Reaction-diffusion equations have been used to describe the dynamical behavior of epidemic models, where the spreading of infectious disease has the same speed in every direction. A natural question is how to describe the dynamical system when the spreading of infectious disease is directed diffusion. We introduce the road diffusion into a Ross epidemic model which describes the spread of infected Mosquitoes and humans. With the comparison principle the system is proved to have a unique global solution. By the approach of upper and lower solutions, we show that the disease-free equilibrium is asymptotically stable if the basic reproduction number is lower than 1 while the endemic equilibrium asymptotically stable if the basic reproduction number is greater than 1.
Citation: Xiaomei Bao, Canrong Tian. Stability in a Ross epidemic model with road diffusion[J]. AIMS Mathematics, 2022, 7(2): 2840-2857. doi: 10.3934/math.2022157
[1] | Shufan Wang, Zhihui Ma, Xiaohua Li, Ting Qi . A generalized delay-induced SIRS epidemic model with relapse. AIMS Mathematics, 2022, 7(4): 6600-6618. doi: 10.3934/math.2022368 |
[2] | Rahat Zarin, Amir Khan, Pushpendra Kumar, Usa Wannasingha Humphries . Fractional-order dynamics of Chagas-HIV epidemic model with different fractional operators. AIMS Mathematics, 2022, 7(10): 18897-18924. doi: 10.3934/math.20221041 |
[3] | 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 |
[4] | Mireia Besalú, Giulia Binotto . Time-dependent non-homogeneous stochastic epidemic model of SIR type. AIMS Mathematics, 2023, 8(10): 23218-23246. doi: 10.3934/math.20231181 |
[5] | Muhammad Altaf Khan, Sajjad Ullah, Saif Ullah, Muhammad Farhan . Fractional order SEIR model with generalized incidence rate. AIMS Mathematics, 2020, 5(4): 2843-2857. doi: 10.3934/math.2020182 |
[6] | Rahat Zarin, Amir Khan, Aurangzeb, Ali Akgül, Esra Karatas Akgül, Usa Wannasingha Humphries . Fractional modeling of COVID-19 pandemic model with real data from Pakistan under the ABC operator. AIMS Mathematics, 2022, 7(9): 15939-15964. doi: 10.3934/math.2022872 |
[7] | Xuechao Zhang, Yuhan Hu, Shichang Lu, Haomiao Guo, Xiaoyan Wei, Jun li . Dynamic analysis and control of online information dissemination model considering information beneficiaries. AIMS Mathematics, 2025, 10(3): 4992-5020. doi: 10.3934/math.2025229 |
[8] | Amir Khan, Abdur Raouf, Rahat Zarin, Abdullahi Yusuf, Usa Wannasingha Humphries . Existence theory and numerical solution of leptospirosis disease model via exponential decay law. AIMS Mathematics, 2022, 7(5): 8822-8846. doi: 10.3934/math.2022492 |
[9] | Ana P. Lemos-Paião, Cristiana J. Silva, Delfim F. M. Torres . A cholera mathematical model with vaccination and the biggest outbreak of world’s history. AIMS Mathematics, 2018, 3(4): 448-463. doi: 10.3934/Math.2018.4.448 |
[10] | Ishtiaq Ali, Sami Ullah Khan . Dynamics and simulations of stochastic COVID-19 epidemic model using Legendre spectral collocation method. AIMS Mathematics, 2023, 8(2): 4220-4236. doi: 10.3934/math.2023210 |
Reaction-diffusion equations have been used to describe the dynamical behavior of epidemic models, where the spreading of infectious disease has the same speed in every direction. A natural question is how to describe the dynamical system when the spreading of infectious disease is directed diffusion. We introduce the road diffusion into a Ross epidemic model which describes the spread of infected Mosquitoes and humans. With the comparison principle the system is proved to have a unique global solution. By the approach of upper and lower solutions, we show that the disease-free equilibrium is asymptotically stable if the basic reproduction number is lower than 1 while the endemic equilibrium asymptotically stable if the basic reproduction number is greater than 1.
Mosquito was firstly discovered a transmission vector of Malaria by Ronald Ross in 1898. Ross was also the first to use differential equation to study epidemiology. In order to build an intra-host epidemic model, the mosquito population is divided into two compartments: healthy mosquitoes (susceptible mosquitoes) and infected mosquitoes, while the human population is divided into two compartments: susceptible humans and infected humans. He constructed a two-component system composed with the infected humans u and the infectious mosquitoes v as the following form:
{dudt=mab1v(1−u)−γ1u,dvdt=ab2(1−v)u−γ2v. | (1.1) |
He proposed out the threshold (the predecessor of basic reproduction number) to determine the asymptotic stability of the disease-free equilibrium and endemic equilibrium.
It is assumed that the densities of mosquitoes and humans are homogeneous in the above model. The heterogeneous spatial distribution has formed a reasonable basis for studying insect dispersal (Lewis [1]). For example, in 1986 Aedes albopictus was found for the first time in northern counties in Florida (Peacock et al. [2]). Then it kept spreading southward, slowly but steadily, and had spread all over the 67 counties (O'Meara et al. [3]) in six years' time. By 2008, Aedes albopictus had spread over to 36 states and was still continuing its expansion (Hawley et al. [4], Enserink [5], and Hahnet et al. [6]). In 2013, Rochlin et al.[7] predicted that, especially in urban areas, in the next 20 years, Aedes albopictus population, will be over three times more. A recent survey on the current distribution of Aedes albopictus (Parker et al. [8]) shows that Aedes albopictus have been tracked in 56 of all 67 Florida counties. The mosquito of Aedes albopictus had also been reported to spread along roads. Bennett et al. [9] found that the infestation rate of Aedes albopictus was high in garages trading used tires along the highways, and this road provided a channel for fast dispersal across Panama.
Considering the effect of highways on the spreading of mosquito, Berestycki, Roquejoffre and Rossi [10] introduced the road diffusion into a reaction-diffusion system. They proposed the model where a two-dimensional environment contains a "line" inside which fast diffusion occurs, while reproduction and the usual diffusion occur only outside. Once "plane" is a field and "line" is a road, this system is a combination of the density of the field with that of the road, between which exists a population exchange satisfied Fickian conservation law. Berestycki et al. have studied the qualitative properties of road-field system (see [11,12]). Since the fact the motion of mosquitoes and humans obey the Gaussian distribution, we take the effect of the dispersal into account by using a Laplacian diffusion. Moreover, we study the spread of mosquitoes along the highways with the help of the road diffusion. Thus we improve Ross epidemic model as the form:
{∂tu−d1Δu=mab1v(1−u)−γ1u,(x,y)∈R×R+,∂tv−d2Δv=ab2(1−v)u−γ2v,(x,y)∈R×R+,∂tw−D∂xxw=νv|y=0−μw,x∈R,−d2∂yv|y=0=μw−νv|y=0,x∈R,∂yu|y=0=0,x∈R. | (1.2) |
where u(x,y,t) represents the density of the infected humans, v(x,y,t) represents the density of the infected mosquitoes, and w(x,t) represents the density of the infected humans living on the road. Meanwhile, R×R+ and R represent field and road respectively. The first equation accounts for the infected humans dynamic on the field, the second and third for the infected mosquitoes dynamic in the field and road respectively, and the fourth for the exchanges between the field and the road. d1, d2 and D denote the diffusion rates. ν represents the ratio moving from field to road, μ for the ratio moving from road to field. m is the rate of the bitten mosquitoes over the whole mosquito community. a is the number of the humans bitted by a mosquito per unit of time. b1 is the probability of a susceptible hunman becoming infectious after a bite of an infected mosquito. b2 is the probability of a susceptible mosquito becoming infectious after it bites a infectd human. γ1 and γ2 are the infection cycles of the mosquitos and humans respectively. Mosquito and human infectious individual remains infectious on a mean time 1γ1 and 1γ2 respectively. Epidemic models without road diffusion have been applied to study the spreading of infectious disease through the investigation of asymptotic behavior, seeing for example Allen et al. [13,14], Yang et al. [15], Lin and Zhu [16], Li et al. [17], Lei et al. [18], and the references therein. Besides the classical Laplacian diffusion in (1.2), the new generalized fractional derivative has been presented in [19,20,21] to study the memory effect of epidemic model dynamics. Recently, some important real examples of epidemic model dynamics have been proposed in [22,23,24].
This paper is mainly aimed to study the existence and stability of the solutions to system (1.2). Section 2 proves the global existence and uniqueness of solutions to the road diffusion problem (1.2). Section 3 investigates the asymptotic stabilities of the disease-free equilibrium and endemic equilibrium based on the theory of the basic reproduction number. Section 4 gives out some discussions and conclusions.
We use the approach of upper and lower solutions to study the existence and uniqueness of solutions.
Definition 2.1. Supposethat˜u:=(˜u,˜v,˜w) and ˆu:=(ˆu,ˆv,ˆw) is continuous for t∈[0,∞). ˆu is called a lower solution of system (1.2) if it satisfies (1.2) with the = signs replaced by ≤ signs. Similarly, ˜u is called a upper solution satisfies(1.2)withthe = signsreplacedby≥signs .
In order to show the uniqueness, we need the following comparison principle.
Theorem 2.1. (Comparison principle) If nonnegative (˜u,˜v,˜w) and (ˆu,ˆv,ˆw) are respectively a upper solution and lower solution of system (1.2) satisfying ˆu≤˜u≤1,ˆv≤˜v≤1,ˆw≤˜w≤1 at t=0. Then ˆu≤˜u,ˆv≤˜v,ˆw≤˜w for all t>0.
Proof. For K>0, we define functions (u_,v_,w_)=(ˆu,ˆv,ˆw)e−Kt,(¯u,¯v,¯w)=(˜u,˜v,˜w)e−Kt. Let χ:R→R be a nonnegative smooth function satisfying:
χ(x)=0 in x∈[0,2x0],χ″≤min{12,12d1,12d2,1D} in R, | (2.1) |
where χ″ is the second derivative of χ, x0=max{x1,x2}, here x1 and x2 are determined in the following (2.8) and (2.12).
For ε>0, set
ˇu(x,y,t):=¯u(x,y,t)+μνε(χ(|x|)+χ(y)+t+1),ˇv(x,y,t):=¯v(x,y,t)+μνε(χ(|x|)+χ(y)+t+1),ˇw(x,t):=¯w(x,t)+ε(χ(|x|)+t+1). | (2.2) |
It is easy to verify that (ˇu,ˇv,ˇw) is strictly above (u_,v_,w_) at t=0. In order to show ˆu≤˜u,ˆv≤˜v,ˆw≤˜w for all t>0, we need to show that u_<ˇu,v_<ˇv,w_<ˇw for all t>0 owing to the arbitrariness of ε. Assume by contradiction that this property is not true for all t>0. Then,
T=sup{τ≥0,u_≤ˇu∈R×R+×[0,τ],v_≤ˇv∈R×R+×[0,τ],w_≤ˇw∈R×[0,τ]}∈(0,+∞). |
According to the continuity of the functions, ˇu, ˇv and ˇw tend to +∞ as the space variable goes to infinity, uniformly in time, implies that T>0 and ˇu−u_=0 or ˇv−v_=0 or ˇw−w_=0 at time T. By choosing
K=2mab1+2ab2, | (2.3) |
we argue differently depending on ˇu−u_=0 or ˇv−v_=0 or ˇw−w_=0 at time T.
Case 1. In the case minR(ˇw−w_)(⋅,T)=0. In view of the definition of ¯w, we have
∂t¯w−D∂xx¯w+(K+μ)¯w≥ν¯v|y=0. |
Applying (2.2) to the above inequality, we have
(∂t−D∂xx+(K+μ))ˇw=(∂t−D∂xx+(K+μ))¯w+ε(1−Dχ″(|x|)+(K+μ)(χ(|x|)+t+1))≥ν¯v|y=0+ε(1−Dχ″(|x|)+(K+μ)(χ(|x|)+t+1))≥ν¯v|y=0+ε(K+μ)(χ(|x|)+t+1), |
where the last inequality comes from (2.1).
By (2.2), it follows from the above inequality that
(∂t−D∂xx+(K+μ))ˇw≥νˇv|y=0+ε(K+μ)(χ(|x|)+t+1)−με(χ(|x|)+t+1)≥νˇv|y=0. | (2.4) |
On the other hand, we have
(∂t−D∂xx+(K+μ))w_≤νv_|y=0. | (2.5) |
Combining (2.4) and (2.5), for x∈R and t∈(0,T], we have
(∂t−D∂xx+(K+μ))(ˇw−w_)≥νˇv|y=0−νv_|y=0≥0. |
Using the parabolic strong maximum principle, we can yield that minR(ˇw−w_)=0 in R×[0,T], which is contrary to ˇw(x,0)−w_(x,0)>0. Thus minR(ˇw−w_)(⋅,T)=0 is not true.
Case 2. In the case minR×R+(ˇu−¯u)(⋅,T)=0. By the definition of u_ and ¯u, we have
(∂t−d1Δ+K)u_≤e−Kt(mab1ˆv(1−ˆu)−γ1ˆu),(∂t−d1Δ+K)¯u≥e−Kt(mab1˜v(1−˜u)−γ1˜u). | (2.6) |
By (2.2), it follows from (2.6) that
(∂t−d1Δ+K)ˇu=(∂t−d1Δ+K)¯u+μνε(1−d1χ″(|x|)−d1χ″(y)+K(χ(|x|)+χ(y)+t+1))≥(∂t−d1Δ+K)¯u+μνεK(χ(|x|)+χ(y)+t+1)≥e−Kt(mab1˜v(1−˜u)−γ1˜u)+μνεK(χ(|x|)+χ(y)+t+1), |
where the second inequality is induced from (2.1).
Thus we have
(∂t−d1Δ+K)(ˇu−u_)≥μνεK(χ(|x|)+χ(y)+t+1)+e−Kt[mab1˜v(1−˜u)−γ1˜u−(mab1ˆv(1−ˆu)−γ1ˆu)],=μνεK(χ(|x|)+χ(y)+t+1)+e−Kt(−γ1−mab1˜v)(˜u−ˆu)+e−Kt(mab1−mab1ˆu)(˜v−ˆv),=μνεK(χ(|x|)+χ(y)+t+1)+(−γ1−mab1˜v)(¯u−u_)+(mab1−mab1ˆu)(¯v−v_),≥(−γ1−mab1˜v)(ˇu−u_)+(mab1−mab1ˆu)(ˇv−v_), |
where the last inequality is due to the definition of K from (2.3).
Owing to ˆu≤1, we have
(∂t−d1Δ+K+γ1+mab1˜v)(ˇu−u_)≥(mab1−mab1ˆu)(ˇv−v_)≥0, | (2.7) |
for (x,y)∈R×R+ and t∈(0,T]. By applying the parabolic strong maximum principle, minR×R+(ˇu−u_)(⋅,T)=0 must be attained on the boundary of R×R+. Thus we have
(ˇu−u_)(x1,0,T)=0, for some x1∈R. | (2.8) |
Moreover, we have
(ˇu−u_)(x1,0,T)≥(¯u−u_)(x1,0,T)=e−Kt(˜u−ˆu)(x1,0,T)≥e−Ktν[∂y(˜u(x1,0,T)−ˆu(x1,0,T))+μ(˜w(x1,0,T)−ˆw(x1,0,T)]≥e−Ktν∂y(˜u(x1,0,T)−ˆu(x1,0,T))=1ν∂y(¯u(x1,0,T)−u_(x1,0,T))=1ν∂y(ˇu(x1,0,T)−u_(x1,0,T))>0, | (2.9) |
where the last inequality is by using the Hopf lemma.
This contradiction of (2.8) and (2.9) implies that minR×R+(ˇu−u_)(⋅,T)=0 is not true.
Case 3. In the case minR×R+(ˇv−v_)(⋅,T)=0. Directly calculations show that
(∂t−d2Δ+K)v_≤e−Kt(ab2(1−ˆv)ˆu−γ2ˆv),(∂t−d2Δ+K)¯v≥e−Kt(ab2(1−˜v)˜u−γ2˜v). | (2.10) |
By (2.2), it follows from (2.10) that
(∂t−d2Δ+K)ˇv=(∂t−d2Δ+K)¯v+μνε(1−d2χ″(|x|)−d2χ″(y)+K(χ(|x|)+χ(y)+t+1))≥(∂t−d2Δ+K)¯v+μνεK(χ(|x|)+χ(y)+t+1)≥e−Kt(ab2(1−˜v)˜u−γ2˜v)+μνεK(χ(|x|)+χ(y)+t+1), |
where the second inequality is induced from (2.1). Inserting (2.10) into the above inequality, we have
(∂t−d2Δ+K)(ˇv−v_)≥μνεK(χ(|x|)+χ(y)+t+1)+e−Kt[ab2(1−˜v)˜u−γ2˜v−(ab2(1−ˆv)ˆu−γ2ˆv)]=μνεK(χ(|x|)+χ(y)+t+1)+e−Kt[(−γ2−ab2˜u)(˜v−ˆv)+(ab2−ab2ˆv)(˜u−ˆu)]=μνεK(χ(|x|)+χ(y)+t+1)+[(−γ2−ab2˜u)(¯v−v_)+(ab2−ab2ˆv)(¯u−u_)]≥(−γ2−ab2˜u)(ˇv−v_)+(ab2−ab2ˆv)(ˇu−u_), |
where the last inequality is because of the definition of K from (2.3).
Owing to ˆv≤1, we have
(∂t−d2Δ+K+γ2+ab2˜u)(ˇv−v_)≥(ab2−ab2ˆv)(ˇu−u_)≥0, | (2.11) |
for (x,y)∈R×R+ and t∈(0,T]. By applying the parabolic strong maximum principle, minR×R+(ˇv−v_)(⋅,T)=0 must be attained on the boundary of R×R+. Thus we have
(ˇv−v_)(x2,0,T)=0, for some x2∈R. | (2.12) |
Using the Hopf lemma yields that
∂y(ˇv−v_)|y=0(x2,0,T)>0, for some x2∈R. | (2.13) |
In view of (2.2) and (2.12), we have
(˜v−ˆv)(x2,0,T)=eKT(¯v−v_)(x2,0,T),=eKT(ˇv−v_)(x2,0,T)−μeKTνε(χ(|x2|)+T+1). | (2.14) |
In view of (2.2), we have
∂y|y=0(˜v−ˆv)(x2,0,T)=eKT∂y|y=0(ˇv−v_)(x2,0,T)−μeKTεν∂y|y=0(χ(|x2|)+χ(y)+T+1),=eKT∂y|y=0(ˇv−v_)(x2,0,T)>0, | (2.15) |
where the last inequality is because of the definition of χ in (2.1). The conclusion (2.15) contradicts to ∂y˜v|y=0≤0≤∂yˆv|y=0. Thus minR×R+(ˇv−v_)(⋅,T)=0 is not true.
In the above three cases, we all reached a contradiction. This completes the proof.
As for a given pair of coupled upper and lower solutions ˜u and ˆu, we denote
Λ:={u∈C(E∗)×C(E∗)×C(E∗∗):ˆu≤u≤˜u,ˆv≤v≤˜v,ˆw≤w≤˜w}. | (2.16) |
where
E∗:={(t,x,y):t∈(0,∞),(x,y)∈R×R+},E∗∗:={(t,x):t∈(0,∞),x∈R}. | (2.17) |
We denote the functions of system (1.2) by
f1(u,v):=mab1v(1−u)−γ1u,f2(u,v):=ab2(1−v)u−γ2v,g(v,w):=νv|y=0−μw. | (2.18) |
We denote the linear operators of system (1.2) by
L1(u):=∂tu−d1Δu,L2(v):=∂tv−d2Δv,L3(w):=∂tw−D∂xxw,L4(v):=−d2∂yv|y=0,L5(u):=−∂yu|y=0. | (2.19) |
By using u_(0)=ˆu and ¯u(0)=˜u as the initial iterations we can construct sequences {u_(m)}∞m=1 and {¯u(m)}∞m=1 satisfying the same initial functions from the iteration process of scalar equations
{L1(¯u(m))+K1¯u(m)=f1(¯u(m−1),¯v(m−1))+K1¯u(m−1),inE∗,L2(¯v(m))+K2¯v(m)=f2(¯u(m−1),¯v(m−1))+K2¯v(m−1),inE∗,L3(¯w(m))+μ¯w(m)=g(¯v(m−1),¯w(m−1))+μ¯w(m−1),inE∗∗,(L4+ν)(¯v(m))=−g(¯v(m−1),¯w(m−1))+ν¯v(m−1),inE∗∗,L5(¯u(m))=0,inE∗∗. | (2.20) |
Here we choose
K1=γ1+mab1,K2=γ2+ab2. | (2.21) |
{u_(m)}∞m=1 satisfies the above equation with the superscripts replaced by subscripts. Indeed, (2.20) is reduced to a linear parabolic equation with half-space homogeneous Neumamm condition with respect to u and v, and a Cauchy problem of linear parabolic equation with respect to w. Thus {¯u(m)}∞m=1 and {u_(m)}∞m=1 are well-defined. By using the monotone dynamical system method (Smith [25]), we have the following lemma.
Lemma 2.1. The sequences {¯u(m)}∞m=1 and {u_(m)}∞m=1 governed by (2.20) possess the monotonicity property
ˆu≤u_(m)≤u_(m+1)≤¯u(m+1)≤¯u(m)≤˜uform=1,2,⋯ | (2.22) |
Moreover, for each m=1,2,⋯, ¯u(m) and u_(m) are coupled upper and lower solutions of (1.2).
Proof. Let s_(1)=u_(1)−u_(0). We necessarily apply the comparison principle of standard parabolic equation to s_(1). By (2.21), s_(1) satisfies
{L1(s_(1))+K1s_(1)=f1(u_(0),v_(0))+K1u_(0)−(L1(u_(0))+K1u_(0))=−L1(ˆu)+f1(ˆu,ˆv)≥0,inE∗,L5(s_(1))=0,inE∗∗,s_(1)(0,x,y)=0,inR×R+. |
Using comparison principle of standard parabolic equation yields s_(1)≥0. It follows that u_(0)≤u_(1). On the other hand, we set q_(1)=v_(1)−v_(0). By (2.20), q_(1) satisfies
{L2(q_(1))+K2q_(1)=f2(u_(0),v_(0))+K2u_(0)−(L2(v_(0))+K2v_(0))=−L2(ˆu)+f2(ˆu,ˆv)≥0,inE∗,(L4+ν)(q_(1))=μw_(0)−νv_(0)|y=0+νv_(0)y=0−L4ˆv−νˆv|y=0≥0,inE∗∗,q_(1)(0,x,y)=0,inR×R+. |
the parabolic comparison yields v_(0)≤v_(1). Meanwhile, using the similar process yields w_(0)≤w_(1). Thus, we have u_(0)≤u_(1). Likewise, we have ¯u(0)≥¯u(1).
Letting z(1)=¯u(1)−u_(1), it follows from (2.21) that
{L1(z(1))+K1z(1)=f1(¯u(0),¯v(0))+K1¯u(0)−(f1(u_(0),v_(0))+K1u_(0))≥0,inE∗,(L4+ν)(z(1))=μ(¯w(1)−w_(1))−ν(¯u(1)−u_(1))|y=0+ν(¯u(1)−u_(1))|y=0≥0,inE∗∗,z(1)(0,x,y)=0,inR×R+. |
It follows again from comparison principle that ¯u(1)≥u_(1), and thus we have v_(1)≤¯v(1). The above conclusions show that
u_(0)≤u_(1)≤¯u(1)≤¯u(0). | (2.23) |
Now we show that ¯u(1) and u_(1) are upper and lower solutions of (1.2).
Next we use an induction method. By choosing ¯u(1) and u_(1) as the coupled upper and lower solutions ˜u and ˆu, after a similar argument as above, we have
u_(1)≤u_(2)≤¯u(2)≤¯u(1), | (2.24) |
so ¯u(2) and u_(2) are coupled upper and lower solutions of (1.2). The conclusion of the lemma follows from the induction principle.
In view of Lemma 2.1, the pointwise limits
limm→∞¯u(m)=¯u,limm→∞u_(m)=u_ | (2.25) |
exist. In the following theorem we show that (¯u,¯v,¯w) and (u_,v_,w_) are respectively the maximal and minimal solutions of system (1.2).
Theorem 2.2. Assume that ˜u and ˆu be upper and lower solutions of system (1.2). Let (¯u,¯v,¯w) and (u_,v_,w_) be given by (2.25). Then (¯u,¯v,¯w) and (u_,v_,w_) are the solutions of system (1.2). Moreover,
ˆu≤u_(m)≤u_(m+1)≤u_=¯u≤¯u(m+1)≤¯u(m)≤˜u. | (2.26) |
Proof. (ⅰ) To show that the limit (¯u,¯v,¯w) in (2.25) is the solution of (1.2), we use an integral representation for the solution of the Cauchy problem and the linear parabolic Eq (2.20) under Neumann boundary condition. Let G(t,x,y;τ,ξ1,ξ2) be the Green function given by
G1(t,x,y;τ,ξ1,ξ2):=e−K1t4πd1(t−τ)(e−(x−ξ1)2+(y−ξ2)24d1(t−τ)+e−(x−ξ1)2+(−y−ξ2)24d1(t−τ)),G2(t,x,y;τ,ξ1,ξ2):=e−K2t4πd2(t−τ)(e−(x−ξ1)2+(y−ξ2)24d2(t−τ)+e−(x−ξ1)2+(−y−ξ2)24d2(t−τ)),G3(t,x;τ,ξ):=e−μt2√πD(t−τ)e−(x−ξ)24D(t−τ) |
We denote
I1(t,x,y):=∫R×R+G1(t,x,y;0,ξ1,ξ2)u0(ξ1,ξ2)dξ1dξ2I2(t,x,y):=∫R×R+G2(t,x,y;0,ξ1,ξ2)v0(ξ1,ξ2)dξ1dξ2I3(t,x):=∫RG3(t,x;0,ξ)w0(ξ)dξ. |
By the integral representation of linear parabolic boundary value problems [26], the solution of (2.20) is given by
¯u(m)(t,x,y)=I1(t,x,y)+∫t0dτ∫R×R+G1(t,x,y;τ,ξ1,ξ2)(f1(¯u(m−1),¯v(m−1))+K1¯u(m−1))dξ1dξ2¯v(m)(t,x,y)=I2(t,x,y)+∫t0dτ∫R×R+G2(t,x,y;τ,ξ1,ξ2)(f2(¯u(m−1),¯v(m−1))+K2¯v(m−1))dξ1dξ2−1d2∫t0dτ∫RG2(t,x,0;τ,ξ1,0)(−g(¯v(m−1),¯w(m−1))+ν(¯v(m−1)−¯v(m)))dξ1¯w(m)(t,x)=I3(t,x)+∫t0dτ∫RG3(t,x;τ,ξ)(g(¯v(m−1),¯w(m−1))+μ¯w(m−1))dξ |
The integrands in the above equation are integrable because the definition of Green function G satisfies the admissible growth in the time variable. Using the dominated convergence theorem yields that the limits (¯u,¯v,¯w) satisfy the relation
¯u(t,x,y)=I1(t,x,y)+∫t0dτ∫R×R+G1(t,x,y;0,ξ1,ξ2)(f1(¯u,¯v)+K1¯u)dξ1dξ2¯v(t,x,y)=I2(t,x,y)+∫t0dτ∫R×R+G2(t,x,y;0,ξ1,ξ2)(f2(¯u,¯v)+K2¯v)dξ1dξ2−1d2∫t0dτ∫RG2(t,x,0;τ,ξ1,0)(−g(¯v,¯w))dξ1¯w(t,x)=I3(t,x)+∫t0dτ∫RG3(t,x;0,ξ)(g(¯v,¯w)+μ¯w)dξ | (2.27) |
Then (¯u,¯v,¯w) is a solution of (1.2). A similar argument shows that (u_,v_,w_) is also a solution of (1.2).
(ⅱ) We denote by
E∗T:={(t,x,y):t∈(0,T],(x,y)∈R×R+},E∗∗T:={(t,x):t∈(0,T],x∈R}. |
By plugging (2.18) into (2.27), we obtain
¯w(t,x)=I3(t,x)+∫t0dτ∫RG3(t,x;τ,ξ)ν¯vdξ,w_(t,x)=I3(t,x)+∫t0dτ∫RG3(t,x;τ,ξ)νv_dξ. |
In view of G3, we have
(¯w−w_)=ν∫t0dτ∫RG3(t,x;τ,ξ)(¯v−v_)dξ≤ν||¯v−v_||L∞(Et)∫t0e−μtdτ∫R12√πD(t−τ)e−(x−ξ)24D(t−τ)dξ=νμ(1−e−μt)||¯v−v_||L∞(E∗t). | (2.28) |
We also obtain
¯u(t,x,y)=I1(t,x,y)+∫t0dτ∫R×R+G1(t,x,y;τ,ξ1,ξ2)(f1(¯u,¯v)+K1¯u)dξ1dξ2u_(t,x,y)=I1(t,x,y)+∫t0dτ∫R×R+G1(t,x,y;τ,ξ1,ξ2)(f1(u_,v_)+K1u_)dξ1dξ2 |
Since the definition of K1 in (2.16), we have |f1(¯u,v_)+K1¯u−f1(u_,¯v)−K1u_|≤2K1(|¯u−u_|+|¯v−v_|). Recalling G2, we have
(¯u−u_)≤2K1∫t0dτ∫R×R+G1(t,x,y;τ,ξ1,ξ2)(|¯u−u_|+|¯v−v_|)dξ1dξ2≤2K1(||¯u−u_||L∞(Et)+||¯v−v_||L∞(Et))∫t0e−K1tdτ∫R×R+14πd1(t−τ)(e−(x−ξ1)2+(y−ξ2)24d1(t−τ)+e−(x−ξ1)2+(−y−ξ2)24d1(t−τ))dξ1dξ2=2(1−e−K1t)(||¯u−u_||L∞(E∗t)+||¯v−v_||L∞(E∗t)). | (2.29) |
By (2.27), we have
¯v(t,x,y)=I2(t,x,y)+∫t0dτ∫R×R+G2(t,x,y;τ,ξ1,ξ2)(f2(¯u,¯v)+K2¯v)dξ1dξ2−1d2∫t0dτ∫RG2(t,x,0;τ,ξ1,0)(−g(¯v,¯w))dξ1v_(t,x,y)=I2(t,x,y)+∫t0dτ∫R×R+G2(t,x,y;τ,ξ1,ξ2)(f2(u_,v_)+K2v_)dξ1dξ2−1d2∫t0dτ∫RG2(t,x,0;τ,ξ1,0)(−g(v_,w_))dξ1 |
Since the definition of K2 in (2.21), we have |f2(¯u,¯v)+K2¯v−f2(u_,v_)−K2v_|≤2K2(|¯u−u_|+|¯v−v_|). We also have |g(¯u,¯w)−g(u_,w_)|≤(μ+ν)(|¯u−u_|+|¯w−w_|). Recalling G2, we have
(¯v−v_)≤2(1−e−K2t)(||¯v−v_||L∞(E∗t)+||¯v−v_||L∞(E∗t))+1d2(μ+ν)(||¯v−v_||L∞(E∗t)+||¯w−w_||L∞(E∗∗t))∫t0dτ∫RG2(t,x,0;τ,ξ1,0)dξ1≤2(1−e−K2t)(||¯v−v_||L∞(E∗t)+||¯v−v_||L∞(E∗t))+1d2(μ+ν)(||¯v−v_||L∞(E∗t)+||¯w−w_||L∞(E∗∗t))1−e−K2tK2 | (2.30) |
Combining (2.28), (2.29) and (2.30), by choosing K:=max{νμ,μ+νd2K2+2} and γ=max{K1,K2,μ}, we obtain
(||¯u−u_||L∞(E∗t)+||¯v−v_||L∞(E∗t)+||¯w−w_||L∞(E∗∗t))≤K(1−e−γt)(||¯u−u_||L∞(E∗t)+||¯v−v_||L∞(E∗t)+||¯w−w_||L∞(E∗∗t)),fort∈(0,∞). | (2.31) |
Thus there exist a constant T:=lnKK−1 such that ||¯u−u_||L∞(E∗t)+||¯v−v_||L∞(E∗t)+||¯w−w_||L∞(E∗∗t)=0, that is ¯u≡u_, ¯v≡v_, and ¯w≡w_ for t∈(0,T]. Owing to the fact the above γ and K do not depend on the initial value, T can be extended to the time ∞.
Notice that the above theorem is also valid when the upper and the lower solutions are constant vectors. Moreover we can induce the global asymptotic convergence of system (1.2). Suppose that there are constant vectors ˜c:=(˜c1,˜c2,˜c3) and ˆc=(ˆc1,ˆc2,ˆc3) such that fi(˜c1,˜c2)≤0 and fi(ˆc1,ˆc2)≥0 for i=1,2. Here ˜c3=νμ˜c2, ˆc3=νμˆc2, and ˆci≤˜ci for i=1,2,3. In view of Definition 2.1, ˜c and ˆc are upper and lower solutions of system (1.2). Thus using ˜c and ˆc as initial iteration, after the procedure (2.20), we can obtain the maximal solution ¯c:=(¯c1,¯c2,¯c3) and minimal solution c_:=(c_1,c_2,c_3). Since the solution of system (2.20) is unique, the sequences {¯c(m)i} and {c_(m)i} are constants. The limits of {¯c(m)i} and {c_(m)i} satisfy
{f1(c_1,c_2)=0=f1(¯c1,¯c2),f2(c_1,c_2)=0=f2(¯c1,¯c2),¯c3=νμ¯c2,c_3=νμc_2. | (2.32) |
We give the convergence result in the next theorem.
Theorem 2.3. Suppose that the initial functions of system (1.2) satisfy ˆc1≤u0(x,y)≤˜c1, ˆc2≤v0(x,y)≤˜c2, and ˆc3≤w0(x)≤˜c3. If ¯ci=c_i:=ci for i=1,2,3, then for (x,y)∈R×R+,
limt→∞u(t,x,y)=c1,limt→∞v(t,x,y)=c2,limt→∞w(t,x)=c3. |
Proof. We denote (˜u1,˜u2,˜u3) by the solution of the following system
{∂tu−d1Δu=f1(u,v),t>0,(x,y)∈R×R+,∂tv−d2Δv=f2(u,v),t>0,(x,y)∈R×R+,∂tw−D∂xxw=νu|y=0−μw,t>0,x∈R,∂yu|y=0=0,t>0,x∈R,−d2∂yv|y=0=μw−νv|y=0,t>0,x∈R,u(0,x,y)=˜c1,v(0,x,y)=˜c2,(x,y)∈R×R+,w(0,x)=˜c3,x∈R. | (2.33) |
It is easy to see that ¯c is a lower solution of (2.33). Then we have ¯ci≤˜ui for i=1,2,3. By using Theorem 2.1, ˜ui is time-nonincreasing. The limit of ˜ui exists as t→∞. We denote by limt→∞˜ui(t,⋅)=¯ui(⋅). Thanks to the Schauder estimates, ¯ui is the stationary solution of system (1.2). That is to say, (¯u1,¯u2,¯u3) satisfies
{−d1Δu=f1(u,v),(x,y)∈R×R+,−d2Δv=f2(u,v),(x,y)∈R×R+,−D∂xxw=νu|y=0−μw,x∈R,∂yu|y=0=0,x∈R,−d2∂yv|y=0=μw−νv|y=0,x∈R. | (2.34) |
Meanwhile, ¯c is also a stationary solution of system (1.2). Moreover, in view of the iterative process of ¯c, (¯c(1)1,¯c(1)2,¯c(1)3) satisfies
{−d1Δ¯c(1)1=f1(˜c1,˜c2)≥f1(¯u1,¯u2),(x,y)∈R×R+,−d2Δ¯c(1)2=f2(˜c1,˜c2)≥f2(¯u1,¯u2),(x,y)∈R×R+,(−D∂xx+μ)¯c(1)3=ν˜c2|y=0≥ν¯u2|y=0,x∈R,∂yc(1)1|y=0=0≥∂y¯u1|y=0,x∈R,(−d2∂y+ν)c(1)2|y=0=μ˜c3≥μ¯u3,x∈R, | (2.35) |
where the righthand inequality sign is due to ¯ui≤˜ci for i=1,2,3. By using comparison principle (Theorem 2.1), we induce that ¯c(1)i≥¯ui for i=1,2,3. By using an induction method, (¯c(2)1,¯c(2)2,¯c(2)3) satisfies
{−d1Δ¯c(2)1=f1(¯c(1)1,¯c(1)2)≥f1(¯u1,¯u2),(x,y)∈R×R+,−d2Δ¯c(2)2=f2(¯c(1)1,¯c(1)2)≥f2(¯u1,¯u2),(x,y)∈R×R+,(−D∂xx+μ)¯c(2)3=ν¯c(1)2|y=0≥ν¯u2|y=0,x∈R,∂yc(2)1|y=0=0≥∂y¯u1|y=0,x∈R,(−d2∂y+ν)c(2)2|y=0=μ¯c(1)3≥μ¯u3,x∈R. | (2.36) |
By using comparison principle (Theorem 2.1), we induce that ¯c(2)i≥¯ui for i=1,2,3. After an induction argument, we have ¯c(m)i≥¯ui for i=1,2,3. Here m=1,2,⋯. By letting m→∞, we have
¯ci≥¯ui=limt→∞˜ui(t,⋅), for i=1,2,3. | (2.37) |
On the other hand, We denote (ˆu1,ˆu2,ˆu3) by the solution of the following system
{∂tu−d1Δu=f1(u,v),t>0,(x,y)∈R×R+,∂tv−d2Δv=f2(u,v),t>0,(x,y)∈R×R+,∂tw−D∂xxw=νu|y=0−μw,t>0,x∈R,∂yu|y=0=0,t>0,x∈R,−d2∂yv|y=0=μw−νv|y=0,t>0,x∈R,u(0,x,y)=ˆc1,v(0,x,y)=ˆc2,(x,y)∈R×R+,w(0,x)=ˆc3,x∈R. | (2.38) |
In a similar way, we can induce
c_i≤u_i=limt→∞ˆui(t,⋅), for i=1,2,3. | (2.39) |
Since ¯ci=c_i:=ci for i=1,2,3, it follows from (2.38) and (2.39) that
limt→∞˜ui(t,⋅)=limt→∞ˆui(t,⋅)=ci, for i=1,2,3. | (2.40) |
For any initial functions satisfying ˆc1≤u0(x,y)≤˜c1, ˆc2≤v0(x,y)≤˜c2, and ˆc3≤w0(x)≤˜c3, the solution (u,v,w) is a lower solution of system (2.33) and a upper solution of system (2.38). By using comparison principle (Theorem 2.1), we have (ˆu1,ˆu2,ˆu3)≤(u,v,w)≤(˜u1,˜u2,˜u3). Thus we induce that
limt→∞(u,v,w)=(c1,c2,c3). |
Theorem 2.4. If the initial functions of system (1.2) satisfy 0≤u0(x,y)≤1, 0≤v0(x,y)≤1, and 0≤w0(x)≤νμ. then system (1.2) possesses a unique solution (u,v,w) for t∈(0,∞). Moreover, for (x,y)∈R×R+,
0≤u(t,x,y)≤1,0≤v(t,x,y)≤1,0≤w(t,x)≤νμ. | (2.41) |
Proof. In order to utilize Theorem 2.2, we need to construct upper and lower solutions of system (1.2). We set
M1=1,M2=1,M3=νμ. |
It is easy to verify that (M1,M2,M3) and (0,0,0) are upper and lower solutions of system (1.2). Using Theorem 2.2, system (1.2) has a unique solution. Moreover we can obtain (2.41).
Based on the basic reproduction R0:=ma2b1b2γ1γ2, we examine the asymptotic stability of the disease-free equilibrium (0,0,0) and endemic equilibrium.
Theorem 3.1. Suppose that the initial functions of system (1.2) satisfy 0≤u0(x,y)≤1, 0≤v0(x,y)≤1 and 0≤w0(x)≤νμ, if R0<1, then the solution to system (1.2) satisfies
limt→∞(u,v,w)=(0,0,0). | (3.1) |
Proof. We use the method of upper and lower solutions. we set
M1=1,M2=1,M3=νμ. |
It is easy to verify that (M1,M2,M3) and (0,0,0) are upper and lower solutions of system (1.2). By using Theorem ???, we can construct the maximal solution ¯c and minimal solution c_, which satisfy
{mab1¯c2(1−¯c1)−γ1¯c1=0=mab1c_2(1−c_1)−γ1c_1,ab2(1−¯c2)¯c1−γ2¯c2=0=ab2(1−c_2)c_1−γ2c_2,¯c3=νμ¯c2,c_3=νμc_2. |
By a directly computation, we have
¯c1¯c2[ma2b1b2γ1γ2(1−¯c1)(1−¯c2)−1]=0. |
Since R0<1, we induce that ¯c1¯c2=0. Combing with mab1¯c2(1−¯c1)−γ1¯c1=0, we have
¯c1=¯c2=¯c3=0. | (3.2) |
On the other hand, a similar argument yields that
c_1=c_2=c_3=0. | (3.3) |
It follows from (3.2) and (3.3) that
c_i=¯ci=0 for i=1,2,3. | (3.4) |
By applying Theorem 2.3, we have
limt→∞(u,v,w)=(c_1,c_2,c_3)=(0,0,0). |
Theorem 3.2. Suppose that δ is an arbitrary small positive constant such that δ≤u0(x,y)≤1, δ≤v0(x,y)≤1 and δ≤w0(x,y)≤νμ. If R0>1, then the solution to system (1.2) satisfies
limt→∞(u,v,w)=(u∗,v∗,w∗), | (3.5) |
where (u∗,v∗) is a unique positive solution of
{f1(u∗,v∗)=0,f2(u∗,v∗)=0. | (3.6) |
and w∗=νμv∗.
Proof. We first consider the nullclines of system (1.2), which is written by
{v=γ1umab1(1−u),u=γ2vab2(1−v). | (3.7) |
By taking the derivatives of (3.6), we have
{dvdu=γ1mab1(1−u)2>0,dudv=γ2ab2(1−v)2>0. | (3.8) |
Hence the two nullclines are monotone increasing. In the domain (u,v)∈[0,1]×[0,∞), they have the same start point (0,0), but the different end points (1,∞) and (1,ab2ab2+γ2). The nullclines has a unique positive cross points (u∗,v∗).
Next we use the method of upper and lower solutions. we set
M1=1,M2=1,M3=νμ. |
It is easy to verify that (M1,M2,M3) and (δ,δ,δ) are upper and lower solutions of system (1.2). By using Theorem 2.3, we can construct the maximal solution ¯c and minimal solution c_, which satisfy ¯c≥c_i>0 for i=1,2,3. Moreover, (¯c1,¯c2) and (c_1,c_2) are the positive solutions of the following nullclines
{v=γ1umab1(1−u),u=γ2vab2(1−v). |
Since the nullclines has a unique positive cross points (u∗,v∗), we obtain that ¯c1=c_1=u∗, ¯c2=c_2=v∗. Then ¯c3=c_3=w∗.
By applying Theorem 2.3, we have
limt→∞(u,v,w)=(u∗,v∗,w∗). |
Since mosquito is a prominent vector of Malaria, it is of crucial importance to study the qualitative spreading behavior of mosquitoes in implementing vector control strategies and preventing mosquito-borne diseases. A large proportion of the current studies on epidemic transmission dynamics, using ordinary differential systems, focuses on the temporal development and control of infectious diseases, thus uncertainty still exists concerning how population mobility affects on epidemic outbreaks. The spatial factor should be taken into consideration in the modeling processes in order to study the geographic spread of infectious diseases. Laplacian diffusion systems with spatially homogeneous parameters (Murray [27], Ruan [28]) and spatially heterogeneous parameters (Allen et al. [14], Lei et al. [18], Li et al. [17], Song et al. [1]) have been proposed to study the spatio-temporal dynamics epidemic models. These studies exhibit an equal probability of mosquitoes' movement to any direction. Yet our road diffusion model is different since mosquitoes move at a faster speed along the highway. We studied the long-time dynamical behaviors of Ross epidemic model. To the best of our knowledge, the road diffusion has never been applied to describe the two compartments spread of infectious diseases in any epidemic model in the literature.
Our results indicate that mosquitoes has an impact on long-time dynamics of infectious diseases in road diffusion models. According to Theorem 3.1, when the basic reproduction number R0<1, system (1.2) admits a globally asymptotically stable disease-free equilibrium, and an asymptotic stable endemic equilibrium when the basic reproduction number R0>1 in view of Theorem 3.2. So we have generalized the threshold dynamics of the classical Ross model to that in a road diffusion model. Although many factors related to a real disease have been simplified in our model, we capture the dynamics of mosquito-borne diseases, which helps us to control and prevent the spread of diseases.
Since Ronald Ross discovers the transmission of Malaria by mosquitoes, the differential equations have been used to study the spread of infectious disease (see [29,30,31]). A natural question is how to describe the dynamical system when the infectious disease spreads along the directed diffusion. We introduce the road-field diffusion into a Ross epidemic model which describes the dynamical behaviors of infected Mosquitoes and humans. Our result reveals that the disease-free equilibrium is asymptotically stable if the basic reproduction number is lower than 1 while the endemic equilibrium asymptotically stable if the basic reproduction number is greater than 1.
The authors would like to thank the anonymous referees for the helpful comments and suggestions. Canrong Tian was was partially supported by NSFC grant (NSFC-61877052), the Jiangsu Province 333 Talent Project, and Jiangsu Province Qinglan Project.
The authors declare that they have no conflict of interest.
[1] |
P. Song, Y. Lou, Y. Xiao, A spatial SEIRS reaction-diffusion model in heterogeneous environment, J. Differ. Equ., 267 (2019), 5084–5114. doi: 10.1016/j.jde.2019.05.022. doi: 10.1016/j.jde.2019.05.022
![]() |
[2] |
B. E. Peacock, J. P. Smith, P. G. Gregory, T. M. Loyless, M. J. Jr, P. R. Simmonds, et al., Aedes albopictus in Florida, J. Am. Mosquito Contr., 4 (1988), 362–365. doi: 10.2307/3495451. doi: 10.2307/3495451
![]() |
[3] |
G. F. O'Meara, L. F. Evans, A. D. Gettman, J. P. Cuda, Spread of Aedes albopictus and decline of Ae. aegypti (Diptera: Culicidae) in Florida, J. Med. Entomol., 32 (1995), 554–562. doi: 10.1093/jmedent/32.4.554. doi: 10.1093/jmedent/32.4.554
![]() |
[4] |
W. A. Hawley, P. Reiter, R. Copeland, C. B. Pumpuni, G. B. Craig, Aedes albopictus in North America: Probable introduction in used tires from Northern Asia, Science, 236 (1987), 1114–1116. doi: 10.1126/science.3576225. doi: 10.1126/science.3576225
![]() |
[5] |
M. Enserink, Entomology: A mosquito goes global, Science, 320 (2008), 864–866. doi: 10.1126/science.320.5878.864. doi: 10.1126/science.320.5878.864
![]() |
[6] |
M. B. Hahn, L. Eisen, J. Mcallister, H. M. Savage, J. P. Mutebi, R. J. Eisen, Updated reported distribution of Aedes (Stegomyia) aegypti and Aedes (Stegomyia) albopictus (Diptera: Culicidae) in the United States, 1995–2016, J. Med. Entomol., 54 (2017), 1420–1424. doi: 10.1093/jme/tjx088. doi: 10.1093/jme/tjx088
![]() |
[7] |
I. Rochlin, D. V. Ninivaggi, M. L. Hutchison, Climate change and range expansion of the Asian tiger mosquito (Aedes albopictus) in Northeastern USA: Implications for public health practitioners, Plos One, 8 (2013). doi: 10.1371/journal.pone.0060874. doi: 10.1371/journal.pone.0060874
![]() |
[8] |
C. Parker, D. Ramirez, C. R. Connelly, State-wide survey of Aedes aegypti and Aedes albopictus (Diptera: Culicidae) in Florida, J. Vector Ecol., 44 (2019), 210–215. doi: 10.1111/jvec.12351. doi: 10.1111/jvec.12351
![]() |
[9] |
K. L. Bennett, C. G. Martínez, A. Almanza, J. R. Rovira, J. R. Loaiza, High infestation of invasive Aedes mosquitoes in used tires along the local transport network of Panama, Parasite. Vector., 12 (2019), 264. doi: 10.1186/s13071-019-3522-8. doi: 10.1186/s13071-019-3522-8
![]() |
[10] |
H. Berestycki, J. M. Roquejoffre, L. Rossi, The influence of a line with fast diffusion on Fisher-KPP propagation, J. Math. Biol., 66 (2013), 743–766. doi: 10.1007/s00285-012-0604-z. doi: 10.1007/s00285-012-0604-z
![]() |
[11] |
H. Berestycki, J. M. Roquejoffre, L. Rossi, Fisher-KPP propagation in the presence of a line: Further effects, Nonlinearity, 26 (2013), 2623–2640. doi: 10.1088/0951-7715/26/9/2623. doi: 10.1088/0951-7715/26/9/2623
![]() |
[12] |
H. Berestycki, R. Monneau, J. Scheinkman, A non-local free boundary problem arising in a theory of financial bubbles, Philos. T. R. Soc. A, 372 (2014), 20130404. doi: 10.1098/rsta.2013.0404. doi: 10.1098/rsta.2013.0404
![]() |
[13] |
L. Allen, B. Bolker, Y. Lou, A. Nevai, Asymptotic profiles of the steady states for an SIS epidemic disease patch model, SIAM J. Appl. Math., 67 (2007), 1283–1309. doi: 10.1137/060672522. doi: 10.1137/060672522
![]() |
[14] |
L. Allen, B. Bolker, Y. Lou, A. Nevai, Asymptotic profiles of the steady states for an SIS epidemic reaction-diffusion model, Discrete Cont. Dyn. Syst., 21 (2008), 1–20. doi: 10.1016/S0927-6505(02)00196-2. doi: 10.1016/S0927-6505(02)00196-2
![]() |
[15] |
F. Yang, W. Li, S. Ruan, Dynamics of a nonlocal dispersal SIS epidemic model with Neumann boundary conditions, J. Differ. Equ., 267 (2019), 2011–2051. doi: 10.1016/j.jde.2019.03.001. doi: 10.1016/j.jde.2019.03.001
![]() |
[16] |
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. doi: 10.1007/s00285-017-1124-7. doi: 10.1007/s00285-017-1124-7
![]() |
[17] |
H. Li, R. Peng, Z. 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. doi: 10.1137/18M1167863. doi: 10.1137/18M1167863
![]() |
[18] |
C. Lei, F. Li, J. Liu, Theoretical analysis on a diffusive SIR epidemic model with nonlinear incidence in a heterogeneous environment, Discrete Cont. Dyn-B., 23 (2018), 4499–4517. doi: 10.3934/dcdsb.2018173. doi: 10.3934/dcdsb.2018173
![]() |
[19] |
K. Hattaf, A new generalized definition of fractional derivative with non-singular kernel, Computation, 8 (2020), 49. doi: 10.3390/computation8020049. doi: 10.3390/computation8020049
![]() |
[20] |
A. A. Lashari, S. Aly, K. Hattaf, G. Zaman, I. H. Jung, X. Z. Li, Presentation of malaria epidemics using multiple optimal controls, J. Appl. Math., 2012 (2012). doi: 10.1155/2012/946504. doi: 10.1155/2012/946504
![]() |
[21] |
K. Hattaf, A. A. Lashari, Y. Louartassi, N. Yousfi, A delayed SIR epidemic model with a general incidence rate, Electron. J. Qual. Theo., 3 (2013), 1–9. doi: 10.14232/ejqtde.2013.1.3. doi: 10.14232/ejqtde.2013.1.3
![]() |
[22] |
A. Bhadra, A. Mukherjee, K. Sarkar, Impact of population density on Covid-19 infected and mortality rate in India, Model. Earth Syst. Env., 7 (2020), 1–7. doi: 10.1101/2020.08.21.20179416. doi: 10.1101/2020.08.21.20179416
![]() |
[23] |
O. Adedire, J. N. Ndam, A model of dual latency compartments for the transmission dynamics of COVID-19 in Oyo state, Nigeria, Eng. Appl. Sci. Lett., 4 (2021), 1–13. doi: 10.30538/psrp-easl2021.0056. doi: 10.30538/psrp-easl2021.0056
![]() |
[24] |
M. Al-Raeei, The basic reproduction number of the new coronavirus pandemic with mortality for India, the Syrian Arab Republic, the United States, Yemen, China, France, Nigeria and Russia with different rate of cases, Clin. Epidemiol. Glob., 9 (2021), 147–149. doi: 10.1016/j.cegh.2020.08.005. doi: 10.1016/j.cegh.2020.08.005
![]() |
[25] | H. L. Smith, Monotone Dynamical Systems: An introduction to the theory of competitive and cooperative systems, Amer. Math. Soc., 1995. doi: 10.1090/surv/041. |
[26] | C. V. Pao, Nonlinear parabolic and elliptic equations, New York: Plenum, 1992. doi: 10.1007/978-1-4615-3034-3. |
[27] | J. D. Murray, Mathematical biology Ⅱ: Spatial models and biomedical applications, Berlin: Springer-Verlag, 2003. doi: 10.1023/A:1025805822749. |
[28] |
S. Ruan, Spatiotemporal epidemic models for rabies among animals, Infect. Dis. Model., 2 (2017), 277–287. doi: 10.1016/j.idm.2017.06.001. doi: 10.1016/j.idm.2017.06.001
![]() |
[29] | Z. Ma, J. Li, Dynamical modeling and analysis of epidemics, London: World Science Publishing, 2009. doi: 10.1007/978-1-4615-3034-3. |
[30] | P. Agarwal, Juan J. Nieto, M. Ruzhansky, Delfim F. M. Torres, Analysis of infectious disease problems (Covid-19) and their global impact, Singapore: Springer, 2021. doi: 10.1007/978-981-16-2450-6. |
[31] | T. Feuring, A. Tenhagen, Stability analysis of neural networks, In: Proceedings of International Conference on Neural Networks, 1997. doi: 10.1109/ICNN.1997.611716. |