
Citation: Yunfeng Liu, Guowei Sun, Lin Wang, Zhiming Guo. Establishing Wolbachia in the wild mosquito population: The effects of wind and critical patch size[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 4399-4414. doi: 10.3934/mbe.2019219
[1] | Luis Almeida, Michel Duprez, Yannick Privat, Nicolas Vauchelet . Mosquito population control strategies for fighting against arboviruses. Mathematical Biosciences and Engineering, 2019, 16(6): 6274-6297. doi: 10.3934/mbe.2019313 |
[2] | Bo Zheng, Wenliang Guo, Linchao Hu, Mugen Huang, Jianshe Yu . Complex wolbachia infection dynamics in mosquitoes with imperfect maternal transmission. Mathematical Biosciences and Engineering, 2018, 15(2): 523-541. doi: 10.3934/mbe.2018024 |
[3] | Bo Zheng, Lihong Chen, Qiwen Sun . Analyzing the control of dengue by releasing Wolbachia-infected male mosquitoes through a delay differential equation model. Mathematical Biosciences and Engineering, 2019, 16(5): 5531-5550. doi: 10.3934/mbe.2019275 |
[4] | Martin Strugarek, Nicolas Vauchelet, Jorge P. Zubelli . Quantifying the survival uncertainty of Wolbachia-infected mosquitoes in a spatial model. Mathematical Biosciences and Engineering, 2018, 15(4): 961-991. doi: 10.3934/mbe.2018043 |
[5] | Yun Li, Hongyong Zhao, Kai Wang . Dynamics of an impulsive reaction-diffusion mosquitoes model with multiple control measures. Mathematical Biosciences and Engineering, 2023, 20(1): 775-806. doi: 10.3934/mbe.2023036 |
[6] | Rajivganthi Chinnathambi, Fathalla A. Rihan . Analysis and control of Aedes Aegypti mosquitoes using sterile-insect techniques with Wolbachia. Mathematical Biosciences and Engineering, 2022, 19(11): 11154-11171. doi: 10.3934/mbe.2022520 |
[7] | Daiver Cardona-Salgado, Doris Elena Campo-Duarte, Lilian Sofia Sepulveda-Salcedo, Olga Vasilieva, Mikhail Svinin . Optimal release programs for dengue prevention using Aedes aegypti mosquitoes transinfected with wMel or wMelPop Wolbachia strains. Mathematical Biosciences and Engineering, 2021, 18(3): 2952-2990. doi: 10.3934/mbe.2021149 |
[8] | Mugen Huang, Moxun Tang, Jianshe Yu, Bo Zheng . The impact of mating competitiveness and incomplete cytoplasmic incompatibility on Wolbachia-driven mosquito population suppressio. Mathematical Biosciences and Engineering, 2019, 16(5): 4741-4757. doi: 10.3934/mbe.2019238 |
[9] | Diego Vicencio, Olga Vasilieva, Pedro Gajardo . Monotonicity properties arising in a simple model of Wolbachia invasion for wild mosquito populations. Mathematical Biosciences and Engineering, 2023, 20(1): 1148-1175. doi: 10.3934/mbe.2023053 |
[10] | Yufei Wang, Huidong Cheng, Qingjian Li . Dynamic analysis of wild and sterile mosquito release model with Poincaré map. Mathematical Biosciences and Engineering, 2019, 16(6): 7688-7706. doi: 10.3934/mbe.2019385 |
Mosquito-borne diseases such as malaria, dengue fever, West Nile virus, chikungunya, and Zika virus have been a great threat to public health. For dengue virus alone, it has been estimated that 3.9 billion people, in 128 countries, are at risk of infection [1]. In the year of 2014, more than 43,000 cases with locally acquired denguelike illness were reported in Guangdong province, China [2]. The human viruses including dengue, Zika, chikungunya and yellow fever are transmitted primarily by Aedes aegypti mosquitoes. Due to the lack of vaccines and efficient clinical cures [3], the only effective control strategy seems to be controlling the population of mosquitoes that transmit human viruses. Since massive spraying of insecticides and elimination of mosquito breeding sites are not sustainable to reduce mosquito density and might also lead to serious environmental problems, a promising strategy is the Wolbachia approach: releasing male and female Aedes aegypti mosquitoes with Wolbachia so that these mosquitoes can breed with the wild mosquito population and pass Wolbachia to the entire mosquito population. On one hand, the ability to transmit viruses to human for mosquitoes with Wolbachia is greatly reduced [4,5]. On the other hand, since the Wolbachia infection often induces cytoplasmic incompatibility (CI), which leads to early embryonic death when Wolbachia-infected males mate with uninfected females [6,7], the Wolbachia approach would greatly reduce the density of the mosquito population and can thus potentially eliminate the mosquito population and thus eradicate the mosquito-born infectious diseases.
To understand the Wolbachia infection dynamics, Zheng, Tang and Yu [8] proposed a delay differential equation model. To be self-contained, we briefly introduce their idea here. We denote by rf and rm the numbers of released female mosquitoes and released males carrying Wolbachia, respectively. Due to strong competition between adults, rf and rm satisfy
{drfdt=−δ1rfT(t),t>0,drmdt=−δ1rmT(t),t>0. | (1.1) |
Here T(t)=rf+rm+If+Im+Uf+Um denotes the total population size, with Uf, Um, If and Im being the numbers of uninfected reproductive females, uninfected reproductive males, and infected reproductive females and males other than those from releasing, respectively. Let bI (resp. bU) be the natural birth rate of the infected (resp. uninfected) mosquitos and 0≤δ≤1 be the proportion of mosquitos born female. Then the proportion of mosquitos born male is 1−δ. With strong CI and high maternal transmission, if the average waiting time from parent mating to the emergence of reproductive progenies for both infected and uninfected mosquitoes is negligible, then we have
{dIfdt=δbI[If+rf]−δ1IfT(t),t>0,dImdt=(1−δ)bI[If+rm]−δ1ImT(t),t>0,dUfdt=δbU[UfUmrm+Im+Um]−δ2UfT(t),t>0,dUmdt=(1−δ)bU[UfUmrm+Im+Um]−δ2UmT(t),t>0. | (1.2) |
Since both rf and rm approach 0 as t→+∞. Let
u(t)=If+Im and v(t)=Uf+Um. | (1.3) |
Assuming equal determination case, which means that δ=1/2, If=Im and Uf=Um, setting b1=bI/2 and b2=bU/2 and considering the spatiotemporal factor, Huang et al. [9] came up with the following reaction-diffusion system:
{∂u∂t=d1Δu+u(b1−δ1(u+v)),t>0,x∈Ω,∂v∂t=d2Δv+v(b2vu+v−δ2(u+v)),t>0,x∈Ω,∂u∂ν=∂v∂ν=0,t>0,x∈∂Ω,u(0,x)=u0(x),v(0,x)=v0(x),x∈Ω. | (1.4) |
In (1.4), d1 and d2 are the diffusion rates, Δ denotes the Laplace operator in the spatial variable x, and ν denotes the unit outward normal vector to the boundary of Ω.
Noticing that the spread of mosquitoes can be greatly affected by the wind speed [10], therefore, the advection due to wind should be incorporated into modeling. Note also practical experience of releasing mosquitoes with Wolbachia to the wild suggests that a minimum release area is needed in order to achieve a stable local establishment and spread in continuous habitats [11]. Motivated by these two aspects, in this paper, we extend the model considered in [9], i.e., system (1.4), to a reaction-diffusion-advection model with spatially heterogeneous environment and flexible boundary conditions. More specifically, our model is described by the following system
{∂u∂t=duxx−αux+u[b1(x)−δ1(u+v)],t>0,x∈(0,L),∂v∂t=dvxx−αvx+v[b2(x)vu+v−δ2(u+v)],t>0,x∈(0,L),dux(0,t)−αu(0,t)=0,t>0,dux(L,t)−αu(L,t)=−βαu(L,t),t>0,dvx(0,t)−αv(0,t)=βαv(0,t),t>0,dvx(L,t)−αv(L,t)=−βαv(L,t),t>0,u(x,0)=u0(x)≥,≢0,x∈[0,L],v(x,0)=v0(x)≥,≢0,x∈[0,L], | (1.5) |
where u(x,t) and v(x,t) represent the population densities of infected and uninfected mosquitoes at location x and time t, respectively. The parameter d denotes the random diffusion rate of u and v. The functions b1(x) and b2(x) denote the halves of the birth rates of the infected and uninfected mosquitoes, respectively ([8]). The parameter δ1 (or δ2) denotes the density dependent death rate for the infected (or uninfected) mosquito species. The advection constant α measures the result of wind transportation. The parameter β<∞ measures the relative rate of population loss at the downstream due to wind flow and replenishment at the upstream.
We suppose that bi(x) satisfies the following hypothesis:
bi(x)∈C1+α([0,L])(α∈(0,1)),i=1,2and0<b_≤b1(x),b2(x)≤ˉb. | (1.6) |
Here b_ and ˉb are two positive constants.
We regard β=+∞ as the Dirichlet boundary case, and system (1.5) becomes
{∂u∂t=duxx−αux+u[b1(x)−δ1(u+v)],t>0,x∈(0,L),∂v∂t=dvxx−αvx+v[b2(x)vu+v−δ2(u+v)],t>0,x∈(0,L),dux(0,t)−αu(0,t)=0=u(L,t),t>0,v(0,t)=0=v(L,t),t>0,u(x,0)=u0(x)≥,≢0,x∈[0,L],v(x,0)=v0(x)≥,≢0,x∈[0,L], | (1.7) |
Ideally, if the entire mosquito population is replaced by mosquitoes with Wolbachia, then the Wolbachia establishment is called successful. This is achieved if the solutions of (1.5) and (1.7) approach a semi-trivial steady state, (˜u,0), where ˜u satisfies the following equations,
{duxx−αux+u[b1(x)−δ1u]=0,x∈(0,L),dux(0)−αu(0)=0,dux(L)−αu(L)=−βαu(L), | (1.8) |
for β<∞ or
{duxx−αux+u[b1(x)−δ1u]=0,x∈(0,L),dux(0)−αu(0)=0,u(L)=0, | (1.9) |
for the case with β=∞.
We point out that there has been several mathematical models formulated to describe the Wolbachia spreading dynamics. These models include differential equations with/without time delays [8,12,13,14], reaction-diffusion equations [9,15], and stochastic equations [16]. These models focused on studying the subtle relation between the threshold releasing level for Wolbachia-infected mosquitoes and several important parameters including the CI intensity and the fecundity cost of Wolbachia infection. We should also point out that besides releasing mosquitoes with Wolbachia, an alternative control strategy is releasing the sterile mosquitoes to the wild mosquito population [17].
We organize the rest of this paper as follows. In Section 2, we give some useful lemmas to establish the relation between two related principal eigenvalues and the domain size, the advection rate as well as the diffusion rate. Our main results are presented in Section 3. To illustrate our results, we also present some numerical simulations in Section 4. We conclude this paper in the last section.
In order to discuss the existence of (˜u,0), corresponding to systems (1.5) and (1.7), we need to study the following subsystems:
{∂u∂t=duxx−αux+u[b1(x)−δ1u],t>0,x∈(0,L),dux(0,t)−αu(0,t)=0,t>0,dux(L,t)−αu(L,t)=−βαu(L,t),t>0,u(x,0)=u0(x)≥,≢0,x∈[0,L], | (2.1) |
and
{∂u∂t=duxx−αux+u[b1(x)−δ1u],t>0,x∈(0,L),dux(0,t)−αu(0,t)=0,t>0,u(L,t)=0,t>0,u(x,0)=u0(x)≥,≢0,x∈[0,L], | (2.2) |
respectively.
This leads to the study of the following linear eigenvalue problem
{dϕxx−αϕx+ϕb1(x)=σϕ,x∈(0,L),dϕx(0)−αϕ(0)=0,dϕx(L)−αϕ(L)=−βαϕ(L). | (2.3) |
In the case that β=+∞, the corresponding eigenvalue problem reads as
{dϕxx−αϕx+ϕb1(x)=σϕ,x∈(0,L),dϕx(0)−αϕ(0)=0,ϕ(L)=0. | (2.4) |
Throughout this paper, we denote the principal eigenvalue of (2.3) or (2.4) by σ1(α,d,L).
First, we shall investigate how σ1(α,d,L) depends on α.
Lemma 2.1. Suppose that (1.6) is satisfied, then
(a) When β∈(0,12), if b′1(x)≤0, then σ1(α,d,L) is a strictly monotone decreasing function of α; When β∈[12,+∞], then σ1(α,d,L) is a strictly monotone decreasing function of α.
(b) 0<b_≤limα→0σ1(α,d,L)≤ˉb provided that β<+∞.
(c) limα→+∞σ1(α,d,L)=−∞.
Proof. Set φ=e−αdxϕ, then (2.3) becomes
{dφxx+αφx+b1(x)φ=σφ,x∈(0,L),φx(0)=0,dφx(L)=−βαφ(L), | (2.5) |
and (2.4) becomes
{dφxx+αφx+b1(x)φ=σφ,x∈(0,L),φx(0)=0=φ(L). | (2.6) |
Let us denote by φα the derivative of φ with respect to α. Multiplying the first equation of (2.5) by eαdxφα, we obtain
d(eαdxφx)xφα+φφαeαdxb1(x)=σφφαeαdx. | (2.7) |
Integrating (2.7) over (0,L) yields
∫L0d(eαdxφx)xφα=∫L0(σ−b1(x))φφαeαdx. | (2.8) |
On the other hand, it follows from (2.5) that
{dφαxx+φx+αφαx+b1(x)φα=σαφ+σφα,dφαx(0)=0,dφαx(L)=−βφ(L)−βαφα(L). | (2.9) |
Multiplying (2.9) by eαdxφ(x) and integrating it over (0,L), we have
∫L0d(eαdxφαx)xφ+∫L0eαdxφxφ=∫L0(σ−b1(x))φφαeαdx+σα∫L0φ2eαdx. | (2.10) |
Using integration by parts and (2.8), we obtain
σα=−eαdLβφ2(L)+∫L0eαdxφxφ∫L0φ2eαdx. | (2.11) |
If β=+∞, (2.11) becomes
σα=∫L0eαdxφxφ∫L0φ2eαdx. | (2.12) |
In general, when β∈(0,12), σ1(α,d,L) may not be monotone in α. To ensure σα<0, we need to show φx<0 in (0,L). To this end, we set W=ϕxϕ, and we can rewrite (2.3) as the following equation
{dWxx+(2dW−φ)Wx=−b′1(x),W(0)=αd,W(L)=(1−β)αd. | (2.13) |
Since b′1(x)≤0, then by the maximum principle we have W<αd, which implies that φx<0 in (0,L) ([18]). This implies that σ1(α,d,L) is a strictly monotone decreasing function of α.
For the case β∈[12,+∞), from (2.11), we have σα=−eαdLβφ2(L)+∫L0eαdxφxφ∫L0φ2eαdx=−eαdLβφ2(L)+∫L0eαdxdφ22∫L0φ2eαdx=−eαdLβφ2(L)+eαdxφ22∣L0−αd∫L0eαdxφ22∫L0φ2eαdx=−(β−12)eαdLφ2(L)−φ2(0)2−αd∫L0eαdxφ22∫L0φ2eαdx.
It is easy to see β∈[12,+∞) implies that σα<0. When β=+∞, Using the same calculation as above and combining with (2.12), we can obtain that σα<0.
If α=0, then the boundary condition of (2.5) is simply a Neumann boundary condition. Furthermore, combining with (1.6), we have
0<b_≤limα→0σ1(α,d,L)≤ˉb. |
Especially, limα→0σ1(α,d,L)=b1, when b1(x)=b1. It follows from Proposition 2.1 of [19] and (1.6) that
σ1(α,d,L)≤(γ2−γ)α2d+ˉb, |
where 0<γ<min{1,β}. Consequently, limα→+∞σ1(α,d,L)=−∞.
Next we investigate the relationship between σ1(α,d,L) and L.
Lemma 2.2. σ1(α,d,L) is a strictly monotone increasing function of L.
Proof. First we consider system (2.5) by fixing the parameters α and d and varying L. For any 0<L1<L2, we show that σ1(α,d,L1)<σ1(α,d,L2). Let φ1 be the positive eigenfunction corresponding to σ1(α,d,L1), and define
ψ={φ1,0<x≤L10,L1<x<L2. |
Since
σ1(α,d,L1)=−βαeαdL1φ21(L1)−β1αφ21(0)−d∫L10eαdx(φ1x)2dx+∫L10b1(x)eαdxφ21dx∫L10eαdxφ21dx=−βαeαdL2ψ2(L2)−β1αψ2(0)−d∫L20eαdx(ψx)2dx+∫L20b1(x)eαdxψ2dx∫L20eαdxψ2dx≤maxψ≠0,ψ∈W1,2−βαeαdL2ψ2(L2)−β1αψ2(0)−d∫L20eαdx(ψx)2dx+∫L20b1(x)eαdxψ2dx∫L20eαdxψ2dx=σ1(α,d,L2). |
Due to the strict positivity of the eigenfunction corresponding to σ1(α,d,L2) in [0,L2], the above inequality should be strict, which implies σ1(α,d,L1)<σ1(α,d,L2). The case with β=+∞ can be dealt with in a similar fashion and we skip the details here.
Lemma 2.3. Assume (1.6) holds. If β=+∞, then σ1(α,d,L)→−∞ as L→0.
Proof. We rewrite σ1(α,d,L) as σ1(α,d,L,b1(x)) to emphasize the dependence of σ1(α,d,L) on the function b1(x). By Corollary 2.2 of [20], one easily sees that σ1(α,d,L,b1(x))≤σ1(α,d,L,ˉb).
Next we consider the following boundary value problem (BVP)
{dϕxx−αϕx+ϕˉb=σϕ,x∈(0,L),dϕx(0)−αϕ(0)=0,ϕ(L)=0. | (2.14) |
Set ψ=e−αx2dϕ, then (2.14) yields
{dψxx+[ˉb−α24d−σ]ψ=0,x∈(0,L),ψx(0)−α2dψ(0)=0,ψ(L)=0, | (2.15) |
From the boundary conditions we find
tan√4d(ˉb−σ1)−α22dL+√4d(ˉb−σ1)−α2α=0, for 4d(ˉb−σ1)>α2. | (2.16) |
Since by Lemma 2.2, we have
limL→0+σ1(ˉb)=σ∗∈R∪{−∞}. |
Assume that σ∗ is finite, then from (2.16), we obtain σ∗=4dˉb−α24d, and hence ψ(x)=Ax+B. The boundary conditions imply that A=B=0. This is impossible. Thus limL→0+σ1(ˉb)=−∞ and hence limL→0+σ1(b1(x))=−∞.
Lemma 2.4. If β=+∞ and α=0, then σ1(α,d,L) is a strictly monotone decreasing function of d.
Proof. Let us denote by φd the derivative of φ with respect to d. Multiplying the first equation of (2.6) by eαdxφd, it holds that
d(eαdxφx)xφd+φφdeαdxb1(x)=σφφdeαdx. | (2.17) |
Integrating (2.17) over (0,L), we obtain
∫L0d(eαdxφx)xφd=∫L0(σ−b1(x))φφdeαdx. | (2.18) |
On the other hand, it follows from (2.6) that
{φxx+dφdxx+αφdx+b1(x)φd=σdφ+σφd,φdx(0)=0orφd(0)=0,φd(L)=0. | (2.19) |
Multiplying (2.19) by eαdxφ(x) and integrating it over (0,L), we have
∫L0d(eαdxφdx)xφ+∫L0eαdxφxxφ=∫L0(σ−b1(x))φφdeαdx+σd∫L0φ2eαdx. | (2.20) |
Using integration by parts and (2.18), we obtain
σd=−∫L0eαdx(φx)2−αd∫L0eαdxφφx∫L0φ2eαdx. | (2.21) |
We find σd<0, when α=0.
If β=+∞, we also have the following remark (See Theorem 3.1 of [21]).
Remark 2.5. (i) σ1(0,d,L)→maxx∈[0,L]b1(x)>0 as d→0;
(ii) σ1(0,d,L)→−∞ as d→+∞.
Our first result is the following theorem.
Theorem 3.1. If system (1.5) admits a semi-trivial steady state (˜u,0), then the semi-trivial steady state (˜u,0) is locally asymptotically stable.
Proof. Linearizing the second equation of system (1.5) at (˜u,0), we obtain the following eigenvalue problem
{dϕxx−αϕx−δ2˜uϕ=ζϕ,in(0,L),dϕx(0)−αϕ(0)=βαϕ(0),dϕx(L)−αϕ(L)=−βαϕ(L). | (3.1) |
Let ˜ϕ=e−αdxϕ, then problem (3.1) becomes
{d(eαdx˜ϕx)x−δ2eαdx˜u˜ϕ=ζeαdx˜ϕ,in(0,L),d˜ϕx(0)=βα˜ϕ(0),d˜ϕx(L)=−βα˜ϕ(L). | (3.2) |
By the variational method, ζ1 can be characterized by
ζ1=−βα~ϕ12(0)−βαeαLd~ϕ12(L)−d∫L0eαdx(˜ϕ1x)2−∫L0δ2eαdx˜u(~ϕ1)2∫L0eαdx(~ϕ1)2 | (3.3) |
and when β=+∞,
ζ1=−d∫L0eαdx(˜ϕ1x)2−∫L0δ2eαdx˜u(~ϕ1)2∫L0eαdx(~ϕ1)2, | (3.4) |
where ~ϕ1 is the eigenfunction associated with ζ1. Thus ζ1<0 and it follows from Proposition 3.1 of [20] that the semi-trivial steady state (˜u,0) is locally asymptotically stable.
Theorem 3.2. If α=0, or α>0 and β=0, then system (1.5) admits a semi-trivial steady state, (˜u,0), which is locally asymptotically stable, where ˜u is the unique positive steady state of problem (2.1).
Proof. Under the assumptions, the linear eigenvalue problem (2.3) reads as
{dϕxx+ϕb1(x)=λϕ,x∈(0,L),ϕx(0)=0=ϕx(L), | (3.5) |
and
{dϕxx−αϕx+ϕb1(x)=ηϕ,x∈(0,L),dϕx(0)−αϕ(0)=0,dϕx(L)−αϕ(L)=0, | (3.6) |
respectively. For (3.5), a standard eigenvalue analysis gives the principle eigenvalue
λ1=maxϕ∈W1,2(0,L),ϕ≠0[−d∫L0(ϕx)2+∫L0b1(x)(ϕ)2∫L0(ϕ)2]. | (3.7) |
Moreover, it is easy to verify that λ1>0 by using the test function ϕ=1. For (3.6), we multiply the equations by e−αdx to get
{d(e−αdxϕx)x+e−αdxϕb1(x)=ηe−αdxϕ,x∈(0,L),(e−αdxϕ)x(x)=0,x=0,L. | (3.8) |
Thus the associated principle eigenvalue is
η1=maxϕ∈W1,2(0,L),ϕ≠0[αe−αdLϕ2(L)−αϕ2(0)−d∫L0e−αdx(ϕx)2+∫L0e−αdxb1(x)(ϕ)2∫L0e−αdx(ϕ)2]. |
Using the test function ϕ≡eαdx, we see that η1>0. The conclusion then follows from Propositions 3.2 and 3.3 of [20] and Theorem 3.1.
For the spatially homogeneous case, by Lemmas 2.1 and 2.2, Theorems 2.1 and 2.3 of [19], we have the following result.
Theorem 3.3. Assume that b1(x)≡b1>0 is a constant.
(a) If β∈(0,12) and 0<α<√db1β(1−β), then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1, where
L∗1={2darctanαβ√4db1−α22db1−α2β√4db1−α2,if0<α≤√4db1,d√α2−4db1ln2db1−βα2+αβ√α2−4db12db1−βα2−αβ√α2−4db1,if√4db1<α<√db1β(1−β). |
Moreover, if α≥√db1β(1−β), then problem (2.1) only has a globally asymptotically stable zero steady state and system (1.5) does not admit a semi-trivial steady state in the form of (˜u,0) satisfying ˜u>0.
(b) When β=12. If 0<α<√4db1, then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1 with
L∗1=2darctanα√4db1−α24db1−α2√4db1−α2. |
If α≥√4db1, then system (1.5) does not admit a semi-trivial steady state in the form of (˜u,0) satisfying ˜u>0.
(c) Suppose β∈(12,+∞). If 0<α<√4db1, then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1 with
L∗1={2darctanαβ√4db1−α22db1−α2β√4db1−α2,if0<α≤√2db1β,2dπ+arctanαβ√4db1−α22db1−α2β√4db1−α2,if√2db1β<α<√4db1. |
If α≥√4db1, then system (1.5) does not admit a semi-trivial steady state in the form of (˜u,0) satisfying ˜u>0.
(d) Suppose that β∈(0,12). If d>α2β(1−β)b1, then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1 with
L∗1={2darctanαβ√4db1−α22db1−α2β√4db1−α2,ifd≥α24b1,d√α2−4db1ln2db1−βα2+αβ√α2−4db12db1−βα2−αβ√α2−4db1,ifα2β(1−β)b1<d<α24b1. |
(e) For β=12, if d>α24b1, then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1=2darctanαβ√4db1−α22db1−α2β√4db1−α2.
(f) Suppose that β∈(12,+∞), when d>α24b1, then system (1.5) admits a semi-trivial steady state, (˜u,0), with ˜u being the unique positive steady state of problem (2.1), if and only if L>L∗1, where L∗1 is given by
L∗1={2darctanαβ√4db1−α22db1−α2β√4db1−α2,d≥α2β2b1,2dπ+arctanαβ√4db1−α22db1−α2β√4db1−α2,α24b1<d<α2β2b1. |
(g) For any α>0, β∈(0,1), if d≤˜d, then system (1.5) does not admit a semi-trivial steady state in the form of (˜u,0) satisfying ˜u>0, where
˜d={α2β(1−β)b1,0<β<12α24b1,12≤β<1. |
Similarly, for system (1.7), we have the following result.
Theorem 3.4. Suppose b1(x)≡b1>0 is a constant. Then we have the following conclusions.
(i) For any d>0, if 0<α<√4b1d, then there exists a critical number
ˆL∗1=2d(π+arctan−√4db1−α2α)√4db1−α2 |
such that for L≥ˆL∗1, system (1.7) admits a stable semi-trivial steady state (˜u,0), where ˜u>0 is the unique positive steady state of system (2.2). Moreover, ˆL∗1 is an increasing function of the advection rate α for 0<α<√4b1d.
(ii) For any α, if d>α24b1, then there exists a critical number
ˆL∗1=2d(π+arctan−√4db1−α2α)√4db1−α2 |
such that for L≥ˆL∗1, system (1.7) admits a stable semi-trivial steady state (˜u,0), where ˜u>0 is the unique positive steady state of system (2.2). Further, ˆL∗1 is an decreasing function of the diffusion rate d, for α24b1<d<α22b1.
(iv) If α≥√4b1d, then system (1.7) does not admit a semi-trivial steady state in the form of (˜u,0), where ˜u>0.
If b1(x) is not a constant, we have the following result.
Theorem 3.5. Suppose that (1.6) is satisfied. Then the following conclusions hold.
(I) If α>0, when β∈(0,12), b′1(x)≤0, or β∈[12,+∞), then there exists α∗>0 such that for 0<α<α∗, system (1.5) admits a semi-trivial steady state, (˜u,0), which is stable, and ˜u is the unique positive steady state of problem (2.1).
(II) For the case with β=∞, we have the following conclusions.
(II.1) If α=0, then there exists d∗ such that for 0<d<d∗, system (1.7) admits a semi-trivial steady state (˜u,0), which is stable and ˜u is the unique positive steady state of system (2.2).
(II.2) If L>π2√db_, then there exists ˜α∗ such that for 0<α<˜α∗, system (1.7) admits a semi-trivial steady state (˜u,0), which is stable and ˜u is the unique positive steady state of system (2.2).
(II.3) If 0<α<√4b_d, then there exists a L∗>0 such that for L>L∗, system (1.7) admits a semi-trivial steady state (˜u,0), which is stable and ˜u is the unique positive steady state of system (2.2).
Proof. The case (I) can be proved by (a), (b) and (c) of Lemma 2.1 and the proof of case (II.1) can be obtained by Lemma 2.4 and Remark 2.5. To prove case (II.2), we just need to verify the following three facts.
(1) σ1(α,b1(x)) is a strictly monotone decreasing function of α;
(2) limα→0σ1(α,b1(x))>0;
(3) limα→+∞σ1(α,b1(x))=−∞.
Note that (1) and (3) follow immediately from Lemma 2.1. Furthermore, (2) follows from noting that σ1(0,b1(x))≥σ1(0,b_) for b1(x)≥b_ and σ1(0,b1(x))≥σ1(0,b_)=b_−dπ24L2>0, when β=+∞ and L>π2√db_.
Next we prove (II.3). Note that \beta = \infty , the associated eigenvalue problem is
\begin{equation} \left\{\begin{array}{ll} { d\phi_{xx}-\alpha \phi_{x}+\phi\underline{b} = \sigma_{1}\phi,}\; & x\in(0,L),\\ d\phi_{x}(0)-\alpha \phi(0) = 0,\\ \phi(L) = 0, \end{array}\right. \end{equation} | (3.9) |
where \phi is corresponding eigenfunction of \sigma_{1} .
If 0 < \alpha < \sqrt{4\underline{b}d} , then it follows from Theorem 3.4 that \sigma_{1}(\underline{b}) > 0 , as L\rightarrow +\infty . Hence, it is easy to know that \sigma_{1}(b_{1}(x)) > 0 , as L\rightarrow +\infty , if 0 < \alpha < \sqrt{4\underline{b}d} . Therefore the proof of case (II.3) follows directly from Lemma 2.2 and Lemma 2.3.
In this section, we use several numerical simulations to illustrate our theoretical results. First we verify Theorem 3.3. We take parameter values: \beta = 0.4 , b_{1}(x) = 0.1077 , b_{2}(x) = 0.1988 , \delta_{1} = \delta_2 = 8.5034\times10^{-6} , d = 1.25\times10^{-2} , \alpha = 1.25\times10^{-2} . For this set of parameters, we find that \alpha < \sqrt{\frac{db_{1}}{\beta(1-\beta)}}\approx 0.0749 and L_{1}^{*}\approx 0.0472 . Theorem 3.3 applies here: if we take L = 0.04 < L_1^* , then as shown in Figure 1 (Left), the solution of (1.5) approaches the trivial steady state (0, 0) , while if we take L = 0.05 > L_1^* , as shown in Figure 1 (Right), the solution of (1.5) approaches the semi-trivial steady state (\widetilde{u}, 0) implying that all mosquitoes are infected with Wolbachia.
Now we take parameter values as: d = 1.25\times10^{-2} , \delta_{1} = \delta_2 = 8.5034\times10^{-6} , b_{1}(x) = e^{-x}+1 , b_2(x) = 1.5+\sin (x) , \beta = 0.4 , L = 0.1 . For this set of parameters, we can numerically find \alpha*\approx 0.261 such that the solutions of (1.5) approach (\widetilde{u}, 0) if \alpha\in [0, \alpha^*) , and approach (0, 0) if \alpha\geq \alpha^* (See Figure 2).
Now we take \beta = \infty , \alpha = 0.1789 , d = 1.25\times10^{-2} , \delta_1 = \delta_2 = 0.85034\times 10^{-6} , b_1(x) = e^{-x}+1 , b_{2}(x) = e^{-x}+1 . Then \underline{b} = 1 , \alpha < \sqrt{4d\underline{b}} = 0.2236 . Numerically we find \hat{L}^{*}\approx 0.21 . As shown in Figure 3, if L > L_1^* , then the solutions of (1.5) approach the semi-trivial steady state (\widetilde{u}, 0) implying a full establishment Wolbachia is achieved, while the establishment of Wolbachia fails if L < L_1^* .
In this paper, we have proposed a reaction-diffusion-advection model in one-dimensional spatially inhomogeneous environment with general boundary conditions. Our results (Theorems 3.2–3.5) show that in order to fully establish Wolbachia in the wild mosquito population, i.e., all mosquitoes eventually carry Wolbachia, the wind related parameter \alpha and the patch size L over which the mosquitoes with Wolbachia are spreading should satisfy certain requirements. Generally speaking, the wind cannot be too strong, and the minimum patch size cannot be too small. The critical values of the advection parameter \alpha and the patch size L for the establishment of Wolbachia depend on the model parameters including the diffusion rate d , the birth rates b_1(x) and b_2(x) , and the death rates \delta_1 and \delta_2 . For instance, Theorem 3.3 (a) indicates, if b_1(x) = b_1 is a positive constant and \beta\in (0, 1/2) , then the full establishment of Wolbachia will not be successful if \alpha is too large such that \alpha\geq\sqrt{\frac{db_{1}}{\beta(1-\beta)}} (i.e., the wind is too strong) or if \alpha is in a proper range ( 0 < \alpha < \sqrt{\frac{db_{1}}{\beta(1-\beta)}} ), but the patch size L is too small satisfying L < L_1^* (See Figure 1).
The authors wish to thank three anonymous reviewers for their very helpful comments. YL and ZG were supported by the National Science Foundation of China (No. 11771104), Program for Chang Jiang Scholars and Innovative Research Team in University (IRT-16R16). LW was partially supported by a Discovery grant from the Natural Sciences and Engineering Research Council of Canada.
The authors declare that they have no competing interests.
[1] | O. J. Brady, P. W. Gething, S. Bhatt, et al., Refining the global spatial limits of dengue virus trans-mission by evidence-based consensus, PLoS Negl. Trop. Dis., 6 (2012), e1760. |
[2] | Dengue Situation Update 453, World Health Organization, (2014), Available from: http://www.wpro.who.int/ emerging diseases/denguebiweekly 02dec2014.pdf. |
[3] | L. M. Schwartz, M. E. Halloran, A. P. Durbin, et al., The dengue vaccine pipeline: Implications for the future of dengue control, Vaccine, 33 (2015), 3293–3298. |
[4] | G. Bian, Y. Xu, P. Lu, et al., The endosymbiotic bacterium Wolbachia induces resistance to dengue virus in Aedes aegypti, PLoS Pathog., 6 (2010), e1000833. |
[5] | H. Dutra, M. Rocha, F. Dias, et al., Wolbachia blocks currently circulating Zika virus isolates Aedes aegypti mosquitoes, Cell Host & Microbe, 19 (2016), 771–774. |
[6] | M. Turelli and A. A. Hoffmann, Rapid spread of an inherited incompatibility factor in California Drosophila, Nature, 353 (1991), 440–442. |
[7] | Z. Xi, C. C. Khoo and S. I. Dobson, Wolbachia establishment and invasion in an Aedes aegypti laboratory population, Science, 310 (2005), 326–328. |
[8] | B. Zheng, M. Tang and J. Yu, Modeling Wolbachia spread in mosquitoes through delay differential equations, SIAM J. Appl. Math., 74 (2014), 743–770. |
[9] | M. Huang, M. Tang and J. Yu, Wolbachia infection dynamics by reaction-diffusion equations, Sci. China Math., 58 (2015), 77–96. |
[10] | L. T. Takahashi, N. A. Maidana, W. C. Ferreira, et al., Mathematical models for the Aedes aegypti dispersal dynamics: travelling waves by wing and wind, Bull. Math. Biol., 67 (2005), 509–528. |
[11] | T. L. Schmidt, N. H. Barton, G. Rašić, et al., Local introduction and heterogeneous spatial spread of dengue-suppressing Wolbachia through an urban population of Aedes aegypti, PLoS Biol., 15 (2017), e2001894. |
[12] | M. Huang, J. Luo, L. Hu, et al., Assessing the efficiency of Wolbachia driven aedes mosquito suppression by delay differential equations, J. Theoret. Biol., 440 (2018), 1–11. |
[13] | J. Yu, Modeling mosquito population suppression based on delay differential equations, SIAM J. Appl. Math., 78 (2018), 3168–3187. |
[14] | B. Zheng, M. Tang, J. Yu, et al., Wolbachia spreading dynamics in mosquitoes with imperfect maternal transmission, J. Math. Biol., 76 (2018), 235–263. |
[15] | M. Huang, J. Yu, L. Hu, et al., Qualitative analysis for a Wolbachia infection model with diffusion, Sci. China Math., 59 (2016), 1249–1266. |
[16] | L. Hu, M. Huang, M. Tang, et al., Wolbachia spread dynamics in stochastic environments, Theor. Popul. Biol., 106 (2015), 32–44. |
[17] | L. Cai, S. Ai and J. Li, Dynamics of mosquitoes populations with different strategies for releasing sterile mosquitoes, SIAM J. Appl. Math., 74 (2014), 1786–1809. |
[18] | P. Zhou and X. Zhao, Evolution of passive movement in advective environments: General boundary condition, J. Differ. Equations, 264 (2018), 4176–4198. |
[19] | Y. Lou and P. Zhou, Evolution of dispersal in advective homogeneous environment: the effect of boundary conditions, J. Differ. Equations, 259 (2015), 141–171. |
[20] | R. Cantrell and C. Cosner, Spatial Ecology via Reaction-diffusion Equations, John Wiley & Sons, (2003). |
[21] | P. Zhou and D. Xiao, The diffusive logistic model with a free boundary in heterogeneous environ-ment, J. Differ. Equations, 256 (2014), 1927–1954. |
1. | Yun Li, Hongyong Zhao, Kai Wang, Dynamics of an impulsive reaction-diffusion mosquitoes model with multiple control measures, 2022, 20, 1551-0018, 775, 10.3934/mbe.2023036 | |
2. | Ruibin Jiang, Zhiming Guo, Dynamics of discrete Ricker models on mosquito population suppression, 2024, 47, 0170-4214, 4821, 10.1002/mma.9840 |