lim_{\xi\rightarrow+\infty}(S(\xi),I(\xi))=(S^{*},I^{*}).

For the case c=cc=c^{*}, we can obtain the existence of traveling wave solutions (S(n+ct),I(n+ct))(S(n+c^{*}t),I(n+c^{*}t)) by an approximation technique used in [9, Section 4]. Since the Lyapunov functional is independent on cc, we can also have that (S(n+ct),I(n+ct))(S(n+c^{*}t),I(n+c^{*}t)) satisfies the asymptotic boundary conditions (2.3) and (2.4). So we omit the details.

4. Nonexistence of traveling wave solutions

In this section, we study the nonexistence of traveling wave solutions. Firstly, we show that c>0c>0 if there exists a nontrivial positive solution (S(ξ),I(ξ))(S(\xi),I(\xi)) of system (2.2) satisfying the asymptotic boundary conditions (2.3) and (2.4).

Lemma 4.1.

Assume that 0>1\Re_{0}>1 and there exists a nontrivial solution (S(ξ),I(ξ))(S(\xi),I(\xi)) of system (1.5) satisfying the asymptotic boundary conditions (2.3) and (2.4). Then c>0c>0, where cc is defined in (2.1).

Proof.

Assume that c0c\leq 0. Since S(ξ)S0andI(ξ)0asξS(\xi)\rightarrow S_{0}\ \ {\rm and}\ \ I(\xi)\rightarrow 0\ \ {\rm as}\ \ \xi\rightarrow-\infty, there exists a ξ<0\xi^{*}<0 such that

cI(ξ)d2[I(ξ+1)+I(ξ1)2I(ξ)]+βS0f(0)+μ22f(0)(f(0)ϵ)I(ξ)μ2I(ξ),cI^{\prime}(\xi)\geq d_{2}[I(\xi+1)+I(\xi-1)-2I(\xi)]+\frac{\beta S_{0}f^{\prime}(0)+\mu_{2}}{2f^{\prime}(0)}(f^{\prime}(0)-\epsilon)I(\xi)-\mu_{2}I(\xi), (4.1)

here we used the condition 0>1\Re_{0}>1. Note that inequality (4.1) is valid for any ϵ(0,f(0))\epsilon\in(0,f^{\prime}(0)), then for ξ<ξ\xi<\xi^{*}, we have

cI(ξ)d2[I(ξ+1)+I(ξ1)2I(ξ)]+βS0f(0)μ22I(ξ).cI^{\prime}(\xi)\geq d_{2}[I(\xi+1)+I(\xi-1)-2I(\xi)]+\frac{\beta S_{0}f^{\prime}(0)-\mu_{2}}{2}I(\xi). (4.2)

Denote ω=βS0f(0)μ22\omega=\frac{\beta S_{0}f^{\prime}(0)-\mu_{2}}{2} and Q(ξ)=ξI(y)dyQ(\xi)=\int_{-\infty}^{\xi}I(y){\rm d}y for ξ\xi\in\mathbb{R}, note that ω>0\omega>0 since 0>1\Re_{0}>1. Integrating inequality (4.1) from -\infty to ξ\xi and using I()=0I(-\infty)=0, one has that

cI(ξ)d2[Q(ξ+1)+Q(ξ1)2Q(ξ)]+ωQ(ξ)forξ<ξ.cI(\xi)\geq d_{2}[Q(\xi+1)+Q(\xi-1)-2Q(\xi)]+\omega Q(\xi)\ \ \textrm{for}\ \ \xi<\xi^{*}. (4.3)

Again, Integrating inequality (4.1) from -\infty to ξ\xi, yields

cQ(ξ)d2(ξξ+1Q(τ)dτξ1ξQ(τ)dτ)+ωξQ(τ)dτforξ<ξ.cQ(\xi)\geq d_{2}\left(\int_{\xi}^{\xi+1}Q(\tau){\rm d}\tau-\int_{\xi-1}^{\xi}Q(\tau){\rm d}\tau\right)+\omega\int_{-\infty}^{\xi}Q(\tau){\rm d}\tau\ \ \textrm{for}\ \ \xi<\xi^{*}. (4.4)

Since Q(ξ)Q(\xi) is strictly increasing in \mathbb{R} and c0c\leq 0, we can conclude that

0cQ(ξ)d2(ξξ+1Q(τ)dτξ1ξQ(τ)dτ)+ωξQ(τ)dτ>0,0\geq cQ(\xi)\geq d_{2}\left(\int_{\xi}^{\xi+1}Q(\tau){\rm d}\tau-\int_{\xi-1}^{\xi}Q(\tau){\rm d}\tau\right)+\omega\int_{-\infty}^{\xi}Q(\tau){\rm d}\tau>0,

which is a contradiction. Hence c>0c>0. The proof is finished. ∎

Now, we are in position to show the nonexistence of traveling wave solutions, we will use two-sided Laplace to prove it (see [2, 46, 52])

Theorem 4.2.

Assume that 0>1\Re_{0}>1 and c<cc<c^{*}. Then there are no nontrivial solution (S(ξ),I(ξ))(S(\xi),I(\xi)) of system (1.5) satisfying the asymptotic boundary conditions (2.3) and (2.4).

Proof.

By way of contradiction, assume that there exists a nontrivial positive solution (S(ξ),I(ξ))(S(\xi),I(\xi)) of system (1.5) satisfying the asymptotic boundary condition (2.3) and (2.4). Then c>0c>0 by Lemma 4.1 and

S(ξ)S0andI(ξ)0asξ.S(\xi)\rightarrow S_{0}\ \ {\rm and}\ \ I(\xi)\rightarrow 0\ \ {\rm as}\ \ \xi\rightarrow-\infty.

Let ω=βS0f(0)μ22\omega=\frac{\beta S_{0}f^{\prime}(0)-\mu_{2}}{2} and Q(ξ)=ξI(y)dyQ(\xi)=\int_{-\infty}^{\xi}I(y){\rm d}y for ξ\xi\in\mathbb{R}. It follows from the proof of Lemma 4.1, there exists a ξ<0\xi^{*}<0, we have

cQ(ξ)d2(ξξ+1Q(τ)dτξ1ξQ(τ)dτ)+ωξQ(τ)dτforξ<ξ.cQ(\xi)\geq d_{2}\left(\int_{\xi}^{\xi+1}Q(\tau){\rm d}\tau-\int_{\xi-1}^{\xi}Q(\tau){\rm d}\tau\right)+\omega\int_{-\infty}^{\xi}Q(\tau){\rm d}\tau\ \ \textrm{for}\ \ \xi<\xi^{*}.

Recalling that Q(ξ)Q(\xi) is strictly increasing in \mathbb{R}, one has that

d2(ξξ+1Q(τ)dτξ1ξQ(τ)dτ)>0.d_{2}\left(\int_{\xi}^{\xi+1}Q(\tau){\rm d}\tau-\int_{\xi-1}^{\xi}Q(\tau){\rm d}\tau\right)>0.

Thus

cQ(ξ)ωξQ(τ)dτforξ<ξ.cQ(\xi)\geq\omega\int_{-\infty}^{\xi}Q(\tau){\rm d}\tau\ \ \textrm{for}\ \ \xi<\xi^{*}. (4.5)

Hence, there exists some constant δ>0\delta>0 such that

ωδQ(ξδ)cQ(ξ)forξ<ξ.\omega\delta Q(\xi-\delta)\leq cQ(\xi)\ \ \textrm{for}\ \ \xi<\xi^{*}. (4.6)

Moreover, there exists a ν>0\nu>0 is large enough and ϵ0(0,1)\epsilon_{0}\in(0,1) such that

Q(ξν)ϵ0Q(ξ)forξ<ξ.Q(\xi-\nu)\leq\epsilon_{0}Q(\xi)\ \ \textrm{for}\ \ \xi<\xi^{*}. (4.7)

Set

μ0:=1νln1ϵ0andV(ξ):=Q(ξ)eμ0ξ.\mu_{0}:=\frac{1}{\nu}\ln\frac{1}{\epsilon_{0}}\ \ \ \textrm{and}\ \ \ V(\xi):=Q(\xi)e^{-\mu_{0}\xi}.

We have

V(ξν)=Q(ξν)eμ0(ξν)<ϵ0Q(ξ)eμ0(ξν)=V(ξ)forξ<ξ,V(\xi-\nu)=Q(\xi-\nu)e^{-\mu_{0}(\xi-\nu)}<\epsilon_{0}Q(\xi)e^{-\mu_{0}(\xi-\nu)}=V(\xi)\ \ \textrm{for}\ \ \xi<\xi^{*},

which implies that V(ξ)V(\xi) is bounded as ξ\xi\rightarrow-\infty. Since I(ξ)dξ<\int_{-\infty}^{\infty}I(\xi){\rm d}\xi<\infty, we obtain that

limξV(ξ)=limξQ(ξ)eμ0ξ=0.\lim_{\xi\rightarrow\infty}V(\xi)=\lim_{\xi\rightarrow\infty}Q(\xi)e^{-\mu_{0}\xi}=0.

From the second equation of (2.2), we have

cI(ξ)d2[I(ξ+1)+I(ξ1)2I(ξ)]+βS0f(0)I(ξ)μ2I(ξ),cI^{\prime}(\xi)\leq d_{2}[I(\xi+1)+I(\xi-1)-2I(\xi)]+\beta S_{0}f^{\prime}(0)I(\xi)-\mu_{2}I(\xi),

integrating over (,ξ)(-\infty,\xi), give us

cI(ξ)d2[Q(ξ+1)+Q(ξ1)2Q(ξ)]+βS0f(0)Q(ξ)μ2Q(ξ).cI(\xi)\leq d_{2}[Q(\xi+1)+Q(\xi-1)-2Q(\xi)]+\beta S_{0}f^{\prime}(0)Q(\xi)-\mu_{2}Q(\xi).

Hence, we can obtain that

supξ{I(ξ)eμ0ξ}<andsupξ{I(ξ)eμ0ξ}<.\sup_{\xi\in\mathbb{R}}\left\{I(\xi)e^{-\mu_{0}\xi}\right\}<\infty\ \ \ \textrm{and}\ \ \ \sup_{\xi\in\mathbb{R}}\left\{I^{\prime}(\xi)e^{-\mu_{0}\xi}\right\}<\infty.

For λ\lambda\in\mathbb{C} with 0<Reλ<μ00<\textrm{Re}\lambda<\mu_{0}, define the following two-sided Laplace transform of I()I(\cdot) by

(λ):=eλξI(ξ)dξ.\mathcal{L}(\lambda):=\int_{-\infty}^{\infty}e^{-\lambda\xi}I(\xi){\rm d}\xi.

Note that

eλξ[I(ξ+1)+I(ξ1)]dξ\displaystyle\ \int_{-\infty}^{\infty}e^{-\lambda\xi}[I(\xi+1)+I(\xi-1)]{\rm d}\xi
=\displaystyle= eλeλ(ξ+1)I(ξ+1)dξ+eλeλ(ξ1)I(ξ1)dξ\displaystyle\ e^{\lambda}\int_{-\infty}^{\infty}e^{-\lambda(\xi+1)}I(\xi+1){\rm d}\xi+e^{-\lambda}\int_{-\infty}^{\infty}e^{-\lambda(\xi-1)}I(\xi-1){\rm d}\xi
=\displaystyle= (eλ+eλ)(λ)\displaystyle\ \left(e^{\lambda}+e^{-\lambda}\right)\mathcal{L}(\lambda)

and

eλξI(ξ)dξ=eλI(ξ)|I(ξ)deλξ=λ(λ).\displaystyle\int_{-\infty}^{\infty}e^{-\lambda\xi}I^{\prime}(\xi){\rm d}\xi=e^{\lambda}I(\xi)\bigg{|}_{-\infty}^{\infty}-\int_{-\infty}^{\infty}I(\xi){\rm d}e^{-\lambda\xi}=\lambda\mathcal{L}(\lambda).

From the second equation of (2.2), we have

d2J[I](ξ)+βS0f(0)I(ξ)μ2I(ξ)cI(ξ)=βS0f(0)I(ξ)βS(ξ)f(I(ξ)).d_{2}J[I](\xi)+\beta S_{0}f^{\prime}(0)I(\xi)-\mu_{2}I(\xi)-cI^{\prime}(\xi)=\beta S_{0}f^{\prime}(0)I(\xi)-\beta S(\xi)f(I(\xi)). (4.8)

Taking two-sided Laplace transform on (4.8), give us

Δ(λ,c)(λ)=eλξ[βS0f(0)I(ξ)βS(ξ)f(I(ξ))]dξ.\Delta(\lambda,c)\mathcal{L}(\lambda)=\int_{-\infty}^{\infty}e^{-\lambda\xi}\left[\beta S_{0}f^{\prime}(0)I(\xi)-\beta S(\xi)f(I(\xi))\right]{\rm d}\xi. (4.9)

It follows from the proof in Lemma 2.6, as ξ\xi\rightarrow-\infty, we have

[βS0f(0)I(ξ)βS(ξ)f(I(ξ))]e2μ0ξ\displaystyle\left[\beta S_{0}f^{\prime}(0)I(\xi)-\beta S(\xi)f(I(\xi))\right]e^{-2\mu_{0}\xi}\leq I2(ξ)e2μ0ξ\displaystyle\ I^{2}(\xi)e^{-2\mu_{0}\xi}
\displaystyle\leq (supξ{I(ξ)eμ0ξ})2\displaystyle\ \left(\sup_{\xi\in\mathbb{R}}\left\{I(\xi)e^{-\mu_{0}\xi}\right\}\right)^{2}
\displaystyle\leq .\displaystyle\ \infty.

Thus, we can obtain that

supξ[βS0f(0)I(ξ)βS(ξ)f(I(ξ))]e2μ0ξ<.\sup_{\xi\in\mathbb{R}}\left[\beta S_{0}f^{\prime}(0)I(\xi)-\beta S(\xi)f(I(\xi))\right]e^{-2\mu_{0}\xi}<\infty. (4.10)

By the property of Laplace transform [39], either there exist a real number μ0\mu_{0} such that (λ)\mathcal{L}(\lambda) is analytic for λ\lambda\in\mathbb{C} with 0<Reλ<μ00<\textrm{Re}\lambda<\mu_{0} and λ=μ0\lambda=\mu_{0} is singular point of (λ)\mathcal{L}(\lambda), or (λ)\mathcal{L}(\lambda) is well defined for λ\lambda\in\mathbb{C} with Reλ>0\textrm{Re}\lambda>0. Furthermore, the two Laplace integrals can be analytically continued to the whole right half line; otherwise, the integral on the left of (4.9) has singularity at λ=μ0\lambda=\mu_{0} and it is analytic for all λ<μ0\lambda<\mu_{0}. However, it follows from (4.10) that the integral on the right of (4.9) is actually analytic for all λ2μ0\lambda\leq 2\mu_{0}, a contradiction. Thus, (4.9) holds for all Reλ>0\textrm{Re}\lambda>0. From Lemma 2.1, Δ(λ,c)>0\Delta(\lambda,c)>0 for all λ>0\lambda>0 and by the definition of Δ(λ,c)\Delta(\lambda,c) in (2.6), we know that Δ(λ,c)\Delta(\lambda,c)\rightarrow\infty as λ\lambda\rightarrow\infty, which is a contradiction with Equation (4.9). This ends the proof. ∎

5. Application and discussion

As an application, we consider the following two discrete diffusive epidemic models. The first one is a model with saturated incidence rate which has been wildly used in epidemic modeling (see, for example, [28, 48, 44, 50, 45]).

Example.

Discrete diffusive epidemic model with saturated incidence rate:

{dSn(t)dt=d1[Sn+1(t)+Sn1(t)2Sn(t)]+ΛβSn(t)In(t)1+αIn(t)μ1Sn(t),dIn(t)dt=d2[In+1(t)+In1(t)2In(t)]+βSn(t)In(t)1+αIn(t)γIn(t)μ1In(t),\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=d_{1}[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]+\Lambda-\frac{\beta S_{n}(t)I_{n}(t)}{1+\alpha I_{n}(t)}-\mu_{1}S_{n}(t),\\ \displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d_{2}[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\frac{\beta S_{n}(t)I_{n}(t)}{1+\alpha I_{n}(t)}-\gamma I_{n}(t)-\mu_{1}I_{n}(t),\end{array}\right. (5.1)

where βIn(t)\beta I_{n}(t) is the force of infection and 11+αIn(t)\frac{1}{1+\alpha I_{n}(t)} measures the inhibition effect which is dependent on the infected individuals.

Setting f(In(t))=βSn(t)In(t)1+αIn(t)f(I_{n}(t))=\frac{\beta S_{n}(t)I_{n}(t)}{1+\alpha I_{n}(t)} in the original system (1.5), we can easily see that (5.1) is a special case of (1.5). In fact, it is obvious that f(In(t))f(I_{n}(t)) satisfy Assumption 1.1. The disease-free equilibrium of system (5.1) is similar to the original one, which is E~0=(S0,0)\tilde{E}_{0}=(S_{0},0). Moreover, we obtain the basic reproduction number of system (5.1) as 1=βS0γ+μ1\Re_{1}=\frac{\beta S_{0}}{\gamma+\mu_{1}} and there exists a positive equilibrium E~=(S~,I~)\tilde{E}^{*}=(\tilde{S}^{*},\tilde{I}^{*}) if 1>1\Re_{1}>1, where

S~=αΛ+γ+μ1β+αμ1andI~=Λβμ1(γ+μ1)(γ+μ1)(β+αμ1).\tilde{S}^{*}=\frac{\alpha\Lambda+\gamma+\mu_{1}}{\beta+\alpha\mu_{1}}\ \ \ {\rm and}\ \ \ \tilde{I}^{*}=\frac{\Lambda\beta-\mu_{1}(\gamma+\mu_{1})}{(\gamma+\mu_{1})(\beta+\alpha\mu_{1})}.

Hence, from Theorems 3.1, 3.7 and 4.2, we obtain the following corollary

Corollary 5.1.

Assume that 1>1\Re_{1}>1. Then there exists some c>0c^{*}>0 such that for any ccc\geq c^{*}, system (5.1) admits a traveling wave solution (S(ξ),I(ξ))(S(\xi),I(\xi)) satisfying

limξ(S(ξ),I(ξ))=(S0,0)andlimξ+(S(ξ),I(ξ))=(S~,I~).\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0)\ \ {\rm and}\ \ \lim_{\xi\rightarrow+\infty}(S(\xi),I(\xi))=(\tilde{S}^{*},\tilde{I}^{*}). (5.2)

Furthermore, system (5.1) admits no traveling wave solutions satisfying (5.2) when c<cc<c^{*}.

The next example was studied in [9], and our results will solve the open problem proposed in [9], which is the traveling wave solutions converge to the endemic equilibrium as ξ+\xi\rightarrow+\infty for discrete diffusive system (1.3).

Example.

Discrete diffusive epidemic model with mass action infection mechanism:

{dSn(t)dt=d1[Sn+1(t)+Sn1(t)2Sn(t)]+ΛβSn(t)In(t)μ1Sn(t),dIn(t)dt=d2[In+1(t)+In1(t)2In(t)]+βSn(t)In(t)γIn(t)μ1In(t).\left\{\begin{array}[]{l}\vspace{2mm}\displaystyle\frac{{\rm d}S_{n}(t)}{{\rm d}t}=d_{1}[S_{n+1}(t)+S_{n-1}(t)-2S_{n}(t)]+\Lambda-\beta S_{n}(t)I_{n}(t)-\mu_{1}S_{n}(t),\\ \displaystyle\frac{{\rm d}I_{n}(t)}{{\rm d}t}=d_{2}[I_{n+1}(t)+I_{n-1}(t)-2I_{n}(t)]+\beta S_{n}(t)I_{n}(t)-\gamma I_{n}(t)-\mu_{1}I_{n}(t).\end{array}\right. (5.3)

Setting f(In(t))=βSn(t)In(t)f(I_{n}(t))=\beta S_{n}(t)I_{n}(t) in the original system (1.5), we can easily see that (5.3) is a special case of (1.5) and this model has been studied in [9]. The disease-free equilibrium of system (5.3) is E¯0=(S0,0)\bar{E}_{0}=(S_{0},0). Moreover, we obtain the basic reproduction number of system (5.3) is the same with (5.1) as 1=βS0γ+μ1\Re_{1}=\frac{\beta S_{0}}{\gamma+\mu_{1}} and there exists a positive equilibrium E¯=(S¯,I¯)\bar{E}^{*}=(\bar{S}^{*},\bar{I}^{*}) if 1>1\Re_{1}>1, where

S¯=γ+μ1βandI¯=Λμ1S¯βS¯.\bar{S}^{*}=\frac{\gamma+\mu_{1}}{\beta}\ \ \ {\rm and}\ \ \ \bar{I}^{*}=\frac{\Lambda-\mu_{1}\bar{S}^{*}}{\beta\bar{S}^{*}}.

Then, from Theorems 3.1, 3.7 and 4.2, we obtain the following corollary

Corollary 5.2.

Assume that 1>1\Re_{1}>1. Then there exists some c>0c^{*}>0 such that for any ccc\geq c^{*}, system (5.3) admits a traveling wave solution (S(ξ),I(ξ))(S(\xi),I(\xi)) satisfying

limξ(S(ξ),I(ξ))=(S0,0)andlimξ+(S(ξ),I(ξ))=(S¯,I¯).\lim_{\xi\rightarrow-\infty}(S(\xi),I(\xi))=(S_{0},0)\ \ {\rm and}\ \ \lim_{\xi\rightarrow+\infty}(S(\xi),I(\xi))=(\bar{S}^{*},\bar{I}^{*}). (5.4)

Furthermore, system (5.3) admits no traveling wave solutions satisfying (5.4) when c<cc<c^{*}.

Note that Corollary 5.2 could answer the open problem proposed in [9], that is the traveling wave solutions for system (1.4) converge to the endemic equilibrium at ++\infty.

Next, we show that how the parameters affect wave speed. Suppose (λ^,c^)(\hat{\lambda},\hat{c}) be a zero root of Δ(λ,c)\Delta(\lambda,c) which defined in (2.6), a direct calculation yields

dc^dβ=S0f(0)λ>0,dc^dd2=eλ+eλ2λ>0anddc^d0=μ2λ>0.\frac{{\rm d}\hat{c}}{{\rm d}\beta}=\frac{S_{0}f^{\prime}(0)}{\lambda}>0,\ \ \frac{{\rm d}\hat{c}}{{\rm d}d_{2}}=\frac{e^{\lambda}+e^{-\lambda}-2}{\lambda}>0\ \ {\rm and}\ \ \frac{{\rm d}\hat{c}}{{\rm d}\Re_{0}}=\frac{\mu_{2}}{\lambda}>0.

that is, c^\hat{c} is an increasing function on β\beta, d2d_{2} and 0\Re_{0}. Biologically, this means that the diffusion and infection ability of infected individuals can accelerate the speed of disease spreading.

Now, we are in a position to make the following summary:

In this paper, a discrete diffusive epidemic model with nonlinear incidence rate has been investigated. When the basic reproduction number 0>1\Re_{0}>1, we proved that there exists a critical wave speed c>0c^{*}>0, such that for each ccc\geq c^{*} the system (1.5) admits a nontrivial traveling wave solution. Moreover, we used a Lyapunov functional to establish the convergence of traveling wave solutions at ++\infty. We also showed the nonexistence nontrivial traveling wave solutions when 0>1\Re_{0}>1 and c<cc<c^{*}. As special example of the model (1.5), we considered two different discrete diffusive epidemic model and apply our general results to show the conditions of existence and nonexistence of traveling wave solutions for the model (5.1). One of the example is studied in [9] and our result solved the open problem proposed in [9], which is the traveling wave solutions converge to the endemic equilibrium as ξ+\xi\rightarrow+\infty for discrete diffusive system (1.3).

Here we mention some functions f(I)f(I) considered in the literature that do not satisfy Assumptions 1.1. For example, the incidence rates with media impact f(I)=IemIf(I)=Ie^{-mI} in [10]; the specific incidence rate f(I)=kI1+αI2f(I)=\frac{kI}{1+\alpha I^{2}} in [42]; and the nonmonotone incidence rate f(I)=kI1+βI+αI2f(I)=\frac{kI}{1+\beta I+\alpha I^{2}} in [43]. In a recent paper, Shu et al. [34] studied a SIR model with non-monotone incidence rates and without constant recruitment, they investigated the existence and nonexistence of traveling wave solutions. What is the condition of existence and nonexistence of traveling wave solution for our model (1.5) with non-monotone incidence rates, which will be an interesting question and we leave this for future work.

References

  • [1] Anderson R.M., May R.M.: Infectious Diseases in Humans: Dynamics and Control. Oxford University Press, Oxford (1991)
  • [2] Bai Z., Zhang S.: Traveling waves of a diffusive SIR epidemic model with a class of nonlinear incidence rates and distributed delay. Commun. Nonlinear Sci. Numer. Simulat. 22, 1370–1381 (2015)
  • [3] Bates P.W., Chmaj A.: A discrete convolution model for phase transitions. Arch. Ration. Mech. Anal. 150, 281–305 (1999)
  • [4] Briggs C.J., Godfray H.C.J.: The dynamics of insect-pathogen interactions in stage-structured populations. Am. Nat. 145, 855–887 (1995)
  • [5] Brucal-Hallare M., Vleck E.V.: Traveling wavefronts in an antidiffusion lattice Nagumo model. SIAM J. Appl. Dyn. Syst. 10, 921–959 (2011)
  • [6] Capasso V., Serio G.: A generalization of the Kermack-Mackendric deterministic model. Math. Biosci. 42, 43–61 (1978)
  • [7] Chang K.-C.: Methods in Nonlinear Analysis. Springer Monographs in Mathematics. Springer-Verlag, Berlin (2005)
  • [8] Chen X., Guo J.-S.: Uniqueness and existence of traveling waves for discrete quasilinear monostable dynamics. Math. Ann. 326, 123–146 (2003)
  • [9] Chen Y.-Y., Guo J.-S., Hamel F.: Traveling waves for a lattice dynamical system arising in a diffusive endemic model. Nonlinearity 30, 2334–2359 (2017)
  • [10] Cui J., Sun Y., Zhu H.: The impact of media on the control of infectious diseases. J. Dynam. Differential Equations 20, 31–53 (2008)
  • [11] Ducrot A., Magal P.: Travelling wave solutions for an infection-age structured epidemic model with external supplies. Nonlinearity 24, 2891–2911 (2011)
  • [12] Erneux T., Nicolis G.: Propagating waves in discrete bistable reaction diffusion systems. Physica D 67, 237–244 (1993)
  • [13] Fang J., Wei J., Zhao X.-Q.: Spreading speeds and travelling waves for non-monotone time-delayed lattice equations. Proc. R. Soc. A-Math. Phys. Eng. Sci. 466, 1919–1934 (2010)
  • [14] Fu S.-C., Guo J.-S., Wu C.-C.: Traveling wave solutions for a discrete diffusive epidemic model. J. Nonlinear Convex Anal. 17, 1739–1751 (2016)
  • [15] Fu S-C.: Traveling waves for a diffusive SIR model with delay. J. Math. Anal. Appl. 435, 20–37 (2016)
  • [16] Guo J.-S., Wu C.-H.: Traveling wave front for a two-component lattice dynamical system arising in competition models. J. Differ. Equ. 252, 4357–4391 (2012)
  • [17] Han X., Kloeden P.E.: Lattice dynamical systems in the biological sciences. In Yin G., Zhang Q. (eds) Modeling, Stochastic Control, Optimization, and Applications. Springer, Cham (2019)
  • [18] He J., Tsai J.-C.: Traveling waves in the Kermark-McKendrick epidemic model with latent period. Z. Angew. Math. Phys. 70: 27, 22 pp. (2019)
  • [19] Heesterbeek J.A.P., Metz J.A.J.: The saturating contact rate in marriage and epidemic models. J. Math. Biol. 31, 529–539 (1993)
  • [20] Hethcote H.W.: The mathematics of infectious diseases. SIAM Rev. 42, 599–653 (2000)
  • [21] Hosono Y., Ilyas B.: Traveling waves for a simple diffusive epidemic model. Math. Models Methods Appl. Sci. 5, 935–966 (1995)
  • [22] Kapral R.: Discrete models for chemically reacting systems. J. Math. Chem. 6, 113–163 (1991)
  • [23] Kermack W., McKendrick A.: A contribution to mathematical theory of epidemics. Proc. R. Soc. A-Math. Phys. Eng. Sci. 115, 700–721 (1927)
  • [24] Korobeinikov A., Maini P.K.: Nonlinear incidence and stability of infectious disease models. Math. Med. Biol. 22, 113–128 (2005)
  • [25] Korobeinikov A.: Lyapunov functions and global stability for SIR and SIRS epidemiological models with non-linear transmission. Bull. Math. Biol. 68, 615–626 (2006)
  • [26] Lam K.-Y., Wang X., Zhang T.: Traveling waves for a class of diffusive disease-transmission models with network structures. SIAM J. Math. Anal. 50, 5719–5748 (2018)
  • [27] Li W.-T., Xu W.-B., Zhang L.: Traveling waves and entire solutions for an epidemic model with asymmetric dispersal. Discret. Contin. Dyn. Syst. 37, 2483–2512 (2017)
  • [28] Li Y., Li W.-T., Yang F.-Y.: Traveling waves for a nonlocal dispersal SIR model with delay and external supplies. Appl. Math. Comput. 247, 723–740 (2014)
  • [29] Liu W. M., Levin S. A., Iwasa X.: Influence of nonlinear incidence rates upon the behaviour of SIRS epidemiological models. J. Math. Biol. 23, 187–204 (1986)
  • [30] Muroya Y., Kuniya T., Enatsu Y.: Global stability of a delayed multi-group SIRS epidemic model with nonlinear incidence rates and relapse of infection. Discret. Contin. Dyn. Syst. Ser. B 20, 3057–3091 (2015)
  • [31] Rudin W.: Principles of Mathematical Analysis. 3nd Edition. International Series in Pure and Applied Mathematics. McGraw-Hill, New York (1976)
  • [32] Rudin W.: Functional Analysis. 2nd Edition. International Series in Pure and Applied Mathematics. McGraw-Hill, New York (1991)
  • [33] San X.F., Wang Z.-C.: Traveling waves for a two-group epidemic model with latent period in a patchy environment. J. Math. Anal. Appl. 475, 1502–1531 (2019)
  • [34] Shu H., Pan X., Wang X.-S., Wu J.: Traveling waves in epidemic models: non-monotone diffusive systems with non-monotone incidence rates. J. Dynam. Differential Equations 31, 883–901 (2019)
  • [35] Thieme H.R.: Global stability of the endemic equilibrium in infinite dimension: Lyapunov functions and positive operators. J. Differ. Equ. 250, 3772–3801 (2011)
  • [36] Tian B., Yuan R.: Traveling waves for a diffusive SEIR epidemic model with standard incidences. Sci. China Math. 60, 813–832 (2017)
  • [37] Wang W., Ma W.: Global dynamics and travelling wave solutions for a class of non-cooperative reaction-diffusion systems with nonlocal infections. Discret. Contin. Dyn. Syst. Ser. B 23, 3213–3235 (2018)
  • [38] Weng P., Huang H., Wu J.: Asymptotic speed of propagation of wave fronts in a lattice delay differential equation with global interaction. IMA J. Appl. Math. 68, 409–439 (2003)
  • [39] Widder D.V.: The Laplace Transform. Princeton Mathematical Series 6, Princeton University Press, Princeton (1941)
  • [40] Wu C.-C.: Existence of traveling waves with the critical speed for a discrete diffusive epidemic model. J. Differ. Equ. 262, 272–282 (2017)
  • [41] Wu S., Weng P., Ruan S.: Spatial dynamics of a lattice population model with two age classes and maturation delay. Euro. J. Appl. Math. 26, 61–91 (2015)
  • [42] Xiao D., Ruan S.: Global analysis of an epidemic model with a nonlinear incidence rate. Math. Biosci. 208, 419–429 (2007)
  • [43] Xiao D., Zhou Y.: Qualitative analysis of an epidemic model. Can. Appl. Math. Q 14, 469–492 (2006)
  • [44] Xu R., Ma Z.: Global stability of a SIR epidemic model with nonlinear incidence rate and time delay. Nonlinear Anal. Real World Appl. 10, 3175–3189 (2009)
  • [45] Xu Z., Guo T.: Traveling waves in a diffusive epidemic model with criss-cross mechanism. Math. Meth. Appl. Sci. 42, 2892–2908 (2019)
  • [46] Yang F.-Y., Li Y., Li W.-T., Wang Z.-C.: Traveling waves in a nonlocal dispersal Kermack-McKendrick epidemic model equation with monostable convolution type nonlinearity. Discret. Contin. Dyn. Syst. Ser. B 18, 1969–1993 (2013)
  • [47] Yang Z., Zhang G.: Stability of non-monotone traveling waves for a discrete diffusion equation with monostable convolution type nonlinearity. Sci. China Math. 61, 1789–1806 (2018)
  • [48] Zhang Q., Wu S.-L.: Wave propagation of a discrete SIR epidemic model with a saturated incidence rate. Int. J. Biomath. 12, 1950029, 18 pp (2009)
  • [49] Zhang S., Xu R.: Travelling waves and global attractivity of an SIRS disease model with spatial diffusion and temporary immunity. Appl. Math. Comput. 224, 635–651 (2013)
  • [50] Zhang Y., Li Y., Zhang Q., Li A.: Behavior of a stochastic SIR epidemic model with saturated incidence and vaccination rules. Physica A 501, 178–187 (2018)
  • [51] Zhao L., Wang Z.-C., Ruan S.: Traveling wave solutions in a two-group epidemic model with latent period. Nonlinearity 30, 1287–1325 (2017)
  • [52] Zhou J., Song L., Wei J.: Mixed types of waves in a discrete diffusive epidemic model with nonlinear incidence and time delay. J. Differ. Equ. Doi:10.1016/j.jde.2019.10.034

Appendix A: Proof of Lemma 3.2

Proof.

From the second equation of (2.2), one has that

cI(ξ)+(2d2+μ2)I(ξ)=d2I(ξ+1)+d2I(ξ1)+βS(ξ)f(I(ξ))>0.cI^{\prime}(\xi)+(2d_{2}+\mu_{2})I(\xi)=d_{2}I(\xi+1)+d_{2}I(\xi-1)+\beta S(\xi)f(I(\xi))>0.

Denote U(ξ)=eνξI(ξ)U(\xi)=e^{\nu\xi}I(\xi), where ν=(2d2+μ2)/c\nu=(2d_{2}+\mu_{2})/c. It follows that

cU(ξ)=eνξ(cI(ξ)+(2d2+μ2)I(ξ))>0,cU^{\prime}(\xi)=e^{\nu\xi}(cI^{\prime}(\xi)+(2d_{2}+\mu_{2})I(\xi))>0,

thus U(ξ)U(\xi) is increasing on ξ\xi. Then U(ξ1)<U(ξ)U(\xi-1)<U(\xi), that is

I(ξ1)I(ξ)<eνforallξ.\frac{I(\xi-1)}{I(\xi)}<e^{\nu}\ \ {\rm for}\ \ {\rm all}\ \ \xi\in\mathbb{R}.

Note that

[eνξI(ξ)]\displaystyle\left[e^{\nu\xi}I(\xi)\right]^{\prime} =1ceνξ[d2I(ξ+1)+d2I(ξ1)+βS(ξ)f(I(ξ))]\displaystyle=\frac{1}{c}e^{\nu\xi}\left[d_{2}I(\xi+1)+d_{2}I(\xi-1)+\beta S(\xi)f(I(\xi))\right]
>d2ceνξI(ξ+1).\displaystyle>\frac{d_{2}}{c}e^{\nu\xi}I(\xi+1). (5.5)

Integrating (Proof.) over [ξ,ξ+1][\xi,\xi+1] and using the fact that eνξI(ξ)e^{\nu\xi}I(\xi) is increasing, we have

eν(ξ+1)I(ξ+1)>\displaystyle e^{\nu(\xi+1)}I(\xi+1)\ > eνξI(ξ)+d2cξξ+1eνsI(s+1)ds\displaystyle\ e^{\nu\xi}I(\xi)+\frac{d_{2}}{c}\int_{\xi}^{\xi+1}e^{\nu s}I(s+1){\rm d}s
>\displaystyle> eνξI(ξ)+d2cξξ+1eν(ξ+1)I(ξ+1)eνds\displaystyle\ e^{\nu\xi}I(\xi)+\frac{d_{2}}{c}\int_{\xi}^{\xi+1}e^{\nu(\xi+1)}I(\xi+1)e^{-\nu}{\rm d}s
=\displaystyle= eν[I(ξ)+d2cI(ξ+1)].\displaystyle\ e^{-\nu}\left[I(\xi)+\frac{d_{2}}{c}I(\xi+1)\right].

By (Proof.), we obtain

[eνξI(ξ)]>(d2c)2e2νeν(ξ+1)I(ξ+1).\left[e^{\nu\xi}I(\xi)\right]^{\prime}>\left(\frac{d_{2}}{c}\right)^{2}e^{-2\nu}e^{\nu(\xi+1)}I(\xi+1). (5.6)

Integrating (5.6) over [ξ12,ξ][\xi-\frac{1}{2},\xi], yields

eνξI(ξ)>\displaystyle e^{\nu\xi}I(\xi)> (d2c)2e2νξ12ξeν(s+1)I(s+1)ds\displaystyle\left(\frac{d_{2}}{c}\right)^{2}e^{-2\nu}\int_{\xi-\frac{1}{2}}^{\xi}e^{\nu(s+1)}I(s+1){\rm d}s
>\displaystyle> (d2c)2e2ν2eν(ξ+12)I(ξ+12),\displaystyle\left(\frac{d_{2}}{c}\right)^{2}\frac{e^{-2\nu}}{2}e^{\nu(\xi+\frac{1}{2})}I\left(\xi+\frac{1}{2}\right),

that is

I(ξ+12)I(ξ)<2(cd2)2e32νforallξ.\frac{I\left(\xi+\frac{1}{2}\right)}{I(\xi)}<2\left(\frac{c}{d_{2}}\right)^{2}e^{\frac{3}{2}\nu}\ \ {\rm for}\ \ {\rm all}\ \ \xi\in\mathbb{R}.

Similarly, integrating (5.6) over [ξ,ξ+12][\xi,\xi+\frac{1}{2}], we have

I(ξ+1)I(ξ+12)<2(cd2)2e32νforallξ.\frac{I(\xi+1)}{I\left(\xi+\frac{1}{2}\right)}<2\left(\frac{c}{d_{2}}\right)^{2}e^{\frac{3}{2}\nu}\ \ {\rm for}\ \ {\rm all}\ \ \xi\in\mathbb{R}.

Thus

I(ξ+1)I(ξ)=I(ξ+12)I(ξ)I(ξ+1)I(ξ+12)<4(cd2)4e3νforallξ.\frac{I(\xi+1)}{I(\xi)}=\frac{I\left(\xi+\frac{1}{2}\right)}{I(\xi)}\frac{I(\xi+1)}{I\left(\xi+\frac{1}{2}\right)}<4\left(\frac{c}{d_{2}}\right)^{4}e^{3\nu}\ \ {\rm for}\ \ {\rm all}\ \ \xi\in\mathbb{R}.

By the second equation of (2.2), it follows that

cI(ξ)I(ξ)=\displaystyle c\frac{I^{\prime}(\xi)}{I(\xi)}= I(ξ+1)I(ξ)+I(ξ1)I(ξ)+βS(ξ)f(I(ξ))I(ξ)(2d2+μ2)\displaystyle\frac{I(\xi+1)}{I(\xi)}+\frac{I(\xi-1)}{I(\xi)}+\beta S(\xi)\frac{f(I(\xi))}{I(\xi)}-(2d_{2}+\mu_{2})
\displaystyle\leq I(ξ+1)I(ξ)+I(ξ1)I(ξ)+βS0f(0)(2d2+μ2),\displaystyle\frac{I(\xi+1)}{I(\xi)}+\frac{I(\xi-1)}{I(\xi)}+\beta S_{0}f^{\prime}(0)-(2d_{2}+\mu_{2}),

which gives us I(ξ)I(ξ)\frac{I^{\prime}(\xi)}{I(\xi)} is bounded in \mathbb{R}. The proof is end. ∎

Appendix B: Proof of Lemma 3.3

Proof.

Assume that there is a subsequence of {ξk}k\{\xi_{k}\}_{k\in\mathbb{N}} again denoted by ξk\xi_{k}, such that Ik(ξk)+I_{k}(\xi_{k})\rightarrow+\infty as k+k\rightarrow+\infty and Sk(ξk)εS_{k}(\xi_{k})\geq\varepsilon in \mathbb{R} for all kk\in\mathbb{N} with some positive constant ε\varepsilon. From the first equation of (2.2), we have

Sk(ξ)2S0+Λc~:=δ0in,S^{\prime}_{k}(\xi)\leq\frac{2S_{0}+\Lambda}{\tilde{c}}:=\delta_{0}\ \ \textrm{in}\ \ \mathbb{R},

where c~\tilde{c} is a positive lower bound of {ck}\{c_{k}\}. It follows that

Sk(ξ)ε2,ξ[ξkδ,ξk]S_{k}(\xi)\geq\frac{\varepsilon}{2},\ \ \ \forall\xi\in[\xi_{k}-\delta,\xi_{k}]

for all kk\in\mathbb{N}, where δ=εδ0\delta=\frac{\varepsilon}{\delta_{0}}. By Lemma 3.2, we can assume that |Ik(ξ)Ik(ξ)|<C0\bigg{|}\frac{I^{\prime}_{k}(\xi)}{I_{k}(\xi)}\bigg{|}<C_{0} for some C0>0C_{0}>0. Then

Ik(ξk)Ik(ξ)=exp{ξξkIk(s)Ik(s)ds}eC0δ,ξ[ξkδ,ξk]\frac{I_{k}(\xi_{k})}{I_{k}(\xi)}=\exp\left\{\int_{\xi}^{\xi_{k}}\frac{I^{\prime}_{k}(s)}{I_{k}(s)}{\rm d}s\right\}\leq e^{C_{0}\delta},\ \ \forall\xi\in[\xi_{k}-\delta,\xi_{k}]

for all kk\in\mathbb{N}. Thus

minξ[ξkδ,ξk]Ik(ξ)eC0δIk(ξk),\min_{\xi\in[\xi_{k}-\delta,\xi_{k}]}I_{k}(\xi)\geq e^{-C_{0}\delta}I_{k}(\xi_{k}),

which give us

minξ[ξkδ,ξk]Ik(ξ)+ask+\min_{\xi\in[\xi_{k}-\delta,\xi_{k}]}I_{k}(\xi)\rightarrow+\infty\ \ \textrm{as}\ \ k\rightarrow+\infty

since Ik(ξk)+I_{k}(\xi_{k})\rightarrow+\infty as k+k\rightarrow+\infty. Recalling we assumed that limx+f(x)=+\lim_{x\rightarrow+\infty}f(x)=+\infty in this case, one has that

maxξ[ξkδ,ξk]Sk(ξ)δ0βε2f(ϖk)ask+.\max_{\xi\in[\xi_{k}-\delta,\xi_{k}]}S^{\prime}_{k}(\xi)\leq\delta_{0}-\frac{\beta\varepsilon}{2}f(\varpi_{k})\rightarrow-\infty\ \ \textrm{as}\ \ k\rightarrow+\infty.

where ϖk:=minξ[ξkδ,ξk]Ik(ξ)\varpi_{k}:=\min\limits_{\xi\in[\xi_{k}-\delta,\xi_{k}]}I_{k}(\xi). Moreover, there exists some K>0K>0 such that

Sk(ξ)2S0δ,kKandξ[ξkδ,ξk].S^{\prime}_{k}(\xi)\leq-\frac{2S_{0}}{\delta},\ \ \forall k\geq K\ \ \textrm{and}\ \ \xi\in[\xi_{k}-\delta,\xi_{k}].

Note that Sk<S0S_{k}<S_{0} in \mathbb{R} for each kk\in\mathbb{N}. Hence Sk(ξk)S0S_{k}(\xi_{k})\leq-S_{0} for all kKk\geq K, which reduces to a contradiction since Sk(ξk)εS_{k}(\xi_{k})\geq\varepsilon in \mathbb{R} for all kk\in\mathbb{N} with some positive constant ε\varepsilon. This completes the proof. ∎

Appendix C: Proof of Lemma 3.5

Proof.

Assume that lim supξ+I(ξ)=+\limsup\limits_{\xi\rightarrow+\infty}I(\xi)=+\infty, then we have limξ+S(ξ)=0\lim\limits_{\xi\rightarrow+\infty}S(\xi)=0 by Lemma 3.3 and Lemma 3.4. Set θ(ξ)=I(ξ)I(ξ)\theta(\xi)=\frac{I^{\prime}(\xi)}{I(\xi)}, from the second equation of (2.2), we have

cθ(ξ)=d2eξξ+1θ(s)ds+d2eξξ1θ(s)ds(2d2+μ2)+B(ξ),c\theta(\xi)=d_{2}e^{\int_{\xi}^{\xi+1}\theta(s){\rm d}s}+d_{2}e^{\int_{\xi}^{\xi-1}\theta(s){\rm d}s}-(2d_{2}+\mu_{2})+B(\xi),

where

B(ξ)=βS(ξ)f(I(ξ))I(ξ).B(\xi)=\beta S(\xi)\frac{f(I(\xi))}{I(\xi)}.

It is easy to have that limξ+B(ξ)=0\lim\limits_{\xi\rightarrow+\infty}B(\xi)=0 since f(I(ξ))I(ξ)f(0)\frac{f(I(\xi))}{I(\xi)}\leq f^{\prime}(0) and limξ+S(ξ)=0\lim\limits_{\xi\rightarrow+\infty}S(\xi)=0. By using [8, Lemma 3.4], θ(ξ)\theta(\xi) has a finite limit ω\omega at ++\infty and satisfies the following equation

h(ω,c):=d2(eω+eω2)cωμ2=0.h(\omega,c):=d_{2}\left(e^{\omega}+e^{-\omega}-2\right)-c\omega-\mu_{2}=0.

By some calculations, we obtain

h(0,c)<0,h(ω,c)ω|ω=0<0,2h(ω,c)ω2>0andlimω+h(ω,c)=.h(0,c)<0,\ \ \frac{\partial h(\omega,c)}{\partial\omega}\bigg{|}_{\omega=0}<0,\ \ \frac{\partial^{2}h(\omega,c)}{\partial\omega^{2}}>0\ \ \textrm{and}\ \ \lim_{\omega\rightarrow+\infty}h(\omega,c)=-\infty.

Thus, there exists a unique positive real root ω0\omega_{0} of h(ω,c)=0h(\omega,c)=0. Recall that λ1\lambda_{1} is the smaller real root of (2.6) and λ2\lambda_{2} is the larger real root of (2.6). From Lemma 2.1, one has

d2(eλ2+eλ22)cλ2μ2=βS0f(0)<0,d_{2}\left(e^{\lambda_{2}}+e^{-\lambda_{2}}-2\right)-c\lambda_{2}-\mu_{2}=-\beta S_{0}f^{\prime}(0)<0,

thus, we have λ2<ω0\lambda_{2}<\omega_{0}. Since limξ+θ(ξ)=ω0\lim\limits_{\xi\rightarrow+\infty}\theta(\xi)=\omega_{0}, then there exists some ξ~\tilde{\xi} large enough such that

I(ξ)Cexp{(λ2+ω02)ξ}forallξξ~I(\xi)\geq C\exp\left\{\left(\frac{\lambda_{2}+\omega_{0}}{2}\right)\xi\right\}\ \ {\rm for}\ \ {\rm all}\ \ \xi\geq\tilde{\xi}

with some constant CC. This is a contradiction since I(ξ)eλ1ξI(\xi)\leq e^{\lambda_{1}\xi} in \mathbb{R} and λ1<ω0\lambda_{1}<\omega_{0}. This ends the proof. ∎

Appendix D: Proof of Lemma 3.6

Proof.

Assume by way of contradiction that there is no such ε0\varepsilon_{0}, that is there exist some sequence {ξk}k\{\xi_{k}\}_{k\in\mathbb{N}} with speed ck(c¯,c¯)c_{k}\in(\underline{c},\overline{c}) such that I(ξk)0I(\xi_{k})\rightarrow 0 as k+k\rightarrow+\infty and I(ξk)0I^{\prime}(\xi_{k})\leq 0, where c¯\underline{c} and c¯\overline{c} are two given positive real number. Denote

Sk(ξ):=S(ξk+ξ)andIk(ξ):=I(ξk+ξ).S_{k}(\xi):=S(\xi_{k}+\xi)\ \ \textrm{and}\ \ I_{k}(\xi):=I(\xi_{k}+\xi).

Thus we have Ik(0)0I_{k}(0)\rightarrow 0 as k+k\rightarrow+\infty and Ik(ξ)0I_{k}(\xi)\rightarrow 0 locally uniformly in \mathbb{R} as k+k\rightarrow+\infty. As a consequence, there also holds that Ik(ξ)0I_{k}^{\prime}(\xi)\rightarrow 0 locally uniformly in \mathbb{R} as k+k\rightarrow+\infty by the second equation of (2.2). From the proof of [9, Lemma 3.8], we can obtain that S=S0S_{\infty}=S_{0}. Let Ψk(ξ):=Ik(ξ)Ik(0)\Psi_{k}(\xi):=\frac{I_{k}(\xi)}{I_{k}(0)}. In the view of

Ψk(ξ)=Ik(ξ)Ik(0)=Ik(ξ)Ik(ξ)Ψk(ξ),\Psi_{k}^{\prime}(\xi)=\frac{I_{k}^{\prime}(\xi)}{I_{k}(0)}=\frac{I_{k}^{\prime}(\xi)}{I_{k}(\xi)}\Psi_{k}(\xi),

we have that Ψk(ξ)\Psi_{k}(\xi) and Ψk(ξ)\Psi_{k}^{\prime}(\xi) are also locally uniformly in \mathbb{R} as k+k\rightarrow+\infty. Letting k+k\rightarrow+\infty, thus

cΨ(ξ)=d2J[Ψ](ξ)+βS0f(Ψ(ξ))μ2Ψ(ξ).c_{\infty}\Psi_{\infty}^{\prime}(\xi)=d_{2}J[\Psi_{\infty}](\xi)+\beta S_{0}f(\Psi_{\infty}(\xi))-\mu_{2}\Psi_{\infty}(\xi).

We claim that Ψ(ξ)>0\Psi_{\infty}(\xi)>0 in \mathbb{R}. In fact, if there exists some ξ0\xi_{0} such that Ψ(ξ0)=0\Psi_{\infty}(\xi_{0})=0, we also have Ψ(ξ0)=0{\Psi}^{\prime}_{\infty}(\xi_{0})=0 since Ψ(ξ)0\Psi_{\infty}(\xi)\geq 0, then

0=d2(Ψ(ξ0+1)+Ψ(ξ01)).0=d_{2}(\Psi_{\infty}(\xi_{0}+1)+\Psi_{\infty}(\xi_{0}-1)).

Thus Ψ(ξ0+1)=Ψ(ξ01)=0\Psi_{\infty}(\xi_{0}+1)=\Psi_{\infty}(\xi_{0}-1)=0, it follows that Ψ(ξ0+m)=0\Psi_{\infty}(\xi_{0}+m)=0 for all mm\in\mathbb{Z}. Recall that cΨ(ξ)μ2Ψ(ξ)c_{\infty}\Psi_{\infty}^{\prime}(\xi)\geq-\mu_{2}\Psi_{\infty}(\xi), then the map ξΨ(ξ)eμ2ξc\xi\mapsto\Psi_{\infty}(\xi)e^{\frac{\mu_{2}\xi}{c_{\infty}}} is nondecreasing. Since it vanishes at ξ0+m\xi_{0}+m for all mm\in\mathbb{Z} and eμ2ξce^{\frac{\mu_{2}\xi}{c_{\infty}}} is increasing, one can concluded that Ψ=0\Psi_{\infty}=0 in \mathbb{R}, which is a contradicts with Ψ(0)=1\Psi_{\infty}(0)=1.

Denote 𝒵(ξ):=Ψ(ξ)Ψ(ξ)\mathcal{Z}(\xi):=\frac{\Psi_{\infty}^{\prime}(\xi)}{\Psi_{\infty}(\xi)}, it is easy to verify that 𝒵(ξ)\mathcal{Z}(\xi) satisfies

c𝒵(ξ)=d2eξξ+1𝒵(s)dsdy+d2eξξ1𝒵(s)dsdy2d2+βS0f(0)μ2,c_{\infty}\mathcal{Z}(\xi)=d_{2}e^{\int^{\xi+1}_{\xi}\mathcal{Z}(s){\rm d}s}{\rm d}y+d_{2}e^{\int^{\xi-1}_{\xi}\mathcal{Z}(s){\rm d}s}{\rm d}y-2d_{2}+\beta S_{0}f^{\prime}(0)-\mu_{2},\emph{} (5.7)

here we use Ik0I_{k}\rightarrow 0 and f(Ik)Ikf(0)\frac{f(I_{k})}{I_{k}}\rightarrow f^{\prime}(0) as k+k\rightarrow+\infty. Recalling [8, Lemma 3.4], 𝒵(ξ)\mathcal{Z}(\xi) has finite limits ω±\omega_{\pm} as ξ±\xi\rightarrow\pm\infty, where ω±\omega_{\pm} are roots of

cω±=d2(eω±+eω±2)+βS0f(0)μ2.c_{\infty}\omega_{\pm}=d_{2}\left(e^{\omega_{\pm}}+e^{-\omega_{\pm}}-2\right)+\beta S_{0}f^{\prime}(0)-\mu_{2}.

By the analogous arguments in Lemma 2.1, we have ω±>0\omega_{\pm}>0. Thus Ψ(ξ)\Psi_{\infty}^{\prime}(\xi) is positive at ±\pm\infty by the definition of 𝒵(ξ)\mathcal{Z}(\xi). Moreover, Ψ(ξ)>0\Psi_{\infty}^{\prime}(\xi)>0 for all ξ\xi\in\mathbb{R}. Indeed, if there exists some ξ\xi^{*} such that 𝒵(ξ)=inf𝒵(ξ)\mathcal{Z}(\xi^{*})=\inf_{\mathbb{R}}\mathcal{Z}(\xi), then 𝒵(ξ)=0\mathcal{Z}(\xi^{*})=0. Differentiating (5.7) gives us

c𝒵(ξ)=d2(𝒵(ξ+1)𝒵(ξ))Ψ(ξ+1)Ψ(ξ)+d2(𝒵(ξ1)𝒵(ξ))Ψ(ξ1)Ψ(ξ),c_{\infty}\mathcal{Z}^{\prime}(\xi)=d_{2}(\mathcal{Z}(\xi+1)-\mathcal{Z}(\xi))\frac{\Psi_{\infty}(\xi+1)}{\Psi_{\infty}(\xi)}+d_{2}(\mathcal{Z}(\xi-1)-\mathcal{Z}(\xi))\frac{\Psi_{\infty}(\xi-1)}{\Psi_{\infty}(\xi)},

it follows that

𝒵(ξ)=𝒵(ξ+1)=𝒵(ξ1).\mathcal{Z}(\xi^{*})=\mathcal{Z}(\xi^{*}+1)=\mathcal{Z}(\xi^{*}-1).

Hence 𝒵(ξ)=𝒵(ξ+m)\mathcal{Z}(\xi^{*})=\mathcal{Z}(\xi^{*}+m) for all mm\in\mathbb{Z}. Then, there is

inf𝒵(ξ)min{𝒵(+),𝒵()}>0.\inf_{\mathbb{R}}\mathcal{Z}(\xi)\geq\min\{\mathcal{Z}(+\infty),\mathcal{Z}(-\infty)\}>0.

So Ψ(ξ)>0\Psi_{\infty}^{\prime}(\xi)>0. From the definition of Ψ(ξ)\Psi_{\infty}(\xi), we have

0<Ψ(0)=limk+Ψk(0)=limk+Ik(0)Ik(0).0<\Psi_{\infty}^{\prime}(0)=\lim_{k\rightarrow+\infty}\Psi_{k}^{\prime}(0)=\lim_{k\rightarrow+\infty}\frac{I_{k}^{\prime}(0)}{I_{k}(0)}.

Thus, I(ξk)=Ik(0)>0I^{\prime}(\xi_{k})=I_{k}^{\prime}(0)>0, which is a contradiction. This completes the proof. ∎