Skip to main content
Infectious Disease Modelling logoLink to Infectious Disease Modelling
. 2022 Sep 27;7(4):605–624. doi: 10.1016/j.idm.2022.09.003

The impact factors of the risk index and diffusive dynamics of a SIS free boundary model

Yachun Tong a, Inkyung Ahn b, Zhigui Lin a,
PMCID: PMC9576570  PMID: 36262268

Abstract

To discuss the impact factors on the spread of infectious diseases, we study a free boundary problem describing a SIS (susceptible-infected-susceptible) model in a heterogeneous environment. Firstly, the existence and uniqueness of the global solution are given. Then the basic reproduction number related to time is defined, and a spreading-vanishing dichotomy of infectious diseases is obtained. The impacts of the diffusion rate of infected individuals, expanding capability, and the scope and scale of initial infection on the spreading and vanishing of infectious disease are analyzed. Numerical simulations are given to show that the large expanding capability is unfavorable to the prevention and control of the disease.

Keywords: SIS model, Free boundary, Risk index, Spreading and vanishing

MSC: primary, 35K57, 92D30, secondary, 35K55

1. Introduction

Infectious diseases are still the major causes of suffering and death in developing countries and developed countries (Hethcote, 2000), and usually have a significant impact on population size and historical events (McNeill, 1976). For example, the West Nile virus caused an encephalitis outbreak in New York City in 1999 (Center for Disease Control and Prevention (CDC), 1999). According to the latest real-time statistics from the World Health Organization (WHO), COVID-19, as of May 2022, has infected more than 516 million individuals, and at least 6.2 million people have died worldwide. However, it is impracticable to study the spreading of infectious diseases by experiments. To understand the spread and control of infectious diseases, mathematical models as an important tool have been attracting much attention (Abdelrazec, Belair, Shan, & Zhu, 2016; Hethcote, 2000; Keeling & Rohani, 2008; Martcheva, 2015; Thomas & Urena, 2001).

To discuss the impact of spatial heterogeneity of environment and movement of individuals on infectious diseases, an SIS epidemic reaction-diffusion system was proposed by Allen et al. in (Allen, Bolker, Lou, & Nevai, 2008), and the giving system is

StdSΔS=β(x)SIS+I+γ(x)I,xΩ,t>0,ItdIΔI=β(x)SIS+Iγ(x)I,xΩ,t>0,Sη=Iη=0,xΩ,t>0, (1.1)

where S(x, t) and I(x, t) denote the density of susceptible and infected individuals at location x and time t, respectively. The positive constants dS and dI account for the diffusion rates of susceptible and infected individuals, the positive bounded Ho¨lder continuous functions β(x) and γ(x) can be interpreted as rates of disease transmission and recovery for x ∈ Ω, respectively.

The authors in (Allen et al., 2008) characterized the risk of the region by the basic reproduction number R0. The DFE (disease-free equilibrium) is always unstable and there exists a unique EE (endemic equilibrium) in the high-risk domain (R0>1), and the DFE is stable if and only if infected individuals have mobility above a threshold value for the low-risk domain (R0<1). In addition, letting N = S + I, adding two equations in (1.1) and then integrating over Ω conclude that for t > 0, tΩ(S+I)dx=0, which implies that the total population size remains unchanged. More work on system (1.1) with constant total population size can be found in the literature (Cui & Lou, 2016; Peng, 2009; Peng & Liu, 2009; Peng & Zhao, 2012).

Considering changeable total population size, Li, Peng and Wang (Li, Peng, & Wang, 2017) studied the following SIS epidemic with a linear external source

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

where dS, dI, β, γ, S and I have the same epidemiological interpretations as in (1.1). The linear function Λ − S represents the external source (or supply) for the susceptible population, for more explanation, please refer to literature (Gao & Ruan, 2011; Hethcote, 2000). The authors in (Li et al., 2017) mainly discussed the global stability of the unique endemic equilibrium when spatial environment is homogeneous, and the asymptotic profile of endemic equilibria if the diffusion rate of the susceptible or infected population is small or large.

Previous studies have focused on SIS epidemic models in a fixed domain. In the real world, the change of biological habitat is caused by the movement of species, which can be described mathematically by free boundary. In fact, the free boundary is widely used in many fields; for example, a single population model with a free boundary for invasive species was proposed by Du and Lin in (Du & Lin, 2010), which was extended to the competition model in (Cao, Li, Wang, & zhao, 2021; Du, Wang, & Zhou, 2017; Du & Lin, 2014; Wang & Zhao, 2014), the predator-prey model in (Lin, 2007; Takhirov & Norov, 2019; Wang & Zhao, 2017; Yousefnezhad, Mohammadi, & Bozorgnia, 2018), and a mutualistic model with advection in (Li & Lin, 2015). In recent years, a free boundary has also been introduced into epidemic models. For example, an SIR epidemic model with free boundary was discussed in (Huang & Wang, 2015; Kim, Lin, & Zhang, 2013; Zhu, Guo, & Lin, 2017), and a SIRS model with free boundary was considered in (Cao et al., 2017a). Ge et al. in (Ge, Kim, Lin, & Zhu, 2015) studied a simplified SIS model with advection and free boundary and gave the spreading speeds when spreading happens; a diffusion-advection simplified SIS epidemic model in a heterogeneous time-periodic environment was studied in (Ge, Lei, & Lin, 2017); a SIS reaction-diffusion model with free boundary and mass action mechanism was discussed in (Wang & Guo, 2019); recently an SIS epidemic reaction-diffusion model with mass-action incidence incorporating spontaneous infection in a spatially heterogeneous environment was examined (Tong, Ahn, & Lin, 2021); a SIS model risk-induced dispersal of infected individuals has been studied (Choi, Lin, & Ahn, 2022); the spreading or vanishing of an SIS epidemic model with free boundary and nonlocal incidence rate was analyzed in (Cao et al., 2017b; Huang & Wang, 2019).

Based on above discussion, in this paper, we will consider the following SIS reaction-diffusion problem

StdΔS=Λ(x)Sβ(x)SIS+I+γ(x)I,xR,t>0,ItdΔI=β(x)SIS+Iγ(x)I,x(g(t),h(t)),t>0,I(x,t)=0,xR/(g(t),h(t)),t>0,g(t)=μIx(g(t),t),g(0)=h0,t0,h(t)=μIx(h(t),t),h(0)=h0,t0,S(x,0)=S0(x),I(x,0)=I0(x),xR, (1.3)

where the interval (g(t), h(t)) is unknown and will be given with (S, I). g′(t) = −μIx(g(t), t) (or h′(t) = −μIx(h(t), t)) is a special case of the well-known Stefan condition, which has been established in (Lin, 2007). Biologically, this system implies that outside of the free boundaries, there is no infectious, only susceptible individuals. The interval (g(t), h(t)), which depends on time t, is the habitat for the infectious individuals. Moreover, we suppose that initial functions S0 and I0 satisfy

S0(x)C2(R)L(R)andS0(x)>c0forxR,wherec0is a positive constant,I0(x)C2([h0,h0]),I0(x)>0forx(h0,h0),I0(x)=0forxR/[h0,h0]. (1.4)

Further, we assume that

(H)limx±β(x)=β±,limx±γ(x)=γ±andβ±>γ±,

which implies that the contact rate (β) and the recovery rate (γ) are similar at the faraway places, and the faraway is high-risk.

It is worth mentioning that in the fixed domain, the authors in (Li et al., 2017) primarily discussed the asymptotic profile of endemic equilibria when the diffusion rate is small or large. However, we discuss how the diffusion coefficient affects the spreading and vanishing of disease described mathematically by the free boundary problem in this paper.

The organization of this article is as below. The existence and uniqueness of the global solution are obtained in Section 2. Section 3 is devoted to the definition and properties of the basic reproduction number. Section 4 establishes a spreading-vanishing dichotomy for problem (1.3). Some sufficient conditions for the disease to spread or vanish are given in Section 5. Finally, we explain our theoretical results by numerical simulations.

2. Global existence and uniqueness

At first, we discuss the existence and uniqueness of the local solution of problem (1.3), and then we show that the solution is global by suitable estimates.

Theorem 2.1

For any given initial value (S0, I0) satisfying (1.4), and any ν ∈ (0, 1), there exists a T > 0 such that problem (1.3) admits a unique solution (S, I; g, h) satisfying

SC1+ν,1+ν2(R×[0,T]),IC1+ν,1+ν2([g(t),h(t)]×[0,T]), (2.1)

and

(g(t),h(t))C1+ν2([0,T])×C1+ν2([0,T]). (2.2)

Furthermore,

SC1+ν,1+ν2(R×[0,T])+IC1+ν,1+ν2([g(t),h(t)]×[0,T])+gC1+ν2([0,T])+hC1+ν2([0,T])C, (2.3)

where C and T only depend on h0,ν,μ,S0C2(R), S0L(R) and I0C2([h0,h0]).

Proof: Applying similar methods as in (Cao et al., 2017b; Chen & Friedman, 2000; Du & Lin, 2010), for system (1.3), we first straighten the free boundaries. Let ξ be a function in C3([0, )) satisfying

ξ(y)=1if|yh0|<h08,ξ(y)=0if|yh0|>h02,|ξ(y)|<5h0for ally.

Consider the transformation (y, t) → (x, t) such that

x=y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0),<y<+.

It is obvious to know that the above transformation xy is a diffeomorphism from (−, + ) onto (−, + ) while |h(t)h0|h08 and |g(t)+h0|h08. Hence, it translates the left free boundary x = g(t) to the fixed line y = −h0 and the right free boundary x = h(t) to the fixed line y = h0. Directing calculations give

yx=11+ξ(y)(h(t)h0)ξ(y)(g(t)+h0)A(g(t),h(t),y),
yt=ξ(y)h(t)ξ(y)g(t)1+ξ(y)(h(t)h0)ξ(y)(g(t)+h0)B(g(t),h(t),y),
2yx2=ξ(y)(h(t)h0)+ξ(y)(g(t)+h0)[1+ξ(y)(h(t)h0)ξ(y)(g(t)+h0)]3C(g(t),h(t),y).

Let

S(x,t)=S(y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0),t)u(y,t),
I(x,t)=I(y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0),t)v(y,t),
β(x)=β(y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0))β~(y),
γ(x)=β(y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0))γ~(y),
Λ(x)=β(y+ξ(y)(h(t)h0)+ξ(y)(g(t)+h0))Λ~(y),

then the free boundary problem (1.3) can be rewritten as

utA2dΔu(CdB)uy=Λ~(y)uβ~(y)uvu+v+γ~(y)v,yR,t>0,vtA2dΔv(CdB)vy=β~(y)uvu+vγ~(y)v,y(h0,h0),t>0,v(y,t)=0,yR/(h0,h0),t>0,g(t)=μvy(h0,t),g(0)=h0,t0,h(t)=μvy(h0,t),h(0)=h0,t0,u(y,0)=u0(y),v(x,0)=v0(y),yR, (2.4)

where A = A(g(t), h(t), y), B = B(g(t), h(t), y) and C = C(g(t), h(t), y). The following proof first finds a suitable complete metric space, and then uses the contraction mapping theorem together with standard Lp theory and the Sobolev imbedding theorem (Ladyzenskaja, Solonnikov, & Ural’ceva, 1968) to obtain the existence and uniqueness of the solution for system (2.4). We are going to omit it here, see Theorem 2.1 in (Du & Lin, 2014) or (Li & Lin, 2015) for more details. □

Theorem 2.2

Let (S, I; g, h) be a bounded solution of problem (1.3) for t ∈ [0, T] with some T ∈ (0, + ), then there exist positive constants C1 and C2 independent of T such that

0<S(x,t)C1forxR,t(0,T],
0<I(x,t)C1forx(g(t),h(t)),t(0,T],
0<g(t)C2,0<h(t)C2fort(0,T].

Proof: For any bounded solution (S, I; g, h) of problem (1.3) for t ∈ [0, T], using the initial value condition together with the strong maximum principle yields

S(x,t)>0forxR,t(0,T]

and

I(x,t)>0forx(g(t),h(t)),t(0,T].

Applying the Hopf's boundary lemma to the equation of I gives that

Ix(g(t),t)>0,Ix(h(t),t)<0fort(0,T].

Therefore g′(t) < 0 and h′(t) > 0 for t ∈ (0, T] by the free boundary condition.

Let N = S + (1 + ϵ0)I with 1ϵ0supxRβ(x)>0, direct computation yields

NtdΔN=Λ(x)S+ϵ0β(x)SIS+Iϵ0γ(x)IΛ(x)S+ϵ0supxRβ(x)Sϵ0infxRγ(x)I=Λ(x)(1ϵ0supxRβ(x))Sϵ0infxRγ(x)1+ϵ0(1+ϵ0)IΛ(x)mN,

where m=min{1ϵ0supxRβ(x),ϵ01+ϵ0infxRγ(x)}.

Recalling that Ix(h(t), t) < 0 = Ix(h(t)+, t) and Ix(g(t), t) = 0 < Ix(g(t)+, t) for 0 < t ≤ T, we have

NtdΔNΛ(x)mN,xR,x{g(t),h(t)},0<tT,N(x,0)=N0(x)S0(x)+(1+ϵ0)I0(x),xR/(h0,h0),Nx(h(t),t)<Nx(h(t)+,t),0<tT,Nx(g(t),t)<Nx(g(t)+,t),0<tT. (2.5)

Consider the following auxiliary problem

Nt=ΛmN,0<tT,N(0)=N0(x)L(R), (2.6)

where Λ=supxRΛ(x). For l > max{h(T), − g(T)}, taking Wl(x,t)=NN+kl2(x2+2dt) for x ∈ [ − l, l] and t ∈ [0, T], k=sup(x,t)(R×[0,T])N(x,t), which satisfies

WltdΔWl(ΛΛ)m(NN).

We claim that

Wl(x,t)0,(x,t)[l,l]×[0,T]. (2.7)

Otherwise, we suppose that min(x,t)(l,l)×[0,T]Wl(x,t)<0, there exists a point (x0, t0) ∈ [ − l, l] × [0, T] so that min Wl = Wl(x0, t0) < 0. Due to N(0)=N0L(R)N0(x)>0, we know that t0 ≠ 0. Moreover, Wll, t) > 0 for t ∈ [0, T], so x0 ≠ ± l. Therefore we have (x0, t0) ∈ (−l, l) × (0, T]. If (x0, t0) ∈ (−l, l) × (0, T] with x0g(t0) and x0h(t0), we have

Wlt(x0,t0)0,dΔWl(x0,t0)0,

which is a contradiction with (Λ∗ − Λ) − m(N∗ − N) > 0.

Assume that x0 = g(t0) or h(t0), without loss of generality, we suppose that x0 = g(t0). Due to I > 0 in (g(t0), h(t0)), and I = 0 in R/(g(t0),h(t0)), we derive

Wlx(g(t),t)0,Wlx(g(t)+,t)0,

that is

Wlx(g(t)+,t)Wlx(g(t),t),

which immediately gives

Nx(g(t)+,t)Nx(g(t),t),

it is a contradiction by Nx(g(t)+, t) > Nx(g(t), t). Hence, (2.7) holds. Let l → +, we have

NNinR×[0,T].

Therefore

S(x,t)N(x,t)N(t)maxΛm,N0L(R)C1forxR,t[0,T],

it follows from I(x, t) ≤ N(x, t) that I(x, t) ≤ C1 for x ∈ (g(t), h(t)) and t ∈ [0, T].

Finally, we show that − g′(t) ≤ C2 and h′(t) ≤ C2 for t ∈ (0, T]. The proof is analogous to that of Lemma 2.2 in (Du & Lin, 2010) with C2 = 2MμC1 and

M=maxsupxRβ(x)2d,4I0C1([h0,h0])3C1.

So we omit it here. □

Theorem 2.3

System (1.3) admits a unique solution (S, I; g, h) for all t ∈ (0, + ).

Proof: Combining all estimates in Theorem 2.2 with the standard continuous extension method, we can prove that the unique solution (S, I; g, h) of system (1.3) exists for all t ∈ (0, + ). The proof is similar to Theorem 2.4 in (Cao et al., 2017b) or Theorem 2.3 in (Du & Lin, 2010), we omit it here. □

3. The basic reproduction number

In general, the basic reproduction number can be used to describe the dynamics of the temporal and spatial spread of the disease. We firstly discuss the basic reproduction number and its properties for the following problem

ItdΔI=β(x)Iγ(x)I,xΩ,t>0,I(x,t)=0,xΩ,t>0, (3.1)

which is given by linearizing the corresponding system of (1.3) in a bounded domain Ω at the disease-free equilibrium.

It is well-known (Allen et al., 2008; Ge et al., 2015) that the basic reproduction number R0D for (3.1) is given by the variational formula

R0D=R0D(Ω,d)=supφH01(Ω),φ0Ωβ(x)φ2dxΩ(d|φ|2+γ(x)φ2)dx.

Due to above definition of the basic reproduction number, the following result was given in (Allen et al., 2008; Ge et al., 2015) and (Huang, Han, & Liu, 2010).

Proposition 3.1

((Ge et al., 2015)) Letλ0is the principal eigenvalue of the principle problem

dΔψ=β(x)ψγ(x)ψ+λψ,xΩ,ψ=0,xΩ,

then 1R0D has the same sign as λ0.

Proposition 3.2

((Ge et al., 2015)) The following assertions hold:

  • (i)R0Dis a positive and monotone decreasing function ofd;

  • (ii)R0DmaxxΩβ(x)γ(x)asd → 0;

  • (iii)R0D0asd;

  • (iv) There exists a threshold value d∗ ∈ [0, + ) such that R0D>1 for d < d∗ and R0D<1 for d > d∗. If all sites in the domain are lower-risk (β(x) < γ(x) for x ∈ Ω), then R0D<1 for all d > 0;

  • (v) Let Bh be a ball in Rn with the radius h. The R0D(Bh) is strictly monotone increasing function of h, that is if h1 < h2, then R0D(Bh1)<R0D(Bh2). Moreover, limhR0D(Bh)=βγ provided that (H) holds;

  • (vi) Assume that βand γare positive constants. If Ω = (−h0, h0), β(x) ≡ βand γ(x) ≡ γ∗, then

R0D=βdπ2h02+γ.

Pay attention to the interval (g(t), h(t)) of the free boundary problem (1.3) is changing with t, so the basic reproduction number is not a constant, and it varies with time t. Now we define the basic reproduction number R0F(t) for the free boundary problem (1.3) by

R0F(t)=R0D((g(t),h(t)),d)=supψH01((g(t),h(t))),ψ0g(t)h(t)β(x)ψ2dxg(t)h(t)(d|ψ|2+γ(x)ψ2)dx, (3.2)

which is called the spatial-temporal risk index in (Lin & Zhu, 2017) or the spatial-temporal basic reproduction number in (Zhu, Ren, & Zhu, 2018). We can define R0F()=limt+R0F(t) by the monotonicity in Theorem 3.3.

It follows from that Propositions 3.1 and 3.2, we have

Theorem 3.3

(a)R0F(t)is strictly monotone increasing function oft, in other words,R0F(t1)<R0F(t2)ift1 < t2;

  • (b) If g(t) → g > − and h(t) → h <  as t → +, we have that R0F()=R0D(g,h);

  • (c) If h(t) − g(t) → as t → +, then limt+R0F(t)min{βγ,β+γ+} provided (H) holds.

Proof: (a) − g′(t) > 0 and h′(t) > 0 for t ∈ (0, + ) in Theorem 2.2 imply that − g(t) and h(t) are strictly increasing functions. It is easy to see that R0F(t) is strictly monotone increasing function of t by (v) of Proposition 3.2.

(b) According to g > − and h < , we consider the problem

φtdΔφ=β(x)φγ(x)φ,x(g,h),t>0,φ(x,t)=0,x={g,h},t>0.

So the basic reproduction number R0D((g,h)) for above problem is given by the variational formula

R0D((g,h))=supφH01((g,h)),φ0ghβ(x)φ2dxgh(d|φ|2+γ(x)φ2)dx.

From (v) of Proposition 3.2 and monotonicity of g(t) and h(t), then R0D((g(t),h(t)))R0D((g,h)) holds for any t > 0, i.e.,

R0F(t)R0D((g,h)).

It follows that

lim supt+R0F(t)R0D((g,h)). (3.3)

On the other hand, duo to the definition of R0D((g,h)), for any given ε > 0, there exists a function φH01((g,h)) such that φH011 and

R0D((g,h))ghβ(x)φ2dxgh(d|φ|2+γ(x)φ2)dx+ε.

By definition of H01((g,h)), for given ε > 0, there exists a function ψC0((g,h)) such that ψφH01<ε.

Note that

suppψ={x(g,h)|ψ(x)0}¯(g,h).

Let hε = dist(suppψ, (g, h)), which means that ψ(x) = 0 in (g, g + hε) and (h − hε, h). In view of g(t) → g and h(t) → h as t → +, there exists a large T > 0 such that for t > T,

g<g(t)<g+hε/2<hhε/2<h(t)<h

and

g+hε/2hhε/2β(x)ψ2dxg+hε/2hhε/2(d|ψ|2+γ(x)ψ2)dx=ghgg+hε/2hhε/2hβ(x)ψ2dxghgg+hε/2hhε/2h(d|ψ|2+γ(x)ψ2)dx=ghβ(x)ψ2dxgh(d|ψ|2+γ(x)ψ2)dxghβ(x)φ2dx(supxRβ(x))3/2ε(2+ε)gh(d|φ|2+γ(x)φ2)dx+ε(2+ε)((supxRγ(x))3/2+d3/2)ghβ(x)φ2dxgh(d|φ|2+γ(x)φ2)dxMεR0D((g,h))(M+1)ε,

where positive constant M only depends on d,supxRβ and supxRγ. According to the monotonicity of g(t) and h(t) yields

R0F(t)R0D((g+hε/2,hhε/2))R0D((g,h))(M+1)εfort>T.

Hence,

lim inftR0F(t)R0D((g,h))(M+1)ε.

In light of the arbitrariness of ε, we can obtain

lim inftR0F(t)R0D((g,h)), (3.4)

which together with (3.3) gives that (b) holds.

(c) Suppose that h(t) → + as t → +. For any given ε > 0, there exists a large and positive constant l such that

β(x)>β+ε,γ(x)<γ++ε

for x > l provided that (H) holds. Moreover, there exists T0 > 0 such that h(t) > l for all t ≥ T0. For each fixed t ∈ [T0, ), let Φ(x) ∈ C2(g(t), h(t)) satisfy

Φ(x)=1forx0+34,h(t)34,
Φ(x)=0forxg(t),0+14h(t)14,h(t),
|Φ(x)|4forx[0,0+1][h(t)1,h(t)].

Due to the definition of R0F(t), for t ∈ [T0, + ), we have

R0F(t)=supψH01((g(t),h(t))),ψ0g(t)h(t)β(x)ψ2dxg(t)h(t)(d|ψ|2+γ(x)ψ2)dxg(t)h(t)β(x)Φ2dxg(t)h(t)(d|Φ|2+γ(x)Φ2)dx=g(t)0βΦ2dx+01βΦ2dx+1lβΦ2dx+lh(t)1βΦ2dx+h(t)1h(t)βΦ2dxg(t)0+01+1l+lh(t)1+h(t)1h(t)(d|Φ|2+γΦ2)dx=01+h(t)1h(t)βΦ2dx+1l+lh(t)1β(x)dx01+h(t)1h(t)d|Φ|2dx+g(t)l+lh(t)1+h(t)1h(t)γΦ2dx1l+lh(t)1β(x)dx01+h(t)1h(t)d|Φ|2dx+g(t)l+lh(t)1+h(t)1h(t)γ(x)dx(h(t)2)(β+ε)32d+(h(t)l)(γ++ε)+(lg(t))maxx[g(t),l]γ(x).

Hence

lim inft+R0F(t)lim inft+(h(t)2)(β+ε)32d+(h(t)l)(γ++ε)+(lg(t))maxx[g(t),l]γ(x)=β+εγ++ε.

In view of the arbitrariness of ε, we can obtain

limt+R0F(t)β+γ+.

If g(t) → − as t → +, we can draw a conclusion that limt+R0F(t)βγ from the above discussion.

Hence, limt+R0F(t)min{β+γ+,βγ} holds provided that (H) holds and h(t) − g(t) → + as t → +.

Remark 3.1

If g(t) → − or h(t) → + as t → +, then limt+R0F(t)>1 provided that (H) holds.

4. Spreading-vanishing dichotomy

In this section, we are going to establish a spreading-vanishing dichotomy for the free boundary problem (1.3). We will prove that if the interval (g(t), h(t)) is finite in the end, then I → 0 uniformly as t → +. Rather, for each solution (S, I) of problem (1.3), lim supt+I(,t)C([g(t),h(t)])>0 if the infected individuals spread to the whole area, that is, (g(t), h(t)) → (−, + ) as t → +.

From Theorem 2.2, we obtain that g′(t) < 0 and h′(t) > 0 for t ∈ (0, + ). Hence g(t) is monotonically decreasing and h(t) is monotonically increasing with t. So their limits exist, i.e.,

limt+g(t)=g,limt+h(t)=h,

where g ∈ [ − , − h0) and h ∈ (h0, + ].

Theorem 4.1

If < g < h < + ∞, then limt+I(,t)C([g(t),h(t)])=0, and limt+S(x,t)=S(x) locally uniformly for xR, where S(x)=1(4dπ)1/20+t1/2et|xy|24dtΛ(y)dydt satisfies

dΔS=Λ(x)S,xR. (4.1)

Proof: At first, we show that limt+I(,t)C([g(t),h(t)])=0. We argue indirectly, and assume that

lim supt+I(,t)C([g(t),h(t)])=δ>0.

Then there exits a sequence {(xk, tk)} with tk ∈ (0, ) and xk ∈ [g(tk), h(tk)] such that

I(xk,tk)>δ2

for all kN, and tk → + as k → +.

Due to −  < g < h < + , then xk ∈ [g(t), h(t)] ⊆ [g, h], there exists a subsequence xki,tki of {(xk, tk)}, denoted by {(xk, tk)}, such that xkx0 as k → +, where x0 ∈ [g, h].

Define

Ik(x,t)=I(x,t+tk),Sk(x,t)=S(x,t+tk)

for x ∈ [g(t + tk), h(t + tk)] and t ∈ [ − tk, + ). Using Theorem 2.1 and the standard parabolic regularity, there exists a subsequence {(Ski,Iki)} of {(Sk, Ik)} such that

(Ski,Iki)(S^,I^)aski+,

where (S^,I^) satisfies

StdΔS=Λ(x)Sβ(x)SIS+I+γ(x)I,x(g,h),t(0,+),ItdΔI=β(x)SIS+Iγ(x)I,x(g,h),t(0,+).

In view of I^(x0,0)=limkiIk(xki,0)=limkiI(xki,tki)δ2, then I^>0 in (g, h) × (0, + ) by the maximum principle.

Note that I^(g,0)=I^(h,0)=0, it follows from the Hopf's boundary lemma that

I^x(g,0)>0,I^x(h,0)<0.

Hence for all large enough ki, we have

Ix(g(tki),tki)=Iki(g(tki),0)x>0,Ix(h(tki),tki)=Iki(h(tki),0)x<0

and

h(tki)=μIx(h(tki),tki)>0,g(tki)=μIx(g(tki),tki)<0.

On the other hand, in view of − < g < h < + , we have obtain

limt+h(t)=limt+g(t)=0,

which is a contradiction. Therefore limt+I(,t)C([g(t),h(t)])=0.

Since limt+I(,t)C([g(t),h(t)])=0, for any ε > 0, there exists a sufficiently large T > 0 such that

0I(x,t)ε,(x,t)(,+)×[T,+). (4.2)

Now, for any given 0 < L < + , in view of (4.2), we have S_(x,t)S(x,t)S¯(x,t) for (x, t) ∈ [ − L, L] × [T, + ), where S¯(x,t) satisfies

S¯tdΔS¯=Λ(x)S¯+γ(x)ε,x(L,L),t>T,S¯(x,T)=S(x,T),x[L,L], (4.3)

and S_(x,t) satisfies

S_tdΔS_=Λ(x)S_β(x)ε,x(L,L),t>T,S_(x,T)=S(x,T),x[L,L]. (4.4)

It is well-known that

S_(x,t)S_(x,ε)andS¯S¯(x,ε)uniformly for[L,L]ast+,

where S_(x,ε) and S¯(x,ε) are unique positive steady state of problems (4.4) and (4.3), respectively. It is easily seen that

S_(x,ε)S(x)andS¯(x,ε)S(x)uniformly forx[L,L]asε0 (4.5)

by the standard parabolic regularity, where S∗(x) is the unique solution to problem (4.1).

Therefore, we obtain that

S_(x,ε)=limt+S_(x,t)lim inft+S(x,t)limt+S(x,t)lim supt+S(x,t)limt+S¯(x,t)=S¯(x,ε)

uniformly for x ∈ [ − L, L]. According to (4.5), we have

limt+S(x,t)=S(x)uniformly forx[L,L].

The arbitrariness of L implies that

limt+S(x,t)=S(x)uniformly in any bounded subset ofR.

Theorem 4.2

Ifh − g = +∞, then lim supt+I(,t)C([g(t),h(t)])>0.

Proof: Assume that limt+I(,t)C([g(t),h(t)])=0 by contradiction. It follows from the proof of Theorem 4.1 that

limt+S(x,t)=S(x)uniformly locally forxR. (4.6)

Since h − g = +, from the Remark 3.1, there exists a large enough T∗ > 0 such that

R0F(t)=R0D((g(t),h(t)))>1fortT.

In view of (4.6), for any 0<ε<minx[g(T),h(T)]S(x), there exists a T∗∗ > T∗ such that

S(x,t)>S(x)ε,x[g(T),h(T)],tT.

Following, we consider the eigenvalue problem

dΔψ=β(x)ψγ(x)ψ+λψ,x(g(T),h(T)),ψ(x)=0,x={g(T),h(T)},

it has the principal eigenvalue λ0 < 0 by Proposition 3.1 and the corresponding eigenvalue function ψ(x) with ψL((g(T),h(T)))=1.

Obviously, I(x, t) satisfies

ItdΔIβ(x)(S(x)ε)IS(x)ε+Iγ(x)I,x(g(T),h(T)),t>T,I(g(T),t)>0,I(h(T),t)>0,t>T,I(x,T)>0,g(T)xh(T).

We consider following auxiliary problem

WtdΔW=β(x)(S(x)ε)WS(x)ε+Wγ(x)W,x(g(T),h(T)),t>T,W(g(T),t)=W(h(T),t)=0,t>T,W(x,T)=I(x,T),g(T)xh(T), (4.7)

and construct a suitable lower solution for problem (4.7).

Define

W_(x,t)=δψ(x),g(T)xh(T),t>T,

where δ > 0 is small enough such that δψ(x) ≤ I(x, T∗∗) for x ∈ [g(T∗), h(T∗)].

For t > T∗∗ and g(T∗) < x < h(T∗), direct computation yields

W_tdΔW_β(x)(S(x)ε)W_S(x)ε+W_+γ(x)W_=β(x)δψ(x)γ(x)δψ(x)+λ0δψ(x)+γ(x)δψ(x)β(x)(S(x)ε)δψ(x)S(x)ε+δψ(x)=β(x)δψ(x)+λ0δψ(x)β(x)(S(x)ε)δψ(x)S(x)ε+δψ(x)=δψ(x)S(x)ε+δψ(x)[β(x)δψ(x)+λ0(S(x)ε)+λ0δψ(x)]δψ(x)S(x)ε+δψ(x)β(x)λ0(minx[g(T),h(T)]S(x)ε)maxx[g(T),h(T)]β(x)+λ0(S(x)ε)+λ0δψ(x)<0

provided that 0<δ<λ0(minx[g(T),h(T)]S(x)ε)maxx[g(T),h(T)]β(x).

Due to the upper and lower solutions method and comparison principle, we obtain

I(x,t)W(x,t)W_(x,t)=δψ(x)in(g(T),h(T))×[T,+).

Hence we have that lim inft+I(x,t)lim inft+W(x,t)δψ(0)>0, which is a contradiction.

Next, we show that the left and right of free boundary are finite or infinite simultaneously in the case where β and γ are positive constants.

Theorem 4.3

Assume thatβandγare positive constants. Ifg > − or h < + ∞, then < g < h < + ∞.

Proof: Without loss of generality, we suppose that h < + . Assume that g = − by a contradiction, then it follows from Theorem 3.1 and (H) that βγ>1 and there exists a large T1 > 0 such that R0F(T1)>1. So R0F(t)>R0F(T1)>1 for t > T1. In view of Proposition 3.1, we have λ1 < 0, where λ1 is a principal eigenvalue of the following problem

dΔφ=βφγφ+λφ,x(g(T1),h(T1)),φ(g(T1))=φ(h(T1))=0

and φ1(x) is corresponding eigenfunction. Hence, there exists a ε > 0 such that

dΔψεψ=βψγψ+λψ,x(g(T1),h(T1)),ψ(g(T1))=ψ(h(T1))=0

has principal eigenvalue λ1ε<0 and eigenfunction ψ1(x) > 0 with ψ1L(g(T1),h(T1))=1 in (g(T1), h(T1)). In addition,

StdΔS=Λ(x)SβSIS+I+γIΛ(x)SβS,xR,t>0.

Obviously, using parabolic comparison principle we have S(x, t) ≥ Sm for (x,t)R×(0,+), where Sm=min{(β+1)1infxRΛ(x),infxRS0(x)}.

From above result, it is easy to see that

ItdΔI=βSIS+IγIβSmISm+IγI,x(g(T1),h(T1)),t>T1.

And the following problem

dΔωεω=βSmωSm+ωγω,x(g(T1),h(T1)),ω(g(T1))=ω(h(T1))=0 (4.8)

admits a unique positive solution satisfying 0<ω<Sm(βγ1) in (g(T1), h(T1)). Using the Hopf's boundary lemma yields ω′(g(T1)) > 0 > ω′(h(T1)). So ω′(x) has at least one zero point in (g(T1), h(T1)). Let the largest zero of ω′(x) in (g(T1), h(T1)) be x0. Then

ω(x)<0in(x0,h(T1)).

Next, we consider the following auxiliary problem

I_tdΔI_=βSmI_Sm+I_γI_,x(g(T1),h(T1)),t>T1,I_(g(T1),t)=I_(h(T1),t)=0,t>T1,I_(x,T1)=I(x,T1),g(T1)xh(T1). (4.9)

In order to obtain 0<ε1φ1(x0)I_(x0,t) for t ≥ T1, we will show that ε1φ1(x) is a lower solution of problem (4.9) in (g(T1), h(T1)) × (T1, + ). Direct calculation gives

ε1φ1tdε1Δφ1βSmε1φ1Sm+ε1φ1+γε1φ1=dε1Δφ1βSmε1φ1Sm+ε1φ1+γε1φ1=ε1(βγ)φ1+λ0ε1φ1βSmε1φ1Sm+ε1φ1+γ(x)ε1φ1=λ0ε1φ1+βε1φ1βSmε1φ1Sm+ε1φ1=ε1φ1λ0+ββSmSm+ε1φ1<0

provided that ε1<λ0Smβ. Meanwhile, ε1φ1(g(T1))=I_(g(T1),t)=0, ε1φ1(h(T1))=I_(h(T1),t)=0. If ε1 is sufficiently small, then

ε1φ1(x)I_(x,T1),x[g(T1),h(T1)].

Hence, we can apply the comparison principle to conclude that

ε1φ1(x)I_(x,t),x[g(T1),h(T1)],t>T1

if ε1 is sufficiently small with ε1<λ0Smβ. Therefore,

0<ε1φ1(x0)I_(x0,t)fortT1,

which implies that I_(x0,t) has a positive lower bound for t ≥ T1.

Finally, based on the above results, we construct a suitable lower solution of I(x, t) in (x0, h(t)) × (T1, + ).

Set

Z=ε2ωx0+h(T1)x0h(t)x0(xx0),x0xh(t),tT1,

and calculate

ZtdΔZ=ε2ωh(T1)x0(h(t)x0)2(x0x)h(t)dε2Δωh(T1)x0h(t)x02=ε2h(T1)x0h(t)x02dΔωxx0h(T1)x0h(t)ω.

Using the fact that h′(t) → 0 and h(t) → h < + as t → +, we can find T2 > T1 such that h(t)<ε(h(T1)x0)hx0 and then

ZtdΔZε2h(T1)x0h(t)x02[dΔωεω]

since that ω′ ≤ 0 and xx0hx01 for (x, t) ∈ [x0, h(t)] × [T2, + ]. Since dΔωεω=βSmωSm+ωγω0 and h(T1)x0h(t)x01, we have

ZtdΔZε2βSmωSm+ωγωβSmZSm+ZγZ

for (x, t) ∈ [x0, h(t)] × [T2, + ].

In addition, we choose ε2 small such that ε2ω(x0) ≤ ε1φ1(x0) and

ε2ωx0+h(T1)x0h(T2)x0(xx0)I(x,T2)in[x0,h(T2)].

Applying the comparison principle yields

Z(x,t)=ε2ωx0+h(T1)x0h(t)x0(xx0)I(x,t)for(x,t)[x0,h(t))×[T2,+).

It follows that

Ix(h(t),t)Zx(h(t),t)=h(T1)x0h(t)x0ε2ωx(h(T1))h(T1)x0hx0ε2ωx(h(T1))<0

as t → +. By the free boundary condition, we have

h(t)=μIx(h(t),t)μZx(h(t),t)μh(T1)x0hx0ε2ωx(h(T1))>0

as t → +, which contradicts with the assumption h < + . Hence, −  < g < h < + holds.

According to Theorems 4.1 and 4.2, we can obtain the following spreading-vanishing dichotomy:

Theorem 4.4

Let (S, I; g, h) be the solution to problem (1.3). Then, the following alternatives hold:

  • (i) Spreading: h − g = + and then lim supt+I(,t)C([g(t),h(t)])>0; or

  • (ii) Vanishing: h − g < + and then limt+I(,t)C([g(t),h(t)])=0, and limt+S(x,t)=S(x) locally uniformly for xR, where S∗(x) satisfies problem (4.1).

Remark 4.1

Theorem 4.3 implies that h = −g =  when spreading happens if β and γ are positive constants.

5. Sufficient condition for vanishing or spreading

First, recalling the definition of R0F(t), we have the following result from the proof of Theorem 4.2.

Theorem 5.1

Ifhg < + ∞, then limt+R0F(t)=R0D((g,h),d)1. Moreover, if R0F(t0)1 for any t0 ≥ 0, theng = h = +∞.

It is well-known in (Allen et al., 2008; Li et al., 2017; Peng, 2009; Peng & Liu, 2009; Peng & Zhao, 2012) that R0D can be used as a threshold to describe the spreading or vanishing of infectious diseases on fixed boundary, that is, if R0D((h0,h0))>1, then infectious diseases is spreading; rather, the infectious diseases is vanishing. For the free boundary problem (1.3), we will show that spreading happens when R0F(0)=R0D((h0,h0))>1; the condition R0F(0)=R0D((h0,h0))<1 can not ensure R0F(t)<1 for any t > 0. We will show that the spreading and vanishing of infectious disease are possible when R0F(0)=R0D((h0,h0))<1, which mainly depends on the variables μ and I0.

Theorem 5.2

SupposeR0F(0)<1, then

limt+I(,t)C([g(t),h(t)])=0

if I0(x)C([h0,h0]) is small enough.

Proof: We prove it by constructing a suitable upper solution for I(x, t), which is similar to that of Lemma 5.3 in (Du & Lin, 2010) or (Ge et al., 2015). Due to R0F(0)=R0D((h0,h0))<1, we consider the following eigenvalue problem

dΔψ=β(x)ψγ(x)ψ+λψ,x(h0,h0),ψ(h0)=ψ(h0)=0,

which has a positive solution ψ(x) with ψL((h0,h0))=1, and principal eigenvalue λ0 > 0. In view of Theorem 2.2, we know that S(x, t) ≤ C1 for (x,t)R×(0,+). Hence,

ItdΔI=β(x)SIS+Iγ(x)Iβ(x)C1IC1+Iγ(x)I,x(g(t),h(t)),t>0.

Set

η(t)=h01+δδ2eδt,t0,

and

W(x,t)=εeδtψxh0η(t),η(t)xη(t),t0.

Clearly W(−η(t), t) = 0 and W(η(t), t) = 0. Direct computations yield that

η(t)=h0δ22eδt,
μWx(η(t),t)=με1+δδ2eδteδtψ(h0)<0,

and

μWx(η(t),t)=με1+δδ2eδteδtψ(h0)>0.

Hence − η′(t) ≤ −μWx(−η(t), t) and η′(t) ≥−μWx(η(t), t) hold if ε is sufficiently small. And then, from the definition of η(t) we know that − η(0) < g(0) = −h0 < h(0) = h0 < η(0). On the other hand, by direct calculations, for t > 0 and x ∈ (−η(t), η(t)), we obtain

WtdΔWβ(x)C1WC1+W+γ(x)W=εδeδtψεeδtψxh0η(t)η2(t)dεeδtΔψh0η(t)2β(x)C1εeδtψC1+εeδtψ+γ(x)εeδtψ=εeδtψxh0η(t)η2(t)+ϖγ(x)1h0η(t)2+ϖδ+β(x)h0η(t)2C1β(x)C1+εeδtψ+λ0h0η(t)2εeδtψxh0η(t)η2(t)+ϖγ(x)1h0η(t)2+ϖδ+β(x)1(1+δ)2β+λ01(1+δ)20

provided that δ is sufficiently small and satisfies δ(1+δ2)+[(1+δ)21]supxRβ(x)λ0.

Finally, we choose I0C([h0,h0]) small enough such that I0L([h0,h0])εψ(h01+δ2). Applying the comparison principle ((Du & Lin, 2010)) for free boundary problem yields

I(x,t)W(x,t),g(t)<x<h(t),t0.

That is

|I(x,t)|εeδt.

So

limt+I(,t)C([g(t),h(t)])=0

if I0(x)C([h0,h0]) is sufficiently small.

From the above proof, taking η(t)=h0(1+δδ2eδt) and ω(x,t)=Meδtψ(xh0η(t)), we can prove similarly that ω(x, t) is still the upper solution of I(x, t) for (x, t) ∈ (g(t), h(t)) × [0, + ) if positive constant M is sufficiently large and μ is small enough. Therefore, the following result holds.

Theorem 5.3

SupposeR0F(0)=R0D((h0,h0),d)<1, then

limt+I(,t)C([g(t),h(t)])=0

if μ is sufficiently small.

When R0F(0)<1, Theorem 5.3 indicates that vanishing happens for small expanding capability μ, the following result shows that spreading occurs for large expanding capability μ.

Theorem 5.4

AssumeR0F(0)<1, thenh − g = + if μ is sufficiently large and (H) holds.

Proof: According to (v) of Proposition 3.2, we have

limhR0D((h,h),d)=βγ>1

if (H) holds, and then there exists a sufficiently large positive constant M > 0 so that R0D((M,M),d)>1. Set

f1(x,t,I)=β(x)SIS+I,xR,t>0,
f2(x,t)=γ(x),xR,t>0.

Using the Lemma 2.5 in (Zhu & Lin, 2018) we can conclude that there exists a μM > 0 such that

lim supt+gμ(t)<Mandlim inft+hμ(t)>M

for μ > μM.

On the other hand, due to the monotonicity of gμ(t) and hμ(t), there exists a T0 > 0 such that

gμ(T0)<Mandhμ(T0)>M.

Moreover,

R0F(T0)=R0D((gμ(T0),hμ(T0)),d)>R0D((M,M),d)>1.

Hence, if μ > μM, then h − g = + holds by Theorem 5.1. □

Due to Theorems 5.3 and 5.4, the following result holds, see details in Theorem 5.5 in (Ge et al., 2015).

Theorem 5.5

Fixh0andI0, there existsμ∗ ∈ [0, + ) such that if μ > μ, then spreading happens; and if 0 < μ ≤ μ, then vanishing happens.

Finally, we will discuss how the diffusion coefficient d affects the vanishing and spreading of the disease.

Theorem 5.6

If the setG = {x ∈ [ − h0, h0], β(x) > γ(x)} ≠ ∅, then there exists d∗ > 0 such thatg = h = + for d ≤ dor d > dand μ is large enough.

Proof: Due to Proposition 3.2 (ii) and G we conclude that

R0D((h0,h0),d)maxx[h0,h0]β(x)γ(x)>1asd0.

There is a positive constant d∗ such that R0D((h0,h0),d)=1, and R0D((h0,h0),d)1 for d ≤ d∗ by Proposition 3.2 (i) − (iii). Theorem 5.1 implies that − g = h = + for d ≤ d∗, while if d > d∗, then Theorem 5.4 gives that − g = h = + provided μ is large enough. □

Similarly, based on Proposition 3.2 (ii) and (iii), and Theorems 5.2 and 5.3, the following result for vanishing holds.

Theorem 5.7

Fixh0 > 0, β(x) and γ(x). There exists a nonnegative constant d^ such that < g < h < + for d>d^ provided that μ or I0(x)C([h0,h0]) is small enough. If β(x) < γ(x) in [ − h0, h0], then d^=0, that is, vanishing happens if μ or I0(x)C([h0,h0]) is small enough.

Proof: It is easily seen that there is a constant d^0 so that R0D((h0,h0),d)<1 for d>d^ by Proposition 3.2 (i) and (iii). If β(x) < γ(x) in [ − h0, h0], then d^=0 by Proposition 3.2 (i) and (ii).

Since R0D((h0,h0),d)<1, then vanishing happens if μ or I0(x)C([h0,h0]) is small enough by Theorems 5.2 and 5.3. □

6. Simulation and discussion

At first, we use numerical simulations to explain impact expanding of the capability μ on spreading of infectious diseases. Taking some functions as follows:

d=1,h0=1,S0(x)=0.5+0.2sinπ2x,I0(x)=0.9cosπ2x,Λ(x)=35(cos(x)+1.05),β(x)=3.6+0.4/(2+x2)sinx,γ(x)=3.55+1/(1+x2)cosx,

we can know that β± = 3.6, γ± = 3.55 and (H) holds. Moreover, according to (3.2), we have

R0F(0)11β(x)ψ2dx11(d|ψ|2+γ(x)ψ2)dxmaxx[1,1]β(x)11ψ2dxminx[1,1]γ(x)11ψ2dx3.83.8202<1.

Next, we discuss the asymptotic behaviors of the solution to problem (1.3) and the changing of free boundaries by choosing different expanding capabilities.

Example 6.1

Taking μ = 50. Theorems 5.3 and 5.5 show that the solution I decays to zero if μ is small. It is easy Fig. 1 that the disease tends to extinction quickly, and the free boundaries expand slowly.

Example 6.2

Letting μ = 60. It follows from Theorems 5.4 and 5.5 that the spreading of the disease happens for large μ. It is easily seen that a spatially inhomogeneous stationary endemic state appears for large μ. The free boundaries expand quickly.

Fig. 1.

Fig. 1

μ = 50. The left graph shows that the solution I decays to zero quickly. The right graph is the corresponding planar graph, which shows the free boundaries expand slowly and will be limited in a long run.

In this paper, we mainly discuss a SIS model with a linear source and variable total population, and the free boundaries are used to describe the spreading left and right fronts. Firstly, we present the local existence and uniqueness of solution for problem (1.3) by the contraction mapping theorem together with standard Lp theory and the Sobolev embedding theorem (see Theorem 2.1), and then we prove that the solution is global solution by some estimates (see Theorems 2.2 and 2.3). Next, the basic reproduction number R0F(t) is defined for the SIS diffusion-reaction system with a free boundary. And then, a spreading-vanishing dichotomy for problem (1.3) is given (see Theorem 4.4). We then prove that R0F(t) is used as the threshold to describe the spreading or vanishing of the disease. If R0F(t0)>1 for any t0 ≥ 0, then the disease is spreading (see Theorems 4.2 and 5.1). When R0F(0)<1, whether the disease is spreading or vanishing depends on the initial value of the infected individuals I0 and the expanding capability μ. That is to say, if R0F(0)=R0D((h0,h0))<1, the disease is vanishing when initial value of the infected individuals I0 is small enough (Theorem 5.2) or the expanding capability μ is small enough (Theorem 5.3); the disease is spreading when the expanding capability μ is large enough and (H) holds (Theorem 5.4). Finally, we deal with the impact of the coefficient d for the disease. There exists a d∗ > 0, for d ≤ d∗ or d > d∗ and μ is sufficiently large, then the disease spreads (Theorem 5.6). On the other hand, there exists a d^0, the disease vanishing provided d>d^ if μ or I0 are sufficiently small (Theorem 5.7). If the diffusion rates dS and dI are different, we believe that complicated dynamical behaviors would appear (Cui & Lou, 2016; Peng, 2009; Peng & Zhao, 2012).

In addition, our simulations indicate that expanding capability μ plays an important role in the spreading and vanishing of disease. If R0F(0)<1, the disease will vanish when the expanding capability μ is comparatively small (Fig. 1), and the disease will spread when the expanding capability μ is comparatively large (Fig. 2).

Fig. 2.

Fig. 2

μ = 60. The solution I stabilizes to an equilibrium in the left graph. The right planar graph shows that the free boundaries expand fast.

When spreading occurs, the spreading speed for the simplified SIS model was given in (Ge et al., 2015). Owing to the coupled system for our reaction-diffusion model, some technical and mathematical difficulties for the spreading speed deserve further study.

Declaration of competing interest

We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work, there is no professional or other personal interest of any nature or kind in any product, service and/or company that could be construed as influencing the position presented in, or the review of, the manuscript entitled, “The impact factors of the risk index and diffusive dynamics of a SIS free boundary model”.

Handling Editor: Dr HE DAIHAI HE

Footnotes

The first author is supported by the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX21-3188), the second author is supported under the framework of international cooperation program managed by the National Research Foundation of Korea (NRF-2019K2A9A2A06025237) and the third author is supported by the National Natural Science Foundation of China (Grant No. 12271470).

Peer review under responsibility of KeAi Communications Co., Ltd.

References

  1. Abdelrazec A., Belair J., Shan C.H., Zhu H.P. Modeling the spread and control of dengue with limited public health resources. Mathematical Bioscience. 2016;271:136–145. doi: 10.1016/j.mbs.2015.11.004. [DOI] [PubMed] [Google Scholar]
  2. Allen L.J.S., Bolker B.M., Lou Y., Nevai A.L. Asymptotic profiles of the steady states for an SIS epidemic reaction-diffusion model. Discrete Contin. Dyn. Syst. 2008;21:1–20. [Google Scholar]
  3. Cao J.F., Li W.T., Wang J., Yang F.Y. A free boundary problem of a diffusive SIRS model with nonlinear incidence. Zeitschrift für Angewandte Mathematik und Physik. 2017;68:1–16. [Google Scholar]
  4. J. F. Cao, W. T. Li, J. Wang, M. zhao, The dynamics of a Lotka-Volterra competition model with nonlocal diffusion and free boundaries, Advances in Differential Equations, 26 (2021), 163-200.
  5. Cao J.F., Li W.T., Yang F.Y. Dynamics of a nonlocal SIS epidemic model with free boundary. Discrete and Continuous Dynamical Systems - Series B. 2017;22:247–266. [Google Scholar]
  6. Center for Disease Control and Prevention (CDC) Update: West Nile-like viral encephalitis-New York, 1999. Morbidity & Mortality Weekly Report. 1999;48:890–892. [PubMed] [Google Scholar]
  7. Chen X.F., Friedman A. A free boundary problem arising in a model of wound healing. SIAM Journal on Mathematical Analysis. 2000;32:778–800. [Google Scholar]
  8. W. Choi, Z. G. Lin, I. Ahn, SIS reaction-diffusion model with risk-induced dispersal under free boundary, Nonlinear Analysis: Real World Applications 21 (2022), 103605.
  9. Cui R.H., Lou Y. A spatial SIS model in advective heterogeneous environments. Journal of Differential Equations. 2016;261:3305–3343. [Google Scholar]
  10. Du Y.H., Lin Z.G. Spreading-Vanishing dichotomy in the diffusive logistic model with a free boundary. SIAM Journal on Mathematical Analysis. 2010;42:377–405. [Google Scholar]
  11. Du Y.H., Lin Z.G. The diffusive competition model with a free boundary: Invasion of a superior or inferior competitor. Discrete and Continuous Dynamical Systems - Series B. 2014;19:3105–3132. [Google Scholar]
  12. Du Y.H., Wang M.X., Zhou M.L. Semi-wave and spreading speed for the diffusive competition model with a free boundary. Journal de Mathematiques Pures et Appliquees. 2017;9:253–287. [Google Scholar]
  13. Gao D.Z., Ruan S.G. An SIS patch model with variable transmission coefficients. Mathematical Biosciences. 2011;232:110–115. doi: 10.1016/j.mbs.2011.05.001. [DOI] [PubMed] [Google Scholar]
  14. Ge J., Kim K.I., Lin Z.G., Zhu H.P. A SIS reaction-diffusion-advection model in a low-risk and high-risk domain. Journal of Differential Equations. 2015;259:5486–5509. [Google Scholar]
  15. Ge J., Lei C.X., Lin Z.G. Reproduction numbers and the expanding fronts for a diffusion-advection SIS model in heterogeneous time-periodic environment. Nonlinear Analysis: Real World Applications. 2017;33:100–120. [Google Scholar]
  16. Hethcote H.W. The mathematics of infectious diseases. SIAM Review. 2000;42:599–653. [Google Scholar]
  17. Huang W.Z., Han M.A., Liu K.Y. Dynamics of an SIS reaction-diffusion epidemic model for disease transmission. Mathematical Biosciences and Engineering. 2010;7:51–66. doi: 10.3934/mbe.2010.7.51. [DOI] [PubMed] [Google Scholar]
  18. Huang H.M., Wang M.X. The reaction-diffusion system for an SIR epidemic model with a free boundary. Discrete and Continuous Dynamical Systems - Series B. 2015;20:2039–2050. [Google Scholar]
  19. Huang H.M., Wang M.X. A nonlocal SIS epidemic problems with double free boundaries. Zeitschrift für Angewandte Mathematik und Physik. 2019;70:1–19. [Google Scholar]
  20. Keeling M.J., Rohani P. Princeton University Press; New Jersey: 2008. Modeling infectious disease in humans and animals. [Google Scholar]
  21. Kim K.I., Lin Z.G., Zhang Q.Y. An SIR epidemic model with free boundary. Nonlinear Analysis: Real World Applications. 2013;14:1992–2001. [Google Scholar]
  22. Ladyzenskaja O.A., Solonnikov V.A., Ural’ceva N.N. Amer. Math. Soc; Providence, RI: 1968. Linear and quasilinear equations of parabolic type. [Google Scholar]
  23. Li M., Lin Z.G. The spreading fronts in a mutualistic model with advection. Discrete and Continuous Dynamical Systems - Series B. 2015;20:2089–2105. [Google Scholar]
  24. Lin Z.G. A free boundary problem for a predator-prey model. Nonlinearity. 2007;20:1883–1892. [Google Scholar]
  25. Lin Z.G., Zhu H.P. Spatial spreading model and dynamics of West Nile virus in birds and mosquitoes with free boundary. Journal of Mathematical Biology. 2017;75:1381–1409. doi: 10.1007/s00285-017-1124-7. [DOI] [PubMed] [Google Scholar]
  26. Li H.C., Peng R., Wang F.B. Varying total population enhances disease persistence: Qualitative analysis on a diffusive SIS epidemic model. Journal of Differential Equations. 2017;262:885–913. [Google Scholar]
  27. Martcheva M. Springer; New York: 2015. An introduction of mathematical epudemiology. [Google Scholar]
  28. McNeill W.H. Blackwell; Oxford: 1976. Plagues and people. [Google Scholar]
  29. Peng R. Asymptotic profiles of the positive steady state for an SIS epidemic reaction-diffusion model. I. Journal of Differential Equations. 2009;247:1096–1119. [Google Scholar]
  30. Peng R., Liu S.Q. Global stability of the steady states of an SIS epidemic reaction- diffusion model. Nonliner Analysis. 2009;71:239–247. [Google Scholar]
  31. Peng R., Zhao X.Q. A reaction-diffusion SIS epidemic model in a time-periodic environment. Nonlinerity. 2012;25:1451–1471. [Google Scholar]
  32. Takhirov J.O., Norov A. On a predator-prey model with free boundary. Uzbek Math. J. 2019;4:162–168. [Google Scholar]
  33. Thomas D., Urena B. A reaction-diffusion SIS epidemic model in a time-periodic environment. Mathematical and Computer Modelling. 2001;34:771–781. [Google Scholar]
  34. Y. Tong, I. Ahn, Z. G. Lin, Effect of diffusion in a spatial SIS epidemic model with spontaneous infection, Discrete and Continuous Dynamical Systems - Series B 34 (2021), 4045-4057.
  35. Wang Y.Z., Guo S.J. A SIS reaction-diffusion model with a free boundary condition and nonhomogeneous coefficients. Discrete and Continuous Dynamical Systems - Series B. 2019;24:1627–1652. [Google Scholar]
  36. Wang M.X., Zhao J.F. Free boundary problems for a Lotka-Volterra competition system. Journal of Dynamics and Differential Equations. 2014;26:655–672. [Google Scholar]
  37. Wang M.X., Zhao J.F. A free boundary problem for the predator-prey model with double free boundaries. Journal of Dynamics and Differential Equations. 2017;29:957–979. [Google Scholar]
  38. Yousefnezhad M., Mohammadi S.A., Bozorgnia F. A free boundary problem for a predator-prey model with nonlinear prey-taxis. Applied Mathematics. 2018;63:125–147. [Google Scholar]
  39. Zhu M., Guo X.F., Lin Z.G. The risk index for an SIR epidemic model and spatial spreading of the infectious disease. Mathematical Biosciences and Engineering. 2017;14:1565–1583. doi: 10.3934/mbe.2017081. [DOI] [PubMed] [Google Scholar]
  40. Zhu M., Lin Z.G. The impact of human activity on the risk index and spatial spreading of dengue fever. Nonlinear Analysis: Real World Applications. 2018;39:424–450. [Google Scholar]
  41. Zhu D.D., Ren J.L., Zhu H.P. Spatial-temporal basic reproduction number and dynamics for a dengue disease diffusion model. Mathematical Methods in the Applied Sciences. 2018;41:5388–5403. [Google Scholar]

Articles from Infectious Disease Modelling are provided here courtesy of KeAi Publishing

RESOURCES