1,491
Views
7
CrossRef citations to date
0
Altmetric
Tianyuan Hengyang Workshop 2020

Discrete dynamical models on Wolbachia infection frequency in mosquito populations with biased release ratios

&
Pages 320-339 | Received 10 Jun 2021, Accepted 24 Aug 2021, Published online: 17 Sep 2021

Abstract

We develop two discrete models to study how supplemental releases affect the Wolbachia spreading dynamics in cage mosquito populations. The first model focuses on the case when only infected males are released at each generation. This release strategy has been proved to be capable of speeding up the Wolbachia persistence by suppressing the compatible matings between uninfected individuals. The second model targets the case when only infected females are released at each generation. For both models, detailed model formulation, enumeration of the positive equilibria and their stability analysis are provided. Theoretical results show that the two models can generate bistable dynamics when there are three positive equilibrium points, semi-stable dynamics for the case of two positive equilibrium points. And when the positive equilibrium point is unique, it is globally asymptotically stable. Some numerical simulations are offered to get helpful implications on the design of the release strategy.

2000 Mathematics Subject Classifications:

This article is part of the following collections:
Tianyuan Hengyang Workshop 2020

1. Introduction

Dengue is a mosquito-borne viral disease which is mainly endemic in tropical and subtropical areas, and has spread rapidly to temperate regions in recent years. About 100–400 million people infect dengue each year and nowadays, almost half of the world's population is at risk of dengue. As a mosquito-borne disease, dengue is transmitted by the bites of female Aedes aegypti and Aedes albopictus which are also vectors of Zika, Chikungunya and yellow fever [Citation11]. The most direct and traditional way to prevent mosquito-borne disease transmission is to kill mosquitoes by spraying insecticides and removing breeding sites, which only has a short-term effect because of the emergence and enhancement of insecticide resistance of mosquitoes and the continual creation of ubiquitous larval sources in the warm and humid seasons [Citation4,Citation26,Citation31]. Although the history of dengue vaccines can be traced back to 1993, dengue vaccine was first applied for use until 2015. However, experiments in [Citation7,Citation8] have proved the phenomenon of antibody dependent enhancement (ADE for short) in dengue serotypes, and further report [Citation3] shows that 130 among 830,000 vaccinated children have died, 19 of those have dengue, meaning that ADE does play a role.

An innovative biological method involves an intracellular bacterium, named Wolbachia, which was first identified by Hertig and Wolbach in 1924 [Citation12]. Wolbachia, which exists in up to 75% of insects, gained widespread attention of scholars in 1956 when Laven [Citation23] revealed its role in cytoplasmic incompatibility (CI for short) in Culex pipiens. Unfortunately, Wolbachia does not exist in Aedes aegypti. Although Aedes albopictus naturally carries two Wolbachia strains, these two strains could not block the replication of the dengue viruses in mosquito. The groundbreaking work is credited to Xi, who established a stable Wolbachia infection in Aedes aegypti for the first time [Citation32].

As a maternally transmitted bacterium, Wolbachia can induce CI when Wolbachia-infected males mate with uninfected females, resulting in an early embryonic death [Citation13,Citation24] and no offspring can be produced from these mated females. Based on these two mechanisms, two release strategies targeting controlling mosquito populations emerge as promising methods to reduce the occurrence of diseases transmitted by mosquitoes. The first one is usually termed as population suppression [Citation46], when a large number of Wolbachia-infected males are released into the wild to suppress, or even eradicate, the wild female mosquitoes through CI. Population replacement, as an alternative release strategy, release both Wolbachia-infected males and females to replace wild mosquito population with infected one, among which females lose their ability in transmitting dengue viruses owing to Wolbachia infection. With promising results to reduce the occurrence of diseases transmitted by mosquitoes, the dynamics of Wolbachia in mosquito population has attracted a lot of attention, and various mathematical models have been developed, including ordinary differential models [Citation16,Citation34,Citation36,Citation38,Citation42,Citation44], delay differential models [Citation18–22,Citation33,Citation35,Citation41], stochastic models [Citation15], reaction–diffusion models [Citation17] and discrete models [Citation6,Citation10,Citation13,Citation14,Citation25,Citation27–30,Citation37,Citation39,Citation40,Citation43,Citation45].

Non-overlapping cage mosquito populations whose dynamics can be monitored by infection frequency rather than number, where the discrete model becomes the first choice for its easy mathematical tractability. The first discrete model was developed by Caspari and Watson [Citation6] to characterize the evolutionary importance of CI sterility in mosquitoes, which reads as (1) xn+1=(1sf)xnshxn2(sf+sh)xn+1,n=0,1,2,,(1) where xn is the frequency of Wolbachia infection at the nth generation, sf(0,1) is the fitness cost of Wolbachia-infected mosquitoes to wild ones, and sh(0,1] is the proportion of unhatched eggs produced from incompatible cross [Citation28,Citation29]. Experimental observations show that Wolbachia can be stably maintained with strong CI and a mild fitness cost [Citation5,Citation24,Citation32]. Hence, infections with sf<sh is widely accepted. Later in 1978, observing that the maternal transmission of Wolbachia is not perfect, Fine [Citation10] introduced the maternal leakage rate μ[0,1) and generalized model (Equation1) to (2) xn+1=(1μ)(1sf)xnshxn2(sf+sh)xn+1,n=0,1,2,,(2) which has also been used [Citation14,Citation27–30] to characterize Wolbachia spreading dynamics in Drosophila simulans during 1990s. Recently, model (Equation2) was revisited in [Citation37]. By introducing the threshold on the maternal leakage rate μ1=(shsf)24sh(1sf),a complete description for the dynamics of model (Equation2) was obtained.

Theorem 1.1

[Citation37]

Model (Equation2) always admits a trivial equilibrium point x0. Furthermore,

(1)

When μ<μ1, model (Equation2) has two positive equilibria x1(μ)=sh+sfD(μ)2sh,x2(μ)=sh+sf+D(μ)2sh,where D(μ)=(shsf)24μsh(1sf). Among which, both x0 and x2(μ) are locally asymptotically stable, while x1(μ) is unstable.

(2)

When μ=μ1, model (Equation2) has a unique positive equilibrium x1(μ1)=x2(μ1)=sh+sf2sh,which is semi-stable: stable from the right side but unstable from the left side, and x0 is locally asymptotically stable.

(3)

When μ>μ1, the unique equilibrium x0 is globally asymptotically stable.

The value μ1 is interpreted as the maximal maternal leakage rate in [Citation37], above which Wolbachia persistence is impossible. For the case when μ<μ1, both models (Equation1) and (Equation2) generate bistable dynamics, with the existence of an unstable equilibrium, which is sf/sh for μ=0, or x1(μ) for μ(0,μ1). When the initial infection frequency x0 is larger than the unstable equilibrium, Wolbachia infection in mosquito population is guaranteed to be persistent. When x0 lies below the unstable equilibrium, wild mosquito populations outcompete the Wolbachia-infected ones. To change the fate of Wolbachia, supplemental releases are needed to guarantee the success of Wolbachia persistence until at some generation n, xn surpasses the unstable equilibrium.

Assume that a proportional release strategy is implemented where both infected females and infected males are released simultaneously at the same ratio r. The next model (3) xn+1=(1μ)(1sf)(1+r)(xn+r)shxn2sf+sh+r(sfsh)xn+1+(2sfsh)r+(1sf)r2,n=0,1,2,(3) was developed in [Citation37] to characterize how supplemental releases affect the Wolbachia infection frequency threshold in [Citation6,Citation10], where r is the constant ratio of infected females/males to the total number of wild females/males at each generation. A release ratio threshold r was found in [Citation37]: for r(0,r), the Wolbachia infection frequency threshold is reduced, and for rr, the threshold is further lowered to 0 which implies that Wolbachia persistence is always successful for any initial infection frequency above 0.

In this paper, we continue to study how supplemental releases affect the Wolbachia spreading dynamics in mosquito populations. Section 2 focuses on the case when only infected males are supplementally released at each generation. This release strategy has been proved to be capable of speeding up the Wolbachia infection by suppressing the compatible matings between uninfected mosquitoes in lab experiments [Citation5]. Detailed model formulation, enumeration of the positive equilibria and their stability analysis are provided. Section 3 studies the case when only infected females are released at each generation. Similar to Section 2, we propose the corresponding discrete model, enumerate the possible equilibria, and analyse their stability. Finally, in Section 4, some numerical simulations are offered to get helpful implications on the design of the release strategy.

2. Releasing infected males with a constant ratio α

Continuous supplemental releases of infected male mosquitos at each generation can promote Wolbachia persistence by suppressing the effective matings between uninfected individuals [Citation5]. In the following, we introduce our first discrete model and give a complete analysis of its dynamics.

2.1. Model formulation

Let InF, InM, UnF and UnM be the numbers of infected females, infected males, uninfected females and uninfected males at the nth generation, respectively. Under the assumption of equal sex determination [Citation2], we have InF=InM,UnF=UnM.Set In=InF+InM and Un=UnF+UnM. Then xn=InIn+Un=InFInF+UnF=InMInM+UnMdefines the infection frequency at the nth generation.

We assume that infected male mosquitoes are released at a ratio α  to the female/male mosquito population size Tn=InF+UnF(=InM+UnM), which means that the number of released Wolbachia-infected males at the nth generation is αTn. Supplemental releases of infected males do not change the infection frequency of females, which is still xn. While the infection frequency of male mosquitos goes from xn to InM+αTnTn+αTn=xn+α1+α.Let Pn+1I and Pn+1U be the proportions of infected and uninfected offspring at the (n+1)th generation, respectively. Then the proportion of infected offspring is Pn+1I=(1μ)(1sf)xn.Under the assumptions of random mating [Citation6] and incomplete CI, the proportion of uninfected offspring Pn+1U contains μ(1sf)xn produced by infected females, (1sh)(1xn)(xn+α)/(1+α) survived from CI, and (1xn)2/(1+α) from matings between uninfected individuals. Hence, we have Pn+1U=μ(1sf)xn+(1sh)(1xn)xn+α1+α+(1xn)1xn1+α.Therefore, a direct computation gives the first discrete model in this paper (4) xn+1=(1μ)(1sf)(1+α)xnshxn2[sf+sh+α(sfsh)]xn+1+α(1sh),n=0,1,2,.(4) Model (Equation4) contains (Equation2) as a special case when α=0. The number of nonnegative equilibria of model (Equation4) and their stability are determined by different combinations of μ and α. In Section 2.2, we divide the parameter region {(μ,α):0μ<1,α>0} into six subregions to study the existence of nonnegative equilibria, respectively. In Section 2.3, we give a complete analysis of the stability of nonnegative equilibria for each case.

2.2. Existence of equilibria

It is easy to see that the origin, denoted by x0, is a boundary equilibrium of (Equation4). For a nontrivial equilibrium of model (Equation4), it satisfies f(x,α)=shx2[sf+sh+α(sfsh)]x+sf+μ(1sf)α[shsfμ(1sf)]=0from (Equation4). Now, we are going to determine the positive roots of f(x,α)=0 lying in (0,1). The discriminant of f(x,α)=0 with respect to x is (5) D(μ,α)=(1+α)[(shsf)24μsh(1sf)+α(shsf)2]=(1+α)[4sh(1sf)(μ1μ)+α(shsf)2]=(1+α)(shsf)2(αα1(μ)),(5) where α1(μ)=μμ11.We have the following result on the sign of D(μ,α).

Lemma 2.1

The following three statements hold:

(i)

D(μ,α)>0 if (μ,α){(μ,α):0<μμ1,α>0}{(μ,α):μ1<μ<1,α>α1(μ)}.

(ii)

D(μ,α1(μ))=0 if μ1<μ<1.

(iii)

D(μ,α)<0 if (μ,α){(μ,α):μ1<μ<1,0<α<α1(μ)}.

Meanwhile, the x-coordinate of the minimum of y=f(x,α) Γx(α)=sh+sfα(shsf)2sh,together with f(1,α)=(1+α)μ(1sf)>0and f(0,α)=sf+μ(1sf)α[shsfμ(1sf)]determine the position and the number of positive solutions of f(x,α)=0 lying in (0,1).

Set α2(μ)=sf+μ(1sf)shsfμ(1sf),α3=sh+sfshsf,μ2=shsf1sf.Then f(0,α) and Γx(α) can be rewritten as f(0,α)=(1sf)(μ2μ)(α2(μ)α) and Γx(α)=shsf2sh(α3α).This leads to the following two lemmas on the signs of f(0,α) and Γx(α).

Lemma 2.2

The following three statements hold:

(i)

f(0,α)>0 if (μ,α){(μ,α):0<μ<μ2,0<α<α2(μ)}{(μ,α):μ2<μ<1,α>0}.

(ii)

f(0,α2(μ))=0 if 0<μ<μ2.

(iii)

f(0,α)<0 if (μ,α){(μ,α):0<μ<μ2,α>α2(μ)}.

Lemma 2.3

Γx(α)>0 for α(0,α3), Γx(α3)=0, and Γx(α)<0 for α>α3.

It's easy to prove that both α1(μ) and α2(μ) are strictly increasing functions, and α1(μ), α2(μ), α3 intersect at point (μ2/2,α3). Figure  divides the μα-plane into six subregions according to the signs of D(μ,α), f(0,α) and Γx(α), from which we can enumerate the positive equilibria of (Equation4) as follows.

Figure 1. The division of the μα-plane depending on the signs of D(μ,α), f(0,α) and Γx(α). It shows that the curves α1(μ), α2(μ) and α3 divide the μα-plane into six subregions: Ω1={(μ,α):D(μ,α)>0,f(0,α)>0,Γx(α)>0}, Ω2={(μ,α):D(μ,α)>0,f(0,α)<0,Γx(α)>0}, Ω3={(μ,α):D(μ,α)>0,f(0,α)<0,Γx(α)<0}, Ω4={(μ,α):D(μ,α)>0,f(0,α)>0,Γx(α)<0}, Ω5={(μ,α):D(μ,α)<0,f(0,α)>0,Γx(α)<0} and Ω6={(μ,α):D(μ,α)<0,f(0,α)>0,Γx(α)>0}. There exist two positive equilibria in subregion Ω1 (yellow), a unique positive equilibrium in subregions Ω2Ω3 and two curves α=α2(μ) for μ(0,μ2/2), and α=α1(μ) for μ(μ1,μ2/2) (red), and no positive equilibria in subregions Ω4Ω5Ω6 together with the curve α=α2(μ) for μ[μ2/2,μ2) (blue).

Figure 1. The division of the μα-plane depending on the signs of D(μ,α), f(0,α) and Γx(α). It shows that the curves α1∗(μ), α2∗(μ) and α3∗ divide the μα-plane into six subregions: Ω1={(μ,α):D(μ,α)>0,f(0,α)>0,Γx(α)>0}, Ω2={(μ,α):D(μ,α)>0,f(0,α)<0,Γx(α)>0}, Ω3={(μ,α):D(μ,α)>0,f(0,α)<0,Γx(α)<0}, Ω4={(μ,α):D(μ,α)>0,f(0,α)>0,Γx(α)<0}, Ω5={(μ,α):D(μ,α)<0,f(0,α)>0,Γx(α)<0} and Ω6={(μ,α):D(μ,α)<0,f(0,α)>0,Γx(α)>0}. There exist two positive equilibria in subregion Ω1 (yellow), a unique positive equilibrium in subregions Ω2∪Ω3 and two curves α=α2∗(μ) for μ∈(0,μ2∗/2), and α=α1∗(μ) for μ∈(μ1∗,μ2∗/2) (red), and no positive equilibria in subregions Ω4∪Ω5∪Ω6 together with the curve α=α2∗(μ) for μ∈[μ2∗/2,μ2∗) (blue).

Theorem 2.1

(1)

Model (Equation4) has two positive equilibria: x1(μ,α)=sh+sfα(shsf)D(μ,α)2sh,x2(μ,α)=sh+sfα(shsf)+D(μ,α)2shif and only if either μ(0,μ1] and α(0,α2(μ)), or μ(μ1,μ2/2) and α(α1(μ),α2(μ)) holds.

(2)

Model (Equation4) has a unique positive equilibrium if one of the next two conditions holds.

(i)

x2(μ,α) when μ(0,μ2/2) and αα2(μ), or μ(μ2/2,μ2) and α>α2(μ).

(ii)

x(μ,α1(μ))=sh+sfα1(μ)(shsf)2sh=shsf2μ(1sf)shsf for μ(μ1,μ2/2).

(3)

Model (Equation4) has no positive equilibria if and only if one of the following four conditions holds. (6) (i)μ(μ1,1) and α(0,α1(μ)).(6) (7) (ii)μ[μ2/2,1) and α=α1(μ).(7) (8) (iii)μ[μ2/2,μ2) and α(α1(μ),α2(μ)].(8) (9) (iv)μ[μ2,1) and α>α1(μ).(9)

2.3. Stability analysis

Before we explore the stability of the nonegative equilibria of model (Equation4), define G(x,α)=a1xshx2b1x+c1,where a1=(1μ)(1sf)(1+α),b1=sf+sh+α(sfsh),c1=1+α(1sh).Then (Equation4) becomes xn+1=G(xn,α). Taking the derivative of G(x,α) with respect to x(0,1), we get (10) G(x,α)x=a1(shx2b1x+c1)a1x(2shxb1)(shx2b1x+c1)2=a1[α(1sh)+(1shx2)](shx2b1x+c1)2>0,(10) which implies that G(x,α) is strictly increasing with respect to x(0,1). We see that equilibria xi(μ,α), i = 1, 2, and x(μ,α1(μ)) satisfy shx2b1x+c1a1=0.Then the derivatives of G(x,α) at x(μ,α1(μ)) and xi(μ,α), i = 1, 2 satisfy (11) G(x,α)x=1x(2shxb1)a1=1±xD(μ,α)a1,(11) which will be used to prove the stability of the positive equilibria of model (Equation4).

Theorem 2.2

If either μ(0,μ1] and α(0,α2(μ)), or μ(μ1,μ2/2) and α(α1(μ),α2(μ)), then both the origin x0 and x2(μ,α) are locally asymptotically stable, while x1(μ,α) is unstable.

Proof.

We first show that x0 is locally asymptotically stable. In fact, since α<α2(μ), we have a1=(1μ)(1sf)(1+α)<1+α(1sh)=c1,which leads to 0<G(x,α)x|x=x0=a1c1<1.Hence, from [Citation1,Citation9], the origin x0 is locally asymptotically stable.

The local asymptotical stability of x2(μ,α) can be obtained from (12) G(x,α)x|x=x2(μ,α)=1x2(μ,α)D(μ,α)a1<1.(12) Meanwhile, G(x,α)x|x=x1(μ,α)=1+x1(μ,α)D(μ,α)a1>1implies the instability of x1(μ,α). This completes the proof.

Theorem 2.3

The following two statements are true.

(1)

If either μ(0,μ2/2) and αα2(μ), or μ(μ2/2,μ2) and α>α2(μ), then the unique positive equilibrium x2(μ,α) is globally asymptotically stable.

(2)

If μ(μ1,μ2/2), then the origin x0 is locally asymptotically stable and the unique positive equilibrium x(μ,α1(μ)) is semi-stable from the right side.

Proof.

(1) The local asymptotical stability of x2(μ,α) is still guaranteed by (Equation10) and (Equation12). To prove the global asymptotical stability of x2(μ,α), we need to show that for any solution of model (Equation4) initiated from x0(0,1), denoted by {xn}n=0={xn(0,x0)}n=0, satisfies (13) limnxn=x2(μ,α).(13) Since x1(μ,α)<0 for μ(0,μ2/2) and αα2(μ), or μ(μ2/2,μ2) and α>α2(μ), from model (Equation4), we have (14) Δxn=xn+1xn=shxn(xnx1(μ,α))shxn2b1xn+c1(xnx2(μ,α)),(14) which further yields xn+1x2(μ,α)=Δxn+xnx2(μ,α)=(b1shx1(μ,α))xn+c1shxn2b1xn+c1(xnx2(μ,α)).Since y(x)=(b1shx1(μ,α))x+c1 is strictly decreasing in x, and y(1)=(1+α)(2sfsh)D(μ,α)2>(1+α)(2sfsh)(1+α)(shsf)2=(1+α)(1sh)>0from (Equation5), we have (xn+1x2(μ,α))(xnx2(μ,α))>0 if xnx2(μ,α). Equation (Equation14) implies that Δxn>0 for x0(0,x2(μ,α)), and Δxn<0 for x0(x2(μ,α),1). Therefore, any solutions {xn}n=0 of (Equation4) initiated from (0,x2(μ,α)) and (x2(μ,α),1) are monotonically increasing and decreasing, respectively. By letting n in (Equation4), we see that (Equation13) holds.

(2) Since μ(μ1,μ2/2), we have D(μ,α1(μ))=0. From (Equation11), we have G(x,α)x|x=x(μ,α1(μ))=1. The asymptotic stability criteria in [Citation1,Citation9] is not applicable. It follows from (Equation4) that Δxn=shxn(xnx(μ,α1(μ)))2shxn2b1xn+c1<0for any xnx(μ,α1(μ)).Similar to the proof above, for any x0(0,x(μ,α1(μ))), solution {xn}n=0 is monotonically decreasing, which proves the local asymptotical stability of the origin x0, as well as the instability of x(μ,α1(μ)) from the left side. Meanwhile, the fact that {xn}n=0 is monotonically decreasing for any x0(x(μ,α1(μ)),1) ensures the stability of x(μ,α1(μ)) from the right side.

Theorem 2.4

The origin x0 is globally asymptotically stable if one of (Equation6)–(Equation9) holds.

Proof.

From the illustration on enumerating the positive equilibria of model (Equation4) in Figure , to prove the global asymptotical stability of the origin x0, we just need to prove that any solution {xn}n=0={xn(0,x0)}n=0 of (Equation4) is monotonically decreasing, where x0(0,1). From model (Equation4), we find that we only need to show that (15) f(x,α)>0 for all x(0,1).(15) In fact, we consider the next four possible cases, μ(μ1,1), α(0,α1(μ)), or μ[μ2/2,1), or μ[μ2/2,μ2), α(α1(μ),α2(μ)], and or μ[μ2,1), α>α1(μ).

For the case when μ(μ1,1) and α(0,α1(μ)), we have D(μ,α)<0. Since f(0,α)>0 and f(x,α) has no zeros for x(0,1), we see that (Equation15) holds.

For the case when μ[μ2/2,1), we get D(μ,α1(μ))=0. Then f(0,α1(μ))0 and Γx(α1(μ))<0 imply that f(x,α1(μ)) has no zeros for x(0,1). Hence (Equation15) holds.

For the case when μ[μ2/2,μ2) and α(α1(μ),α2(μ)], or μ[μ2,1) and α>α1(μ), we obtain D(μ,α)>0. Together with f(0,α)>0 and Γx(α)<0, we find that (Equation15) is also true. This completes the proof.

3. Releasing infected females with a constant ratio β

Population replacement aims to replace a local mosquito population with Wolbachia-infected ones so that their capacity in transmitting disease is reduced, whose implementation requires the release of infected females. For this purpose, we formulate the second discrete model and then analyse its dynamics.

3.1. Model formulation

When supplemental infected females are released with a constant ratio β to the total number of male/female mosquitoes Tn=InM+UnM(=InF+UnF), the infection frequency of males is still xn, while the infection frequency of females increases from xn to InF+βTnTn+βTn=xn+β1+β.The proportion of infected mosquitoes at the (n+1)th generation is Pn+1I=(1μ)(1sf)xn+β1+β,since Pn+1I does not depend on the parental infection status. On Pn+1U, taking the imperfect maternal transmission and incomplete CI into consideration, we have Pn+1U=μ(1sf)xn+β1+β+(1sh)xn1xn1+β+(1xn)21+β,where μ(1sf)(xn+β)/(1+β) counts the proportion from infected females owing to maternal leakage, (1sh)xn(1xn)/(1+β) is the proportion survived from CI, and (1xn)2/(1+β) represents the proportion from uninfected matings. Therefore, the second discrete model in this paper is expressed as (16) xn+1=(1μ)(1sf)(β+xn)shxn2(sf+sh)xn+1+β(1sf),n=0,1,2,.(16)

3.2. Existence of equilibria

For model (Equation16), a positive equilibrium lying in [0,1) satisfies g(x,β)=shx3(sf+sh)x2+[sf+(1sf)(μ+β)]x(1μ)(1sf)β=0.We now investigate the zeros of g(x,β)=0 in [0,1). For any β>0, we get g(0,β)=(1μ)(1sf)β<0  and  g(1,β)=μ(1sf)(1+β)>0,which imply that equation g(x,β)=0 has at least one solution in (0,1), i.e.

Lemma 3.1

Model (Equation16) has at least one equilibrium in (0,1).

To determine the number of the solutions of equation g(x,β)=0 lying in (0,1), we explore the monotonicity of function g(x,β) with respect to β. It follows from gβ(x,β)=(1sf)[x(1μ)]that g(x,β) is strictly decreasing with respect to β for x<1μ. Let x3(μ,β) be the largest positive equilibrium of model (Equation16) lying in (0,1), we claim that (17) x3(μ,β)<1μ for β0.(17) In fact, let H(x,β)=(1μ)(1sf)(β+x)shx2(sf+sh)x+1+β(1sf).Taking the derivative of H(x,β) with respect to x(0,1), we have H(x,β)x=(1μ)(1sf)1shx2+β[1shx+sh(1x)][shx2(sf+sh)x+1+β(1sf)]2>0,which implies that H(x,β) is strictly increasing with respect to x(0,1). Hence, x3(β)=H(x3(β),β)<H(1,β)=1μ for β0 and (Equation17) holds. To sum up, we get

Lemma 3.2

Let x3(μ,β)<1μ be the largest positive equilibrium of model (Equation16) lying in (0,1). Then function g(x,β) is strictly decreasing with respect to β for x<x3(μ,β).

Particularly, since g(x,0)=x[shx2(sf+sh)x+sf+μ(1sf)]=shxxsf+sh2sh2+4shμ(1sf)(shsf)24sh2=shxxsf+sh2sh2+4sh(1sf)(μμ1)4sh2,there are three possible cases to consider.

3.2.1. The case when μ(0,μ1)

For the case when μ(0,μ1), function g(x,0) has three zeros lying in [0,1): x0=0, x1(μ)=sh+sfD(μ)2sh, and x2(μ)=sh+sfD(μ)2sh,where D(μ) is defined in Theorem 1.1 and positive. From Lemma 3.2, there exists a unique β1 (see Figure  for illustration) such that

Figure 2. Given sf=0.1 and sh=0.9, we have μ10.1975. For the case μ=0.15<μ1, we get x1(μ)0.3375, x2(μ)0.7736. Numerical trials imply that β10.0255. Taking β=0.01<β1, we have x1(μ,β)0.0367, x2(μ,β)0.2987 and x3(μ,β)0.7758. At β1, x1(μ,β1)=x2(μ,β1)0.1661 and x3(μ,β1)0.7788. Furthermore, when increasing β to 0.004, both x1(μ,β) and x2(μ,β) vanish, and x3(μ,β)0.7815.

Figure 2. Given sf=0.1 and sh=0.9, we have μ1∗≈0.1975. For the case μ=0.15<μ1∗, we get x1∗(μ)≈0.3375, x2∗(μ)≈0.7736. Numerical trials imply that β1∗≈0.0255. Taking β=0.01<β1∗, we have x1∗(μ,β)≈0.0367, x2∗(μ,β)≈0.2987 and x3∗(μ,β)≈0.7758. At β1∗, x1∗(μ,β1∗)=x2∗(μ,β1∗)≈0.1661 and x3∗(μ,β1∗)≈0.7788. Furthermore, when increasing β to 0.004, both x1∗(μ,β) and x2∗(μ,β) vanish, and x3∗(μ,β)≈0.7815.

Theorem 3.1

On enumerating the positive equilibria of model (Equation16) for μ(0,μ1), we have

(1)

If β(0,β1), then model (Equation16) admits three positive equilibria, denoted by x1(μ,β), x2(μ,β) and x3(μ,β). Furthermore, we have 0<x1(μ,β)<x2(μ,β)<x1(μ)<x2(μ)<x3(μ,β)<1μ.

(2)

If β=β1, then model (Equation16) admits two positive equilibria x1(μ,β1)=x2(μ,β1) and x3(μ,β1).

(3)

If β>β1, then model (Equation16) admits a unique positive equilibrium x3(μ,β).

3.2.2. The case when μ=μ1

For the case when μ=μ1, function g(x,0) has two zeros lying in [0,1) which are x0=0, x1(μ1)=x2(μ1)=sh+sf2sh.Again, from Lemma 3.2, there exists a unique β2 (see Figure  for illustration) such that

Figure 3. Given sf=0.1 and sh=0.9, we take μ=μ10.1975. When β=0, x1(μ1) and x2(μ1) coincide to x1(μ1)=x2(μ1)0.5556. Numerical trials offer β20.0426. When β=0.02<β2, x1(μ1,β)0.0606, x2(μ1,β)0.4209 and x3(μ1,β)0.6296. When β=β2, both x1(μ1,β2) and x2(μ1,β2) coincide to x1(μ1,β2)=x2(μ1,β2)0.2280 and x3(μ1,β2)0.6539. For β=0.08>β2, x3(μ1,β)0.6772.

Figure 3. Given sf=0.1 and sh=0.9, we take μ=μ1∗≈0.1975. When β=0, x1∗(μ1∗) and x2∗(μ1∗) coincide to x1∗(μ1∗)=x2∗(μ1∗)≈0.5556. Numerical trials offer β2∗≈0.0426. When β=0.02<β2∗, x1∗(μ1∗,β)≈0.0606, x2∗(μ1∗,β)≈0.4209 and x3∗(μ1∗,β)≈0.6296. When β=β2∗, both x1∗(μ1∗,β2∗) and x2∗(μ1∗,β2∗) coincide to x1∗(μ1∗,β2∗)=x2∗(μ1∗,β2∗)≈0.2280 and x3∗(μ1∗,β2∗)≈0.6539. For β=0.08>β2∗, x3∗(μ1∗,β)≈0.6772.

Theorem 3.2

Assume that μ=μ1. Then the following three statements are true.

(1)

If β(0,β2), then model (Equation16) admits three positive equilibria x1(μ1,β), x2(μ1,β) and x3(μ1,β) with 0<x1(μ1,β)<x2(μ1,β)<x1(μ1)=x2(μ1)<x3(μ1,β)<1μ1.

(2)

If β=β2, then model (Equation16) admits two positive equilibria x1(μ1,β2)=x2(μ1,β2) and x3(μ1,β2).

(3)

If β>β2, then model (Equation16) admits a unique positive equilibrium x3(μ1,β).

3.2.3. The case when μ>μ1

For the case when μ>μ1, we have g(x,0)>0 for x(0,1). As β  increases, from Lemma 3.2, there exist β3 and β4 (see Figure  for illustration) such that

Figure 4. Given sf=0.1 and sh=0.9, we take μ=0.2>μ10.1975. Numerical simulations show that β30.0057, β40.0437. The number of zeros of g(x,β) lying in (0,1) goes from 1, passing 2, 3, 2, and finally to 1 as β increases from 0 to the β with β>β4.

Figure 4. Given sf=0.1 and sh=0.9, we take μ=0.2>μ1∗≈0.1975. Numerical simulations show that β3∗≈0.0057, β4∗≈0.0437. The number of zeros of g(x,β) lying in (0,1) goes from 1, passing 2, 3, 2, and finally to 1 as β increases from 0 to the β with β>β4∗.

Theorem 3.3

Assume that μ>μ1. Then the following three statements are true.

(1)

If β(0,β3)(β4,+), then model (Equation16) admits a unique positive equilibrium. The unique positive equilibrium is x1(μ,β) for β(0,β3), or x3(μ,β) for β(β4,+).

(2)

If β=β3 (or β=β4), then model (Equation16) admits two positive equilibria x1(μ,β3)=x2(μ,β3), x3(μ,β4) (or x1(μ,β), x2(μ,β4)=x3(μ,β4)).

(3)

If β(β3,β4), then model (Equation16) admits three positive equilibria x1(μ,β), x2(μ,β) and x3(μ,β) satisfying 0<x1(μ,β3)<x1(μ,β)<x1(μ,β4)=x2(μ,β4)<x2(μ,β)<x2(μ,β3)=x3(μ,β3)<x3(μ,β)<x3(μ,β4)<1μ.

3.3. Stability analysis

In this section, we investigate the stability of the positive equilibria of model (Equation16). The following first result generates a bistable dynamics for the case when there exist three positive equilibria.

Theorem 3.4

If (μ,β){(μ,β):0<μ<μ1,β(0,β1)}{(μ,β):μ=μ1,β(0,β2)}{(μ,β):μ>μ1,β3<β<β4}, then model (Equation16) has three positive equilibria x1(μ,β), x2(μ,β) and x3(μ,β), where both x1(μ,β) and x3(μ,β) are locally asymptotically stable, while x2(μ,β) is unstable.

Proof.

We first show that x1(μ,β) is locally asymptotically stable. For any initial value x0(0,x1(μ,β)), from (Equation16), it is easy to see that Δx0>0 and hence x1>x0. Therefore, we reach xn+1>xn by induction for all n0, which means that solution {xn}n=0 monotonically increases to x1(μ,β) if x0(0,x1(μ,β)). Similarly, we can prove that any solutions initiated from (x1(μ,β),x2(μ,β)) monotonically decrease to x1(μ,β), and any solutions initiated from (x2(μ,β),x3(μ,β)) (or (x3(μ,β),1)) monotonically increase (or decrease) to x3(μ,β). While the instability of x2(μ,β) is obvious. The proof is finished.

For the case when (μ,β){(μ,β):0<μ<μ1,β=β1}{(μ,β):μ=μ1,β=β2}{(μ,β):μ>μ1,β=β4}, equilibria x1(μ,β) and x2(μ,β) shrink to one, which is semi-stable. While for (μ,β){(μ,β):μ>μ1,β=β3}, x2(μ,β) and x3(μ,β) coincide, which is also semi-stable. To sum up, we have

Theorem 3.5

The following two statements are true.

(1)

If (μ,β){(μ,β):0<μ<μ1,β=β1}{(μ,β):μ=μ1,β=β2}{(μ,β):μ>μ1,β=β4}, then x3(μ,β) is locally asymptotically stable, and x1(μ,β)=x2(μ,β) is semi-stable from the right side.

(2)

If (μ,β){(μ,β):μ>μ1,β=β3}, then x2(μ,β)=x3(μ,β) is semi-stable from the left side, and x1(μ,β) is locally asymptotically stable.

The following theorem indicates that the unique positive equilibrium is globally asymptotically stable.

Theorem 3.6

If (μ,β){(μ,β):0<μ<μ1,β>β1}{(μ,β):μ=μ1,β>β2}{(μ,β):μ>μ1,0<β<β3}{(μ,β):μ>μ1,β>β4}, then the unique positive equilibrium is globally asymptotically stable.

4. Discussions

4.1. Dynamics driven by (4) and (16)

Two discrete models (Equation4) and (Equation16) are formulated to study the dynamics of Wolbachia infection frequency in cage mosquito populations. Model (Equation4) aims to the infection frequency when only infected males are released at every generation with a constant release ratio α. For given sf and sh, enumeration of the positive equilibria of (Equation4) is offered in Theorem 2.1, depending on the values of the maternal leakage rate μ and α relative to μ1, μ2, α1(μ) and α2(μ). Theorem 2.2 shows that (Equation4) generates a bistable dynamics when there exist two positive equilibria. When the positive equilibrium is unique, Theorem 2.3 shows that it is either globally asymptotically stable or semi-stable. When there is no positive equilibria, Theorem 2.4 proves the global asymptotical stability of the origin x0.

Regarding the situation when only infected females are supplementally released, model (Equation16) introduced the constant release ratio β. By using the maximal leakage rate μ1 deduced in [Citation37]. The existence of the positive equilibria of model (Equation16) is characterized in Theorems 3.1–3.3, along the existence of four thresholds of β, denoted by β1 in Theorem 3.1, β2 in Theorem 3.2, as well as β3 and β4 with β3<β4 in Theorem 3.3. Different from model (Equation4), model (Equation16) has no the origin x0. Bistable dynamics occurs when there are three positive equilibria. Let xi(μ,β), i = 1, 2, 3 satisfying x1(μ,β)x2(μ,β)x3(μ,β) denote the three positive equilibria of (Equation16). Theorem 3.4 manifests that both x1(μ,β) and x3(μ,β) are locally asymptotically stable, while x2(μ,β) is unstable. Theorem 3.5 shows that if x1(μ,β) equals to x2(μ,β) at β=βi, i = 1, 2, 4, then x1(μ,β)=x2(μ,β) is semi-stable and x3(μ,β) is locally asymptotically stable. Also, when x2(μ,β) equals to x3(μ,β) at β=β3, x1(μ,β3) is locally asymptotically stable and x2(μ,β3)=x3(μ,β3) is semi-stable.

We take sf=0.1 and sh=0.9 as an example to illustrate our theoretical results. The parameters μ, α and β are chosen so that both models (Equation4) and (Equation16) generate bistable dynamics. In this case, we have μ10.1975,α20.3534,β10.0073.From Theorems 2.1 and 3.1, when we take μ=0.15(0,μ1), α=0.01(0,α2), and β=0.01(0,β1), both model (Equation4) and model (Equation16) admit three equilibria in [0,1), which are shown in Figure . Model (Equation16) generates a lower infection frequency threshold with x2(μ,β)0.2987<0.3275x1(μ,α), and a slightly higher polymorphic infection frequency with x3(μ,β)0.7758>0.7747x2(μ,α). This observation implies that releasing infected females is more efficient than releasing infected males at the same constant ratio.

Figure 5. Distable dynamics driven by model (Equation4) and model (Equation16). Panel (A) is for model (Equation4) and Panel (B) is for model (Equation16).

Figure 5. Distable dynamics driven by model (Equation4(4) xn+1=(1−μ)(1−sf)(1+α)xnshxn2−[sf+sh+α(sf−sh)]xn+1+α(1−sh),n=0,1,2,….(4) ) and model (Equation16(16) xn+1=(1−μ)(1−sf)(β+xn)shxn2−(sf+sh)xn+1+β(1−sf),n=0,1,2,….(16) ). Panel (A) is for model (Equation4(4) xn+1=(1−μ)(1−sf)(1+α)xnshxn2−[sf+sh+α(sf−sh)]xn+1+α(1−sh),n=0,1,2,….(4) ) and Panel (B) is for model (Equation16(16) xn+1=(1−μ)(1−sf)(β+xn)shxn2−(sf+sh)xn+1+β(1−sf),n=0,1,2,….(16) ).

4.2. Comparisons on three release strategies introduced in (3), (4) and (16)

The above observation that model (Equation16) performs better than model (Equation4) for μ=0.15 and α=β=0.01 is not a special case, but a general one. To see this, we plot the infection frequency thresholds driven by model (Equation3) with r = 0.0005, model (Equation16) with β=0.0005 and model (Equation4) with α=0.0005 for μ(0,μ1), sf=0.1 and sh=0.9 in Figure (A), the infection frequency threshold generated from model (Equation3) is the smallest, while the release strategy with only infected males released requires the largest threshold for Wolbachia fixation. Meanwhile, Figure (B) plots the polymorphic states (the largest positive equilibria) for μ under the three release strategies driven by (Equation3), (Equation16) and (Equation4), respectively. It shows that releasing both infected females and males brings the Wolbachia to fix at the highest infection level. And when only infected males are supplementally released, the Wolbachia infection frequency will fix at the lowest one. For the three release strategies, the increase of μ pulls the infection frequency thresholds higher, and drags down the Wolbachia fixation frequencies.

Figure 6. Comparisons on the infection frequency thresholds (A) and the polymorphic states (B) driven by (Equation3), (Equation16) and (Equation4) on different maternal leakage rates μ lying in (0,μ1).

Figure 6. Comparisons on the infection frequency thresholds (A) and the polymorphic states (B) driven by (Equation3(3) xn+1=(1−μ)(1−sf)(1+r)(xn+r)shxn2−sf+sh+r(sf−sh)xn+1+(2−sf−sh)r+(1−sf)r2,n=0,1,2,…(3) ), (Equation16(16) xn+1=(1−μ)(1−sf)(β+xn)shxn2−(sf+sh)xn+1+β(1−sf),n=0,1,2,….(16) ) and (Equation4(4) xn+1=(1−μ)(1−sf)(1+α)xnshxn2−[sf+sh+α(sf−sh)]xn+1+α(1−sh),n=0,1,2,….(4) ) on different maternal leakage rates μ lying in (0,μ1∗).

4.3. Implications on the design of release strategy

Regarding the Wolbachia persistence in cage experiments, the two most crucial concerns are: (1) how fast the persistence is when the initial infection frequency is larger than the infection frequency threshold. (2) how supplemental releases of infected mosquitoes lower the infection frequency threshold to make the persistence achievable. To answer such two questions numerically, let sf=0.1, sh=0.9, μ=0.05.Numerical trials indicate that α20.1921 in Theorem 2.1, β10.0073 in Theorem 3.1, and r0.00726 in Theorem 3.1 in [Citation37]. Take α=β=r=0.005, we have x1(μ,α)0.1670,x2(μ,β)0.1335,x2(μ,r)0.1347,x2(μ,α)0.9397,x3(μ,β)0.9397,x3(μ,r)0.9340.Hence, if we define x2=max{x1(μ,α),x2(μ,β),x2(μ,r)}0.1670,x3=min{x2(μ,α),x3(μ,β),x3(μ,r)}0.9340,then solutions initiated from (0.18,0.92)(x2,x3) will eventually go to Wolbachia fixation under three release strategies modelled by (Equation3), (Equation16) and (Equation4). To estimate the persistence speed, we numerically find the generation, denoted by N, at which the infection frequency surpasses 0.93 for the first time. Take model (Equation16) as an example, if we let {xn}n=0 be the solution of (Equation16) initiated from x0(0.18,0.92), then N satisfies xn0.93for n=1,2,,N1, but xN>0.93.Following this procedure, we plot the curves of N by randomly selecting initial values in (0.18,0.92) in Figure (A), which shows that among these three release strategies driven by (Equation3), (Equation16) and (Equation4), the fastest to reach persistence is the release of both infected females and males, followed by the release of only infected females, and the lowest is the release of only infected males.

Figure 7. Implications on the design of release strategy. (A) The generation N  to reach xN>0.93 shows a step-like decrease as the increase of the initial values. (B) Under three release strategies, the infection frequency thresholds are monotonically decreasing with respect to the release ratios.

Figure 7. Implications on the design of release strategy. (A) The generation N  to reach xN>0.93 shows a step-like decrease as the increase of the initial values. (B) Under three release strategies, the infection frequency thresholds are monotonically decreasing with respect to the release ratios.

We end the whole manuscript with numerical trials for answering the second question, i.e. how supplemental releases of infected mosquitoes lower the infection frequency threshold to make the persistence achievable. To this end, still letting sf=0.1, sh=0.9, μ=0.05, we plot the infection frequency thresholds for the release ratios lying in (0,0.007) in Figure (B). Here we take 0.007 to guarantee the existence of the thresholds under three release strategies. And numerical observation agree with our theoretical results that higher release ratios lead to lower infection frequency thresholds to guarantee the success of Wolbachia persistence.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by National Natural Science Foundation of China [grant numbers 12071095,11971127, 11631005], Program for Changjiang Scholars and Innovative Research Team in University [grant number IRT_16R16] and Innovative Research Grant for the Postgraduates of Guangzhou University [grant number 2020GDJC-D10].

References

  • C. Ahlbrandt and A.C. Peterson, Discrete Hamiltonian System, Springer, Boston, MA, 2010.
  • H.N. Aida, H. Dieng, A.T. Nurita, M.C. Salmah, F. Miake, and B. Norasmah, The biology and demographic parameters of Aedes albopictus in northern peninsular Malaysia, Asian Pac. J. Trop. Biomed. 1(6) (2011), pp. 472–477.
  • F. Arkin, Dengue researcher faces charges in vaccine fiasco, Science 364 (2019), pp. 320.
  • F. Baldacchino, B. Caputo, F. Chandre, A. Drago, A. Della Torre, F. Montarsi, and A. Rizzolli, Control methods against invasive Aedes mosquitoes in Europe: A review, Pest Manag. Sci. 71(11) (2015), pp. 1471–1485.
  • G. Bian, D. Joshi, Y. Dong, P. Lu, G. Zhou, X. Pan, Y. Xu, G. Dimopoulos, and Z. Xi, Wolbachia invades Anopheles stephensi populations and induces refractoriness to plasmodium infection, Science340 (2013), pp. 748–751.
  • E. Caspari and G.S. Watson, On the evolutionary importance of cytoplasmic sterility in mosquitoes, Evolution 13(4) (1959), pp. 568–570.
  • J. Cohen, Dengue may bring out the worst in Zika, Science 356(6332) (2017), pp. 175–180.
  • W. Dejnirattisai, P. Supasa, W. Wongwiwat, A. Rouvinski, G. Barba-Spaeth, T. Duangchinda, A.Sakuntabhai, V.M. Cao-Lormeau, P. Malasit, and F.A. Rey, Dengue virus sero-cross-reativity drives antibody-dependent enhancement of infection with Zika virus, Nat. Immunol. 17 (2016), pp. 1102–1109.
  • S. Elaydi, An Introduction to Difference Equations, Springer, New York, 2005.
  • P.E.M. Fine, On the dynamics of symbiote-dependent cytoplasmic incompatibility in Culicine mosquitoes, J. Invertebr. Pathol. 31(1) (1978), pp. 10–18.
  • J. Guarnera and G.L. Hale, Four human diseases with significant public health impact caused by mosquito-borne flaviviruses: West Nile, Zika, dengue and yellow fever, Semin. Diagn. Pathol. 36(3) (2019), pp. 170–176.
  • M. Hertig and S.B. Wolbach, Studies on Rickettsia-like micro-organisms in insects, J. Med. Res. 44(3) (1924), pp. 329–374.
  • A.A. Hoffmann, B.L. Montgomery, J. Popovici, I. Iturbe-Ormaetxe, P. Johnson, F. Muzzi, M. Greenfield, M. Durkan, Y.S. Leong, Y. Dong, H. Cook, J. Axford, A.G. Callahan, N. Kenny, C. Omodei, E.A. McGraw, P.A. Ryan, S.A. Ritchie, M. Turelli, and S.L. O'Neill, Successful establishment of Wolbachia in Aedes populations to suppress dengue transmission, Nature 476(7361) (2011), pp. 454–457.
  • A.A. Hoffmann, M. Turelli, and L.G. Harshman, Factors affecting the distribution of cytoplasmic incompatibility in Drosophila simulans, Genetics 126(4) (1990), pp. 933–948.
  • L. Hu, M. Huang, M. Tang, J. Yu, and B. Zheng, Wolbachia spread dynamics in stochastic environments, Theor. Popul. Biol. 106 (2015), pp. 32–44.
  • L. Hu, M. Tang, Z. Wu, Z. Xi, and J. Yu, The threshold infection level for Wolbachia invasion in radom environment, J. Differ. Equ. 266 (2019), pp. 4377–4393.
  • M. Huang, L. Hu, and J. Yu, Wolbachia infection dynamics by reaction-diffusion equations, Sci. China Math. 58(1) (2015), pp. 77–96.
  • M. Huang, L. Hu, and B. Zheng, Comparing the efficiency of Wolbachia driven Aedes mosquito suppression strategies, J. Appl. Anal. Comput. 9(1) (2019), pp. 211–230.
  • M. Huang, J. Luo, L. Hu, B. Zheng, and J. Yu, Assessing the efficiency of Wolbachia driven Aedes mosquito suppression by delay differential equations, J. Theor. Biol. 440(7) (2018), pp. 1–11.
  • M. Huang, M. Tang, J. Yu, and B. Zheng, The impact of mating competitiveness and incomplete cytoplasmic incompatibility on Wolbachia-driven mosquito population suppression, Math. Biosci. Eng.16(5) (2019), pp. 4741–4757.
  • M. Huang, M. Tang, J. Yu, and B. Zheng, A stage structured model of delay differential equations for Aedes mosquito population suppression, Discrete Contin. Dyn. Syst. 40(6) (2020), pp. 3467–3484.
  • Y. Hui, G. Lin, J. Yu, and J. Li, A delayed differential equation model for mosquito population suppression with sterile mosquitoes, Discrete Contin. Dyn. Syst. Ser. B 25(12) (2020), pp. 4659–4676.
  • H. Laven, Cytoplasmic inheritance in Culex, Nature 177 (1956), pp. 141–142.
  • C.J. Mcmeniman, R.V. Lane, B.N. Cass, A. Fong, M. Sidhu, Y. Wang, and S. O'Neill, Stable introduction of a life-shortening Wolbachia infection into the mosquito Aedes aegypi, Science 323(5910) (2009), pp. 141–144.
  • Y. Shi and J. Yu, Wolbachia infection enhancing and decaying domains in mosquito population based on discrete models, J. Biol. Dyn. 14(1) (2020), pp. 679–695.
  • P. Somwang, J. Yanola, W. Suwan, C. Walton, N. Lumjuan, L. Prapanthadara, and P. Somboon, Enzymes-based resistant mechanism in pyrethroid resistant and susceptible Aedes aegypti strains from northern Thailand, Parasitol. Res. 109(3) (2011), pp. 531–537.
  • M. Turelli, Evolution of incompatibility-inducing microbes and their hosts, Evolution 48(5) (1994), pp. 1500–1513.
  • M. Turelli, Cytoplasmic incompatibility in populations with overlapping generations, Evolution 64(1) (2010), pp. 232–241.
  • M. Turelli and A.A. Hoffmann, Rapid spread of an inherited incompatibility factor in California Drosophila, Nature 353(6343) (1991), pp. 440–442.
  • M. Turelli and A.A. Hoffmann, Cytoplasmic incompatibility in Drosophila simulans: Dynamics and parameter estimates from natural populations, Genetics 140(4) (1995), pp. 1319–1338.
  • Y. Wang, X. Liu, C. Li, T. Su, J. Jin, Y. Guo, D. Ren, Z. Yang, Q. Liu, and F. Meng, A survey of insecticide resistance in Aedes albopictus (Diptera: Culicidae) during a 2014 dengue fever outbreak in Guangzhou, China, J. Econ. Entomol. 110(1) (2017), pp. 239–244.
  • Z. Xi and S.L. Dobson, Characterization of Wolbachia transfection efficiency by using microinjection of embryonic cytoplasm and embryo homogenate, Appl. Environ. Microbiol. 71(6) (2005), pp. 3199–3204.
  • J. Yu, Modelling mosquito population suppression based on delay differential equations, SIAM J. Appl. Math. 78(6) (2018), pp. 3168–3187.
  • J. Yu, Existence and stability of a unique and exact two periodic orbits for an interactive wild and sterile mosquito model, J. Differ. Equ. 269(12) (2020), pp. 10395–10415.
  • J. Yu and J. Li, Dynamics of interactive wild and sterile mosquitoes with time delay, J. Biol. Dyn.13(4) (2019), pp. 1–15.
  • J. Yu and J. Li, Global asymptotic stability in an interactive wild and sterile mosquito model, J. Differ. Equ. 269(7) (2020), pp. 6193–6215.
  • J. Yu and B. Zheng, Modeling Wolbachia infection in mosquito population via discrete dynamical models, J. Differ. Equ. Appl. 25(11) (2019), pp. 1549–1567.
  • B. Zheng, W. Guo, L. Hu, M. Huang, and J. Yu, Complex Wolbachia infection dynamics in mosquitoes with imperfect maternal transmission, Math. Biosci. Eng. 15(2) (2018), pp. 523–541.
  • B. Zheng, M. Tang, and J. Yu, Modeling Wolbachia spread in mosquitoes through delay differential equations , SIAM J. Appl. Math. 74(3) (2014), pp. 743–770.
  • B. Zheng, X. Liu, M. Tang, Z. Xi, and J. Yu, Use of age-stage structural models to seek optimal Wolbachia-infected male mosquito releases for mosquito-borne disease control, J. Theor. Biol. 472(7) (2019), pp. 95–109.
  • B. Zheng, M. Tang, and J. Yu, Modeling Wolbachia spread in mosquitoes through delay differential equations, SIAM J. Appl. Math. 74(3) (2014), pp. 743–770.
  • B. Zheng, M. Tang, J. Yu, and J. Qiu, Wolbachia spreading dynamics in mosquitoes with imperfect maternal transmission, J. Math. Biol. 76(1–2) (2018), pp. 235–263.
  • B. Zheng and J. Yu, Existence and uniqueness of periodic orbits in a discrete model on Wolbachia infection frequency, Adv. Nonlinear Anal. 11 (2022), pp. 212–224.
  • B. Zheng, J. Yu, and J. Li, Modeling and analysis of the implementation of the Wolbachia incompatible and sterile insect technique for mosquito population suppression, SIAM J. Appl. Math. 81(2) (2021), pp. 718–740.
  • B. Zheng, J. Yu, Z. Xi, and M. Tang, The annual abundance of dengue and Zika vector Aedes albopictus and its stubbornness to suppression, Ecol. Model. 387(10) (2018), pp. 38–48.
  • X. Zheng, D. Zhang, Y. Li, C. Yang, Y. Wu, X. Liang, Y. Liang, X. Pan, L. Hu, Q. Sun, X. Wang, Y. Wei, J. Zhu, W. Qian, Z. Yan, A. Parker, J. Giles, K. Bourtzis, J. Bouyer, M. Tang, B. Zheng, J. Yu, J. Liu, J. Zhuang, Z. Hu, M. Zhang, J. Gong, X. Hong, Z. Zhang, L. Lin, Q. Liu, Z. Hu, Z. Wu, L. Baton, A. Hoffmann, and Z. Xi, Incompatible and sterile insect techniques combined eliminate mosquitoes, Nature 572(7767) (2019), pp. 56–61.