In this paper, we consider a reaction-diffusion epidemic model with nonlocal diffusion and free boundaries, which generalises the free-boundary epidemic model by Zhao et al. [
Citation: Ting-Ying Chang, Yihong Du. Long-time dynamics of an epidemic model with nonlocal diffusion and free boundaries[J]. Electronic Research Archive, 2022, 30(1): 289-313. doi: 10.3934/era.2022016
[1] | Meng Zhao . The longtime behavior of the model with nonlocal diffusion and free boundaries in online social networks. Electronic Research Archive, 2020, 28(3): 1143-1160. doi: 10.3934/era.2020063 |
[2] | Liping Wang, Runqi Liu, Yangyang Shi . Dynamical analysis of a nonlocal delays spatial malaria model with Wolbachia-infected male mosquitoes release. Electronic Research Archive, 2025, 33(5): 3177-3200. doi: 10.3934/era.2025139 |
[3] | Xiao Chen, Fuxiang Li, Hairong Lian, Peiguang Wang . A deep learning framework for predicting the spread of diffusion diseases. Electronic Research Archive, 2025, 33(4): 2475-2502. doi: 10.3934/era.2025110 |
[4] | Yijun Chen, Yaning Xie . A kernel-free boundary integral method for reaction-diffusion equations. Electronic Research Archive, 2025, 33(2): 556-581. doi: 10.3934/era.2025026 |
[5] | Meng-Xue Chang, Bang-Sheng Han, Xiao-Ming Fan . Global dynamics of the solution for a bistable reaction diffusion equation with nonlocal effect. Electronic Research Archive, 2021, 29(5): 3017-3030. doi: 10.3934/era.2021024 |
[6] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[7] | Wenbin Zhong, Yuting Ding . Spatiotemporal dynamics of a predator-prey model with a gestation delay and nonlocal competition. Electronic Research Archive, 2025, 33(4): 2601-2617. doi: 10.3934/era.2025116 |
[8] | Lin Shen, Shu Wang, Yongxin Wang . The well-posedness and regularity of a rotating blades equation. Electronic Research Archive, 2020, 28(2): 691-719. doi: 10.3934/era.2020036 |
[9] | Xiaoju Zhang, Kai Zheng, Yao Lu, Huanhuan Ma . Global existence and long-time behavior of solutions for fully nonlocal Boussinesq equations. Electronic Research Archive, 2023, 31(9): 5406-5424. doi: 10.3934/era.2023274 |
[10] | Xiang Zhang, Tingting Zheng, Yantao Luo, Pengfei Liu . Analysis of a reaction-diffusion AIDS model with media coverage and population heterogeneity. Electronic Research Archive, 2025, 33(1): 513-536. doi: 10.3934/era.2025024 |
In this paper, we consider a reaction-diffusion epidemic model with nonlocal diffusion and free boundaries, which generalises the free-boundary epidemic model by Zhao et al. [
To model the 1973 cholera epidemic in the European Mediterranean region, Capasso and Paveri-Fontana [3] proposed the following ODE system
u′(t)=−au(t)+cv(t),v′(t)=−bv(t)+G(u(t)),t>0, | (1.1) |
where
● u(t) and v(t) represent respectively the average population concentration of the infectious agents and the infective humans in the infected area at time t,
● a, b, c are all positive constants such that 1/a represented the mean lifetime of the agents in the environment, 1/b the mean infectious period of the infective humans, c the multiplicative factor of the infectious agents due to the infective humans, and
● the function G(u) is the infection rate of the human population, assuming that the total number of susceptible humans remain constant during the epidemic. The function G is assumed to satisfy the following:
(G1)G∈C1([0,∞]),G(0)=0,G′(z)>0 for all z≥0;
(G2)(G(z)z)′<0 for z>0 and limz→+∞G(z)z<abc.
A simple example of such a function is given by G(z)=αz1+z with α∈(0,ab/c).
They were able to establish the following result for the long-time dynamics of (1.1): Let R0:=cG′(0)ab; then regardless of the positive initial values u(0) and v(0),
(i) the epidemic tends to extinction if R0<1, namely limt→∞(u(t),v(t))→(0,0) if R0<1,
(ii) the epidemic tends to a positive equilibrium state if R0>1, namely limt→∞(u(t),v(t))→(u∗,v∗) if R0>1, where u∗,v∗ are uniquely determined by
G(u∗)u∗=abcandv∗=acu∗. | (1.2) |
To include the mobility of the infectious agents (assuming the mobility of the infective human population is small and thus ignored), Capasso and Maddalena [4] proposed the following spatial reaction-diffusion model with Robin (or Neumann) boundary conditions
{u′(t)=dΔu−au+cv,t>0,x∈Ω,v′(t)=−bv+G(u),t>0,x∈Ω,∂u∂n+αu=0,t>0,x∈∂Ω,u(x,0)=u0(x),v(x,0)=v0(x),x∈¯Ω, | (1.3) |
where d denotes the diffusion rate of u, the epidemic region Ω⊂RN is a smooth bounded domain and α≥0. They proved that the long-time behaviour of (1.3) is similar to the ODE model (1.1) with R0 there replaced by ˜R0:=cG′(0)(a+dλ1)b, where λ1 is the first eigenvalue of the eigenvalue problem
{Δϕ=λϕin Ω,∂ϕ∂n+αϕ=0on ∂Ω. |
In the literature, R0 and ˜R0 are often called the reproduction number of the epidemic being modelled.
To describe the spatial spreading of an epidemic, it is important to know how the front of the epidemic propagates. In [5], Ahn, Baek, and Lin regarded the epidemic region as a changing interval and used the following free boundary problem to model the evolution and spreading of the epidemic:
{ut=duxx−au+cv,t>0, x∈(g(t),h(t)),vt=−bv+G(u),t>0, x∈(g(t),h(t)),u(x,t)=v(x,t)=0,t>0, x∈{g(t),h(t)},h′(t)=−μux(h(t),t),t>0,g′(t)=−μux(g(t),t),t>0,h(0)=−g(0)=h0,u(x,0)=u0(x),v(x,0)=v0(x),x∈[−h0,h0], | (1.4) |
where h(t) and g(t) are the moving boundaries of the infected region, μ is a positive constant and the initial data (u0,v0) satisfy
u0,v0∈C([−h0,h0]),u0(±h0)=v0(±h0)=0, and u0,v0>0 in (−h0,h0). | (1.5) |
The equations for h′(t) and g′(t) mean that the expanding rate of the infected region is proportional to the spatial gradient of u at the front. This is known as the Stefan condition which was first used to describe the melting of ice (see, e.g., [6]). It has been extensively used in the study of the spread of population since Du and Lin [7].
The long-time dynamics of (1.4) can be described by a spreading-vanishing dichotomy; more precisely, Ahn et al. showed that the unique solution (u,v,g,h) to (1.4) satisfies either
(i) (vanishing)
{limt→∞(g(t),h(t))=(g∞,h∞)is a finite interval, andlimt→∞(u(x,t),v(x,t))=(0,0) uniformly for x∈[g(t),h(t)], or |
(ii) (spreading)
{limt→∞(g(t),h(t))=R,andlimt→∞(u(x,t),v(x,t))=(u∗,v∗) locally uniformly for x∈R. |
Furthermore, using the reproduction number of (1.1), namely
R0:=cG′(0)ab, | (1.6) |
the dichotomy is determined as follows: If R0≤1, then vanishing always happens; in the case where R0>1, there exists a critical length, l∗:=π2√d1a(R0−1), such that
● if h0≥l∗ (i.e., the initial size of the infected region is no less than 2l∗), then spreading always happens, and
● if h0<l∗, then vanishing (resp. spreading) happens if the initial functions (u0,v0) are sufficiently small (resp. large).
In the case of an epidemic spreading predicted by (1.4), it was shown by Zhao, Li, and Ni [8] that there exists a uniquely determined c0>0 such that
limt→∞h(t)t=limt→∞g(t)−t=c0, |
which means the epidemic region [g(t),h(t)] expands with asymptotic speed c0.
In (1.4) (as well as in (1.3)), the spatial dispersal of the infectious agents is assumed to follow the rules of random walk, which ignores any nonlocal effect in the dispersal process. Such nonlocal effect can be included if the local diffusion operator is replaced by a nonlocal diffusion operator of the form
d∫RJ(x−y)[u(y,t)−u(x,t)]dy |
with an appropriate kernel function J. Here J(x) can be interpreted as the probability that an individual of the species moves from location 0 to x. A widely used class of kernel functions consists of J:R→R satisfying
(J):J∈C(R)∩L∞(R),J is even and nonnegative, J(0)>0,∫RJ(x)dx=1. |
One recent paper by Cao, Du, Li, and Li [9] extended many basic results of [7] to the corresponding nonlocal model with the above kernel. Following the fashion of [9], Zhao, Zhang, Li, and Du [1] considered the corresponding nonlocal version of (1.4), which has the following form
{ut=d∫RJ(x−y)[u(y,t)−u(x,t)]dy−au+cv,t>0, x∈(g(t),h(t)),vt=−bv+G(u),t>0, x∈(g(t),h(t)),u(x,t)=v(x,t)=0,t>0, x∉(g(t),h(t)),h′(t)=μ∫h(t)g(t)∫∞h(t)J(x−y)u(x,t)dydx,t>0,g′(t)=−μ∫h(t)g(t)∫g(t)−∞J(x−y)u(x,t)dydx,t>0,u(x,0)=u0(x),v(x,0)=v0(x),h(0)=−g(0)=h0,|x|≤h0, | (1.7) |
It was shown in [1] that (1.7) has a unique solution defined for all t>0, and its long-time dynamics is determined by a spreading-vanishing dichotomy, in a similar fashion to (1.4) (with some subtle differences though). A striking difference of (1.7) to (1.4) is revealed by [2], which shows that the spreading determined by (1.7) may have infinite asymptotic spreading speed, a phenomenon known as "accelerated spreading". More precisely, if the kernel function J satisfies
∫∞0xJ(x)=∞, |
and spreading happens, then
limt→∞h(t)t=limt→∞g(t)−t=∞. |
Moreover, if J(x)≃|x|−γ for some γ∈(1,2] and all large |x|>0, then for all large t>0,
{−g(t),h(t)≃tlntif γ=2,−g(t),h(t)≃t1/(γ−1)if γ∈(1,2). |
Here, and in what follows, η(t)≃ξ(t) means C1ξ(t)≤η(t)≤C2ξ(t) for some positive constants C1≤C2 and all t in the specified range.
In this paper, to understand the effect of the mobility of the infective host on the epidemic spreading, we examine a full version of (1.7) *, where the dispersal of infective host is included. Before giving this full version, let us note that, in (1.7), since u(x,t)=0 for x∉(g(t),h(t)) and ∫RJ(x)dx=1,
*A full version of the local diffusion model (1.4) was recently investigated in [10], which showed that its long-time dynamics is similar to that of (1.4) though some differences occur in the criteria governing the spreading-vanishing dichotomy. In particular, when spreading happens, there exists a finite asymptotic spreading speed.
∫RJ(x−y)[u(y,t)−u(x,t)]dy=∫h(t)g(t)J(x−y)u(y,t)dy−u(x,t) for x∈(g(t),h(t)). |
For i=1,2, suppose Ji:R→R satisfy (J). Let a,b,c,d1,d2,μ1,μ2 and h0 be constants, all positive except μ1 and μ2, which are assumed to be nonnegative with μ1+μ2>0, and let the initial functions u0(x) and v0(x) satisfy (1.5). Then the full version of (1.7) can be written in the following form
{ut=d1∫h(t)g(t)J1(x−y)u(y,t)dy−d1u−au+cv,t>0,x∈(g(t),h(t)),vt=d2∫h(t)g(t)J2(x−y)v(y,t)dy−d2v−bv+G(u),t>0,x∈(g(t),h(t)),u(x,t)=v(x,t)=0,t>0,x∈{g(t),h(t)},h′(t)=∫h(t)g(t)∫∞h(t)[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx,t>0,g′(t)=−∫h(t)g(t)∫g(t)−∞[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx,t>0,h(0)=−g(0)=h0, u(x,0)=u0(x),v(x,0)=v0(x),|x|≤h0, | (1.8) |
We will prove the following results.
Theorem 1.1 (Existence and Uniqueness). The problem (1.8) admits a unique positive solution (u,v,g,h) defined for t≥0.
Theorem 1.2 (Spreading-Vanishing Dichotomy). Let (u,v,g,h) be the solution to (1.8) and denote h∞:=limt→∞h(t) and g∞:=limt→∞g(t). Then either
(i) (vanishing) (g∞,h∞) is a finite interval and
limt→∞(u(x,t),v(x,t))=(0,0)uniformly forx∈[g(t),h(t)], or | (1.9) |
(ii) (spreading) (g∞,h∞)=R and
limt→∞(u(x,t),v(x,t))=(u∗,v∗)locally uniformly forx∈R. |
Theorem 1.3 (Spreading-Vanishing Criteria). Let (u,v,g,h) be the solution of (1.8) and R0 be given by (1.6).
(a) If R0≤1, then vanishing always occurs.
(b) If R0>1, then spreading always occurs if one of the following holds:
(I)cG′(0)(d1+a)(d2+b)≥1,(II)cG′(0)(d1+a)(d2+b)<1andh0≥L∗, |
where L∗>0 is a certain critical length depending on a,b,c,d1,d2,J1,J2 but independent of the initial data (u0,v0).
(c) If R0>1 and
cG′(0)(d1+a)(d2+b)<1andh0<L∗, |
then
(i) for any given initial datum (u0,v0) satisfying (1.5), and any given constants σ01,σ02 nonnegative satisfying σ01+σ02>0, there exists μ∗>0 such that
(α) if (μ1,μ2)=(μσ01,μσ02) and 0<μ≤μ∗, then vanishing occurs, and
(β) if (μ1,μ2)=(μσ01,μσ02) and μ>μ∗, then spreading occurs.
(ii) for fixed (μ1,μ2) and sufficiently small initial datum (u0,v0), vanishing occurs.
Remark 1.4. (i) The constant L∗ in Theorem 1.3 is uniquely determined by an eigenvalue problem; see Proposition 3.4(iii) below.
(ii) In the case of (1.7) considered in [1], which is equivalent to (1.8) with d2=μ2=0, the long-time dynamics is also governed by a spreading-vanishing dichotomy, and the spreading-vanishing criteria coincide with those in Theorem 1.3 but with d2 and μ2 replaced by 0.
(iii) If μ2=0 and all the other parameters in (1.8) are positive and fixed except d2, which is allowed to vary in [0,∞), then from Lemmas 3.2 and 3.3 below it is easily seen that L∗=L∗(d2) is strictly increasing in d2. Therefore part (b) of Theorem 1.3 and the result in [1] indicate that the range of parameters (a,b,c,d1,h0) for which spreading happens regardless of the size of the initial function pare (u0,v0), i.e., (I) or (II) above holds, is enlarged as d2 is decreased, and such a range is maximized when d2=0. Biologically, this means that reducing the mobility of the infective host increases the chance of successful spreading of the disease, which appears counter-intuitive at first look. However, such a phenomenon is not new; it arises in the local diffusion models considered in [4,10] already.
When spreading happens, the spreading profile of (1.8) can be determined by using the general results in [2]. For this purpose, we will need the following condition
(J1):∫∞0xJi(x)dx<∞ for i=1,2 with μi>0.
Theorem 1.5 (Spreading Speed). In Theorem 1.2, if spreading happens, then
limt→∞g(t)−t=limt→∞h(t)t={c0if(J1)holds,∞if (J1)doesnothold, | (1.10) |
where c0>0 is uniquely determined by the associated semi-wave problem to (1.8) (see [2,Section 1.2]) .
Furthermore, in the case of accelerated spreading, we can determine the rate of accelerated spreading for a rather general class of kernel functions.
Theorem 1.6 (Rate of Accelerated Spreading). In Theorem 1.5 suppose additionally that for i=1,2 with μi>0, the kernel function Ji satisfies Ji(x)≃|x|−γ for some γ∈(1,2] and |x|≫1. Then for t≫1, we have
−g(t),h(t)≃tlntifγ=2,−g(t),h(t)≃t1/(γ−1)ifγ∈(1,2). | (1.11) |
Let us note that when Ji satisfies Ji(x)≃|x|−γ for |x|≫1 and for i∈{1,2} such that μi>0, (J1) holds if and only if γ>2. Thus Theorem 1.6 covers the exact range of γ such that accelerated spreading is possible. Note also that in condition (J1) as well as in Theorem 1.6, the condition only applies to the kernel function Ji where μi>0. For example, if μ2=0, then no extra condition on J2 is needed apart from satisfying (J).
Problem (1.8) has an entire space version where no free boundary is involved, which has the form
{ut=d1∫RJ1(x−y)u(y,t)dy−d1u−au+cv,t>0, x∈R,vt=d2∫RJ2(x−y)v(y,t)dy−d2v−bv+G(u),t>0, x∈R,u(x,0)=u0(x), v(x,0)=v0(x),x∈R. | (1.12) |
Problem (1.12) has been successfully used to determine the spreading speed of the epidemic; see [11] and the references therein for many interesting results on this and related problems. For the entire space version of (1.7), see [2,12] and the references therein for more details. The local diffusion counterparts of these entire space problems have been studied more extensively; see, for example, [13,14,15]. As mentioned above, the corresponding free boundary models have the advantage of providing the exact location of the spreading front of the concerned epidemics.
The rest of the paper is organized as follows. In Section 2, we introduce the preparatory results relating to (1.8) and use them to prove Theorem 1.1. In Section 3, we gather the necessary results associated with the corresponding fixed boundary problems, which will be used to determine the long-time dynamical behavior of (1.8). In Section 4, we use the results of the previous sections to establish the vanishing-spreading dichotomy as related to the reproduction number R0, proving Theorems 1.2 and 1.3. Finally, Section 5 is devoted to proving the assumptions required in [2] for Theorems 1.5 and 1.6.
We would like to point out that, although (1.8) has some significantly different features from the West Nile virus model studied in [16], for example, the nature of the reaction terms in (1.8) makes any nonnegative initial function admissible while the model in [16] only allows initial functions taken from a certain bounded order interval, but many techniques of [16] can be adapted to treat (1.8), which has helped to considerably reducing the length of the current paper. Here we only provide the details of the proofs when they are very different from [16].
In this section, we prove the well-posedness of (1.8) and some associated comparison principles.
We first recall a maximum principle from [16,Lemma 3.1], which is more general than needed in this paper, but in view of possible applications elsewhere, we state it in the general form as in [16].
Let T>0 and ξ∈C([0,T]). We define the set of strict local semi-maximum points of ξ by
Σξmax:={t∈(0,T]:∃ϵ>0 such that ξ(t)>ξ(s) for s∈[t−ϵ,t)}. |
Similarly, the set of strict local semi-minimum points of ξ is given by
Σξmin:={t∈(0,T]:∃ϵ>0 such that ξ(t)<ξ(s) for s∈[t−ϵ,t)}. |
If ξ is strictly increasing, then Σξmax=(0,T]. If ξ is nondecreasing, then Σξmin=∅.
Lemma 2.1 (Maximum Principle). Let T,h0>0,g,h∈C([0,T]) satisfy g(t)<h(t) and −g(0)=h(0)=h0. Denote DT:={(x,t):t∈(0,T],g(t)<x<h(t)} and suppose that for i,j∈{1,2,...,n}, ϕi,∂tϕi∈C(¯DT),di,cij∈L∞(DT), di≥0, and
{(ϕi)t≥di∫h(t)g(t)Ji(x−y)ϕi(y,t)dy−diϕi+n∑j=1cijϕj,(x,t)∈DT,ϕi(g(t),t)≥0,t∈Σgmin,ϕi(h(t),t)≥0,t∈Σhmax,ϕi(x,0)≥0,|x|≤h0, |
where Ji satisfies (J). Then the following holds:
(i) If cij≥0 on DT for i,j∈{1,2,...n} and i≠j, then ϕi≥0 on ¯DT for i∈{1,2,...,n}.
(ii) If for some i0∈{1,...,n} we assume additionally that di0>0 in DT and ϕi0(x,0)≢0 in [−h0,h0], then ϕi0(x,t)>0 in DT.
Lemma 2.2 (Comparison Principle I). For T∈(0,+∞), suppose that ¯g,¯h∈C([0,T]), D={(x,t):t∈(0,T],¯g(t)<x<¯h(t)}, \ ¯u,¯v∈C(¯D), ¯u,¯v≥0. If (¯u,¯v,¯g,¯h) satisfies
{¯ut≥d1∫h(t)g(t)J1(x−y)¯u(y,t)dy−d1¯u−a¯u+c¯v,t>0,x∈(¯g(t),¯h(t)),¯vt≥d2∫h(t)g(t)J2(x−y)¯v(y,t)dy−d2¯v−b¯v+G(¯u),t>0,x∈(¯g(t),¯h(t)),¯h′(t)≥∫h(t)g(t)∫∞h(t)[μ1J1(x−y)¯u(x,t)+μ2J2(x−y)¯v(x,t)]dydx,t>0,¯g′(t)≤−∫h(t)g(t)∫g(t)−∞[μ1J1(x−y)¯u(x,t)+μ2J2(x−y)¯v(x,t)]dydx,t>0,¯g(0)≤−h0,¯h(0)≥h0, ¯u(x,0)≥u0(x),¯v(x,0)≥v0(x),|x|≤h0, | (2.1) |
then the unique solution (u,v,g,h) of (1.8) satisfies
u(x,t)≤¯u(x,t),v(x,t)≤¯v(x,t),g(t)≥¯g(t),h(t)≤¯h(t), | (2.2) |
for 0<t≤T and g(t)≤x≤h(t).
Proof. By (G1), we have that G(¯u)=G′(ξ)¯u with ξ=ξ(x,t)∈(0,¯u(x,t)]. From (1.5) and Lemma 2.1, we infer that ¯u,¯v>0, for 0<t≤T and ¯g(t)<x<¯h(t). Therefore, −¯g and ¯h are strictly increasing.
For small ϵ>0, let (uϵ,vϵ,gϵ,hϵ) denote the unique solution of (1.8) with h0 replaced by hϵ0:=h0(1−ϵ), μi by μϵi:=μi(1−ϵ) for i=1,2, and (u0,v0) by (uϵ0,vϵ0) satisfying
{0<uϵ0(x)<u0(x),0<vϵ0(x)<v0(x)in (−hϵ0,hϵ0),uϵ0(±hϵ0)=vϵ0(±hϵ0)=0, and(uϵ0(h0hϵ0x),vϵ0(h0hϵ0x))→(u0(x),v0(x))as ϵ→0 in the C([−h0,h0]) norm. | (2.3) |
We claim that hϵ(t)<¯h(t) and gϵ(t)>¯g(t) for all t∈(0,T]. It is clear that these hold for small t>0. Suppose that there exists t1≤T such that
hϵ(t)<¯h(t), gϵ(t)>¯g(t) for t∈(0,t1) and [hϵ(t1)−¯h(t1)][gϵ(t1)−¯g(t1)]=0. |
Without loss of generality, assume that hϵ(t1)=¯h(t1) and gϵ(t1)≥¯g(t1). Let w:=¯u−uϵ and z:=¯v−vϵ; then (w,z) satisfies
{wt≥d1∫hϵ(t)gϵ(t)J1(x−y)w(y,t)dy−d1w−aw+cz,0<t≤t1, x∈(gϵ(t),hϵ(t)),zt≥d2∫hϵ(t)gϵ(t)J2(x−y)z(y,t)dy−d2z−bz+G′(η)w,0<t≤t1, x∈(gϵ(t),hϵ(t)),w(x,t)≥0, z(x,t)≥0,0<t≤t1, x=gϵ(t) or hϵ(t),w(x,0)>0, z(x,0)>0,x∈[gϵ(0),hϵ(0)], | (2.4) |
where η=η(x,t) is between ¯u(x,t) and uϵ(x,t). Therefore we can apply Lemma 2.1 to conclude that w(x,t)>0 and z(x,t)>0 for 0<t≤t1 and gϵ(t)<x<hϵ(t).
However, by definition of t1, we have h′ϵ(t1)≥¯h′(t1), giving us that
0≥¯h′(t1)−h′ϵ(t1)≥∫¯h(t1)¯g(t1)∫∞¯h(t1)[μ1J1(x−y)¯u(x,t1)+μ2J2(x−y)¯v(x,t1)]dydx−∫hϵ(t1)gϵ(t1)∫∞hϵ(t1)[μϵ1J1(x−y)uϵ(x,t1)+μϵ2J2(x−y)vϵ(x,t1)]dydx≥∫hϵ(t1)gϵ(t1)∫∞hϵ(t1)[μϵ1J1(x−y)w(x,t1)+μϵ2J2(x−y)z(x,t1)]dydx>0. |
This contradiction proves our claim, namely, hϵ(t)<¯h(t) and gϵ(t)>¯g(t) for all t∈(0,T]. Hence (2.4) holds with t1 replaced by T, which yields that ¯u(x,t)>uϵ(x,t) and ¯v(x,t)>vϵ(x,t) for 0<t≤T and gϵ(t)<x<hϵ(t). Letting ϵ→0, we obtain the desired result from the continuous dependence of (uϵ,vϵ,gϵ,hϵ) on ϵ.
We introduce a second comparison principle where the boundaries are regarded as given.
Lemma 2.3 (Comparison Principle II). Assume (J) holds, T>0, g,h∈C([0,T]) satisfying g(t)<h(t), and DT defined as in Lemma 2.1. If for i=1,2, ui,˜ui∈C(¯DT) satisfy the following conditions:
(i)ϕt∈C(¯DT) for ϕ∈{u1,u2,˜u1,˜u2},
(ii) for (x,t)∈DT,
{(˜u1)t≥d1∫h(t)g(t)J1(x−y)˜u1(y,t)dy−d1˜u1−a˜u1+c˜u2,(˜u2)t≥d2∫h(t)g(t)J2(x−y)˜u2(y,t)dy−d2˜u2−b˜u2+G(˜u1), | (2.5) |
(iii) for (x,t)∈DT, (u1,u2) satisfies (2.5) but with the inequalities reversed,
(iv) at the boundary,
{ui(g(t),t)≤˜ui(g(t),t)fort∈Σgmin,ui(h(t),t)≤˜ui(h(t),t)fort∈Σhmin, |
(v) and at the initial time, ui(x,0)≤˜ui(x,0) for x∈[g(0),h(0)] and i=1,2.
Then for i=1,2, we must have
ui(x,t)≤˜ui(x,t)for(x,t)∈DT. |
Proof. For i=1,2, define ϕi:=~ui−ui and
c11:=−a,c12:=c,c21:=G(˜u1)−G(u1)˜u1−u1,c22:=−b. |
Then by the maximum principle in Lemma 2.1, we obtain that ϕi≥0 in DT for i=1,2.
Lemma 2.4 (A Priori Bound). For T∈(0,+∞), let (u,v,g,h) be a solution of (1.8) for t∈(0,T]. Then there exists constants C1 and C2 independent of T such that
u(x,t)≤C1andv(x,t)≤C2forg(t)<x<h(t),t∈(0,T]. |
Proof. By assumption (G2), there exist C1 and C2 such that
G(C1)C1<abcandC1≥u0(x) in [−h0,h0] | (2.6) |
and
G(C1)b<C2<acC1andC2≥v0(x) in [−h0,h0]. | (2.7) |
Let (U1(x,t),U2(x,t))≡(C1,C2). For i=1,2, let us denote
Li[w](x,t):=∫h(t)g(t)Ji(x−y)w(y,t)dy−w(x,t). |
Clearly Li[Ui](x,t)≤0 and (Ui)t=0. It now follows from (2.6) and (2.7) that
{(U1)t>d1L1[U1]−aU1+cU2,t>0,x∈(g(t),h(t)),(U2)t>d2L2[U2]−bU2+G(U1),t>0,x∈(g(t),h(t)),U1(x,t)>u(x,t),U2(x,t)>v(x,t),t>0,x∈{g(t),h(t)},U1(x,0)≥u0(x), U2(x,0)≥v0(x),|x|≤h0. |
By Lemma 2.3, we obtain u(x,t)≤C1 and v(x,t)≤C2 for all 0<t≤T and g(t)≤x≤h(t).
Now we are ready to prove Theorem 1.1.
Proof of Theorem 1.1. Let us first consider the following slightly modified problem, with C2 taken from Lemma 2.4:
{ut=d1∫h(t)g(t)J1(x−y)u(y,t)dy−d1u−au+cmin{v,C2},t>0,x∈(g(t),h(t)),vt=d2∫h(t)g(t)J2(x−y)v(y,t)dy−d2v−bv+G(u),t>0,x∈(g(t),h(t)),u(x,t)=v(x,t)=0,t>0,x∈{g(t),h(t)},h′(t)=∫h(t)g(t)∫∞h(t)[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx,t>0,g′(t)=−∫h(t)g(t)∫g(t)−∞[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx,t>0,u(x,0)=u0(x),v(x,0)=v0(x),h(0)=−g(0)=h0,|x|≤h0. | (2.8) |
By taking f1(u,v)=−au+cmin{v,C2} and f2(u,v)=−bv+G(u), in view of the conditions on G, we see that they satisfy the conditions of Theorem 4.1 in [16]. Therefore (2.8) has a unique solution (u,v,g,h) defined for all t>0.
Since f1(u,v)≤−au+cv, from the proof of Lemma 2.4 we see that u(x,t)≤C1 and v(x,t)≤C2, and therefore f1(u(x,t),v(x,t))=−au(x,t)+cv(x,t). Thus (u,v,g,h) solves (1.8).
Conversely, by Lemma 2.4 any solution of (1.8) satisfies v(x,t)≤C2 and hence it solves (2.8). Thus global existence and uniqueness holds for (1.8).
For any L>0, we consider the eigenvalue problem
{λϕ=−d1∫L−LJ1(x−y)ϕ(y)dy+d1ϕ+aϕ−cψ,x∈(−L,L),λψ=−d2∫L−LJ2(x−y)ψ(y)dy+d2ψ+bψ−G′(0)ϕ,x∈(−L,L). | (3.1) |
By Theorems 2.2 and 2.3 of [17], we have the following result:
Proposition 3.1. The eigenvalue problem (3.1) has a principal eigenvalue λ=λ1(L) with positive eigenfunction pair (ϕ,ψ)=(ϕ1,ψ1)∈C([−L,L])×C([−L,L]).
Then by Lemma 2.2 and Proposition 2.3 of [16], we have the following two results on the properties of the eigenvalue λ1(L).
Lemma 3.2. Let λ1(L) be the principal eigenvalue of (3.1). Let Φ,Ψ∈C([−L,L]) be two functions such that Φ,Ψ≥0 and Φ,Ψ≢0 in [−L,L], and ˜λ be a constant such that
{−d1∫L−LJ1(x−y)Φ(y)dy+d1Φ+aΦ−cΨ≥(≤)˜λΦ,x∈(−L,L),−d2∫L−LJ2(x−y)Ψ(y)dy+d2Ψ+bΨ−G′(0)Φ≥(≤)˜λΨ,x∈(−L,L), | (3.2) |
then λ1(L)≥(≤)˜λ. Moreover, λ1(L)=˜λ only if equalities hold in (3.2).
Lemma 3.3. Let λ1(L) be the principal eigenvalue of (3.1). Then
(i)λ1(L) is strictly decreasing with respect to L∈(0,∞),
(ii)λ1(L) is continuous for L∈(0,∞).
The following proposition is essential for establishing the spreading and vanishing criteria in Theorem 1.3.
Proposition 3.4. The principal eigenvalue λ1(L) of (3.1) has the following properties:
(i) If R0≤1, then λ1(L)>0 for any L>0.
(ii) If R0>1 and
cG′(0)(d1+a)(d2+b)≥1, | (3.3) |
then λ1(L)<0 for any L>0.
(iii) If R0>1 and
cG′(0)(d1+a)(d2+b)<1, | (3.4) |
then there exists L∗ such that
λ1(L∗)=0and(L−L∗)λ1(L)<0forL∈(0,L∗)∪(L∗,∞). |
Proof. The proof follows that of [16,Proposition 2.4], and the details are omitted.
Corollary 3.5. Let l1<l2 and λ1(l1,l2) be the principal eigenvalue of (3.1) with [−L,L] replaced by [l1,l2]. Then
(i)λ1(l1,l2) is strictly decreasing with respect to l2−l1 and is continuous in l1 and l2.
(ii) If R0≤1, then λ1(l1,l2)>0 for any l1 and l2.
(iii) If R0>1 and (3.3) holds, then λ1(l1,l2)<0 for any l1 and l2.
(iv) If R0>1 and (3.4) holds, then λ1(l1,l2)=0 for l2−l1=2L∗ and
λ1(l1,l2)>0forl2−l1<2L∗, λ1(l1,l2)<0forl2−l1>2L∗, | (3.5) |
where L∗>0 is given by Proposition 3.4(iii).
For L>0, we define QL=(−L,L)×(0,∞) and consider the corresponding fixed boundary problem of (1.8):
{ut=d1∫L−LJ1(x−y)u(y,t)dy−d1u−au+cv,(x,t)∈QL,vt=d2∫L−LJ2(x−y)v(y,t)dy−d2v−bv+G(u),(x,t)∈QL,u(x,0)=u0(x),v(x,0)=v0(x),x∈[−L,L], | (3.6) |
where u0,v0∈C([−L,L]) are nonnegative and not identically 0 simultaneously. It is well-known that fixed boundary problems such as (3.6) has a unique positive solution which is defined for all t>0 (see, for example, Remark 4.3 in [16]).
The corresponding steady state problem of (3.6) is
{−d1∫L−LJ1(x−y)˜u(y)dy+d1˜u=−a˜u+c˜v,x∈(−L,L),−d2∫L−LJ2(x−y)˜v(y)dy+d2˜v=−b˜v+G(˜u),x∈(−L,L). | (3.7) |
Definition 3.6. A function pair (ϕ,ψ)∈C([−L,L])×C([−L,L]) is said to be an upper solution of (3.7) if
{−d1∫L−LJ1(x−y)ϕ(y)dy+d1ϕ≥−aϕ+cψ,x∈(−L,L),−d2∫L−LJ2(x−y)ψ(y)dy+d2ψ≥−bψ+G(ϕ),x∈(−L,L). |
It is called a lower solution of (3.7) if these inequalities are reversed.
Proposition 3.7. Suppose that R0>1 and (u,v) is the unique positive solution of (3.6). Let λ1(L) be the principal eigenvalue of (3.1) and (u∗,v∗) be as defined in (1.2). Then the following conclusions hold:
(i) The fixed boundary problem (3.6) has a unique positive steady state solution (˜u,˜v)∈C([−L,L])×C([−L,L]) if λ1(L)<0, and (0,0) is the only nonnegative steady state when λ1(L)≥0. Moreover, 0<˜u(x)≤u∗ and 0<˜v(x)≤v∗ in [−L,L] when λ1(L)<0.
(ii) If λ1(L)≥0, then (u(x,t),v(x,t)) converges to (0,0) as t→∞ uniformly for x∈[−L,L].
(iii) If λ1(L)<0, then (u(x,t),v(x,t)) converges to (˜u,˜v) as t→∞ uniformly for x∈[−L,L].
Proof. Due to the different nature of the reaction terms in (1.8) from the model in [16], our proof here uses rather different techniques. In particular, we will make use of the monotonicity (in time t) of the to-be-constructed lower and upper solutions and Dini's theorem.
(i) Suppose that λ1(L)<0. Then we easily see that (Mu∗,Mv∗) and (ϵϕ1,ϵψ1) are respectively upper and lower solutions of (3.7) for small enough ϵ>0 and any M≥1, where (ϕ1,ψ1) is a positive eigenfunction pair corresponding to λ1(L).
Let (u_,v_) be the unique positive solution of (3.6) with initial function pair (ϵϕ1,ϵψ1). Using (ϵϕ1,ϵψ1) as a lower solution of (3.6), we can use the comparison principle in Lemma 2.3 with (g(t),h(t))≡(−L,L) to conclude that (ϵϕ1(x),ϵψ1(x))≤(u_(x,t),v_(x,t))≤(u∗,v∗) for (x,t)∈QL. In particular, for any fixed s>0,
(ϵϕ1(x),ϵψ1(x))=(u_(x,0),v_(x,0))≤(u_(x,s),v_(x,s)) for x∈[−L,L]. |
It is easily seen that (ˆu(x,t),ˆv(x,t):=(u_(x,s+t),v_(x,s+t)) is a solution of (3.6) with initial data (u_(x,s),v_(x,s)). Therefore we can use the comparison principle to deduce
(u_(x,t),v_(x,t))≤(ˆu(x,t),ˆv(x,t))=(u_(x,s+t),v_(x,s+t)) for (x,t)∈QL. |
Since s>0 is arbitrary, this implies that (u_(x,t),v_(x,t)) is nondecreasing in t and hence
(U(x),V(x)):=limt→∞(u_(x,t),v_(x,t)) exists, |
and (ϵϕ1(x),ϵψ1(x))≤(U(x),V(x))≤(u∗,v∗) in [−L,L]. Moreover, it is easily seen that (U,V) solves (3.7). Thus there exists at least one positive steady state solution.
We show next that (U,V) as well as any other positive solution of (3.7) are continuous in [−L,L]. Indeed, from the continuity of J1 and J2, we easily see that
G1(x):=d1∫L−LJ1(x−y)U(y)dy, G2(x):=d2∫L−LJ2(x−y)V(y)dy |
are continuous in [−L,L]. From (3.7), we obtain
{V(x)=a+d1cU(x)−G1(x)c,(a+d1)(b+d2)cU(x)−G(U(x))=G2(x)+b+d2cG1(x). |
From the conditions on G, we see that F(z):=(a+d1)(b+d2)cz−G(z) satisfies
F′(z)=(a+d1)(b+d2)c−G′(z)≥(a+d1)(b+d2)c−abc>0 for z>0. |
Thus from F(U(x))=G2(x)+b+d2cG1(x), F′(z)>0 and the fact that G1(x) and G2(x) are continuous, we obtain U(x) is continuous, which in turn implies that V(x)=a+d1cU(x)−G1(x)c is continuous.
To prove uniqueness, let (ˆU,ˆV) be another positive solution of (3.7). By choosing ϵ>0 sufficiently small, we may assume that (ϵϕ1(x),ϵψ1(x))≤(ˆU(x),ˆV(x)) in [−L,L]. Thus the above obtained (U,V) satisfies (U,V)≤(ˆU,ˆV).
We define
k∗:=inf{k>0:k(U,V)≥(ˆU,ˆV) in [−L,L]}. |
From (U,V)≤(ˆU,ˆV) and the definition of k∗, we have immediately k∗(U,V)≥(ˆU,ˆV) and k∗≥1. If k∗=1, then we immediately obtain (U,V)=(ˆU,ˆV) and the uniqueness is proved. If k∗>1, we show that a contradiction arises. So suppose k∗>1. Since G(z)/z is decreasing by (G2), it is easily checked that (k∗U,k∗V) is an upper solution of (3.7). We now consider (Φ(x),Ψ(x)):=(k∗U(x)−ˆU(x),k∗V(x)−ˆV(x)). It can be shown that it satisfies a pair of inequalities of the form in Lemma 2.1, and Φ(x)≢0, Ψ(x)≢0 (due to k∗>1 and (U,V)≤(ˆU,ˆV)). Thus, by Lemma 2.1 we deduce Φ(x)>0 and Ψ(x)>0 in [−L,L]. Since they are continuous functions on [−L,L], this implies that k(U,V)≥(ˆU,ˆV) in [−L,L] for some k<k∗, which is a contradiction to the definition of k∗. Thus we must have that k∗=1, and uniqueness is proved.
Now let us assume that (˜u,˜v) is a positive steady state solution of (3.6) and λ1(L)≥0. By (G2), we see that (˜u,˜v) satisfies
{−d1∫L−LJ1(x−y)˜u(y,t)dy+d1˜u+a˜u−c˜v=0,x∈(−L,L),−d2∫L−LJ2(x−y)˜v(y,t)dy+d2˜v+b˜v−G′(0)˜u<0x∈(−L,L). | (3.8) |
By applying Lemma 3.2 with ˜λ=0, we obtain that λ1(L)<0. This contradicts our assumption, proving the nonexistence.
Next we prove (ii) and (iii) simultaneously. By choosing ϵ>0 sufficiently small and M>1 sufficiently large, we can ensure that (ϵϕ1(x),ϵψ1(x))≤(u(x,1),v(x,1))≤(Mu∗,Mv∗) in [−L,L]. Let (¯u(x,t),¯v(x,t)) be the unique solution of (3.6) with initial data (Mu∗,Mv∗). Then an analogous reasoning to that for (u_(x,t),v_(x,t)) shows that (¯u(x,t),¯v(x,t)) is non-increasing in t and hence
(˜U(x),˜V(x)):=limt→∞(¯u(x,t),¯v(x,t)) exists and is a nonnegative solution of (3.7). |
If λ1(L)≥0, then by (i) we know that necessarily (˜U,˜V)=(0,0). By Dini's theorem, the monotonicity in t implies that the convergence in the above limit is uniform in x∈[−L,L]. The comparison principle implies that (0,0)≤(u(x,t+1),v(x,t+1))≤(¯u(x,t),¯v(x,t)). Letting t→∞, we thus obtain limt→∞(u(x,t),v(x,t))=(0,0) uniformly in [−L,L] when λ1(L)≥0. This proves (ii).
Note that the comparison principle implies
(u_(x,t),v_(x,t))≤(u(x,t+1),v(x,t+1))≤(¯u(x,t),¯v(x,t)). |
When λ1(L)<0, letting t→∞, we deduce (U,V)≤(˜U,˜V), and the uniqueness result obtained in (i) implies (U,V)=(˜U,˜V). This in turn implies, by the above inequalities, limt→∞(u(x,t),v(x,t))=(U(x),V(x)). Moreover, this convergence is uniform in [−L,L] (by Dini's theorem again) since the convergences of (u_(x,t),v_(x,t)) and (¯u(x,t),¯v(x,t)) to (U(x),V(x)) are uniform due to their monotonicity in t and the continuity of U(x) and V(x).
Remark 3.8. By Proposition 3.4, if R0>1, then λ1(L)<0 for large enough L. Thus the steady state problem (3.7) has a unique positive solution for all large L, which we will denote by (˜uL,˜vL) to stress its dependence on L.
Following the proof of [16,Proposition 3.5], we have the following result.
Proposition 3.9. Assume (J) holds and that R0>1. Then
limL→+∞(˜uL(x),˜vL(x))=(u∗,v∗)locally uniformly inR, |
where (u∗,v∗) is defined in (1.2).
In this section, we prove Theorems 1.2 and 1.3. It is clear that h(t) and g(t) are respectively monotonically increasing and decreasing. Therefore, their limits
limt→∞h(t)=h∞∈(h0,+∞]andlimt→∞g(t)=g∞∈[−∞,−h0) |
are well-defined.
Here we look at cases where vanishing happens: either when the reproduction number R0≤1 or for sufficiently small initial data (u0,v0).
Lemma 4.1. If h∞−g∞<∞, then
limt→∞‖u‖C[g(t),h(t)]=limt→∞‖v‖C[g(t),h(t)]=0 | (4.1) |
Proof. We claim that λ1(g∞,h∞)≥0 where λ1(g∞,h∞) is the principal eigenvalue of (3.1) with [−L,L] replaced by [g∞,h∞].
Suppose by contradiction that λ1(g∞,h∞)<0. Then by Lemma 3.3, there exists a constant T>0 large enough such that λ1(g(T),h(T))<0. We may assume that g(T) and h(T) satisfy |g(T)−g∞|<ϵ and |h(T)−h∞|<ϵ with ϵ∈(0,h0) small enough such that J1(x),J2(x)>0 for x∈[−4ϵ,4ϵ]. Clearly [h(t)−2ϵ,h(t)−ϵ]⊂[g(T),h(T)] for t≥T.
Let (u1(x,t),v1(x,t)) be the solution of (3.6) with QL replaced by (g(T),h(T))×(0,∞), and initial functions (u0,v0)=(u(x,T),v(x,T)). By the comparison principle, we obtain that
u1(x,t)≤u(x,t+T),v1(x,t)≤v(x,t+T)for (x,t)∈(g(T),h(T))×(0,∞). |
By Proposition 3.7(iii), we obtain uniform convergence of (u1,v1) for x∈[g(T),h(T)], giving
0<˜u1(x):=limt→∞u1(x,t)≤lim inft→∞u(t,x) and 0<˜v1(x):=limt→∞v1(x,t)≤lim inft→∞v(t,x). |
Therefore, there exists T1≥T such that
0<12˜u1(x)<u(x,t)and0<14˜v1(x)<v(x,t)for t≥T1,x∈[g(T),h(T)]. |
Let c1,c2,c3 and c4 be constants as defined below
c1:=minx∈[−4ϵ,4ϵ]J1(x)>0,c2:=minx∈[−4ϵ,4ϵ]J2(x)>0,c3:=minx∈[g(T),h(T)]˜u1(x)>0,c4:=minx∈[g(T),h(T)]˜v1(x)>0. | (4.2) |
Then from the above, we can calculate that
h′(t)=∫h(t)g(t)∫∞h(t)[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx≥∫h(t)h(t)−2ϵ∫h(t)+2ϵh(t)[μ1J1(x−y)u(x,t)+μ2J2(x−y)v(x,t)]dydx≥∫h(t)h(t)−2ϵ[2μ1c1ϵu(x,t)+2μ2c2ϵv(x,t)]dx≥∫h(t)−ϵh(t)−2ϵ[μ1c1ϵ˜u(x)+μ2c2ϵ˜v(x)]dx≥ϵ2(μ1c1c3+μ2c2c4)>0for t≥T1, |
which implies h∞=∞, contradicting to h∞<∞. Therefore λ1(g∞,h∞)≥0.
Now, let (u2(x,t),v2(x,t)) be the solution of (3.6) with QL replaced by (0,∞)×(g∞,h∞) and with the same initial data (u0,v0) as (u,v). By the comparison principle, we have 0≤u(x,t)≤u2(x,t) and 0≤v(x,t)≤v2(x,t) for t>0 and x∈[g(t),h(t)]. Since λ1(g∞,h∞)≥0, Proposition 3.7(ii) gives that
limt→∞(u2(x,t),v2(x,t))=(0,0)uniformly for x∈[g∞,h∞], | (4.3) |
which implies the vanishing result.
Lemma 4.2. If R0≤1, then
h∞−g∞≤2h0+μ1+μ2m0∫h0−h0[u0(x)+cbv0(x)]dx, | (4.4) |
and hence vanishing happens, where m0:=min{d1,d2cb}.
Proof. Since ∫RJi(x)dx=1 and Ji(x) is even for i=1,2, a straightforward calculation shows that
−[h′(t)−g′(t)]=μ1(∫h(t)g(t)∫h(t)g(t)J1(x−y)u(x,t)dydx−∫h(t)g(x)u(x,t)dx)+μ2(∫h(t)g(t)∫h(t)g(t)J2(x−y)v(x,t)dydx−∫h(t)g(x)v(x,t)dx). |
Moreover, we can calculate that
∫h(t)g(t)∫h(t)g(t)J1(x−y)u(x,t)dydx−∫h(t)g(x)u(x,t)dx=−∫h(t)g(t)∫∞h(t)J1(x−y)u(x,t)dydx−∫h(t)g(t)∫g(t)−∞J1(x−y)u(x,t)dydx≤0, |
and the same holds when (u,J1) is replaced by (v,J2). By the above, we obtain that
ddt∫h(t)g(t)[u(x,t)+cbv(x,t)]dx=∫h(t)g(t)[ut(x,t)+cbvt(x,t)]dx+h′(t)[u+cbv]|(h(t),t)+g′(t)[u+cbv]|(g(t),t)=∫h(t)g(t)[d1(∫h(t)g(t)J1(x−y)u(y,t)dy−u(x,t))−au(x,t)+cd2b(∫h(t)g(t)J2(x−y)v(y,t)dy−v(x,t))+cbG(u(x,t))]dx≤−min{d1,d2c/b}μ1+μ2[h′(t)−g′(t)]+∫h(t)g(t)[−au(x,t)+cbG(u(x,t))]dx. |
Using (G2), it follows from R0≤1 that −au(x,t)+cbG(u(x,t))≤0 for x∈[g(t),h(t)] and t≥0. Hence,
ddt∫h(t)g(t)[u(x,t)+cbv(x,t)]dx≤−m0μ1+μ2[h′(t)−g′(t)] for t>0. |
Integrating the above from 0 to t gives us (4.4). Then by Lemma 4.1, we obtain the vanishing result.
Now for initial data (u0,v0) small enough, we show that vanishing also occurs.
Lemma 4.3. Let λ1(h0) be the principal eigenvalue of (3.1) with L=h0. If R0>1, λ1(h0)>0 and ‖u0‖C([−h0,h0])+‖v0‖C([−h0,h0]) is sufficiently small, then vanishing happens.
Proof. Since λ1(h0)>0, there exists h1>h0 but close to h0 such that λ1(h1)>0. Let (ϕ,ψ) be a positive eigenfunction pair corresponding to λ1(h1) and
δ:=λ1(h1)2,c:=h1−h0,and M:=δc(μ1∫h1−h1ϕ(x)dx+μ2∫h1−h1ψ(x)dx)−1. |
Then define, for t≥0, x∈[−h1,h1],
¯h(t):=h0+c[1−e−δt],¯g(t):=−¯h(t),¯u(x,t):=Me−δtϕ(x),¯v(x,t):=Me−δtψ(x). |
We see that ¯h(t)∈[h0,h1) for t≥0 and if we let σ:=min{minx∈[−h0,h0]ϕ(x),minx∈[−h0,h0]ψ(x)} and
‖u0‖C[−h0,h0]+‖v0‖C[−h0,h0]≤σM, |
then we have
u0(x)≤Mϕ(x)=¯u(x,0),v0(x)≤Mψ(x)=¯v(x,0)for x∈[−h0,h0]. | (4.5) |
Clearly, we can calculate
¯ut−d1∫¯h(t)¯g(t)J1(x−y)¯u(y,t)dy+d1¯u+a¯u−c¯v≥−δ¯u−d1∫h1−h1J1(x−y)¯u(y,t)dy+d1¯u+a¯u−c¯v=Me−δt[λ1(h1)−δ]ϕ≥0, for t>0, x∈[¯g(t),¯h(t)]. |
By (G2), we obtain G(¯u)≤G′(0)¯u and hence
¯vt−d2∫¯h(t)¯g(t)J2(x−y)¯v(y,t)dy+d2¯v+b¯v−G(¯u)≥−δ¯v−d2∫h1−h1J2(x−y)¯v(y,t)dy+d2¯v+b¯b−G′(0)¯u=Me−δt[λ1(h1)−δ]ψ≥0 for t>0, x∈[¯g(t),¯h(t)]. |
Moreover, for x∈{¯g(t),¯h(t)}, we have that (¯u(x,t),¯v(x,t))≥(0,0) and
μ1∫¯h(t)¯g(t)∫∞¯h(t)J1(x−y)¯u(x,t)dydx+μ2∫¯h(t)¯g(t)∫∞¯h(t)J2(x−y)¯v(x,t)dydx≤μ1∫¯h(t)¯g(t)¯u(x,t)dx+μ2∫¯h(t)¯g(t)¯v(x,t)dydx≤Me−δt[μ1∫h1−h1ϕ(x)dx+μ2∫h1−h1ψ(x)dx]=δce−δt=¯h′(t) for t>0. |
In view of ¯g(t)=−¯h(t), we can now use the the comparison principle in Lemma 2.2 to conclude that h(t)≤¯h(t)≤h1 for all t>0, and hence vanishing happens.
Remark 4.4. If (μ1,μ2)=(μσ01,μσ02) with σ01 and σ02 fixed, nonnegative and σ01+σ02>0, then by the proof of Lemma 4.3, we see that M→∞ when μ→0. Thus, for any given initial data (u0,v0), there exists μ0>0 such that (4.5) holds for all μ∈(0,μ0). Thus if 0<μ≤μ0, then vanishing must happen for (1.8) for this given initial data (u0,v0).
We note that the following lemma implies that if h∞−g∞=∞ holds, then we must have that h∞=−g∞=+∞.
Lemma 4.5. The inequality h∞<+∞ if and only if g∞>−∞.
Proof. The proof of this lemma is similar to the proof of Lemma 4.10 in [16]. Since the modifications are obvious, we omit the details.
In this section, we look at cases where spreading happens.
Lemma 4.6. If λ1(g(t0),h(t0))<0 for some t0≥0, then h∞=−g∞=+∞ and
limt→∞u(x,t)=u∗andlimt→∞v(x,t)=v∗locally uniformly forx∈R, | (4.6) |
where λ1(g(t0),h(t0)) is the eigenvalue of (3.1) with [−L,L] replaced by [g(t0),h(t0)], and (u∗,v∗) are as defined in (1.2).
Remark 4.7. If for R0>1 and for some t0>0, we have h(t0)−g(t0)≥2L∗, then by Proposition 3.4, we obtain that λ1(h(t0),g(t0))≤0. Hence for any t1>t0 we have λ1(h(t1),g(t1))<0. Then by Lemma 4.6, we have that spreading occurs. This implies that when R0>1 and vanishing happens, we must have h(t)−g(t)<2L∗ for all t≥0.
Proof of Lemma 4.6. We see that h∞=−g∞=∞. Suppose by contradiction that h∞−g∞<∞. Since [g(t0),h(t0)]⊂[g(t),h(t)] for some t≥t0, by Corollary 3.5, we have λ1(g(t),h(t))<0 for t≥t0. Then we derive the contradiction as in Lemma 4.1. By Lemma 4.5, we further obtain that −g∞=h∞=∞. Then by Lemma 4.2, we must have R0>1, which guarantees the existence of the positive equilibrium (u∗,v∗). It remains to show (4.6).
Let us first consider the limit superior of the solution. Let (¯u,¯v) be the unique positive solution of the following ODE problem:
{¯u′=−a¯u+c¯v,t>0,¯v′=−b¯v+G(¯u),t>0,¯u(0)=‖u0‖L∞([−h0,h0]),¯v(0)=‖v0‖L∞([−h0,h0]). | (4.7) |
Since R0>1, we have limt→∞(¯u(t),¯v(t))=(u∗,v∗). We then note that
d1∫h(t)g(t)J1(x−y)¯u(t)dy−d1¯u(t)≤0,d2∫h(t)g(t)J2(x−y)¯v(t)dy−d2¯v(t)≤0, |
and ¯u(0)≥u0(x),¯v(0)≥v0(x); so by the comparison principle in Lemma 2.3, we have
(u(x,t),v(x,t))≤(¯u(t),¯v(t))for g(t)<x<h(t) and t>0. |
Thus we have that
lim supt→∞(u(x,t),v(x,t))≤(u∗,v∗)uniformly for x∈[g(t),h(t)]. |
Then following [16,Lemma 4.11], we can make use of Propositions 3.7 and 3.9 to show
lim inft→∞(u(x,t),v(x,t))≥(u∗,v∗)locally uniformly for x∈R. |
Thus (4.6) holds.
Lemma 4.8. If R0>1, h0<L∗ and (μ1,μ2)=(μσ01,μσ02) with σ01, σ02 nonnegative and σ01+σ02>0, then there exists μ0>0 depending on the initial data (u0,v0) such that spreading happens if μ>μ0.
Proof. Similar to the calculations in the proof of Lemma 4.2, by setting m0:=max{d1,d2cb}, we obtain for t>0,
∫h(t)g(t)[u(x,t)+cbv(x,t)]dx≥∫h0−h0[u0(x)+cbv0(x)]dx+m0μ(σ01+σ02)(2h0−[h(t)−g(t)]). |
Suppose that h∞−g∞<∞; then in view of Lemma 4.1 and Remark 4.7, by letting t→∞ in the above inequality, we get
∫h0−h0[u0(x)+cbv0(x)]dx≤m0μ(σ01+σ02)(2L∗−2h0). |
However, this is patently false in the case
μ>μ0:=2m0(L∗−h0)(σ01+σ02)∫h0−h0[u0(x)+cbv0(x)]dx. |
This completes the proof.
Proof of Theorem 1.2. If h∞−g∞<∞, then (1.9) holds by Lemma 4.2. On the other hand, if h∞−g∞=∞, then R0>1 by Lemma 4.2. By Corollary 3.5, we find that λ1(g(t0),h(t0))<0 for some large t0>0. Hence (4.6) holds by Lemma 4.6.
Proof of Theorem 1.3. (a) This follows from Lemma 4.2.
(b) By Proposition 3.4, we obtain that λ1(h0)≤0. Since h(t) is strictly increasing in t and λ1(L) is strictly decreasing in L, we obtain λ1(h(1))<λ1(h0)≤0. Thus by Lemma 4.6, we have that spreading occurs.
(c)(i) It follows from Remark 4.7 that if vanishing occurs, then h∞−g∞≤2L∗. Define
Γ:={μ>0:h∞−g∞≤2L∗}. |
Then by Remark 4.4 and Lemma 4.8, we respectively have that (0,μ0]⊂Γ and Γ∩(μ0,∞)=∅. Denoting by μ∗:=supΓ∈[μ0,μ0], we have by definition that h∞−g∞>2L∗ for μ>μ∗ and hence spreading happens for μ>μ∗ by Theorem 1.2.
Suppose that μ∗∉Γ. Then we have h∞−g∞=∞ when μ=μ∗ and there exists a T>0 such that h(T)−g(T)>2L∗. Let us emphasis the dependence of the solution (u,v,g,h) of (1.8) on μ by rewriting it as (uμ,vμ,gμ,hμ). Then we have hμ∗(T)−gμ∗(T)>2L∗. By the continuity of the solution in μ, hence there exists ϵ>0 such that for |μ−μ∗|<ϵ, we have hμ(T)−gμ(T)>2L∗. Then for every μ such that |μ−μ∗|<ϵ, by the monotonicity of h(t) and −g(t) in t, we have that limt→∞hμ(t)−gμ(t)>hμ(T)−gμ(T)>2L∗. Thus we get the contradiction that supΓ≤μ∗−ϵ. Hence we must have μ∗∈Γ.
It remains to show that vanishing also occurs for μ<μ∗. For every μ∈(0,μ∗), (uμ∗,vμ∗,gμ∗,hμ∗) is an upper solution to (1.8). Thus by the comparison principle, we see that hμ(t)≤hμ∗(t) and gμ(t)≥gμ∗(t) for t>0. It follows that limt→∞(hμ(t)−gμ(t))≤limt→∞(hμ∗(t)−gμ∗(t))≤2L∗. This proves our assertion.
(ii) From the assumptions, we obtain that λ1(h0)>0. Thus the assertion follows directly from Lemma 4.3.
The proof is now complete.
In this section, we consider the asymptotic spreading speed when spreading happens in our system (1.8). As such, we necessarily have that R0>1.
Let F(u,v)=(f1(u,v),f2(u,v)) with f1(u,v):=−au+cv and f2(u,v):=−bv+G(u). We now check that F satisfies the assumptions (f1)−(f4) in [2] with ˆu=∞ and m=n=2, namely
(f1): F(u,v)=(0,0) has only two nonnegative solutions (0,0) and (u∗,v∗), and the Jacobian matrix of F evaluated at (0, 0), denoted by ∇F(0,0), is irreducible with principal eigenvalue positive.
(f2): F(ku,kv)≥kF(u,v) for k∈[0,1] and all u,v≥0.
(f3): ∇F(u∗,v∗) is invertible, (u∗,v∗)∇F(u∗,v∗)≤(0,0) component wise, and for i∈{1,2},
either ∂ufi(u∗,v∗)u∗+∂vfi(u∗,v∗)v∗<0 or {∂ufi(u∗,v∗)u∗+∂vfi(u∗,v∗)v∗=0 and fi(u,v) is linear foru<u∗close tou∗andv<v∗close tov∗. |
(f4): The solution of the corresponding problem (1.12) with initial function pair (u0,v0) nonnegative, bounded and not identically (0,0) is positive and globally defined, and as time t→∞, it converges to (u∗,v∗) locally uniformly for x∈R.
It is straightforward to check that (f1), (f2) and (f3) are satisfied. It now remains for us to prove (f4), namely the following lemma.
Lemma 5.1. Let (U(x,t),V(x,t)) satisfy
{Ut=d1∫RJ1(x−y)U(y,t)dy−dU−aU+cVfor allt>0, x∈R,Vt=d2∫RJ2(x−y)V(y,t)dy−dV−bV+G(U)for allt>0, x∈R. | (5.1) |
If (U(⋅,0),V(⋅,0))∈L∞(R)2∩C(R)2 is nonnegative, then (U(x,t),V(x,t))∈[0,∞)×[0,∞) for every t>0 and x∈R. Moreover, it holds that limt→∞(U(x,t),V(x,t))=(u∗,v∗) in L∞loc(R) if additionally (U(x,0),V(x,0))≢(0,0).
Proof. Let (U(⋅,0),V(⋅,0))∈L∞(R)2∩C(R)2 be nonnegative. If (U(x,0),V(x,0))≡(0,0), then clearly (U,V)≡(0,0) is the unique solution of (5.1). In the following we assume that (U(x,0),V(x,0))≢(0,0). For L>0, let (uL(x,t),vL(x,t)) be the solution to (3.6) with initial data (uL(x,0),vL(x,0))=(U(x,0),V(x,0))|[−L,L]. By the comparison principle in Lemma 2.3, we get that
(0,0)≤(uL(x,t),vL(x,t))≤(U(x,t),V(x,t))for (x,t)∈[−L,L]×(0,∞). |
Let (¯u(t),¯v(t)) be the unique positive solution of the following system of ordinary differential equations:
{¯u′=−a¯u+c¯v,t>0,¯v′=−b¯v+G(¯u),t>0,¯u(0)=‖U(⋅,0)‖L∞(R),¯v(0)=‖V(⋅,0)‖L∞(R). | (5.2) |
By the comparison principle in Lemma 2.3, we obtain (U(x,t),V(x,t))≤(¯u(t),¯v(t)) for (x,t)∈[−L,L]×(0,∞). Since L≥L0 is arbitrary, this and the earlier estimates imply that
(U(x,t),V(x,t))∈[0,∞)×[0,∞)for every t>0 and x∈R. |
Moreover, it follows from R0>1 that limt→∞(¯u(t),¯v(t))=(u∗,v∗). Therefore we must have
lim supt→∞(U(x,t),V(x,t))≤(u∗,v∗) uniformly for x∈R. | (5.3) |
Since (U(x,0),V(x,0))≢(0,0) for x∈R, there exists L0>L∗ large enough such that
(U(x,0),V(x,0))|[−L,L]≢(0,0)forx∈[−L,L]whenL≥L0. |
By Proposition 3.7(iii), we obtain
limt→∞(uL(x,t),vL(x,t))=(˜uL(x),˜vL(x)) uniformly for x∈[−L,L], L≥L0>L∗, |
where (˜uL,˜vL) is the unique positive steady-state of (3.6). It follows that
lim inft→∞(U(x,t),V(x,t))≥(˜uL(x),˜vL(x)) uniformly for x∈[−L,L], L≥L0>L∗. |
Letting L→∞, by Proposition 3.9, we obtain
lim inft→∞(U(x,t),V(x,t))≥(u∗,v∗)locally uniformly in R. |
This and (5.3) imply
limt→∞(U(x,t),V(x,t))=(u∗,v∗)locally uniformly in R. |
The proof is complete. Since F=(f1,f2) satisfies (f1)−(f4) in [2], Theorems 1.3 and 1.5 (as well as two results on the associated semi-wave problem: Theorems 1.1 and 1.2) in [2] can be applied to obtain Theorems 1.5 and 1.6 here.
Y. Du's research was supported by the Australian Research Council. We thank the referees for their useful suggestions which helped to improve the presentation of the paper.
All authors declare no conflicts of interest in this paper.
[1] |
M. Zhao, Y. Zhang, W.-T. Li, Y. Du, The dynamics of a degenerate epidemic model with nonlocal diffusion and free boundaries, J. Diff. Equ., 269 (2020), 3347–3386. https://doi.org/10.1016/j.jde.2020.02.029 doi: 10.1016/j.jde.2020.02.029
![]() |
[2] | Y. Du, W. Ni, Semi-wave, traveling wave and spreading speed for monostable cooperative systems with nonlocal diffusion and free boundaries, (2020), arXiv: 2010.01244. |
[3] | V. Capasso, S. L. Paveri-Fontana, A mathematical model for the 1973 cholera epidemic in the European Mediterranean region, Revue d'epidemiologie et de sante publique, 27 (1979), 121–132. |
[4] | V. Capasso, L. Maddalena, Convergence to equilibrium states for a reaction-diffusion system modelling the spatial spread of a class of bacterial and viral diseases, J. Math. Biol., 13 (1981/82), 173–184. https://doi.org/10.1007/BF00275212 |
[5] |
I. Ahn, S. Baek, Z. Lin, The spreading fronts of an infective environment in a man-environment-man epidemic model, Appl. Math. Model., 40 (2016), 7082–7101. https://doi.org/10.1016/j.apm.2016.02.038 doi: 10.1016/j.apm.2016.02.038
![]() |
[6] | L. I. Rubinšteĭn, The Stefan Problem, American Mathematical Soc., Providence, RI, 1971. |
[7] |
Y. Du, Z. Lin, Spreading-vanishing dichotomy in the diffusive logistic model with a free boundary, SIAM J. Math. Anal., 42 (2010), 377–405. https://doi.org/10.1137/090771089 doi: 10.1137/090771089
![]() |
[8] |
M. Zhao, W.-T. Li, W. Ni, Spreading speed of a degenerate and cooperative epidemic model with free boundaries, Discrete Contin. Dyn. Syst. B, 25 (2020), 981–999. https://doi.org/10.3934/dcdsb.2019199 doi: 10.3934/dcdsb.2019199
![]() |
[9] |
J.F. Cao, Y. Du, F. Li, W.T. Li, The dynamics of a Fisher-KPP nonlocal diffusion model with free boundaries, J. Funct. Anal., 277 (2019), 2772–2814. https://doi.org/10.1016/j.jfa.2019.02.013 doi: 10.1016/j.jfa.2019.02.013
![]() |
[10] |
R. Wang, Y. Du, Long-time dynamics of a diffusive epidemic model with free boundaries, Discrete Contin. Dyn. Syst.Ser. B, 26 (2021), 2201–2238. https://doi.org/10.3934/dcdsb.2020360 doi: 10.3934/dcdsb.2020360
![]() |
[11] |
W.B. Xu, W.-T. Li, S. Ruan, Spatial propagation in an epidemic model with nonlocal diffusion: The influences of initial data and dispersals, Sci. China Math., 63 (2020), no. 11, 2177–2206. https://doi.org/10.1007/s11425-020-1740-1 doi: 10.1007/s11425-020-1740-1
![]() |
[12] |
C. Hu, Y. Kuang, B. Li, H. Liu, Spreading speeds and traveling wave solutions in cooperative integral-differential systems, Discrete Contin. Dyn. Syst. Ser. B, 20 (2015), 1663–1684. https://doi.org/10.3934/dcdsb.2015.20.1663 doi: 10.3934/dcdsb.2015.20.1663
![]() |
[13] |
J. Fang, X-Q. Zhao, Monotone wavefronts for partially degenerate reaction-diffusion systems, J. Dyn. Differ. Equ., 21 (2009), 663–680. https://doi.org/10.1007/s10884-009-9152-7 doi: 10.1007/s10884-009-9152-7
![]() |
[14] |
B. Li, H. F. Weinberger, M. A. Lewis, Spreading speeds as slowest wave speeds for cooperative systems, Math. Biosci., 196 (2005), 82–98. https://doi.org/10.1016/j.mbs.2005.03.008 doi: 10.1016/j.mbs.2005.03.008
![]() |
[15] |
X.-Q. Zhao, W. Wang, Fisher waves in an epidemic model, Discrete Contin. Dyn. Syst. Ser. B, 4 (2004), 1117–1128. https://doi.org/10.3934/dcdsb.2004.4.1117 doi: 10.3934/dcdsb.2004.4.1117
![]() |
[16] |
Y. Du, W. Ni, Analysis of a West Nile virus model with nonlocal diffusion and free boundaries, Nonlinearity, 33 (2020), 4407–4448. https://doi.org/10.1088/1361-6544/ab8bb2 doi: 10.1088/1361-6544/ab8bb2
![]() |
[17] |
X. X. Bao, W. Shen, Criteria for the existence of principal eigenvalues of time periodic cooperative linear systems with nonlocal dispersal, Proc. Amer. Math. Soc., 145 (2017), 2881–2894. https://doi.org/10.1090/proc/13602 doi: 10.1090/proc/13602
![]() |
[18] |
Y. Du, F. Li, M. Zhou, Semi-wave and spreading speed of the nonlocal Fisher-KPP equation with free boundaries, J. Math. Pure Appl., 154 (2021), 30–66. https://doi.org/10.1016/j.matpur.2021.08.008 doi: 10.1016/j.matpur.2021.08.008
![]() |
1. | Yihong Du, Propagation and reaction–diffusion models with free boundaries, 2022, 12, 1664-3607, 10.1142/S1664360722300018 | |
2. | Rong Wang, Yihong Du, Long-time dynamics of a nonlocal epidemic model with free boundaries: Spreading-vanishing dichotomy, 2022, 327, 00220396, 322, 10.1016/j.jde.2022.04.031 | |
3. | Rong Wang, Yihong Du, Long-time dynamics of an epidemic model with nonlocal diffusion and free boundaries: Spreading speed, 2023, 43, 1078-0947, 121, 10.3934/dcds.2022143 | |
4. | Yihong Du, Wenjie Ni, Rong Wang, Rate of accelerated expansion of the epidemic region in a nonlocal epidemic model with free boundaries, 2023, 36, 0951-7715, 5621, 10.1088/1361-6544/acf63c | |
5. | Yachun Tong, Inkyung Ahn, Zhigui Lin, Threshold dynamics of a nonlocal dispersal SIS epidemic model with free boundaries, 2025, 18, 1793-5245, 10.1142/S179352452350095X |