Skip to content
BY 4.0 license Open Access Published by De Gruyter December 31, 2020

Analysis of a diffusive host-pathogen model with standard incidence and distinct dispersal rates

  • Jinliang Wang and Renhao Cui EMAIL logo

Abstract

This paper concerns with detailed analysis of a reaction-diffusion host-pathogen model with space-dependent parameters in a bounded domain. By considering the fact the mobility of host individuals playing a crucial role in disease transmission, we formulate the model by a system of degenerate reaction-diffusion equations, where host individuals disperse at distinct rates and the mobility of pathogen is ignored in the environment.We first establish the well-posedness of the model, including the global existence of solution and the existence of the global compact attractor. The basic reproduction number is identified, and also characterized by some equivalent principal spectral conditions, which establishes the threshold dynamical result for pathogen extinction and persistence. When the positive steady state is confirmed, we investigate the asymptotic profiles of positive steady state as host individuals disperse at small and large rates. Our result suggests that small and large diffusion rate of hosts have a great impacts in formulating the spatial distribution of the pathogen.

MSC 2010: 35K57; 92A15; 37N25; 35B40

1 Introduction

In recent years, the studies of some reaction-diffusion host-pathogen models have received much attentions, as the investigation of these systems allow us to get better understanding the interactions between host and pathogens and the mechanisms of the disease spread. Let u1(x, t), u2(x, t) and u3(x, t) be the densities of susceptible hosts, infected hosts and pathogen particles at the spatial location x and time t. Laplacian operator accounts for the host movement (d > 0 is the diffusion coefficient). In [6], under a one-dimensional and unbounded domain, the authors considered the following model,

(1.1) u1t=dΔu1+r1u1+u2Ku1βu1u3,xR,t>0,u2t=dΔu2+βu1u3αu2ru1+u2Ku2,xR,t>0,u3t=λu2δu3,xR,t>0.

where r and K represent respectively the reproductive rate and the carrying capacity of the susceptible and infected hosts; β and α are the transmission coefficient and disease-induced mortality rate; λ and δ are respectively the pathogens production rate from infected hosts and the pathogens’ decay rate. All above mentioned coefficients in (1.1) are assumed to be positive constants. With the consideration of spatial spread of pathogens, the authors in [6] investigated the existence problem of traveling wave solution.

In reality, the habitats where hosts live should be a spatially bounded domain. Wang et al. [29] took the model (1.1) as a basis and extended it to a more general model in bounded spatial domain Ω ∈ ℝn with smooth boundary ∂Ω,

(1.2) u1t=dΔu1+r1u1+u2K(x)u1β(x)u1u3,(x,t)Ω×(0,),u2t=dΔu2+β(x)u1u3αu2ru1+u2K(x)u2,(x,t)Ω×(0,),u3t=λ(x)u2δu3β(x)(u1+u2)u3,(x,t)Ω×(0,),u1n=u2n=0,(x,t)Ω×(0,),

where u1n stands for the differentiation along the unit outward normal n to ∂Ω; β(x)(u1 + u2)u3 represents the consumption of the pathogen due to the interaction with the hosts. Unlike in model (1.1), where parameters β, K, λ are constants, model (1.2) allows the space-dependent parameter functions, β(·), K(·) and λ(·), which are used to obey the spatial heterogeneity arising from the variance in environmental conditions (for example, temperature and humidity etc). Compared to model (1.1), the space-dependent functions β(x), K (x) and λ(x) are assumed to be continuous and positive in Ω. Since the mobility of pathogen is ignored in the domain, the semiflow induced by solution lacks of compactness. The authors in [29] overcame this difficulty by verifying the k-contracting condition. The basic reproduction number (BRN) is proved to be a threshold index for the dynamics of (1.2), and defined by adopting the concept of next generation operator (NGO). Bifurcation analysis for steady state solution are also carried out when space-dependent parameters are used.

Note that the susceptible and infected hosts in (1.1) and (1.2) share the same diffusion rate. Wu and Zou [31] further generalized the model (1.2) with distinct diffusion rates d1 and d2,

(1.3) u1t=d1Δu1+y(x)μ(x)u1β(x)u1u3,(x,t)Ω×(0,),u2t=d2Δu2+β(x)u1u3υ(x)u2,(x,t)Ω×(0,),u3t=α(x)u2δ(x)u3,(x,t)Ω×(0,),u1n=u2n=0,(x,t)Ω×(0,),

Here, d1 and d2 are not necessarily equal. The simplest growth term for susceptible host, y(x) − μ(x)u1, is used. As pointed in [31] that distinct dispersal rates for hosts brings difficulty in proving the boundedness of the solution. The existence of global attractor needs appealing the general result in [8, Theorem 2.4.6] and Arezelà-Ascoli Theorem, so that the the asymptotic smoothness of semiflow is used instead of weak compactness by verifying the k-contraction condition. The authors studied the threshold dynamics of (1.3) and the effects of the spatial heterogeneity on disease dynamics. Wu and Zou in [31] also explored the asymptotic profiles of positive steady state for the case where d1 → 0 and d2 → 0.

Subsequently, inspired by the work [31], Shi et al. [23] revisited the model (1.2) by incorporating the horizontal disease transmission. The authors in [23] established the threshold dynamics, and performed the bifurcation analysis for steady state solution. Based on framework of [31], Wang and Wang [32] further extended the model (1.3) by adding the horizontal transmission term to (1.3). In [32], u1, u2 and u3 are used to stand for respectively the density of susceptible, infected individuals and the concentration of vibrios in the environment. The horizontal transmission is termed as direct person-to-person transmission route in cholera dynamics. The complete analysis of (1.3) with horizontal transmission term demonstrated the threshold dynamics of (1.3) and the effect of the spatial heterogeneity on disease dynamics, and also revealed that ignoring horizontal transmission will underestimate the risk of disease transmission.

Our current work is also inspired by a series of works on diffusive models for disease dynamics in the spatial heterogeneous environment. These models are in the form of reaction-diffusion susceptible-infected-susceptible (SIS) equations in spatially bounded domain and the main concern are how the spatial heterogeneity and the diffusion affect the disease spread and control [1, 9, 10, 21, 31,32,33]. In an earlier article [1], the authors studied a SIS model with frequency-dependent interaction,

(1.4) St=dSΔSβ(x)SIS+I+y(x)I,(x,t)Ω×(0,),It=dIΔI+β(x)SIS+Iy(x)I,(x,t)Ω×(0,),Sn=In=0,(x,t)Ω×(0,),Ω(S(x,0)+I(x,0))dxN>0,

where S(x, t) and I(x, t) represent respectively the density of susceptible and infected individuals. y(x) and β(x) are respectively the space-dependent recovery rate and disease transmission rate. dS and dI represent respectively the diffusion rates of susceptible and infected individuals. The total number of human individuals remains constant N. The main concern in the aspect of biological implication is: limiting the flow of susceptible individuals (dS → 0) can eliminate the disease, provided that the disease is of low risk (i.e., β(·) < y(·) for xΩ). This pioneering work start up the investigation that how the spatial heterogeneity and the diffusion affect the disease spread and control. Subsequently, the result that limiting the flow of the infected individuals can not eliminate the disease (see in [21]) revealed that dS and dI play different role in disease control. Motivated by meaningful and important aspect of spatial heterogeneity of environment and distinct dispersal rates, Wu and Zou [33] further modified the model (1.4) by replacing the frequency-dependent interaction with mass action mechanisms. They showed that an additional condition on the total population is needed for disease control if dS → 0. Additionally, disease can not be controlled when dS → 0 and the total population accounts large, and inversely, disease disappears in certain area when dI → 0. In contrast with [1, 21, 33], Li et al. [9, 10] analyzed an spatial SIS model with linear source and logistic source (which allows varying total population). With small and large diffusion rates, both of the works revealed that disease can not be controlled, which is not a good situation in disease control.

Based on the above mentioned works, we continue to explore how diffusion rates and the spatial heterogeneity affect the dynamics of (1.3) by incorporating the frequency-dependent interaction used in (1.4), and thus, can be considered as a continuation of the work [1, 9, 21, 31,32,33]. We shall explore the following system,

(1.5) u1t=d1Δu1+b(x)β1(x)u1u2u1+u2β2(x)u1u3ζ(x)u1,(x,t)Ω×(0,),u2t=d2Δu2+β1(x)u1u2u1+u2+β2(x)u1u3y(x)u2,(x,t)Ω×(0,),u3t=ϱ(x)u2δ(x)u3,(x,t)Ω×(0,),u1n=u2n=0,(x,t)Ω×(0,),

with the initial condition

(1.6) ui(x,0)=ui0(x),xΩ,i=1,2,3,

where ζ (x) and y(x) represent the death rates of susceptible and infected host; b(x) is the recruitment rate; βi(x)(i = 1, 2) represent the disease transmission rate; δ(x) is the pathogens’ decay rate; ϱ(x) is the pathogens’ production rate. By replacing the frequency-dependent interaction with mass action mechanisms (direct person-to-person transmission route), model (1.5) reduced to the model in [32]. Even though (1.5) bears a resemblance to that in [32], there is one major difference: frequency-dependent interaction β1(x)u1u2u1+u2 assumes bounded infection force, while unbounded infection mechanism for mass action term.

Note that when d1 = d2 = 0 and all parameters are space-independent, the reduced system of (1.5) is the same as the model [24] with standard incidence function for the direct disease transmission and bilinear incidence function for indirect disease transmission. It is mentioned in [24] that disease transmission within and without groups may be different. In this sense, it is natural to assume that the contact probability between a susceptible and an infected host is decreasing function of total host population (standard incidence function) and the contact probability between a susceptible host and pathogen is a constant (bilinear incidence function). Meanwhile, the model studied in [24] can also provide us a biological interpretation for our model (1.5) that: 1) In Zika virus transmission, u1, u2 can respectively stands for uninfected individuals, infected individuals, and u3 stands for the infected mosquitoes in the local landscape; 2) In H1N1 and seasonal influenzas, u1, u2 represent respectively the uninfected and infected individuals, and u3 represents the contaminated environment such as classrooms, or other public places; 3) In the transmission of avian influenza, u1, u2 can respectively stand for the uninfected migratory birds, infected migratory birds, and u3 stands for infected domestic poultry. On the other hand, (1.5) can be used to describe the transmission of cholera in the sense that u1, u2 and u3 stand for respectively the density of susceptible, infected individuals and the concentration of vibrios in the environment. However, it is well-known that those who recovered from cholera do lose immunity. A realistic excuse to justify the hypothesis that infected individuals do not lose immunity may be that if we care about this model during one outbreak, then those who recovered are very likely to remain immune throughout the outbreak.

We plan to proceed this paper as follows. Section 2 shall pay attention to the well-posedness of (1.5) with (1.6) such as, the global existence and uniqueness and ultimate boundedness of solution of (1.5), the asymptotic smoothness condition of semiflow, the existence of global compact attractor. In Section 3, we identify the basic reproduction number, ℜ0. We also establish that ℜ0 can be equivalently characterized as the principal spectral conditions. In section 4, with the ℜ0, detailed analysis are carried out on the threshold dynamics of (1.5), that is, ℜ0 predicts whether or not the disease persist. In a critical case that ℜ0 = 1, the global asymptotic stability of disease free steady state is also addressed. Section 5 is spent on the dynamics of (1.5) in homogeneous case. We addressed the existence of unique positive equilibrium and local stability. The global attractivity of positive equilibrium with additional condition is achieved by the technique of Lyapunov function. Section 6 is devoted to exploring the asymptotic profiles of the positive steady state for the case that d1 → 0, d1 → ∞, d2 → 0 and d2 → ∞. Finally, detailed conclusions are drawn and some discussion is presented.

2 Well-posedness of the problem

The main aim of this section is to confirm that (1.5) has a global compact attractor. Throughout of the paper,

  • X:=C(Ωˉ,R3),Y:=C(Ωˉ,R2),Z:=C(Ωˉ,R) are respectively the Banach space with the supremum norm ∥ · ∥, and their positive cone are denoted by X+:=C(Ωˉ,R+3),Y+:=C(Ωˉ,R+2),Z+:=C(Ωˉ,R+) , respectively.

G:=maxxΩˉG(),G=minxΩˉG(),

where G(·) = y(·), δ(·), ϱ(·), ζ (·), β1(·), β2(·), respectively.

  • Γ is the Green function of ut = Δu in Ω with the Neumann boundary condition.

  • A1(t),A2(t):C(Ωˉ,R)C(Ωˉ,R) are respectively the C0 semigroups of d1Δζ (x) and d2Δy(x) with the Neumann boundary condition, i.e., for any φC(Ωˉ,R) ,

(2.1) (A1(t)φ)()=eζ()tΩΓ(d1t,,y)φ(y)dy,

and

(2.2) (A2(t)φ)()=ey()tΩΓ(d2t,,y)φ(y)dy.

We further, denote

(A3(t)φ)(x)=eδ(x)tφ(x).

According to [22, Section 7.1], A1(t),A2(t):C(Ωˉ,R)C(Ωˉ,R),t>0, are strong positive and compact. Hence, for any φ ∈ 𝕏+,

(A(t)φ)():=((A1(t)φ)(),(A2(t)φ)(),(A3(t)φ)())T=eζ()tΩΓ(d1t,,y)φ1(y)dyey()tΩΓ(d2t,,y)φ2(y)dyeδ()tφ3(),t0,

forms a C0 semigroup on 𝕏 preserving 𝕏+, namely, 𝒜(t)𝕏+ ⊂ 𝕏+, t ≥ 0 (see, for example, [19]). Further, let

F(φ):=(F1(φ)(),F2(φ)(),F3(φ)())=b()β1()φ1φ2φ1+φ2β2()φ1φ3β1()φ1φ2φ1+φ2+β2()φ1φ3ϱ()φ2.

By these settings, we can rewrite (1.5) as

(2.3) u(t)=A(t)φ+0tA(ts)F(u(s))ds.

We can easily check that

limh0+dist(φ+hF(φ),X+)=0,φX+.

In fact, it is easy to verify that for any φ ∈ 𝕏+ and small enough h ≥ 0,

φ+hF(φ)=φ1+hb()β1()φ1φ2φ1+φ2β2()φ1φ3φ2+hβ1()φ1φ2φ1+φ2+β2()φ1φ3φ3+hϱ()φ2φ11hβ1φ2ϕ1+φ2+β2φ3φ2φ3.

As just a consequence of [22, Corollary 4], we have

Lemma 2.1

For any ϕ ∈ 𝕏+, (1.5) with (1.6) admits a unique nonnegative solution u(·, t) := (u1(·, t), u2(·, t), u3(·, t)) on Ωˉ × [0, tmax) with tmax ≤ ∞. Furthermore, u(·, t; ϕ) ∈ 𝕏+, t ∈ [0, tmax).

In what follows, we shall confirm tmax = ∞by verifying the boundedness of u(·, t) in Ω × (0, tmax).

2.1 Existence of the global solution

Theorem 2.1

For any ϕ ∈ 𝕏+, (1.5) with (1.6) admits a unique nonnegative solution defined on Ωˉ × [0,∞). Furthermore, the semiflow Y(t) : 𝕏+ → 𝕏+ defined by

(2.4) Y(t)ϕ=(u1(,t;ϕ),u2(,t;ϕ),u3(,t;ϕ)),(x,t)Ωˉ×[0,),

is ultimately bounded.

Proof. We first prove that the boundedness of u1. From (1.5), u1(x, t) is governed by

(2.5) uˉ1t=d1Δuˉ1+b()m()uˉ1,(x,t)Ω×(0,),uˉ1ν=0,(x,t)Ω×(0,).

By the comparison principle, we directly have

(2.6) lim suptu1(,t)limtuˉ1(,t)=uˉ1(),uniformly forxΩˉ,

where uˉ1() is a unique global asymptotic stable steady state of (2.5) in C(Ωˉ,R). It follows that

u1(,t)∥≤M0,

where M0=∥uˉ1(), independent of initial conditions.

Next, we verify that the solution u2 and u3 of (1.5) are ultimately bounded. For this purpose, we first establish the uniform estimate in L1(Ω) × L1(Ω) and then pass it to the uniform estimate in C(Ω) × C(Ω).

  • u(·, t) satisfies the L1 bounded estimate, i.e.,

lim supt(u1(,t)L1+u2(,t)L1+u3(,t)L1)M1,

where 𝕄1 is a positive constant.

By (1.5), we have

ddtu1(,t)+u2(,t))L1=Ωb()dxΩζ()u1(,t)dxΩy()u2(,t)dx|Ω|bhu1(,t)+u2(,t)L1,

where h=minxΩˉ{ζ,y}and|Ω| is the volume of Ω. Hence,

lim supt(u1(,t)L1+u2(,t)L1)M1,whereM1=|Ω|b/h.

Further, we have

ddtu3(,t)L1ϱM1δu3(,t)L1.

It follows that

lim suptu3(,t)L1M2,whereM2=ϱM1/δ.

Hence, by taking 𝕄1 = max{M1, M 2}, the assertion directly follows.

  • Fork ≥ 0, u2 and u3 satisfies the L2kˉ bounded estimate, that is,

(2.7) lim suptu2(,t)2k+u3(,t)2kM2k,

where M 2k is a positive constant.

We shall prove (2.7) by induction. Obviously, k = 0 holds. Suppose that (2.7) holds for k−1. Then forM2k1>0, we have

(2.8) lim suptu2(,t)2k1+u3(,t)2k1M2k1.

We multiply u2 by u22k1 , and then integrate it over Ω,

(2.9) 12ktΩu22kdxd2Ωu22k1u2dx+Ωβ2()u1u3u22k1dx+Ωβ1()u22kdxΩy()u22kdx.

Recall that

d2Ωu22k1u2dx=d2Ωu2u22k1dx=(2k1)d2Ω(u2u2)u22k2dx=2k122k2d2Ω|u22k1|2dx.

Hence, (2.9) becomes

(2.10) 12ktΩu22kdxDkΩ|u22k1|2dx+Ωβ2()u1u3u22k1dx+Ωβ1()u22kdxΩy()u22kdx.

where Dk=2k122k2d2.

By lim supt→u1(·, t)∥L1M0, there is t0 > 0 such that for tt0,

(2.11) Ωβ1()u22kdxβ1Ωu22kdx

and

(2.12) Ωβ2()u1u22k1u3dxβ2(M0+1)Ωu3u22k1dx.

Applying Young’s inequality: abϵap+1q(ϵp)qpbq, where a, b, ϵ > 0, 1 < p, q < ∞and 1p+1q=1. One can estimate (2.12) by setting ϵ1=δ4β2(M0+1),p=2kandq=2k/(2k1) as follows,

(2.13) Ωu3u22k1dxδ4β2(M0+1)Ωu32kdx+Cϵ1Ωu22kdx,fortt0,whereCϵ1=2k12k(2kϵ1)12k1.

Thus, (2.10) can be estimated by

(2.14) 12ktΩu22kdxDkΩ|u22k1|2dx+δ4Ωu32kdx+CkΩu22kdx,

where Ck=β1+β2(M0+1)Cϵ1.

We multiply u3byu32k1, and then integrate it over Ω,

(2.15) 12ktΩu32kdx=Ωϱ(x)u32k1u2dxΩδ(x)u32kdxϱΩu32k1u2dxδΩu32kdx.

Again applying Young’s inequality (by setting ϵ2=δ4ϱ,p=2k/(2k1)andq=2k), we have

(2.16) Ωu32k1u2dxδ4ϱΩu32kdx+Cϵ2Ωu22kdx,whereCϵ2=ϵ212k(2k1)2k1(2k)2k.

Hence (2.15) becomes

(2.17) 12ktΩu32kdx3δ4Ωu32kdx+ϱCϵ2Ωu22kdx.

Combined with (2.14) and (2.17), we obtain

(2.18) 12ktΩ(u22k+u32k)dxDkΩ|u22k1|2dx+EkΩu22kdxδ2Ωu32kdx,fortt0,

where Ek=Ck+ϱCϵ2.

Applying interpolation inequality:

ξ22ϵξ22+Cϵξ12,whereϵ>0,Cϵ>0,ξW1,2(Ω).

Let ϵ3=Dk/(2Ek),ξ=u22k1, then

DkΩ|u22k1|2dx2EkΩu22kdx+2EkCϵ3Ωu22k1dx2.

Thus, (2.18) becomes

(2.19) 12ktΩ(u22k+u32k)dxrΩ(u22k+u32k)dx+2EkCϵ3Ωu22k1dx2,fortt0,

where r* = minr=min{Ek,δ2}.

It then follows from (2.8) that lim suptΩu22k1dxM2k12k1, which in turn implies that

lim suptu2(,t)2k+u3(,t)2kM2k,withM2k=2EkCϵ3r2kM2k1.

Thus, according to continuous embedding Lq(Ω) ⊂ Lp(Ω), qp ≥ 1, we have

(2.20) lim supt(u2(,t)Lp+u3(,t)Lp)Mp,

where M p > 0, independent of initial conditions. Denote by Ya, 0 ≤ a ≤ 1 the fractional power space. By [31, Lemma 2.4], we obtain that YaC(Ωˉ) by selecting p > n/2 and an/(2p). Hence lim supt→u2(·, t)∥ ≤ M, where M > 0. Further, lim suptu3(,t)ϱMδ. This proves the ultimate boundedness of the solution of (1.5). The proof is cpmplete.

2.2 Asymptotic smoothness of Y(t)

We refer the readers to consult [8, 36] for the definition of κ(·), the Kuratowski measure of noncompactness.

Lemma 2.2

The semigroup Y(t) is a κ-contraction on 𝕏+, that is, for any bounded set 𝔹 ⊆ 𝕏+,

κ(Y(t)B)eδtκ(B),

where δ=minxΩˉδ(x).

Proof. For t ≥ 0, let Y(t) = Y1(t) + Y2(t), where

Y1(t)ϕ=u1(,t;ϕ),u2(,t;ϕ),0teδ()(ts)ϱ()u2(,s;ϕ)ds,

and

Y2(t)ϕ=0,0,eδ()tu30(x).

By [31, Lemma 2.5], Y1(t)𝔹 is precompact. Thus, we have κ(Y1(t)𝔹) = 0. We estimate Y2(t) as

Y2(t)=supψXY2(t)ψXψXeδtsupψXψXψX=eδt.

Hence,

κ(Y(t)B)κ(Y1(t)B)+κ(Y2(t)B)0+Y2(t)κ(B)eδtκ(B),

that is, Y(t) is κ-contraction on 𝕏+. This completes the proof. □

Theorem 2.2

(1.5) possesses a global compact attractor in 𝕏+, denoted by 𝒜0.

Proof. By Theorem 2.1, the solution of (1.5) exists globally and Y(t) is point dissipative (see Lemma 2.2). Asymptotic smoothness condition of Y(t) is implied by κ-contraction condition. Hence, by [8, Theorem 2.4.6], system (1.5) possesses a global compact attractor, which attracts every bounded set in 𝕏+.  □

3 Basic reproduction number

Obviously, (1.5) has a disease-free steady state E0=(u10(x),0,0),whereu10(x)=b(x)ζ(x) and satisfies dΔu10(x)+b(x)m(x)u10(x)=0,forxΩandu10(x)ν=0,forxΩ. The linearization of (1.5) at E0 leads to the following cooperative system,

(3.1) u2tu3t=Bu2u3,(x,t)Ω×(0,),u2n=0,(x,t)Ω×(0,),

where

(3.2) B=d2Δ+β1()y()β2()u10()ϱ()δ().

Substituting (u2(·, t), u3(·, t)) := eλt(ϕ2(·), ϕ3(·)) into (3.1) gets

(3.3) λϕ2ϕ3=Bϕ2ϕ3,(x,t)Ω×(0,),ϕ2n=0,(x,t)Ω×(0,).

We rewrite

B:=B+F=d2Δ+β1()y()0ϱ()δ()+0β2()u10()00.

It is easy to see that both 𝓑 and B are resolvent-positive operators [27]. Denote by T(t)(resp. T˜(t)):YY the positive semigroup generated by 𝓑 (resp. B). Since both 𝓑 and B are cooperative for any xΩ, we get that T(t)𝕐+ ⊆ 𝕐+ (resp. (t)𝕐+ ⊆ 𝕐+). Throughout of the paper, we denote by s(Q) = sup{Reλ, λσ(Q)} the spectral bound of Q and r(Q) = sup{|λ|, λσ(Q)}, the the spectral radius of Q.

Following the standard procedures in [27, 30], we define the NGO 𝓛 as

L[ϕ]()=0F()T˜(t)ϕ()dt=F()0T˜(t)ϕ()dt,ϕY,xΩˉ,

that is, within the infection period, 𝓛 [ϕ](·) represents the total new infections distribution from initial distribution ϕ. Then, the BRN ℜ0 is defined as

(3.4) 0:=r(L)

The following result is a consequence of [27, 30].

Lemma 3.1

Let 𝓑 and0 be defined in (3.4) and (3.2), respectively. s(𝓑) has the same sign as0 − 1.

Proof. Due to the fact that B is resolvent-positive operator, we then have

(3.5) (λIB)1ϕ=0eλtT˜(t)ϕdt,λ>s(B),ϕX+.

Due to s(B) < 0, we can let λ = 0 in (3.5), leading to

(3.6) B1ϕ=0T˜(t)ϕdt,λ>s(B),ϕX+.

That is, 𝓛 = −FB−1. Further, 𝓑 = B + F can be viewed as the perturbation of B. According to [27, Theorem 3.5], s(𝓑) has the same sign as r(−FB−1) − 1 = ℜ0 − 1. □

For the convenience of forthcoming discussions, we next claim that ℜ0 has relationship with other important indicators: λ˜0 andη0.

Lemma 3.2

Let0 be defined by (3.4). Then we have

(i) 0=1/λ˜0,whereλ˜0 is the principal eigenvalue of

(3.7) d2Δφy()φ+λ˜β1()+β2()u10()ϱ()δ()φ=0,xΩ,φν=0,xΩ;

(ii)0 − 1 and s(𝓑) have the same sign as η0, where η0 is the principal eigenvalue of

(3.8) d2Δϕ+β1()+β2()u10()ϱ()δ()y()ϕ=ηϕ,xΩ,ϕν=0,xΩ.

Proof. We first prove (i). We rewrite

F=β1()β2()u10()00=F11F12F21F22,B=diag(d2Δ,0)V,

where

V=y()0ϱ()δ()=V11V12V21V22.

As F21 = 0, F22 = 0, by [30, Theorem 3.3], we know that 0=r(B1F)=r(B11F2), where

B1=d2ΔV11+V12V221V21=d2Δy()andF2=F11F12V221V21=β1()+β2u10()ϱ()/δ().

Hence,

(B11F2)φ=(d2Δy())1β1()+β2()u10()ϱ()δ()φ.

Therefore, ℜ0 satisfies

(d2Δy())1β1()+β2()u10()ϱ()δ()φ=0φ,φC2(Ωˉ),

that is,

(3.9) d2Δφy()φ+β1()+β2()u10()ϱ()δ()10φ=0,φC2(Ωˉ).

This proves (i).

We next prove (ii). In fact, eigenvalue problem (3.8) has a principle eigenvalue η0, associated with a positive eigenfunction ϕ* on Ω, i.e.,

(3.10) d2Δϕy()ϕ+β1()+β2()u10()ϱ()δ()ϕ=η0ϕ,xΩ,ϕν=0,xΩ.

We then multiply the first equation of (3.10) and (3.9) by φ and ϕ*, respectively, then integrate and subtract the equation, obtaining

110Ωβ1()+β2()u10()ϱ()δ()ϕφ=η0Ωϕφ.

Since both Ωβ1(x)+β2(x)u10(x)ϱ(x)δ(x)ϕΦandΩϕΦ are positive, we arrive at the conclusion that 110 and η0 have the same sign, i.e., ℜ0 > 1 when η0 > 0 and ℜ0 < 1 when η0 < 0. This proves (ii). □

Due to the assertion in (i) of Lemma 3.2 and variational formula,

(3.11) 0=1λ˜0=supϕH1(Ω),ϕ0Ωβ1()+β2()u10()ϱ()δ()ϕ2dxΩd2ϕ2+y()ϕ2dx,

which indicate how ℜ0 depends on the diffusion coefficient d2 (see also in [1, Theorem 2] and [31]).

Theorem 3.1

Let0 be defined by (3.11). Then we have

(i) Fix d1 > 0, then0 is decreasing with respect to d2, and satisfies

limd200=maxβ1()+β2()u10()ϱ()δ()y():xΩˉandlimd20=Ωβ1()+β2()u10()ϱ()δ()dxΩy()dx.

(ii) Fix d2 > 0, then

limd100=0,0:=supϕH1(Ω),ϕ0Ω(β1()+β2()ϱ()b()δ()ζ())ϕ2dxΩd2ϕ2+y()ϕ2dx

and

limd10=0,:=supϕH1(Ω),ϕ0Ω(β1()+β2()ϱ()Ωb()dxδ()Ωζ()dx)ϕ2dxΩd2ϕ2+y()ϕ2dx.

(iii) If Ωβ1()+β2()u10()ϱ()δ()dx/Ωy()dx>1,wehave0>1foralld1,d2>0 .

(iv) If Ωβ1()+β2()u10()ϱ()δ()dx/Ωy()dx<1andβ1()+β2()u10()e()δ()/y()>1forsomexΩˉ , then there exists d2 ∈ (0,∞) such that0 > 1 when d2<d2, and0 < 1 when d2<d2.

The following statements come from [13, Theorem 1.1].

Remark 1

It is clear that the asymptotic behavior of the unique positive solution u10 with d1:

  • u10()b()/ζ()inC1(Ω)asd10;

  • u10()Ωb()dx/Ωζ()dxinC1(Ω)asd1.

Remark 2

Fix d2 > 0, η0 of the eigenvalue problem (3.8) satisfies

  • η0η00asd10;

  • η0η0asd1,

where η00andη0 are respectively 0the principal eigenvalue of (3.8) with u10() = b(·)/ζ (·), and the principal eigenvalue of

(3.12) d2Δϕ+β1()+β2()ϱ()Ωb(x)dxδ()Ωζ()dxy()ϕ=ηϕ,xΩ,ϕν=0,xΩ.

Remark 3

0,0 − 1 has the same sign as η00,and0,1 has the same sign as η0.

Theorem 3.2

If0 ≥ 1 (or s(𝓑) ≥ 0), s(𝓑) is the principal eigenvalue of (3.3).

Proof. In fact, we see from (3.1) that

u2(,t,ϕ)=A2(t)ϕ2+0tA2(ts)q(u2(,s,ϕ),u3(,s,ϕ))ds,u3(,t,ϕ)=A3(t)ϕ3+0tA3(t)(ts)(ϱ()u2(,s,ϕ))ds,

where q(u2,u3)=β1()u2+β2()u10()u3, and A1 and A2 are respectively defined in (2.1) and (2.2). We decompose T(t) as the sum of T2(t) and T3(t), T(t) = T2(t) + T3(t), where

(3.13) T2(t)ϕ=0,A3(t)ϕ3,ϕ=ϕ2,ϕ3Y,T3(t)ϕ=u2(,t,ϕ),0tA3(t)(ts)ϱ()u2(,s,ϕ)ds,ϕ=ϕ2,ϕ3Y.

By [31, Lemma 2.5], the compactness of T3(t) directly follows. Further,

supϕC(Ωˉ,R2),ϕ0T2(t)ϕϕsupϕC(Ωˉ,R2),ϕ0eδ()tϕ3ϕsupϕC(Ωˉ,R2),ϕ0eδtϕ3ϕeδt,

that is, ∥T2(t)∥ ≤ eδ* t.

Hence, for any bounded set 𝕊 in 𝕐,

κ(T(t)S)T2(t)κ(S)eδtκ(S),t>0.

Thus, T(t) is a κ-contraction on 𝕐, that is, the essential growth bound, ωess(Π(t)) ≤ −δ and the essential spectral radius

re(Π(t))eδt<1,t>0.

It is well-known that ω(T(t)), the exponential growth bound (defined by ω(T(t)):=limtlnT(t)t such that T(t)≤ Meω(T(t))t , for some M > 0), satisfies

ω(T(t))=max{s(B),ωess(T(t))}.

With the assumption that s(𝓑) ≥ 0, the spectral radius of T(t),

r(T(t))=es(B)t1,t>0.

Consequently, re(T(t)) < r(T(t)). With the help of the generalized Krein-Rutman Theorem (see, for example, [18]), we complete the proof. □

The following result indicates that for a special case, s(𝓑) is the principal eigenvalue of (3.3) without any limitations.

Lemma 3.3

Suppose that δ(x) ≡ δ. s(𝓑) is the principal eigenvalue of (3.3).

Proof. Let

Lλϕ=d2Δϕ+β1()+β2()u10()ϱ()λ+δϕy()ϕ,λ>δ,

Denote C1 := minxΩˉ {β1(·)} and C2:=minxΩˉ{ϱ()β2()u10()}. Note that the eigenvalue problem

(3.14) ηφ()=d2Δφ()y()φ(),xΩφ(x)ν=0,xΩ.

admits one principal eigenvalue, ηˆ, (associated with a positive eigenvector φ ⨠ 0). Denote by λ* the larger root of

λ2+(δC1ηˆ)λ(C2+δ(C1+ηˆ))=0.

Then λ=12[(ηˆδ+C1)+(δ+C1+ηˆ)2+4C2]>δ. Hence,

Lλφ=d2Δφ+β1()φ+ϱ()β2()u10()λ+δφy()φ(ηˆ+C1+C2λ+δ)φ=λφ.

According to [30, Theorem 2.3], the assertion directly follows. We complete the proof. □

4 Threshold dynamics

We now study the threshold dynamics of model (1.5) as ℜ0 ≤ 1 and ℜ0 > 1.

Theorem 4.1

If0 ≤ 1 (or s(𝓑) ≤ 0), then E0 is globally asymptotically stable.

Proof. We divide the proof into two parts. One is ℜ0 < 1, the other is ℜ0 = 1. When ℜ0 < 1, we know from [30, Theorem 3.1] that E0 is locally asymptotically stable. We are mow in a position to consider the global attractivity of E0. Fix ϵ0 > 0. By (2.6), there is t1 > 0 such that for all (x, t) ∈ Ω‾× [t1,∞),

0u1(,t)u10()+ϵ0.

Let (û2(·, t), û3(·, t)) is the solution of the following problem

(4.1) uˆ2tuˆ3t=Bϵ0uˆ2uˆ3,(x,t)Ωˉ×[t1,),uˆ2n=0,(x,t)Ωˉ×[t1,),

where

Bϵ0=d2Δ+β1()y()β2()(u10()+ϵ0)ϱ()δ().

By the comparison principle (see, e.g. [16]), (u2(,t),u3(,t))(uˆ2(,t),uˆ3(,t))onΩˉ×[t1,).

Let Tϵ0 (t) be the positive semigroup induced by 𝓑ϵ0 . We next aim to prove that the exponential growth bound of Tε0 (t), ωϵ0 is negative. Let ωess(Tε0(t)):=limtα(Tε0(t))t be the essential growth bound of Tε0 (t), where α(·) is the measure of non-compactness. From Lemma 2.2, we have ωess(Tϵ0) ≤ −δ*. Recall that ωϵ0 = max{s(𝓑ϵ0 ), ωess(Tε0 (t))}. This allow us to conclude that ωϵ0 has the same sign as s(𝓑ϵ0). From Lemma 3.2, we directly obtain that s(𝓑ϵ0) has the same sign to the principle eigenvalue ηϵ0 of

(4.2) d2Δφy()φ+β1()+β2()(u10()+ϵ0)ϱ()δ()φ=ηφ,xΩ,φν=0,xΩ.

It follows from (ii) of Lemma 3.2,ℜ0 < 1 and the continuous dependence of ηϵ00 on ϵ0 that ηϵ00<0 if ϵ0 is small. Hence, ωϵ0 < 0. This together with Tε0(t)Meωϵ0t, for some M > 0, imply that (uˆ2(,t),uˆ3(,t))(0,0) as t → ∞ uniformly for xΩˉ. Consequently, (u2(·, t), u3(·, t)) → (0, 0) as t → ∞ uniformly for x Ωˉ. Further, from (1.5), u1(,t)u10() as t → ∞uniformly for xΩ‾. This completes the first part.

When ℜ0 = 1, the global asymptotic stability result of E0 needs to combine with the local stability and global attractivity of E0 (see also in [3, 4, 31, 32]). We first confirm the local asymptotic stability of E0. Let ϵ > 0 be given. Choosing ρ > 0 such that for any u0 = (u10, u20, u30), we have ∥ u0E0 ∥≤ ρ.

Let us introduce

Σ(x,t):=u1(,t)u10()1andτ(t):=maxxΩˉ{Σ(,t),0}.

Hence, the first equation of (1.5) becomes

(4.3) Σtd1ΔΣ2d1u10()Σu10()+b()u10()Σ=β1()u1u2u1+u2+β2()u1u3u10(x)

Solving (4.3) gives

(4.4) Σ(,t)=Tˆ(t)(t)Σ00tTˆ(t)(ts)β1()u1u2u1+u2+β2()u1u3u10()ds,

where Σ0=u10u101 andTˆ(t)Mert for some r,𝕄 > 0, is the positive semigroup induced by d1Δ+2d1u10()u10()+ b()u10(). It follows that

(4.5) τ(t)=maxxΩˉTˆ(t)Σ00tTˆ(ts)β1()u1u2u1+u2+β2()u1u3u10(x)ds,0maxxΩˉTˆ(t)Σ0,0≤∥Tˆ(t)Σ0Mertu10u10()1ρMert/u_10,

where u_10=minxΩˉ{u10()}.

By a zero trick and the definition of T(t) (positive semigroup generated by 𝓑), we have

(4.6) u2(,t)u3(,t)=T(t)u20()u30()+0tT(ts)β1()u1(,s)u2(,s)u1(,s)+u2(,s)β1()u2(,s)+β2()(u1(,s)u10)u3(,s)0dsT(t)u20()u30()+0tT(ts)β2(u1(,s)u10())u3(,s)0ds.

Under the condition that ℜ0 = 1 (or s(𝓑) = 0), we know that ∥ T(t) ∥≤ Mτ for t ≥ 0 and Mτ > 0. Hence, by (4.6), we have

(4.7) max{u2(,t),u3(,t)}Mτρ+Mτβ2u10()0tτ(s)u3(,s)dsMτρ+Mτβ2u10()ρMu_100tersu3(,s)ds=Mτρ+Lρ0tersu3(,s)ds,

where L=Mτβ2u10()M/u_10. With the help of the Gronwall’s inequality, it is easy to get

(4.8) u3(x,t)MτρeLρ/r.

Further from (4.7) and (4.8), we know that u2(x,t)Mτρ+LρMτρeLρ/r/r. This together with the first equation of system (1.5) and (4.8) imply that

u1td1Δu1+b()ζ()u1β1u1β2MτρeLρ/ru1,(x,t)Ω×(0,).

Thus, we consider the system

(4.9) ϑt=d1Δϑ+b()ζ()ϑβ1ϑβ2MτρeLρ/rϑ,(x,t)Ω×(0,),ϑn=0,(x,t)Ω×(0,),ϑ(,0)=u10(),xΩ.

Obviously, u1(·, t) ≥ ϑ(·, t), for all (x,t)Ω×(0,).Letϑˆ(,t)=ϑ(,t)ϑρ()andϑρ() is the positive steady state of (4.9). Hence, ϑˆ(,t) satisfies

(4.10) ϑˆt=d1Δϑˆ(ζ()+β1+β2MτρeLρ/r)ϑˆ,(x,t)Ω×(0,),ϑˆn=0,(x,t)Ω×(0,),ϑˆ(,0)=u10()ϑρ(),xΩ.

Solving (4.10) yields

ϑˆ(,t)=A1(t)(u10()ϑρ())0tA1(ts)(β1+β2MτρeLρ/r)ϑˆ(,s)ds.

where A1(t) is defined in (2.1) and satisfies ∥A1(t)∥ ≤ K0eζ*t for some constant K0. Hence,

ϑˆ(,t)K0eζtu10()u1ρ()+0tK0eζ(ts)(β1+β2MτρeLρ/r)ϑˆ(,s)ds.

Let L1=K0(β1+β2MτρeLρ/r). This together with the Gronwall’s inequality imply that

ϑˆ(,t)=ϑ(,t)ϑρ()K0u10()ϑρ()eL1t+ζt.

Choose ρ > 0 small enough that L1<ζ/2 and

ϑ(,t)ϑρ()K0u10()ϑρ()eζmt/2.

By a zero trick,

(4.11) u1(,t)u10()ϑ(,t)u10()=ϑˆ(,t)ϑρ()+ϑρ()u10()K0u10()ϑρ()eζmt/2+ϑρ()u10()K0(u10()u10()+u10()ϑρ())ϑρ()u10()K0ρ(K0+1)ϑρ()u10().

From τ(t)ρMert/u_10, we directly have

(4.12) u1(,t)u10()=u10()u1(,t)u10()1ρMu10()/u_10.

Combined with (4.11) and (4.12), one can get

(4.13) u1(,t)u10()max{K0ρ+(K0+1)ϑρ()u10(),ρMu10()/u_10}.

By choosing ρ sufficiently small, limρ0ϑρ()=u10(), the inequalities (4.11) and (4.13) imply that

u1(,t)u10(),u2(,t),u3(,t)ε,forallt>0,

where ε can be chosen as a small number. This proves the local stability of E0.

Secondly, we confirm the global stability of E0. For any u0 ∈ 𝕏+, we set

X1={(u1,u2,u3)X+:u2=u3=0},

and

σ(t;u0):=inf{c˜R:u2(,t)c˜ϕ2andu3(,t)c˜ϕ3}.

where (ϕ2, ϕ3) is the positive eigenfunction of s(𝓑) = 0. Recall that (1.5) has a connected global compact attractor in 𝕏+, denoted by 𝒜0 (see Theorem 2.2).

We claim that for any u0 ∈ 𝒜0, ω(u0) ∈ 𝕏1. It is clear that for all t > 0, σ(t; u0) > 0 is strictly decreasing. In fact, for t˜0>0, let

u2(,t)=σ(t˜0;u0)ϕ2andu3(,t)=σ(t˜0;u0)ϕ3,fort>t˜0.

such that

(u2(,t),u3(,t))(u2(,t),u3(,t)),for(x,t)Ωˉ×[t˜0,).

where (u2(x,t),u3(x,t)) satisfies

(4.14) u2t=d2Δu2+β1()u1u2u1+u2+β2()u1u3γ()u2,(x,t)Ω×[t˜0,),u3t=ϱ(x)u2δ(x)u3,(x,t)Ω×[t˜0,),u2n=0,(x,t)Ω×[t˜0,),u2(,t˜0)u2(,t˜0),u3(,t˜0)u3(,t˜0),xΩ.

Furthermore, by (4.14) and comparison principle, one can see that

σ(t;u0)ϕ2=u2(,t)>u2(,t)andσ(t;u0)ϕ3=u3(,t)>u3(,t),for(x,t)Ωˉ×[t˜0,).

Since 0 is arbitrary, σ(t; u0) is strictly decreasing.

Next, we pay attention to the the semiflow Υ(t), generated by (1.5) (see Theorem 2.1). Let {t˜k} be a sequence with t˜k,σ=limtσ(t;u0) and v = (v1, v2, v3) ∈ ω(u0). We can conclude that

limt˜kY(t+t˜k)u0=Y(t)limt˜kY(t˜k)u0=Y(t)v.

This means that σ(t; v) = σ*. Repeating the previous arguments, σ(t; v) is strictly decreasing when v2 ≠0 or v3 ≠0. This gives a contradiction. So, we have v2 = v3 = 0. When t → ∞, we get (u2(·, t), u3(·, t)) → (0, 0). It follows that u1(,t)u10() eventually. This proves the second part. The proof is complete. □

The following result indicates that (1.5) is uniformly persistent when ℜ0 > 1. Before going into details, we set

X0:=ϕ()=ϕ1,ϕ2,ϕ3()X+:ϕ2()0 and ϕ3()0,X0:=X+X0=ϕ()=ϕ1,ϕ2,ϕ3()X+:ϕ2()0 or ϕ3()0.

and

M:={ϕX0:Y(t)ϕX0,t0},

where Y(t) is the semiflow generated by (1.5).

Theorem 4.2

If0 > 1, then (1.5) is uniformly persistent, i.e., for any ϕ ∈ 𝕏+ with ϕ2(·)/ ≡ 0, or ϕ3(·)/ ≡ 0, there exists σ > 0,

(4.15) lim inftui(,t)σ0,uniformlyforxΩˉ.

Moreover, (1.5) possesses at least one positive steady state.

Proof. Theorem 4.2 is achieved by the following claims. The first two claims are the direct consequences as those in [31, Theorem 3.10] with slight modifications. We present them here without proof.

Claim 1 Υ(t)𝕏0 ⊆ 𝕏0 for all t ≥ 0.

Claim 2 ω(u0) = {E0}, for any u0M, where ω(u0) is the ω limit set of u0.

Claim 3 lim supt→Υ(t)ϕE0∥ ≥ σ, ∀ϕ ∈ 𝕏0, where σ is a positive constant.

We prove Claim 3 by the way of contradiction. Assume to the contrary that for σ > 0, there exists 1 > 0 such that

u1(,t,ϕ)u10σ,u2(,t;ϕ)<σ,u3(,t;ϕ)<σ,tt1.

Hence (u2(·, t; ϕ), u3(·, t; ϕ)) is the upper solution of

(4.16) u˜2tu˜3t=Bσu˜2u˜3,(x,t)Ω×[tˉ1,),u˜2n=0,(x,t)Ω×[tˉ1,),

where

Bσ=d2Δ+β1(x)y(x)β2(x)(u10(x)σ)ϱ(x)δ(x).

Combined with Lemma 3.2 and the continuous dependence of s(𝓑) ≥ 0 on σ, choosing σ > 0 small enough that s(𝓑σ) > 0, associated with positive eigenvalue function (ϕ2σ(),ϕ3σ()). Choose ξ > 0 small enough that

(u2(,tˉ1;ϕ),u3(,tˉ1;ϕ)ξ(ϕ2σ(),ϕ3σ()).

It is easy to see that for (u˜2(,tˉ1),u˜3(,tˉ1)=ξ(ϕ2σ(),ϕ3σ()), (4.16) has a solution

(u˜2(,t;ϕ),u˜3(,t;ϕ)=ξes(Bσ)(ttˉ1)(ϕ2σ(),ϕ3σ()),t>tˉ1.

Since s(𝓑σ) > 0, u2(·, t; ϕ), u3(·, t; ϕ) → ∞as t → ∞, a contradiction.

Similar to [25, Theorem 3] (see also in [31, Theorem 3.10]), let us define a distance function ρ(·) : 𝕏+ → [0,∞) by

ρ(ϕ)=min{minxΩˉϕ2(),minxΩˉϕ3()},ϕX+.

we conclude that

lim inftu2(,t,ϕ)σ1andlim inftu3(,t,ϕ)σ1,ϕX0.

Recall from Theorem 2.1 that there exist M > 0 and t2 > 0 such that u2(·, t, ϕ), u3(·, t, ϕ) ≤ M, (x, t) ∈ Ω × [t2,∞). It then follows that

u1td1u1+b(ζ+β1M+β2M),(x,t)Ω×[t2,)u˜1n=0,(x,t)Ω×[t2,),

Thus, there exists a constant σ2, lim inft→ u1(·, t, ϕ) ≥ σ2. Let σ0 = min{σ1, σ2}. The uniform persistence is obtained. Further from [15, Theorem 4.7], (1.5) possesses at least one positive steady state. This completes the proof. □

5 Spatially homogeneous case

When

b(x)b,βi(x)βi,ζ(x)ζ,y(x)y,δ(x)δ,ϱ(x)ϱ,

system (1.5) becomes

(5.1) u1t=d1Δu1+bβ1u1u2u1+u2β2u1u3ζu1,(x,t)Ω×(0,),u2t=d2Δu2+β1u1u2u1+u2+β2u1u3yu2,(x,t)Ω×(0,),u3t=ϱu2δu3,(x,t)Ω×(0,),

subject to the same initial and boundary condition as in (1.5). Clearly, system (5.1) possesses a disease-free equilibrium E0=(u10,0,0)withu10:=b/ζ. From (3.4), we have the concrete formula of ℜ0,

(5.2) [0]=β1δ+β2ρu10yδ.

In this circumstance, Theorems 4.1 and 4.2 still hold for (5.1).

Theorem 5.1

If [ℜ0] < 1, then E0 is globally asymptotically stable; while if [ℜ0] > 1, then (5.1) is uniformly persistent.

The positive equilibrium of (5.1) (whenever it exists) should satisfy

u1=byδ/bϱβ2+β1yδ,y=ζ;bϱβ2+β1yδ+yδζy2δ+bϱβ2+β1yδ+yδζy2δ2+4byδϱβ2(yζ)2ϱβ2(yζ),y>ζ,u2=bζu1y and u3=bζu1ϱyδ.

To make u2>0andu3>0, we need u1<b/ζ, which is equivalent with [ℜ0] > 1. The equivalence is obvious when y = ζ. When y > ζ, we rewrite u1 < b/ζ as

B+B2+4AC2A<bζ,

where 𝔵 = ϱβ2(yζ) > 0, 𝔹 = bϱβ2 + β1 + yδζy2δ and ℂ = byδ > 0. Isolating the square root and squaring both sides gives 𝔹 + 2𝔵b/ζ > 0 and

B2+4AC<B2+4BAbζ+4A2b2ζ2.

Simplifying the last inequality yields ℂζ 2 < 𝔹 +𝔵b2. We make use of the expressions of 𝔵, 𝔹, ℂ and rewrite the inequality as

byδζ2<bζ(bϱβ2+β1yδ+yδζy2δ)+b2ϱβ2(yζ).

By expanding and canceling, we obtain yδζ < bϱβ2 + β1δζ, which is the same as [ℜ0] > 1. Furthermore, coupling [ℜ0] > 1 and y > ζ implies 𝔹 + 2𝔵b/ζ > 0. Thus, if [ℜ0] > 1 and yζ, (5.1) possesses a unique positive equilibrium.

Remark 4

Note that y and ζ are the death rates of infected and susceptible host, respectively. Due to the disease burden, it is realistic to assume that yζ. Combined with the analysis above, (5.1) possesses a unique positive equilibrium.

Theorem 5.2

Suppose that yζ hold. If [ℜ0] > 1, then unique positive equilibrium E(5.1)=(u1,u2,u3) is locally asymptotically stable.

Proof. Linearizing system (5.1) at E(5.1) gives

χt=DΔχ(,t)+Qχ(,t),

where χ = (u1, u2, u3), D = diag(d1, d2, 0) and

Q=ζβ1u22u1+u22β2u3β1u12u1+u22β2u1β1u22u1+u22+β2u3β1u12u1+u22yβ2u10ϱδ.

Hence, the characteristic equation of E(5.1) is

(5.3) λ3+a1(l2)λ2+a2(l2)λ+a3(l2)=0,

where l is the wavenumber,

a1(l2):=l2d1+ζ+β2u3+β1u22(u1+u2)2+l2d2+β1u1u22+β2u1u3(u1+u2)2u2(u1+u2)2+δ,a2(l2):=l2d2+β1u1u22(u1+u2)2δ+β2u3+β1u22(u1+u2)2β1u12(u1+u2)2+l2d1+ζ+β2u3+β1u22(u1+u2)2l2d2+β1u1u22+β2u1u3(u1+u2)2u2(u1+u2)2+δ,

and

a3(l2):=(l2d1+ζ)l2d2+β1u1u22(u1+u2)2δ+β1u12(u1+u2)2β1u22(u1+u2)2+β2u3δ+l2d2+β1u1u22+β2u1u3(u1+u2)2u2(u1+u2)2β1u22(u1+u2)2+β2u3δ.

It then follows that a1(l2) > 0, a2(l2) > 0 and a3(l2) > 0. Further, after elementary calculations, a1(l2)a2(l2)− a3(l2) > 0. Hence, the local stability of E(5.1) directly follows from the Routh-Hurwitz criterion. □

Theorem 5.3

Assume that (H1):β1u2u124ζu1u1+u2.If0>1, then E(5.1) is globally asymptotically stable.

Proof. Define

V(t):=Ω(u1u1lnu1+u2u2lnu2+β2u1u3ρu2(u3u3lnu3))dx,

Since

β1u1u2u1+u2+β2u1u3+ζu1=b,β1u1u2u1+u2+β2u1u3=yu2andρu2=δu3,

after elementary calculations, we can obtain

dV(t)dt=d1Ω1u1u1Δu1dx+d2Ω1u2u2Δu2dx+Ωζ(u1u1)2u1β1u2(u1u1)2u1(u1+u2)+β1(u1u1)[u2(u1u1)u1(u2u2)](u1+u2)(u1+u2)dx+Ωβ1(u2u2)[u2r(u1u1)u1(u2u2)](u1+u2)(u1+u2)dx+β2u1u3Ω3u1u1u2u1u3u2u1u3u2u3u2u3dx.

By using u1u1 + u2, we have

dV(t)dtΩβ1(u1+u2)(u1+u2)ζ(u1+u2)β1(u1u1)2(u2u1)(u1u1)(u2u2)+u1(u2u2)2dx+β2u1u3Ω3u1u1u2u1u3u2u1u3u2u3u2u3dx.

By assumption (H1), we have

ζ(u1+u2)β1(u1u1)u1(u2u2)2ζ(u1+u2)β1(u1u1)2(u2u1)(u1u1)(u2u2)+u1(u2u2)2.

Hence,

dV(t)dtΩβ1(u1+u2)(u1+u2)ζ(u1+u2)β1(u1u1)u1(u2u2)2dx0.

Hence, the largest invariant subset {(u1,u2,u3)X+:dV(t)dt=0} consists just one singleton {E(5.1)}. According to the LaSalle’s invariance principle, the global stability of E(5.1) is confirmed. This completes the proof. □

Remark 5

In Theorem 5.3, (H1) is a technical condition such that the derivation of Lyapunov function is less than zero. It is significant to establish the global asymptotic stability of endemic equilibrium by constructing a suitable Lyapunov function in epidemiology. Our model (1.5) includes two types of infection: frequency dependent and mass action, which leads to challenging issue in proving global asymptotic stability of endemic equilibrium by Lyapunov function. Similar arguments can be found in (C1)-(C3) in the proof of Theorem 2.3 of [24].

6 Asymptotic profiles of the positive steady state

Theorem 4.2 indicates that when ℜ0 > 1, (6.1) admits at least one positive steady state E* , which is the positive solution of

(6.1) d1Δu1+b(x)β1(x)u1u2u1+u2β2(x)u1u3ζ(x)u1=0,xΩ,d2Δu2+β1(x)u1u2u1+u2+β2(x)u1u3y(x)u2=0,xΩ,ϱ(x)u2δ(x)u3=0,xΩ,u1ν=u2ν=0,xΩ.

This section is devoted to the investigation of the asymptotic profiles of E* for the cases that d1 → 0, d1 → ∞, d2 → 0 and d2 → ∞. As exploration of such problem can achieve better understanding the spatial distribution of disease.

We begin with the preliminary estimates of solutions of (6.1).

Lemma 6.1

Let (u1, u2, u3) be any positive solution of (6.1). We then directly have

(i) For any d1, d2 > 0, we have the upper bound of u1:

(6.2) u1(x)bζ,xΩ;

(ii) Fix d2 > 0, for any d1 > 0, we have the upper bound of u2 and lower bound of u1:

(6.3) u2(x)C1,u1(x)C2,xΩ,

where the positive constants Ci (i = 1,2) do not depend on d1 > 0.

Proof. (i) Set u1(x0)=maxxΩu1(x). By applying the maximum principle (see, e.g. [14, Proposition 2.2]) to u1-equation of (6.1), we get

b(x0)β1(x0)u1(x0)u2(x0)u1(x0)+u2(x0)β2(x0)ϱ(x0)δ(x0)u1(x0)u2(x0)ζ(x0)u1(x0)0.

Thus, we know that

maxxΩu1()=u1(x0)bζ,

this gives the upper bound of u1.

(ii) Making the sum of u1, u2 of (6.1) and integrating over Ω lead to

(6.4) yΩu2dxΩζu1dx+Ωyu2dx=Ωbdxb|Ω|.

Notice that u2 solves

(6.5) Δu2+[β1()u1d2(u1+u2)+β2()ϱ()d2δ()u1y()d2]u2=0,xΩ,u2ν=0,xΩ.

Fix d2 > 0, it follows from statement (i) that

β1()u1(u1+u2)d2+β2()ϱ()d2δ()u1y()d2β1d2+β2ϱbd2δζ+yd2.

By using [20, Lemma 2.2] (see also [12]), we get a Harnack-type inequality of u2 as follows

(6.6) maxΩu2CminΩu2.

where C > 0 does not depend on d1. In what follows, we permit it changing from place to place.

Based on (6.4) and (6.6), we have

(6.7) u2()CminΩu2CΩΩu2dxC,forxΩ.

Set u1(x1)=minxΩu1(x). By applying the maximum principle, we have

b(x1)β1(x1)u1(x1)u2(x1)u1(x1)+u2(x1)β2(x1)ϱ(x1)δ(x1)u1(x1)u2(x1)ζ(x1)u1(x1)0.

From (6.7), we have

bβ1+β2ϱCδ+ζu1(x1).

Hence, we get the lower bound of u1. □

Remark 6

Note that the statement (ii) hold for any d1 > 0and d2 ≥ 1. Actually, we can get (6.6) for any d2 ≥ 1.

6.1 Profile as d1 → 0

In view of Theorems 3.1 and 4.2, (6.1) possesses at least a positive solution for sufficiently small d1 when ℜ0,0 > 1. This subsection is spent on exploring the asymptotic profile of the positive solution of (6.1) with d1 → 0.

Theorem 6.1

Fix d2 > 0, and let d1 → 0. If0,0 > 1, then up to a subsequence of d1, every positive solution (u1(·; d1), u2(·; d1), u3(·; d1)) of (6.1) satisfies

(u1(;d1),u2(;d1),u3(;d1))(Φu1,Φu2,ρδΦu2)uniformlyforxΩ,

where

Φu1()=G(x,Φu2()):=bδβ1δ+β2ϱΦu2+ζδΦu2+bδβ1δ+β2ϱΦu2+ζδΦu22+4bδΦu2β2ϱΦu2+ζδ2β2ϱΦu2+ζδ,

and Φu2 is a positive solution to

(6.8) d2ΔΦu2+β1(x)G(x,Φu2)Φu2G(x,Φu2)+Φu2+β2(x)ϱ(x)δ(x)G(x,Φu2)Φu2y(x)Φu2=0,xΩ,Φu2ν=0,xΩ.

Proof. We deal with the proof by two steps.

Step 1. Convergence of u2. Observe that u2 satisfies

(6.9) d2Δu2+y()u2=β1()u1u2u1+u2+β2()ϱ(x)δ()u1u2,xΩ,u2ν=0,xΩ.

From the estimates in Lemmma 6.1, we know that ∀ p > 1,

β1(x)u1u2u1+u2+β2(x)ϱ(x)δ(x)u1u2Lp(Ω)C.

Combined with the standard Lp-estimate for elliptic equations and Sobolev embedding theorem (see, e.g. [7]),

u2C1+α(Ω)C,forsome0<α<1.

From the third equation of (6.1), it is clear that

u3C1+α(Ω)C,forsome0<α<1.

Denoted by d1n the subsequence of d1 → 0, satisfying d1n → 0 as n → ∞. With this subsequence, (u1n , u2n , u3n) := (u1(x; d1n), u2(x; d1n), u3(x; d1n)) of (6.1) with d1 = d1n satisfies

(6.10) (u2n,u3n)(Φu2,ρδΦu2)uniformlyforxΩandn,

where Φu2C 1(Ω) and Φu2 ≥ 0. Taking into account of Harnack-type inequality of u2 (6.6), it is clear that

(6.11) eitherΦu20onΩorΦu2>0onΩ.

We next prove that Φu2 > 0 on Ω. Assume to the contrary that Φu20onΩ, i.e., as n → ∞,

(6.12) u2n0uniformlyonΩ.

Choose 0<ϵ<b/β1 small enough that

0u2n()ϵonΩ,for all largen.

Hence, u1n satisfies

d1nΔu1nb()ζ()u1n,xΩ,

and

d1nΔu1nb()β1ϵβρδϵu1nζ()u1n,xΩ,

respectively, with u1nν=0, x∂Ω. Fix sufficiently large n, denoted by Un and Vn respectively the unique positive solution of the following two auxiliary systems:

(6.13) d1nΔU=b()ζ()U,xΩ;Uν=0,xΩ,

and

(6.14) d1nΔV=b()β1ϵβρδϵVζ()V,xΩ;Vν=0,xΩ.

It follows from the super-subsolution argument that

(6.15) Vnu1nUnonΩ.

By applying the singular perturbation theory technique (see, e.g. [5, Lemma 2.4] or [13, Lemma 2.1]), we know that as n → ∞,

Un(x)b()ζ()andVn()b()β1ϵβρδϵ+ζ()uniformlyonΩ.

Hence,

(6.16) b()β1ϵβρδϵ+ζ()lim infnu1n()lim supnu1n()b()ζ()onΩ.

Since ϵ is arbitrary, (6.16) further implies that

(6.17) u1n(x)b()ζ()uniformlyonΩ,asn.

We now pay attention to u2 equation of (6.1) with u2n:

(6.18) d2Δu2n=β1()u1nu2nu1n+u2n+β2()ϱ()δ()u1nu2ny()u2n,xΩ;u2nν=0,xΩ.

For all n ≥ 1, define u˜2n:=u2nu2nL(Ω)thatu˜2nL(Ω)=1. In this setting, u˜2n satisfies

(6.19) d2Δu˜2n=β1()u1nu1n+u2n+β2()ϱ()δ()u1ny()u˜2n,xΩ;u˜2nν=0,xΩ.

As above, with the aid of standard compactness argument, as n → ∞,

u˜2nu˜2inC1(Ω),

where u˜20 belongs to C1(Ω), and satisfies u˜2L(Ω)=1. From (6.12) and (6.17), by sending n → ∞in (6.19), it is clear that u˜2 satisfies

d2Δu˜2=β1()y()+β2()ϱ()b()δ()ζ()u˜2,xΩ;u˜2ν=0,xΩ.

From [20, Lemma 2.2] (see also [12]), it follows that the Harnack-type inequality of u˜2 remains true and then ũ2 > 0 on Ω. Therefore, 0 is the principal eigenvalue of (3.8) with u10(x)=b(x)/ζ(x). This contradicts with our assumption that ℜ0,0 > 1. Thus, Φu2 > 0 on Ω , which means that

(6.20) u2nΦu2>0uniformlyforxΩ.

Step 2. Convergence of u1. From (6.1), we know that u1n solves:

d1nΔu1n=b()β1()u1nu2nu1n+u2nβ2()ρ()δ()u1nu2nζ()u1n,xΩ,u1nν=0,xΩ.

By (6.20), choose ϵ > 0 small enough that 0 < Φu2ϵu2nΦu2 + ϵ on Ω, for all large n. Then,

b()β1()u1nu2nu1n+u2nβ2()ρ()δ()u1nu2nζ()u1nb()β1()u1n(Φu2ϵ)u1n+(Φu2ϵ)β2()ϱ()δ()u1n(Φu2ϵ)ζ()u1n=(h+ϵ(x,Φu2)u1n)(u1nhϵ(x,Φu2))u1n+(Φu2ϵ),

where

h±ϵ(x,Φu2)=b()(β1()+β2()ϱ()δ()(u2ϵ)+ζ())(u2ϵ)±D12[β2()ϱ()δ()(u2ϵ)+ζ()]andD1=[b()(β1()+β2()ϱ()δ()(u2ϵ)+ζ())(u2ϵ)]2+4b(u2ϵ)(β2()ϱ()δ()(u2ϵ)+ζ()).Obviously,h+ϵ(x,Φu2)>0andhϵ(x,Φu2)<0onΩ¯.

Consider the following problem

(6.21) dnΔz=(h+ϵ(x,Φu2)z)(zhϵ(x,Φu2))(z+(Φu2ϵ)),xΩ,zν=0,xΩ.

Then, for sufficiently large C > 0, (u1n , C) is a pair of sub-supersolution of (6.21) on Ω. Hence, (6.21) admits at least one positive solution, zn, and zn satisfies u1nznConΩ. Further by the singular perturbation theory technique,

znh+ϵ(x,Φu2(x))uniformlyonΩ,asn.

Hence,

(6.22) lim supnu1n(x)h+ϵ(x,Φu2(x))uniformlyonΩ.

On the other hand,

b()β1()u1nu2nu1n+u2nβ2()ϱ()δ()u1nu2nζ()u1nb()β1()u1n(Φu2+ϵ)u1n+(Φu2+ϵ)β2()ϱ()δ()u1n(Φu2+ϵ)ζ()u1n=(hϵ+(x,Φu2)u1n)(u1nhϵ(x,Φu2))u1n+(Φu2+ϵ),

with

hϵ±(x,Φu2)=b()(β1()+β2()ϱ()δ()(Φu2+ϵ)+ζ())(Φu2+ϵ)±D22[β2()ϱ()δ(Φu2+ϵ)+ζ()],

where D2=[b()(β1()+β2()ϱ()δ()(Φu2+ϵ)+ζ())(Φu2+ϵ)]2+4b(Φu2+ϵ)(β2()ϱ()δ()(Φu2+ϵ)+ζ()). By a similar argument as before, we further have

(6.23) lim infnu1n(x)hϵ+(x,Φu2(x))uniformlyonΩ.

Since

limϵ0h+ϵ(x,Φu2)=G(x,Φu2(x))=limϵ0hϵ+(x,Φu2),

then as n → ∞,

u1n(x)G(x,Φu2(x))uniformlyonΩ.

As a result, by (6.18), Φu2 satisfies (6.9). This proves Theorem 6.1. □

6.2 Profile as d2 → 0

This subsection is spent on exploring the asymptotic profile of the positive solution of (6.1) with d2 → 0. In this case, because of mathematical difficulty, we only consider the limiting profile of positive solutions in 1-D space. Without loss of generality, we take Ω = (0, 1).

Theorem 6.2

Assume that {x[0,1]:δ()β1()+β2()ϱ()u10()>δ()y()} is non-empty. Fix d1 > 0, and let d2 → 0, then every positive solution (u1(·; d2), u2(·; d2), u3(·; d2)) of (6.1) satisfies

(u1(;d2),u2(;d2),u3(;d2))(Ψu1,Ψu2,ρδΨu2)uniformlyon[0,1],

where

(6.24) Ψu2(x):=(β1()δ()+β2()ρ()Ψu1()y()δ())Ψu1()y()δ()β2()ρ()Ψu1()+,

and Ψu1 solves

d1Ψu1′′+b()β1()Ψu1Ψu2Ψu1+Ψu2β2()ρ()δ()Ψu1Ψu2ζ()Ψu1=0,x(0,1),Ψu1(0)=Ψu1(1)=0.

Proof. From the estimats (6.2) and (6.4), together with the elliptic L1-estimate theory (see, e.g. [2] or [12, Lemma 2.2]),

u1W1,q((0,1))C,q>0,

Choose p large enough, combined with the Sobolev embedding theorem, that

u1Cα([0,1])Cforsome0<α<1.

Denoted by d2k the subsequence of d2 → 0, satisfying d2k → 0 as k → ∞. With this subsequence, (u1k , u2k , u3k) := (u1(·; d2k), u2(·; d2k), u3(·; d2k)) of (6.1) with d2 = d2k satisfies u1kΨu1 in C([0, 1]) as k → ∞.

We note that u2k satisfies

(6.25) d2ku2k′′+β1()u1ku1k+u2k+β2()ϱ()δ()u1ky()u2k=0,x(0,1),u2k(0)=u2k(1)=0.

By a similar super-subsolution argument (see Theorem 6.1), we get

u2kΨu2uniformly on [0,1]ask,

where Ψu2 is given by (6.24). Moreover, from the expression of Ψu2, it is clear that Ψu1 solves (6.2). The proof is complete.

6.3 Profile as d1 → ∞

In view of Theorems 3.1 and 4.2, (6.1) possesses at least a positive solution for sufficiently small d1 when ℜ0,∞ > 1. This subsection is spent on exploring the asymptotic profile of the positive solution of (6.1) with d1 → ∞.

Theorem 6.3

Assume that0,∞ > 1. Fix d2 > 0, and let d1 → ∞, then every positive solution (u1(·; d1), u2(·; d1), u3(·; d1)) of (6.1), up to a subsequence of d1, satisfies

(u1(;d1),u2(;d1),u3(;d1))(u1,u2,ρδu2)uniformlyonΩ,

where (u1,u2) solves

(6.26) Ω[b()ζ()u1y()u2]dx=0,xΩ,d2Δu2=β1()u1u2u1+u2+β2()ϱ()δ()u1u2y()u2,xΩ,u2ν=0,xΩ,

and u1 is a positive constant and u2>0onΩ.

Proof. Rewriting u1 equation of (6.1) as

(6.27) Δu1=1d1b()β1()u1u2u1+u2β2()ϱ()δ()u1u2ζ()u1,xΩ,u1ν=0,xΩ.

From the estimates of u1 and u2 in Lemma 6.1, we know that

1d1b()β1()u1u2u1+u2β2()ϱ()δ()u1u2ζ()u1Lp(Ω)C,p1,

where C > 0 does not dependent on d1 ≥ 1. Combined with the standard Lp-estimate for elliptic equations and Sobolev embedding theorem,

u1C1+α(Ω)Cforsome0<α<1.

Denoted by d1i the subsequence of d1 → ∞, satisfying d1i → 0 as i → ∞. With this subsequence, (u1i, u2i, u3i) := (u1(·; d1i), u2(·; d1i), u3(·; d1i)) of (6.1) with d1 = d1i satisfies u1iu1 in C1(Ω) as i → ∞, where u1 > 0 on Ω due to the estimate of u1. Moreover, u1∞ solves Δu1 = 0, xΩ with homogeneous Neumann boundary condition implies that u1 > 0 is a positive constant.

Similar arguments as before, as i → ∞,

(u2i,u3i)(u2,ρδu2)inC1(Ω),

where u2 C1(Ω) and u2 ≥ 0 onΩ. Since the Harnack-type inequality of u2, then either u2 > 0 on Ω or u2 0 on Ω. Similar to the arguments in Section 6.1, we suppose that u2 ≡ 0 on Ω. Define ũ2i := u2i /∥u2iL∞(Ω) that ∥ũ2iL∞(Ω) = 1 for all i ≥ 1, and ũ2i satisfies

d2Δu˜2i+β1()u1iu1i+u2i+β2()ϱ()δ()u1iy()u˜2i=0,xΩ,u˜2iν=0,xΩ.

As a result, ũ2iuˆ2 in C1(Ω) as i → ∞, where uˆ2 ≥ 0 on Ω , uˆ2L(Ω) = 1. As u2i → 0 as i → ∞, it follows from a super-subsolution argument that u1i → ∫Ω b(x) dx/ ∫Ω ζ (x) dx. Thus, uˆ2 solves

d2Δuˆ2+β1()+β2()ϱ()Ωb()dxδ()Ωζ()dxy()uˆ2=0,xΩ,uˆ2ν=0,xΩ.

It can be seen that the Harnack-type inequality of uˆ2 also remains true and then uˆ2 > 0 on Ω. Consequently, 0 is the principal eigenvalue of (3.12), a contradiction with ℜ0,∞ > 1. Hence, u2 > 0 satisfies (6.26). This proves Theorem 6.3. □

6.4 Profile as d2

This subsection is spent on exploring the asymptotic profile of the positive solution of (6.1) with d2 → ∞.

Theorem 6.4

Assume Ωβ1()+β2()u10()ϱ()δ()dx/Ωy()dx>1.Fixd1>0, and let d2 → ∞, then

every positive solution (u1(·; d2), u2(·; d2), u3(·; d2)) of (6.1), up to a subsequence of d2, satisfies

(u1(;d2),u2(;d2),u3(;d2))(u1,u2,ρδu2)uniformlyonΩ,

where (u1∞, u2∞) solves

(6.28) d1Δu1=b()β1()u1u2u1+u2β2()ϱ()δ()u1u2ζ()u1,xΩ,Ω[b()ζ()u1y()u2]dx=0,xΩ,u1ν=0,xΩ,

and u2∞ is a positive constant and u1∞ > 0 on Ω.

Proof. From the discussion in Lemma 6.1 and Remark 6, it is clear that the estimats (6.6) and (6.7) hold for all large d2, and C > 0 does not depend on d1 and d2 ≥ 1.

Denoted by d2j the subsequence of d2 → ∞, satisfying d2j → 0 as j → ∞. With this subsequence, combined with the standard Lp-estimate for elliptic equations and Sobolev embedding theorem, (u1j , u2j, u3j) := (u1(·; d2j), u2(·; d2j), u3(·; d2j)) of (6.1) with d2 = d2j satisfies

(u1j,u2j,u3j)(u1,u2,ρ()δ()u2)inC1(Ω),asj,

where u1∞ > 0 on Ω and u2∞ is a nonnegative constant.

We next show that u2∞ is a positive constant. By a similar argument as before, suppose that u2∞ ≡ 0. For all j ≥ 1, define ũ2j := u2j/∥u2jL∞(Ω) that ∥ũ2jL∞(Ω) = 1, and ũ2j satisfies

(6.29) d2jΔu˜2j+β1()u1ju1j+u2j+β2()ϱ()δ()u1jy()u˜2j=0,xΩ,u˜2jν=0,xΩ.

Thus, if necessary, by passing to a further subsequence of d2j, one can get

(6.30) u˜2j1inC1(Ω),asj.

On the other hand, since u2j → 0 as j → ∞, it follows from a super-subsolution argument that u1ju10 as j → ∞. Then, integrating (6.29) and sending j → ∞, we have

Ωβ1()+β2()ϱ()u10()δ()y()dx=0,

a contradiction. Therefore, u2∞ is a positive constant, and (u1∞, u2∞) satisfies (6.28). The proof is complete. □

7 Conclusion and Discussion

In current work, we explore the global dynamics and asymptotic profiles of a host-pathogen model in a spatially bounded domain. We formulate the model by a reaction-diffusion system with space dependent parameters, where host individuals disperse at distinct rates and the mobility of pathogen is ignored. Complete analysis of the system allow us to investigate how large or small flows of hosts affect the spatial spread of disease, and what is the role of spatial heterogeneity on disease transmission.

We firstly establish that the existence of global solution, which is achieved by extending the local solution to a global one (see Lemma 2.1 and Theorem 2.1). To cope with the non-compactness of solution semiflow y(t), we utilize the Kuratowski measure of noncompactness to verify the asymptotic smoothness condition. Hence, by [8, Theorem 2.4.6], (1.5) possesses a global compact attractor in 𝕏+, denoted by 𝒜0 (see Theorem 2.2).

The BRN, ℜ0, is identified as the spectral radius of NGO, and also characterized by some equivalent principal spectral conditions, which establishes the threshold dynamical result for pathogen extinction and persistence (see Theorem 4.1 and 4.2). Specifically, we demonstrate that howℜ0 depends on the diffusion coefficient d1 for d1 → 0 and d1 → ∞ (see Theorem 3.1). We have also confirmed the global stability of E0 in a critical case that ℜ0 = 1. It should be pointed here that the method used in Theorem 4.1 can also be applied in Avian influenza dynamical model [28] and Ebola transmission model [35], where the dynamic behaviors for the case that ℜ0 = 1 are still open. We left it as future investigation. In a homogeneous case and additional condition, we explore the global attractivity of PSS (positive equilibrium) by the technique of Lyapunov function.

When ℜ0 > 1, (1.5) possesses at lease one positive steady state. To achieve better understanding the effects of the host’s movements on the spatial distribution of pathogen, we explore the asymptotic profiles of positive steady state for the cases that d1 → 0, d1 → ∞, d2 → 0 and d2 → ∞. When d1 → 0, our result (Theorem 6.1) demonstrate that host individuals distribute on Ω in a non-homogeneous way. Under the assumption that 1-D space, Ω = (0, 1), and the condition that {x ∈ [0, 1] : δ(·)β1(·) + β2()ϱ()u10()> δ(·)y(·)} is non-empty (of which locations can be termed as the favorite or not favorite sites for pathogens), we see from Theorem 6.2 that the infected hosts will vanish in some place and distribute in the remaining place (due to 6.24) and the susceptible hosts stay inhomogeneously on the whole habitat. When d1 → ∞ (d2 → ∞), susceptible (resp. infected) hosts distribute eventually over Ω, and infective (resp. susceptible) hosts distribute on Ω in a non-homogeneous way (see Theorems 6.3 and Theorem 6.4). The condition in Theorem 6.4 that ∫Ω[β1(·) + β2()ϱ()u10()/δ()]dx > ∫Ω y(·)dx is usually termed as the favorable domain for pathogens and also ensure the existence of the positive solution in Theorem 6.4). In summary, our result suggests that slow or fast movement of host individuals have a great impacts on the spatial distribution of the pathogens, which may help to design strategies for disease control and prevention.

On the other hand, our results on asymptotic profiles are established when positive steady state exists. With the same arguments as those in [31, 32], u10(x) b(x)/ζ (x) in C1(Ω) as d1 → 0 and η0 defined in 3.8 ensures that η0 < 0. It follows that ℜ0 < 1if d1 < 1, for some 1 > 0. This together with Theorem 4.1 indicate that pathogens still can be eliminated with d1 → 0. Compared to the results in [31, 32], our results enrich the dynamical results on the asymptotic profiles. In fact, our results also reveal that disease can not be eliminated by fast movement of host individuals. It also remains an challenging problem to revisit Theorem 6.2 as d2 → 0 without spatial dimension limitations. In a homogeneous case, it also remains an interesting problem to perform the bifurcation analysis on steady state solutions with specific bifurcation parameter.

Acknowledgments

J. Wang was supported by National Natural Science Foundation of China (nos. 12071115, 11871179), Natural Science Foundation of Heilongjiang Province (nos. LC2018002, LH2019A021) and Heilongjiang Provincial Key Laboratory of the Theory and Computation of Complex Systems. R. Cui was supported by Natural Science Foundation of Heilongjiang Province (no. LH2020A012), Heilongjiang Postdoctoral Science Foundation (no. LBH-Q17086) and the Fundamental Research Funds for Heilongjiang Provincial Universities (no. 2018-KYYWF-0996).

References

[1] L.J.S. Allen, B.M. Bolker, Y. Lou, A.L. Nevai, Asymptotic profiles of the steady states for an SIS epidemic reaction-diffusion model, Disc. Cont. Dyn. Syst. 21 (2008) 1-20.10.3934/dcds.2008.21.1Search in Google Scholar

[2] H. Brezis, W. A. Strauss, Semi-linear second-order elliptic equations in L1 J. Math. Soc. Japan 25(4) (1973) 565-590.10.2969/jmsj/02540565Search in Google Scholar

[3] R. Cui, K.-Y. Lam and Y. Lou, Dynamics and asymptotic profiles of steady states of an epidemic model in advective environments, J. Differential Equations, 263 (2017) 2343–2373.10.1016/j.jde.2017.03.045Search in Google Scholar

[4] R. Cui, Y. Lou, A spatial SIS model in advective heterogeneous environments, J. Differential Equations 261 (2016) 3305-3343.10.1016/j.jde.2016.05.025Search in Google Scholar

[5] Y. Du, R. Peng, M. Wang, Effect of a protection zone in the diffusive Leslie predator-prey model, J. Differential Equations 246 (2009) 3932-3956.10.1016/j.jde.2008.11.007Search in Google Scholar

[6] G. Dwyer, Density dependence and spatial structure in the dynamics of insect pathogens, The American Naturalist 143 (1994) 533-562.10.1086/285619Search in Google Scholar

[7] D. Gilbarg, N.S. Trudinger, Elliptic Partial Differential Equation of Second Order, Springer, 2001.10.1007/978-3-642-61798-0Search in Google Scholar

[8] J.K. Hale, Asnmptotic Behavior of Dissipative Systems, in: Mathematical Surveys and Monographs, vol. 25, American Mathematical Society, Providence (RI), 1988.Search in Google Scholar

[9] H. Li, R. Peng and F.-B. Wang, Varying total population enhances disease persistence: Qualitative analysis on a diffusive SIS epidemic model, J. Differential Equations 262 (2017) 885-913.10.1016/j.jde.2016.09.044Search in Google Scholar

[10] B. Li, H. Li and T. Tong, Analysis on a diffusive SIS epidemic model with logistic source, Z. Angew. Math. Phys. 68 (2017) 96.10.1007/s00033-017-0845-1Search in Google Scholar

[11] H. Li, R. Peng, Z.-A. Wang, On a diffusive susceptible-infected-susceptible epidemic model with mass action mechanism and birth-death effect: analysis, simulations, and comparison with other mechanisms, SIAM J. Appl. Math. 78 (2018) 2129-2153.Search in Google Scholar

[12] C.-S. Lin, W.-M. Ni, I. Takagi, Large amplitude stationary solutions to a chemotaxis system, J. Differential Equations 72 (1988) 1-27.10.1016/0022-0396(88)90147-7Search in Google Scholar

[13] Y. Lou, On the effects of migration and spatial heterogeneity on single and multiple species, J. Differential Equations 223 (2006) 400-426.10.1016/j.jde.2005.05.010Search in Google Scholar

[14] Y. Lou, W.-M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Differential Equations 131 (1996) 79-131.10.1006/jdeq.1996.0157Search in Google Scholar

[15] P. Magal, X-Q. Zhao, Global attractors and steady states for uniformly persistent dynamical systems, SIAM. J. Math. Anal. 37 (2005) 251-275.10.1137/S0036141003439173Search in Google Scholar

[16] R.H. Martin, H.L. Smith, Abstract functional differential equtions and reaction-diffusion systems, Trans. Amer. Math. Soc. 321 (1990) 1-44.Search in Google Scholar

[17] W.-M. Ni, I. Takagi, On the Neumann problem for some semilinear elliptic equations and systems of activator-inhibitor type, Trans. Amer. Math. Soc. 297 (1986) 351-368.10.1090/S0002-9947-1986-0849484-2Search in Google Scholar

[18] R. D. Nussbaum, Eigenvectors of nonlinear positive operator and the linear Krein-Rutman theorem, in: E. Fadell, G. Fournier (Eds.), Fixed Point Theory, Lecture Notes in Mathe- matics, Springer, New York/Berlin, 886 (1981), 309-331.Search in Google Scholar

[19] A. Pazy, Semigroups of Linear Operators and Application to Partial Differential Equations, Springer-Verlag, New York, 1983.10.1007/978-1-4612-5561-1Search in Google Scholar

[20] R. Peng, J. Shi, M. Wang, On stationary patterns of a reaction-diffusion model with autocatalysis and saturation law, Nonlinearity 21 (2008) 1471-1488.10.1088/0951-7715/21/7/006Search in Google Scholar

[21] R. Peng, Asymptotic profiles of the positive steady state for an SIS epidemic reaction-diffusion model. Part I, J. Differential Equations 247 (2009) 1096-1119.10.1016/j.jde.2009.05.002Search in Google Scholar

[22] H. L. Smith, Monotone dynamical systems: an introduction to the theory of competitive and cooperative systems, 41, American Mathematical Soc., 1995.Search in Google Scholar

[23] Y. Shi, J. Gao, J. Wang, Analysis of a reaction-diffusion host-pathogen model with horizontal transmission, J. Math. Anal. Appl. 481 (2020) 123481.10.1016/j.jmaa.2019.123481Search in Google Scholar

[24] H. Shu, X. Wang, Global dynamics of a coupled epidemic model, Discrete Continu Dyna. Syst. B 22(4) (2017) 1575-1585.10.3934/dcdsb.2017076Search in Google Scholar

[25] H.L. Smith, X.-Q. Zhao, Robust persistence for semidynamical systems, Nonlinear Anal. 47 (2001) 6169-6179.10.1016/S0362-546X(01)00678-2Search in Google Scholar

[26] H.R. Thieme, Convergence results and a Poincaré-Bendixson trichotomy for asymptotically autonomous differential equations, J. Math. Biol. 30 (1992) 755-763.10.1007/BF00173267Search in Google Scholar

[27] H.R. Thieme, Spectral bound and reproduction number for intinite-dimensional population structure and time heterogeneity, SIAM J. Appl. Math. 70 (2009) 188-211.10.1137/080732870Search in Google Scholar

[28] N. K. Vaidya, F.-B. Wang, X. Zou, Avian influenza dynamics in wild birds with bird mobility and spatial heterogeneous environment, Contin. Dyn. Syst. Ser. B 17 (2012) 2829-2848.10.3934/dcdsb.2012.17.2829Search in Google Scholar

[29] F.-B. Wang, J. Shi, X. Zou, Dynamics of a host-pathogen system on a bounded spatial domain, Commun. Pure Appl. Anal. 14 (6) (2015) 2535-2560.10.3934/cpaa.2015.14.2535Search in Google Scholar

[30] W. Wang, X.-Q. Zhao, Basic reproduction number for reaction-diffusion epidemic models, SIAM J. Appl. Dyn. Syst. 11 (2012) 1652-1673.10.1137/120872942Search in Google Scholar

[31] Y. Wu, X. Zou, Dynamics and profiles of a diffusive host-pathogen system with distinct dispersal rates, J. Differential Equations 264 (2018) 4989-5024.10.1016/j.jde.2017.12.027Search in Google Scholar

[32] J. Wang, J. Wang, Analysis of a reaction-diffusion cholera model with distinct dispersal rates in the human population, J. Dyn. Diff. Equat. https://doi.org/10.1007/s10884-019-09820-8.10.1007/s10884-019-09820-8Search in Google Scholar

[33] Y. Wu and X. Zou, Asymptotic profiles of steady states for a diffusive SIS epidemic model with mass action infection mechanism, J. Differential Equations 261 (2016) 4424-4447.10.1016/j.jde.2016.06.028Search in Google Scholar

[34] G.F. Webb, Theory of Nonlinear Age-Dependent Population Dynamics, Marcel Dekker, New York, 1985.Search in Google Scholar

[35] K. Yamazaki, Threshold dynamics of reaction-diffusion partial differential equations model of Ebola virus disease, Int. J. Biomath. 11 (2018), 1850108.10.1142/S1793524518501085Search in Google Scholar

[36] X.-Q. Zhao, Dynamical Systems in Population Biology, Springer-Verlag, New York, 2003.10.1007/978-0-387-21761-1Search in Google Scholar

Received: 2020-05-27
Accepted: 2020-11-11
Published Online: 2020-12-31

© 2021 Jinliang Wang and Renhao Cui, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 11.5.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0161/html
Scroll to top button