Parameter | Value | Parameter | Value | Parameter | Value | Parameter | Value |
b1 | 1.6 | Nb | 150 | μm | 0.15 | Λm | 6.5 |
μb | 0.01 [25] | Λb | 1.5 | d1 | 0.025 | d2 | 0.025 |
In this study, we investigated the threshold dynamics of a spatially heterogeneous nonlocal diffusion West Nile virus model. By employing semigroup theory and continuous Fréchet-differentiable, we established the well-posedness of the solution. The expression for the basic reproduction number derived using the next-generation matrix method. The authors demonstrated the threshold dynamics of the system by constructing a Lyapunov function and applying the comparison principle. Finally, numerical simulations were used to validate the theorem results. It can be suggested that to control disease development rapidly, measures should be taken to reduce the spread of mosquitoes and birds.
Citation: Kangkang Chang, Zhenyu Zhang, Guizhen Liang. Threshold dynamics of a nonlocal diffusion West Nile virus model with spatial heterogeneity[J]. AIMS Mathematics, 2023, 8(6): 14253-14269. doi: 10.3934/math.2023729
[1] | Folashade Agusto, Daniel Bond, Adira Cohen, Wandi Ding, Rachel Leander, Allis Royer . Optimal impulse control of West Nile virus. AIMS Mathematics, 2022, 7(10): 19597-19628. doi: 10.3934/math.20221075 |
[2] | Shaymaa H. Salih, Nadia M. G. Al-Saidi . 3D-Chaotic discrete system of vector borne diseases using environment factor with deep analysis. AIMS Mathematics, 2022, 7(3): 3972-3987. doi: 10.3934/math.2022219 |
[3] | Liping Wang, Peng Wu, Mingshan Li, Lei Shi . Global dynamics analysis of a Zika transmission model with environment transmission route and spatial heterogeneity. AIMS Mathematics, 2022, 7(3): 4803-4832. doi: 10.3934/math.2022268 |
[4] | Liye Wang, Wenlong Wang, Ruizhi Yang . Stability switch and Hopf bifurcations for a diffusive plankton system with nonlocal competition and toxic effect. AIMS Mathematics, 2023, 8(4): 9716-9739. doi: 10.3934/math.2023490 |
[5] | Xiong Zhang, Zhongyi Xiang . Piecewise immunosuppressive infection model with viral logistic growth and effector cell-guided therapy. AIMS Mathematics, 2024, 9(5): 11596-11621. doi: 10.3934/math.2024569 |
[6] | Xin Jiang . Threshold dynamics of a general delayed HIV model with double transmission modes and latent viral infection. AIMS Mathematics, 2022, 7(2): 2456-2478. doi: 10.3934/math.2022138 |
[7] | Frithjof Lutscher, Thomas Hillen . Correlated random walks in heterogeneous landscapes: Derivation, homogenization, and invasion fronts. AIMS Mathematics, 2021, 6(8): 8920-8948. doi: 10.3934/math.2021518 |
[8] | Chenxuan Nie, Dan Jin, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and generalist predator. AIMS Mathematics, 2022, 7(7): 13344-13360. doi: 10.3934/math.2022737 |
[9] | Kamel Guedri, Rahat Zarin, Ashfaq Khan, Amir Khan, Basim M. Makhdoum, Hatoon A. Niyazi . Modeling hepatitis B transmission dynamics with spatial diffusion and disability potential in the chronic stage. AIMS Mathematics, 2025, 10(1): 1322-1349. doi: 10.3934/math.2025061 |
[10] | J. P. Ndenda, J. B. H. Njagarah, S. Shaw . Influence of environmental viral load, interpersonal contact and infected rodents on Lassa fever transmission dynamics: Perspectives from fractional-order dynamic modelling. AIMS Mathematics, 2022, 7(5): 8975-9002. doi: 10.3934/math.2022500 |
In this study, we investigated the threshold dynamics of a spatially heterogeneous nonlocal diffusion West Nile virus model. By employing semigroup theory and continuous Fréchet-differentiable, we established the well-posedness of the solution. The expression for the basic reproduction number derived using the next-generation matrix method. The authors demonstrated the threshold dynamics of the system by constructing a Lyapunov function and applying the comparison principle. Finally, numerical simulations were used to validate the theorem results. It can be suggested that to control disease development rapidly, measures should be taken to reduce the spread of mosquitoes and birds.
West Nile virus disease is zoonotic and is becoming increasingly prevalent in Africa, Europe, the Americas, Australia, the Middle East, the Indian subcontinent and other regions. This poses a significant risk to human and animal health, with severe cases resulting in encephalitis, and even death. In nature, birds act as reservoirs for the West Nile virus, which circulates through a mosquito-bird cycle. Mosquitoes carrying the virus bite susceptible birds, leading to the development of toxemia in the birds. Subsequently, susceptible mosquitoes bite infected birds, contributing to the spread of the virus. Over 60 mosquito species and hundreds of bird species have been implicated in the transmission of the virus in the United States of America [1]. There is no specific treatment for the disease, and mathematical models can help us understand the pathogenesis of disease.
There is currently no specific treatment for the disease, but mathematical models can help us understand its pathogenesis. Numerous models have been developed to study the dynamic behavior of the West Nile virus [2,3,4,5,6,7,8,9,10,11,12,13]. Bai and Zhang [2] examined the dynamics of a periodic West Nile virus model, and concluded that using time-averaged extrinsic incubation period might underestimate the risk of disease outbreaks. Ge et al. [6] investigated the spatial transmission of reaction-diffusion West Nile virus with free boundaries in a heterogeneous environment, providing sufficient conditions for virus persistence and extinction based on the spatial risk index. Cheng and Zheng [9] explored the spatiotemporal spread and asymptotic behavior of the West Nile virus through a reaction-advection-diffusion system with free boundaries. They considered the impact of advection terms on the virus's extinction and persistence. Tarboush and Zhang [13] analyzed the effects of a periodically evolving domain on the dynamics of the West Nile virus spread. They introduced a spatiotemporal basic reproduction number that varies with the rate of periodic evolution to study the virus long-term behavior.
The existing literature West Nile virus models predominantly focus on reaction-diffusion approaches, in which the Laplacian operator accounts for the local behavior of random diffusion at adjacent spatial locations. However, the reaction-diffusion model represented by the Laplacian operator may not accurately capture the long-distance effects of disease transmission [14]. Consequently, to better represent the spread of the disease, integral operators are employed to describe the diffusion process, which accounts for the movement between non-adjacent spatial locations. Generally, ∫ΩJ(x−y)φ(y)dy−φ(x) (convolution operator) is utilized to model the nonlocal diffusion [15]. For instance, Du and Nie [16] explored the long-term dynamic behavior of a West Nile virus model with non-local spread and free boundaries. Additionally, Pu et al. [17] incorporated the influence of seasonal variation into the model based on Reference [16]. Some studies, e.g., [16] and [17] only considered the impact of non-local diffusion on the infected population; thus, we introduce a model [18] to facilitate the analysis of non-local diffusion's influence on both susceptible and infected populations.
{dSmdt=Λm−μmSm−b1βmNbSmIb,dImdt=b1βmNbSmIb−μmIm,dSbdt=Λb−μbSb−b1βbNbSbIm,dIbdt=b1βbNbSbIm−μbIb−dbIb, | (1.1) |
where Sm and Im represent the populations of uninfected and infected female mosquitoes, respectively. Sb and Ib denote the population of susceptible and infected birds, respectively. Λm and Λm are the recruitment rates of susceptible mosquitoes and birds, respectively. b1 is the per capita biting rate of mosquitoes on the bird. βm and βb are the West Nile virus transmission probabilities from infected birds to uninfected mosquitoes and from mosquitoes to birds, respectively. μm and μb are the natural death rates of mosquitos and birds, respectively; db denotes the disease-induced death rate of infected birds. The parameters in this model are all positive constants.
It is worth noting that the activity range of mosquitoes is generally limited, in general, with most only flying tens to hundreds of meters, and the longest flight distance being one to two kilometers. Therefore, we disregard the non-local spread of mosquitoes and we consider the following model.
{∂Sm∂t=Λm(x)−μm(x)Sm−b1(x)βm(x)NbSm(x,t)Ib(x,t),∂Im∂t=b1(x)βm(x)NbSm(x,t)Ib(x,t)−μm(x)Im(x,t),∂Sb∂t=d1∫ΩJ(x−y)Sb(y,t)dy−d1Sb(x,t)+Λb(x)−μb(x)Sb(x,t)−b1(x)βb(x)NbSb(x,t)Im(x,t),∂Ib∂t=d2∫ΩJ(x−y)Ib(y,t)dy−d2Ib(x,t)+b1(x)βb(x)NbSb(x,t)Im(x,t)−μb(x)Ib(x,t)−dbIb(x,t),Sm(x,0)=Sm,0(x),Im(x,0)=Im,0(x),Sb(x,0)=Sb,0(x)Ib(x,0)=Ib,0(x),x∈Ω,t>0. | (1.2) |
with the boundary condition
∂Sm∂ν=∂Im∂ν=∂Sb∂ν=∂Ib∂ν=0,x∈∂Ω,t>0, | (1.3) |
and initial condition
Sm(x,0)=Sm,0(x),Im(x,0)=Im,0(x),Sb(x,0)=Sb,0(x),Ib(x,0)=Ib,0(x),x∈Ω. | (1.4) |
d1>0 and d2>0 represent the diffusion coefficient of susceptible and infectious for birds, respectively. μh(x), μ(x), βH(x), b(x), γH(x), βv(x) and ν(x) are positive continuous functions on ¯Ω. The dispersal kernel function J is continuous and satisfies the following properties
J(0)>0,∫RJ(x)dx=1,J(x)>0on¯Ω,J(x)=J(−x)≥0onR. | (1.5) |
This paper is organized as follows. In Section 2, we prove the existence, uniqueness, positivity, and boundedness of solutions, and derive the expression for the basic reproduction number using the next-generation matrix method. In Section 3, we establish the global asymptotic stability and uniform persistence of the system by constructing Lyapunov functions and applying the comparison principle. Section 4 presents numerical simulations, and Section 5 concludes the paper with some final remarks.
Let us consider the following function spaces and positive cones.
X:=C(¯Ω),X+:=C+(¯Ω),Y:=C(¯Ω)×C(¯Ω)×C(¯Ω)×C(¯Ω),Y+:=C+(¯Ω)×C+(¯Ω)×C+(¯Ω)×C+(¯Ω). |
The norms in X and Y are defined as follows, respectively: ‖ϑ‖X:=supX∈¯Ω|ϑ(x)|,ϑ∈X, and
‖(z1,z2,z3,z4)‖Y:=supx∈¯Ω√|z1(x)|2+|z2(x)|2+|z3(x))|2+|z4(x))|2,(z1,z2,z3,z4)∈Y. |
Next, we define the linear operators on X.
A1ϑ1(x):=−μm(x)ϑ1(x),A2ϑ2(x):=−μm(x)ϑ2(x),A3ϑ3(x):=d1∫ΩJ(x−y)ϑ3(y)dy−d1ϑ3(x)−μb(x)ϑ3(x),A4ϑ4(x):=d2∫ΩJ(x−y)ϑ4(y)dy−d2ϑ4(x)−μb(x)ϑ4(x)−db(x)ϑ4(x). | (2.1) |
From the above, we know that A1, A2, A3 and A4 are bounded linear operators; by virtue of [19, Theorem 1.2], we obtain that A1(t)t≥0, A2(t)t≥0, A3(t)t≥0 and A4(t)t≥0 are uniformly continuous semigroups on X. Furthermore, according to [20, Section 2.1.1], the semigroups A1(t)t≥0, A2(t)t≥0, A3(t)t≥0 and A4(t)t≥0 are positive.
Next, we will prove the existence and uniqueness of the solution for system (1.2).
Lemma 2.1. For any initial data (Sm,0,Im,0,Sb,0,Ib,0) and t∈[0,t0), the solution (Sm(x,t),Im(x,t),Sb(x,t),Ib(x,t)) of system (1.2) satisfies that
limsupt→∞∫Ω[Sm(x,t)+Im(x,t)+Sb(x,t)+Ib(x,t)]dx<∞. |
Here |Ω| denotes the volume of Ω.
Proof. By (1.2), we have
ddt∫Ω[Sm(x,t)+Im(x,t)+Sb(x,t)+Ib(x,t)]dx=∫ΩΛm(x)dx−∫Ωμm(x)Smdx−∫Ωb1(x)βm(x)NbSm(x,t)Ib(x,t)dx∫Ωb1(x)βm(x)NbSm(x,t)Ib(x,t)dx−∫Ωμm(x)Im(x,t)dx+d1∫Ω∫ΩJ(x−y)Sb(y,t)dydx−d1∫ΩSb(x,t)dx+∫ΩΛb(x)dx−∫Ωμb(x)Sb(x,t)dx−∫Ωb1(x)βb(x)NbSb(x,t)Im(x,t)dx+d2∫Ω∫ΩJ(x−y)Ib(y,t)dydx−d2∫ΩIb(x,t)dx+∫Ωb1(x)βb(x)NbSb(x,t)Im(x,t)dx−∫Ωμb(x)Ib(x,t)dx−∫Ωdb(x)Ib(x,t)dx=∫ΩΛm(x)dx−∫Ωμm(x)(Sm+Im(x,t))dx+d1∫Ω∫ΩJ(x−y)Sb(y,t)dydx−d1∫ΩSb(x,t)dx+∫ΩΛb(x)dx−∫Ωμb(x)(Sb(x,t)+Ib(x,t))dx+d2∫Ω∫ΩJ(x−y)Ib(y,t)dydx−d2∫ΩIb(x,t)dx−∫Ωdb(x)Ib(x,t)dx. |
Moreover, according to (1.5), we obtain
ddt∫Ω[Sm(x,t)+Im(x,t)+Sb(x,t)+Ib(x,t)]dx≤∫ΩΛm(x)dx−∫Ωμm(x)(Sm(x,t)+Im(x,t))dx+d1∫Ω∫ΩJ(x−y)dySb(y,t)dx−d1∫ΩSb(x,t)dx |
+∫ΩΛb(x)dx−∫Ωμb(x)(Sb(x,t)+Ib(x,t))dx+d2∫Ω∫ΩJ(x−y)dyIb(y,t)dx−d2∫ΩIb(x,t)dx≤(ˉΛm+ˉΛb)|Ω|−∫Ωmin{μ_m,μ_b}(Sm+Im(x,t)+Sb(x,t)+Ib(x,t))dx, |
where |Ω| denotes the volume of Ω; ˉΛm and ˉΛb represent the upper bound of Λm and Λb, respectively. μ_m and μ_b denote the lower bound of μm and μb, respectively. By calculating, we have
limsupt→∞∫Ω[Sm(x,t)+Im(x,t)+Sb(x,t)+Ib(x,t)]dx≤(ˉΛm+ˉΛb)|Ω|min{μ_m,μ_b}. |
Lemma 2.2. If (Sm(⋅,t),Im(⋅,t),Sb(⋅,t)Ib(⋅,t))∈Y is the solution of system (1.2) with (Sm,0,Im,0,Sb,0Ib,0)∈Y+, then (Sm(⋅,t),Im(⋅,t),Sb(⋅,t)Ib(⋅,t))∈Y+ for all t∈[0,t0).
Proof. By calculation, we have
Sm(x,t)=Sm,0(x)e−∫t0(μm(x)+b1(x)βm(x)NbIb(x,u))du+∫t0Λme∫tτ(−μm(x)+b1(x)βm(x)NbIb(x,u))dudτ,Im(x,t)=Im,0(x)e−μ(x)t+∫t0b1(x)βm(x)NbSm(x,u)Ib(x,u)e−μ(x)(t−τ)dτ, | (2.2) |
and
Sb(x,t)=Sb,0(x)e−∫t0(d1+μb(x)+b1(x)βb(x)NbIm(x,u))du+∫t0(d1∫ΩJ(x−y)Sb(y,t)dy+Λb(x))×e−∫tτ(d1+μb(x)+b1(x)βb(x)NbIm(x,u))dudτ,Ib(x,t)=Ib,0(x)e−(d2+db+μb(x))t+∫t0(d2∫ΩJ(x−y)Ib(y,t)dy+b1(x)βb(x)NbSb(x,u)Im(x,u))×e−(d2+μb(x))(t−τ)dτ | (2.3) |
for all t∈[0,t0) and x∈ˉΩ. Due to (Sm,0,Im,0,Sb,0Ib,0)∈Y+ and J(x)>0 on R, it means that Sm(x,t)≥0, Im(x,t)≥0, Sb(x,t)≥0 and Ib(x,t)≥0; further, Sm(⋅,t)>0, Im(⋅,t)>0, Sb(⋅,t)>0 and Ib(⋅,t)>0 for t∈[0,t0).
Theorem 2.1. Suppose that (Sm,0,Im,0,Sb,0,Ib,0)∈Y. Then, there exists a t0>0 such that system (1.2) has the unique solution (Sm(⋅,t),Im(⋅,t),Sb(⋅,t)Ib(⋅,t)) for all t∈[0,t0), and either t0=+∞ or lim supt→t0−0‖(Sm(⋅,t),Im(⋅,t),Sb(⋅,t),Ib(⋅,t)‖=+∞.
Proof.
ϝ(ς1,ς2,ς3,ς4)(x)=(Λm(x)−βm(x)b1(x)Nbς1ς4βm(x)b1(x)Nbς1ς4Λb(x)−βb(x)b1(x)Nbς2ς3βb(x)b1(x)Nbς2ς3). |
Let ϝ′[ϱ1,ϱ2,ϱ3,ϱ4] be a linear operator on Y defined as follows:
ϝ′[ϱ1,ϱ2,ϱ3,ϱ4](ς1,ς2,ς3,ς4)(x)=(−βm(x)b1(x)Nbϱ4ς1−βm(x)b1(x)Nbϱ1ς4βm(x)b1(x)Nbϱ4ς1+βm(x)b1(x)Nbϱ1ς4−βb(x)b1(x)Nbϱ3ς2−βb(x)b1(x)Nbϱ2ς3βb(x)b1(x)Nbϱ3ς2+βb(x)b1(x)Nbϱ2ς3). |
By calculating, we have
ϝ(ς1,ς2,ς3,ς4)(x)=F(ϱ1,ϱ2,ϱ3,ϱ4)(x)+ϝ′[ϱ1,ϱ2,ϱ3,ϱ4](ς1−ϱ1,ς2−ϱ2,ς3−ϱ3,ς4−ϱ4)(x)+(−βm(x)b1(x)Nb(ς1−ϱ1)(ς4−ϱ4)βm(x)b1(x)Nb(ς1−ϱ1)(ς4−ϱ4)−βb(x)b1(x)Nb(ς2−ϱ2)(ς3−ϱ3)βb(x)b1(x)Nb(ς2−ϱ2)(ς3−ϱ3)); |
due to the coefficients being positive and bounded, we have that the last term on the right-hand side of this equation is o{(ς1,ς2,ς3,ς4)T−(ϱ1,ϱ2,ϱ3,ϱ4)T}. It means that ϝ is Fréchet-differentiable for (ϱ1,ϱ2,ϱ3,ϱ4)T on Y. Moreover, we have
‖ϝ′[ϱ1,ϱ2,ϱ3,ϱ4]−ϝ′[˜ϱ1,˜ϱ2,˜ϱ3,˜ϱ4]‖=sup‖(ς1,ς2,ς3,ς4)T‖Y≤1‖{ϝ′[ϱ1,ϱ2,ϱ3,ϱ4]−ϝ′[˜ϱ1,˜ϱ2,˜ϱ3,˜ϱ4]}(ς1,ς2,ς3,ς4)‖Y=sup‖(ς1,ς2,ς3,ς4)T‖Y≤1‖{−βm(x)b1(x)Nb(ϱ4−˜ϱ4)ς1−βm(x)b1(x)Nb(ϱ1−˜ϱ1)ς4,βm(x)b1(x)Nb(ϱ4−˜ϱ4)ς1+βm(x)b1(x)Nb(ϱ1−˜ϱ1)ς4,−βb(x)b1(x)Nb(ϱ3−˜ϱ3)ς2−βb(x)b1(x)Nb(ϱ2−˜ϱ2)ς3,βb(x)b1(x)Nb(ϱ3−˜ϱ3)ς2+βb(x)b1(x)Nb(ϱ2−˜ϱ2)ς3}‖≤2ˉβˉb1Nb‖(ϱ1−˜ϱ1,ϱ2−˜ϱ2,ϱ3−˜ϱ3,ϱ4−˜ϱ4)T‖Y, |
where (˜ϱ1,˜ϱ2,˜ϱ3,˜ϱ4)T∈Y; this implies that ϝ is continuously Fréchet-differentiable [21, Lemma 3.1] on Y.
Due to A1(t)t≥0, A2(t)t≥0, A3(t)t≥0 and A4(t)t≥0 being generators of uniformly continuous semigroups A1(t)t≥0, A2(t)t≥0, A3(t)t≥0 and A4(t)t≥0, the solution (Sm(x,t),Im(x,t),Sb(x,t),Ib(x,t)) of system (1.2) can be written as follows:
w(x,t)=eA(t)w(⋅,t)(x)+∫t0EA(t−s)F(w(⋅,τ))(x)ds,t≥0,x∈ˉΩ, |
where
w(x,t)=(Sm(x,t)Im(x,t)Sb(x,t)Ib(x,t)),A(t)=(A1(t)A2(t)A3(t)A4(t)),ϝ(w(x,t))=(Λm(x)−b1(x)βm(x)NbSm(x,t)Ib(x,t)b1(x)βm(x)NbSm(x,t)Ib(x,t)Λb(x)−b1(x)βb(x)NbSb(x,t)Im(x,t)b1(x)βb(x)NbSb(x,t)Im(x,t)). |
Due to A being the infinitesimal generator of etAt≥0 and ϝ being continuously Frˊechet-differentiable on Y, from [22, Proposition 4.16], the result holds.
To give the expression of the basic reproduction number in an abstract way using the next-generation matrix method [23], we consider the following linearized equations around the disease-free equilibrium E0=(S0m(x),0,S0b(x),0):
{∂Im∂t=b1(x)βm(x)NbS0mIb(x,t)−μm(x)Im(x,t),∂Ib∂t=d2∫ΩJ(x−y)Ib(y,t)dy−d2Ib(x,t)+b1(x)βb(x)NbS0bIm(x,t)−μb(x)Ib(x,t)−dbIb(x,t). | (3.1) |
System (3.1) is equivalent to
∂Ξ∂t=(M−N)Ξ+QΞ,x∈Ω,t>0, |
where
Ξ=(ImIb),M=(000d2∫ΩJ(x−y)dy), |
and
N=(μm00d2+μb+db)Q=(0b1(x)βm(x)NbS0mb1(x)βb(x)NbS0b0). |
Let K(t) be the solution semigroup with respect to the following linear reaction-diffusion equation
∂Ξ∂t=(M−N)Ξ,x∈Ω,t>0. | (3.2) |
Define
K(ς)(x):=∫∞0Q[K(t)(ς)](x)dt. |
By virtue of the next-infection operator, the spectral radius of K can be defined as the basic reproduction number
R0:=r(K). |
Moreover, by virtue of [24], we can obtain the following lemma:
Lemma 3.1. sign(R0−1)=signλ0.
The principal eigenvalue λ0 is related to eigenfunction (Θ0(x),Υ0(x)), which satisfies the following equation:
{λΘ(x)=b1(x)βm(x)NbS0mΥ(x)−μm(x)Θ(x),λΥ(x)=d2∫ΩJ(x−y)Υ(x)dy−d2Υ(x)+b1(x)βb(x)NbS0bΘ(x))−μb(x)Υ(x)−dbΥ(x). | (3.3) |
Before proving the global asymptotic stability of the disease-free equilibrium point, some lemmas need to be given; we first consider the following eigenvalue problem, which has been studied by Garcˊia-Meliˊan and Rossi [15].
{∫RNJ(x−y)(Ψ(y)−Ψ(x))dy=−λeΨ(x),inΩ,Ψ(x)=0,onRN∖Ω. | (3.4) |
Lemma 3.2. For system (3.4), there exists a unique principal eigenvalue λ1 corresponding to eigenfunction Ψ(x). Furthermore, 0<λ1<1 and
λ1=infψ∈L2(Ω),Ψ≠0∫ΩΨ2(x)dx−∫Ω∫ΩJ(x−y)Ψ(y)Ψ(x)dydx∫ΩΨ2(x). |
Now, we have the following global stability result.
Theorem 3.1. If R0<1, the solution (Sm(x,t),Im(x,t),Sb(x,t)Ib(x,t)) of system (1.2) satisfies that
limt→+∞Sm(x,t)=S0m(x),limt→+∞Im(x,t)=0,limt→+∞Sh(x,t)=S0h(x),limt→+∞Ih(x,t)=0. |
Proof. By virtue of Eqs (2.2) and (2.3), we can obtain that Im→0 and Ib→0 as t→+∞.
Moreover, we prove that Sm(x,t)→S0m(x) on x as t→+∞; let p1(x,t)=Sm(x,t)−S0m(x); then, we have
∂p1(x,t)∂t=−μm(x)p1(x,t)−b1(x)βm(x)NbSm(x,t)Ib(x,t),x∈Ω. | (3.5) |
By calculating, we have
p1(x,t)=p1,0(x)e−μm(x)t−1μm(x)b1(x)βm(x)NbSm(x,t)Ib(x,t), |
as t→∞, we obtain that p1(x,t)→0; hence, Sm(x,t)→S0m(x).
Next, we prove that Sb(x,t)→S0b(x) on x as t→+∞; let p2(x,t)=Sb(x,t)−S0b(x). Furthermore, we have
∂p2(x,t)∂t=d1∫ΩJ(x−y)p2(y,t)dy−d1p2(x,t)−μb(x)p2(x,t)−b1(x)βb(x)NbSb(x,t)Im(x,t),x∈Ω. | (3.6) |
Let P(t)=∫Ωp22(x,t)dx; we can obtain
dP(t)dt=2∫Ωp2(x,t)∂p2(x,t)∂tdx=2∫Ωp2(x,t){d1∫ΩJ(x−y)p2(y,t)dy−d1p2(x,t)−μb(x)p2(x,t)−b1(x)βb(x)NbSb(x,t)Im(x,t)}dx=2{d1∫Ω∫ΩJ(x−y)p2(y,t)p2(x,t)dydx−∫Ωp22(x,t)dx}−2∫Ω{μb(x)p2(x,t)+b1(x)βb(x)NbSb(x,t)Im(x,t)}p2(x,t)dx≤−2d1λ1P(t). | (3.7) |
By calculation we have that
P(t)≤c0e−2d1λ1t. |
Hence, there exists a constant c0; we have
‖p2(⋅,t)‖L2(Ω)≤c0e−d1λ1t. |
By virtue of Eq (3.6), we can obtain
p2(x,t)=p0(x)e−(d1+μb)t+e−(d1+μb)t∫t0e(d1+μb)s(d1∫ΩJ(x−y)p2(y,t)dy−b1(x)βb(x)NbSb(x,t)Im(x,t))ds. | (3.8) |
Applying the H¨older inequality to the following equation, there exists some positive constant satisfying that
∫ΩJ(x−y)p2(y,t)dy≤C‖p2(⋅,s)‖L2(Ω). | (3.9) |
Combine (3.8) and (3.9); there exists some positive constants ci(i=1,2) we have
|p2(x,t)|≤c1e−(d1+μb)t+c2e−d1λ1t. |
Hence, as t→∞, p2(x,t)→0 uniformly on x∈Ω. Furthermore, we obtain that Sb(x,t)→S0b(x).
In this section, we consider the uniform persistence of system (1.2). To achieve these goals, we first consider the following problem.
Theorem 3.2. If R0>1, then there exists a function Π(x), such that
limt→∞inf(Sm(x,t)+Im(x,t)+Sb(x,t)+Ib(x,t))≥Π(x); |
hence, the disease is uniformly persistent.
Proof. Due to R0>1, there exists a κ>0 such that λ(S∗m−κ,S∗b−κ)>0 (where (S∗m,I∗m,S∗b,I∗b) represents that the endemic equilibrium). It means that there exists a ˜t1>0 satisfying that Sm(x,t)>Sm,0−κ and Sb(x,t)>Sb,0−κ for t≥˜t1 and x∈¯Ω. For x∈Ω,t>˜t1; according to the comparison principle, we can obtain
{∂Im∂t≥b1(x)βm(x)Nb(Sm,0−k)Ib(x,t)−μm(x)Im(x,t),∂Ib∂t≥d2∫ΩJ(x−y)Ib(y,t)dy−d2Ib(x,t)+b1(x)βb(x)Nb(Sb,0−κ)Im(x,t)−μb(x)Ib(x,t)−dbIb(x,t). |
Define (˜Im(x,t),˜Ib(x,t))=(Ne˜λt˜ψ1(x),Ne˜λt˜ψ2(x)), (˜Im(x,t),˜Ib(x,t)) to satisfy the following equation
{∂˜Im∂t=b1(x)βm(x)Nb(Sm,0−k)Ib(x,t)−μm(x)Im(x,t),∂˜Ib∂t=d2∫ΩJ(x−y)Ib(y,t)dy−d2Ib(x,t)+b1(x)βb(x)Nb(Sb,0−κ)Im(x,t)−μb(x)Ib(x,t)−dbIb(x,t), |
where (˜ψ1(x),˜ψ2(x)) is the eigenfunction with respect to ˜λ<0. According to the comparison principle, we know that Im(x,t)≥˜Im(x,t), Ib(x,t)≥˜Ib(x,t) for x∈Ω,t>˜t1. Therefore, Im(x,t)≥Ne˜λt˜ψ1(x), Ib(x,t)≥Ne˜λt˜ψ2(x) such that
limt→∞infIH(x,t)≥N˜ψ1(x),limt→∞infIV(x,t)≥N˜ψ2(x). |
On the basis of Lemma (2.1), we know that there exists constants K>0 and ˜t2 such that
Im(x,t)≤K,Ib(x,t)≤K,t≥˜t2,x∈¯Ω. |
Then, SH and SV satisfy the following equation
{∂Sm∂t≥Λm(x)−μm(x)Sm−b1(x)βm(x)KNbSm(x,t),∂Sb∂t≥d1∫ΩJ(x−y)Sb(y,t)dy−d1Sb(x,t)+Λb(x)−μb(x)Sb(x,t)−b1(x)βb(x)KNbSb(x,t),x∈Ω,t>˜t2. |
By virtue of Eq (3.4), we obtain
{∂Sm∂t≥Λm(x)−μm(x)Sm−b1(x)βm(x)KNbSm(x,t),∂Sb∂t≥−d1λeSb(x,t)+Λb(x)−μb(x)Sb(x,t)−b1(x)βb(x)KNbSb(x,t),x∈Ω,t>˜t2. |
Hence
limt→∞infSm(x,t)≥(Λm(x))/(μm(x)+b1(x)βm(x)KNb),limt→∞infSb(x,t)≥(Λb(x))/(d1λe+μb(x)+b1(x)βb(x)KNb). |
Let Π(x):=min{(Λm(x))/(μm(x)+b1(x)βm(x)KNb),(Λb(x))/(d1λe+μb(x)+b1(x)βb(x)KNb),N˜ψ1(x),N˜ψ2(x)}. The disease uniformly persistent is obtained.
In order to present the theoretical results, some numerical simulations are presented in this section. The parameter values (Table 1) and initial value are selected as follows (all rates are per capita per day (day−1)):
Parameter | Value | Parameter | Value | Parameter | Value | Parameter | Value |
b1 | 1.6 | Nb | 150 | μm | 0.15 | Λm | 6.5 |
μb | 0.01 [25] | Λb | 1.5 | d1 | 0.025 | d2 | 0.025 |
Initial value:
(Sm,0(x),Im,0(x),Sb,0(x),Ib,0(x))=(0.03sinx+0.05cosx,0.02cosx,0.01sinx+0.03cosx,0). |
Moreover, the nonlocal kernel function [20] is selected as follows:
Jx={Aexp(1x2−1),−1<x<1,0,otherwise. |
Here, A=2.6423, x∈[−1,1]⊂R and ∫RJ(x)dx=∫1−1J(x)dx≈1.
In Figure 1, let βm=0.12,βb=0.15. In Figure 3, set βm=0.075(1−0.85cosx),βb=0.15. As t→∞, the density of the infected mosquitoes and infected bird will converge to 0, that is, the disease will become extinct.
In Figure 2, let βm=1.2,βb=0.88. In Figure 4, set βm=0.75(1−0.85cosx),βb=0.88 [18]. It can be seen that the solution of system (1.2) finally converges to a steady state, indicating the persistence of the diseases.
For R0<1, we can see that when the transmission rate is constant, the number of infected drops to 0 relatively quickly; on the contrary, when the transmission rate is spatially dependent, the number of infected tends to 0 for a long time. As R0>1, spatial heterogeneity can increase in the number of susceptible mosquitoes and birds, and the number of infected mosquitoes and birds become more evenly distributed in space.
In this section, we aim to examine the impact of the diffusion coefficient on disease persistence considering that the spread of extinct diseases is meaningless. For Figures 4–7, we set d1=d2=0.025, d1=d2=0.10, d1=d2=0.20 and d1=d2=0, respectively. The results indicate that an increase in the diffusion coefficient leads to an increase in the number of suspected mosquitoes and birds while decreasing the number of infected mosquitoes and birds. However, the differences in mosquito and bird populations across in different spatial distributions become more apparent, resulting in a higher persistence of the disease and making it challenging to control. Therefore, in the event of a disease outbreak, we recommend employing physical or chemical methods to kill mosquitoes in order to better control the spread of the disease.
We investigated the threshold dynamics of a nonlocal diffusion West Nile virus model with spatial heterogeneity. The well-posedness of the solution is proven using the semigroup theory, and it is continuously Fréchet differentiable. The basic reproduction number is expressed using the next-generation matrix method. Global stability and uniform persistence of the system are proven by constructing a Lyapunov function and applying the comparison principle. Numerical simulations were performed to verify the theorem. The values of the transmission rate were adjusted to examine extinction and persistence of the disease, and the impact of diffusion was considered. Simulation results demonstrated that an increase in the diffusion coefficient leads to an increase in the number of susceptible mosquitoes and birds and a decrease in the number of infected mosquitoes and birds. However, differences in mosquito and bird populations in different spatial distributions become more pronounced, leading to enhanced disease persistence and reduced chances of rapid disease control. Therefore, to better control the spread of the disease, physical or chemical methods for mosquito control are recommended when a disease outbreak occurs.
The authors declare that they have no conflict of interest.
[1] | Chinese Center for Disease Control and Prevention. Available from: https://www.chinacdc.cn/. |
[2] |
Z. Bai, Z. Zhang, Dynamics of a periodic West Nile virus model with mosquito demographics, Commun. Pure Appl. Anal., 21 (2022), 3755–3775. http://doi.org/10.3934/cpaa.2022121 doi: 10.3934/cpaa.2022121
![]() |
[3] |
J. Ge, Z. Lin, A. K. Tarboush, H. Zhu, Dynamics of West Nile virus driven by seasonal fluctuations in a spatially variable habitat, Discrete Contin. Dyn. Syst. Ser. B, 28 (2023), 2081–2103. http://doi.org/10.3934/dcdsb.2022159 doi: 10.3934/dcdsb.2022159
![]() |
[4] |
S. A. Moon, L. W. Cohnstaedt, D. S. McVey, C. M. Scoglio, A spatio-temporal individual-based network framework for West Nile virus in the USA: spreading pattern of West Nile virus, PLoS Comput. Biol., 15 (2019), e1006875. http://doi.org/10.1371/journal.pcbi.1006875 doi: 10.1371/journal.pcbi.1006875
![]() |
[5] |
A. K. Tarboush, J. Ge, Z. Lin, Coexistence of a cross-diffusive West Nile virus model in a heterogenous environment, Math. Biosci. Eng., 15 (2018), 1479–1494. http://doi.org/10.3934/mbe.2018068 doi: 10.3934/mbe.2018068
![]() |
[6] |
J. Ge, Z. Lin, H. Zhu, Modeling the spread of West Nile virus in a spatially heterogeneous and advective environment, J. Appl. Anal. Comput., 11 (2021), 1868–1897. http://doi.org/10.11948/20200258 doi: 10.11948/20200258
![]() |
[7] | C. Cheng, Z. Zheng, Spatial and temporal dynamics of an almost periodic reaction-diffusion system for West Nile virus, arXiv: 2012.11789. |
[8] |
Z. Lin, H. Zhu, Spatial spreading model and dynamics of West Nile virus in birds and mosquitoes with free boundary, J. Math. Biol., 75 (2017), 1381–1409. https://doi.org/10.1007/s00285-017-1124-7 doi: 10.1007/s00285-017-1124-7
![]() |
[9] |
C. Cheng, Z. Zheng, Dynamics and spreading speed of a reaction-diffusion system with advection modeling West Nile virus, J. Math. Anal. Appl., 493 (2021), 124507. https://doi.org/10.1016/j.jmaa.2020.124507 doi: 10.1016/j.jmaa.2020.124507
![]() |
[10] |
M. J. Wonham, T. De-Camino-Beck, M. A. Lewis, An epidemiological model for West Nile virus: invasion analysis and control applications, Proc. R. Soc. Lond. B, 271 (2004), 501–507. https://doi.org/10.1098/rspb.2003.2608 doi: 10.1098/rspb.2003.2608
![]() |
[11] |
A. Abdelrazec, S. Lenhart, H. Zhu, Transmission dynamics of West Nile virus in mosquitoes and corvids and non-corvids, J. Math. Biol., 68 (2014), 1553–1582. https://doi.org/10.1007/s00285-013-0677-3 doi: 10.1007/s00285-013-0677-3
![]() |
[12] |
N. A. Maidana, H. M. Yang, Spatial spreading of West Nile virus described by traveling waves, J. Theor. Biol., 258 (2009), 403–417. https://doi.org/10.1016/j.jtbi.2008.12.032 doi: 10.1016/j.jtbi.2008.12.032
![]() |
[13] |
A. K. Tarboush, Z. Zhang, The diffusive model for West Nile virus on a periodically evolving domain, Complexity, 2020 (2020), 6280313. https://doi.org/10.1155/2020/6280313 doi: 10.1155/2020/6280313
![]() |
[14] | J. D. Murray, Mathematical biology II: spatial models and biomedical applications, 3 Eds., New York: Springer, 2003. https://doi.org/10.1007/b98869 |
[15] |
J. Garcˊia-Meliˊan, J. D. Rossi, On the principal eigenvalue of some nonlocal diffusion problems, J. Differ. Equations, 246 (2009), 21–38. https://doi.org/10.1016/j.jde.2008.04.015 doi: 10.1016/j.jde.2008.04.015
![]() |
[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] |
L. Pu, Z. Lin, Y. Lou, A West Nile virus nonlocal model with free boundaries and seasonal succession, J. Math. Biol., 86 (2023), 25. https://doi.org/10.1007/s00285-022-01860-x doi: 10.1007/s00285-022-01860-x
![]() |
[18] |
J. Jiang, Z. Qiu, J. Wu, H. Zhu, Threshold conditions for West Nile virus outbreaks, Bull. Math. Biol., 71 (2009), 627–647. https://doi.org/10.1007/s11538-008-9374-6 doi: 10.1007/s11538-008-9374-6
![]() |
[19] | A. Pazy, Semigroups of linear operators and applications to partial differential equations, New York: Springer, 1983. https://doi.org/10.1007/978-1-4612-5561-1 |
[20] |
C. Y. Kao, Y. Lou, W. Shen, Random dispersal vs non-local dispersal, Discrete Contin. Dyn. Syst., 26 (2010), 551–596. https://doi.org/10.3934/dcds.2010.26.551 doi: 10.3934/dcds.2010.26.551
![]() |
[21] |
T. Kuniya, J. Wang, Lyapunov functions and global stability for a spatially diffusive SIR epidemic model, Appl. Anal., 96 (2017), 1935–1960. https://doi.org/10.1080/00036811.2016.1199796 doi: 10.1080/00036811.2016.1199796
![]() |
[22] | G. F. Webb, Theory of nonlinear age-dependent population dynamics, CRC Press, 1985. |
[23] |
O. Diekmann, J. A. P. Heesterbeek, J. A. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol., 28 (1990), 365–382. https://doi.org/10.1007/BF00178324 doi: 10.1007/BF00178324
![]() |
[24] |
V. Hutson, S. Martinez, K. Mischaikow, G. T. Vickers, The evolution of dispersal, J. Math. Biol., 47 (2003), 483–517. https://doi.org/10.1007/s00285-003-0210-1 doi: 10.1007/s00285-003-0210-1
![]() |
[25] |
M. Maliyoni, Probability of disease extinction or outbreak in a stochastic epidemic model for West Nile virus dynamics in birds, Acta Biotheor., 69 (2021), 91–116. https://doi.org/10.1007/s10441-020-09391-y doi: 10.1007/s10441-020-09391-y
![]() |
1. | Kangkang Chang, Zhenyu Zhang, Guizhen Liang, Dynamics analysis of a nonlocal diffusion dengue model, 2023, 13, 2045-2322, 10.1038/s41598-023-42440-3 | |
2. | Arturo J. Nic‐May, Eric J. Avila‐Vales, Dynamics of a Two‐Strain Model With Vaccination, General Incidence Rate, and Nonlocal Diffusion, 2025, 0170-4214, 10.1002/mma.10680 |
Parameter | Value | Parameter | Value | Parameter | Value | Parameter | Value |
b1 | 1.6 | Nb | 150 | μm | 0.15 | Λm | 6.5 |
μb | 0.01 [25] | Λb | 1.5 | d1 | 0.025 | d2 | 0.025 |
Parameter | Value | Parameter | Value | Parameter | Value | Parameter | Value |
b1 | 1.6 | Nb | 150 | μm | 0.15 | Λm | 6.5 |
μb | 0.01 [25] | Λb | 1.5 | d1 | 0.025 | d2 | 0.025 |