986
Views
3
CrossRef citations to date
0
Altmetric
2019 Guangzhou Workshop

Modeling the suppression dynamics of Aedes mosquitoes with mating inhomogeneity

&
Pages 656-678 | Received 09 Sep 2019, Accepted 13 Jul 2020, Published online: 04 Aug 2020

Abstract

A novel strategy for controlling mosquito-borne diseases, such as dengue, malaria and Zika, involves releases of Wolbachia-infected mosquitoes as Wolbachia cause early embryo death when an infected male mates with an uninfected female. In this work, we introduce a delay differential equation model with mating inhomogeneity to discuss mosquito population suppression based on Wolbachia. Our analyses show that the wild mosquitoes could be eliminated if either the adult mortality rate exceeds the threshold δA or the release amount exceeds the threshold r uniformly. We also present the nonlinear dependence of δA and r on the parameters, respectively, as well as the effect of pesticide spraying on wild mosquitoes. Our simulations suggest that the releasing should be started at least 5 weeks before the peak dengue season, taking into account both the release amount and the suppression speed.

2010 Mathematics Subject Classifications:

This article is part of the following collections:
Mathematical Modeling and Analysis of Populations and Infectious Diseases

1. Introduction

With the global aggressive invading of mosquitoes by urbanization, international trades and travels, mosquito-borne diseases, such as dengue, malaria and Zika have become a major public health concern. Dengue, increasing more than 30-fold globally in the past few decades, infects over 390 million people annually in the tropical and subtropical regions [Citation4,Citation42]. The mechanism of antibody-dependent enhancement greatly increases the difficulty of vaccine development. It was reported that 130 vaccinated children had died after using Dengvaxia, the first registered dengue vaccine [Citation1]. Thus the key measure to prevent mosquito-borne diseases is to reduce the population of mosquito vectors. Pesticide application has already been the primary method for mosquito eradication in the world. To defeat the unprecedented dengue outbreak in Guangzhou in 2014, more than 27,000 kg pyrethroids were sprayed in over 3291 km2 of land [Citation41]. However, pesticide spraying is not a perfect solution as it also leads to environmental pollution and insecticide resistance of mosquitoes [Citation25,Citation34].

An innovative technique for the control of mosquito-borne diseases involves Wolbachia, which is a genus of common intracellular bacterial endosymbionts and was first identified in 1924 [Citation13,Citation35]. The infection of Wolbachia can reduce host's ability to reproduce and transmit disease-causing viruses [Citation5,Citation39,Citation40,Citation43]. Due to the mechanism cytoplasmic incompatibility (CI), the insects cannot produce offspring when infected males cross with field females which are not infected with the same strain [Citation11,Citation27]. Thus repeated release of Wolbachia-infected males in the field could suppress the wild mosquitoes, and a preparatory work is to clarify the suppression dynamics, i.e. the dynamical changes of wild mosquito population under the intervention of infected males.

Mosquito population suppression has become a very active research topic in recent years and various mathematical models have been formulated and analysed for studying the wild mosquito population suppression effects on the releases of sterile male mosquitoes as well as Wolbachia-infected male mosquitoes, for example, difference equation models [Citation47,Citation51], ordinary differential equation models [Citation7,Citation12,Citation50], delay differential equation models [Citation2,Citation6,Citation19,Citation20,Citation23], and reaction-diffusion equation models [Citation17,Citation18,Citation32], etc. By combining sterile insect techniques and Wolbachia, Zheng et al. have successfully eradicated more than 90% of Aedes albopictus in Shazai island in Guangzhou recently [Citation52], which verifies the feasibility of mosquito population suppression in the field. Because in the interaction dynamics, the fundamental and only role for sterile male mosquitoes or Wolbachia-infected male mosquitoes is to mate with wild female mosquitoes, so that the mating wild female mosquitoes either fail to reproduce or lay eggs that do not hatch. The author in [Citation44] proposed a new modelling idea of mosquito population inhibition that is completely different from the existing modelling approaches, where R(t), the number of sterile male mosquitoes or Wolbachia-infected male mosquitoes released at time t, is just regarded as a given function instead of an independent variable satisfying a dynamic equation. This idea has been well developed and used recently, see for example [Citation20–22,Citation28,Citation45,Citation46]. In [Citation45], the authors only considered those sexually active sterile mosquitoes in the model and ignored the death of those sterile mosquitoes, and clearly explained why the case when R(t)r, a constant, is possible, we also refer to [Citation28]. We know that once sterile mosquitoes lose their mating competitiveness, they have no effect on the wild mosquito population suppression, even if they are still alive. In this work, we continue to follow such an idea in [Citation44] and focus on the classical Allee effect to formulate and analyse a mosquito population suppression model with time delay. Let A(t) be the size of wild adult mosquitoes, distributing equally in sex [Citation14,Citation15,Citation51], and uniformly in space [Citation17,Citation18]. The wild mosquitoes are interfered by releasing Wolbachia-infected males with size R(t) which induce complete CI to wild females [Citation16,Citation49]. By random mating, for a single female, the chance of incompatible mating equals to the number of infected males R(t) over the total number of males A(t)/2+R(t). Let τ be the average development period from oviposition to the eclosion of adults, and δI be the natural mortality rate of the immature stage. Hence the birth rate of a female adult is given by (1) eδIτA(tτ)/2A(tτ)/2+R(tτ)χ(A(tτ)/2+R(tτ)),(1) where eδIτ denotes the mean survival probability from egg to adult and χ(A/2+R) the average number of adult offsprings by a female. As the opportunity of a female to mate with a male in a small population is lower than that in a large population, we follow the idea of the classical Allee effect to describe the mating inhomogeneity [Citation7]. Let β be the average number of larvae by a female and assume (2) χ(A/2+R)=βA/2+Rα+A/2+R,(2) where α is a positive constant. We use the logistic model to describe the mortality rate of adults with natural mortality rate m, and a constraint constant K associated with density-dependent competition [Citation19,Citation20,Citation44] (3) f(A)=m1+A+RKA.(3) By combining (Equation1)–(Equation3) and setting b=βeδIτ/2, we derive the following delay differential equation: (4) dA(t)dt=bA2(tτ)2α+A(tτ)+2R(tτ)m1+A(t)+R(t)KA(t).(4) We let A(t)=A(t,t0,φ) denote the solution of (Equation4) subjecting to the initial function φC([t0τ,t0],(0,)), for some t00.

Our discussions in the following parts focus on system (Equation4). In Section 2, we present complete descriptions of the dynamics of wild mosquitoes by delicate analyses based on the Fluctuation Lemma [Citation36]. By setting R(t)>0 (or R(t)0) in (Equation4), we obtain the dynamics of wild mosquitoes with (or without) the intervention of released infected males. The results summarized in Theorems 2.1–2.6 present sufficient conditions for the extinction of wild mosquitoes in different situations. We identify an adult mortality rate threshold δA and a release amount threshold r. The wild mosquitoes could be eliminated if either the adult mortality rate exceeds δA or the release amount exceeds r uniformly. When the releasing number is kept constant less than r, (Equation4) displays a bistable dynamics with two asymptotically stable equilibria A0 and A2 and an unstable equilibrium A1. When R(t)r, the two positive equilibria A1 and A2 coincide into A~1, which is semistable, i.e. asymptotically stable from right-side and unstable from left-side. In Section 3, by letting the parameters in (Equation4) be temperature-dependent on a daily basis, we use simulations to examine how the temporal profiles of wild mosquito population respond to the variation of temperature. By using the daily average temperatures from 2011 to 2017 in Guangzhou, the simulated curves capture several critical features of wild Aedes albopictus population in Guangzhou. We employ the model to design the most cost-efficient releasing policy, targeting at more than 95% reduction at the highest mosquito density. Our simulation shows that the releasing should be started no less than 5 weeks before the peak period of wild mosquito density.

2. Threshold dynamics of population suppression

By studying the global dynamics of (Equation4), we can obtain the precise expression of the releasing threshold over which the population suppression is guaranteed. For each initial function φC([t0τ,t0],(0,)), the positiveness and boundedness of the solution A(t)=A(t,t0,φ) to (Equation4) can be proved by a similar method in [Citation21,Citation22].

As b=βeδIτ/2 counts the fertility of adult females and m measures its natural mortality rate, it holds naturally that b>m. We assume throughout the rest of this paper that (5) b0=bm1>0.(5)

2.1. A classification of the equilibria

In the case R(t)0, (Equation4) reduces to (6) dA(t)dt=bA2(tτ)2α+A(tτ)m1+A(t)KA(t),(6) which characterizes the dynamics of the wild mosquito population. Obviously, A0=0 is the unique constant steady-state of (Equation4) when the release number of infected males R(t) changes in time t. If the loss of crossable infected males in the field is compensated by a new release such that R(t) is maintained almost at a constant level, then it can be modelled by setting R(t)r0. Besides the trivial equilibrium A0, we enumerate the nonnegative equilibria of (Equation4) in the following lemma.

Lemma 2.1

Let b0>0, R(t)r0, and (7) K=2αb02(b0+1+1)2,r=(3b0+4)K2α22(b0+1)[(b0+2)Kα]K.(7) Then the positive equilibria of (Equation4) can be classified as follows:

  1. For the case K>K, we have r>0 and (Equation4) has no positive equilibrium when r>r, a unique positive equilibrium (8) A~1=b0K2α3r2(8) when r=r, two positive equilibria (9) A1=b0K2α3rD(r,K)2,A2=b0K2α3r+D(r,K)2(9) when r<r, where (10) D(r,K)=r22[(3b0+4)K2α]r+(b0K2α)28αK.(10) Specially, if r=0, then (Equation6) has two positive equilibria (11) A1=b0K2αF(K)2,A2=b0K2α+F(K)2,(11) where F(K)=(b0K2α)28αK.

  2. For the case K=K, we have r=0, and (Equation4) has no positive equilibrium when r>0, a unique positive equilibrium (12) A~2=b02Kα(12) when r = 0.

  3. For the case K<K, we have r<0, and (Equation4) or (Equation6) has no positive equilibrium for all r0.

Proof.

We only present the proof for Part (1) as the proof for the other two cases is similar. Besides A0=0, a positive constant steady-state of (Equation4) is a positive zero of quadratic form (13) g(x,r)=x2(b0K2α3r)x+2(r+α)(r+K),(13) with discriminant D(r,K) defined in (Equation10). If D(r,K)>0, then g(x,r)=0 has two solutions (14) x1=b0K2α3rD(r,K)2andx2=b0K2α3r+D(r,K)2,(14) satisfying x1+x2=b0K2α3r and x1x2=2(r+α)(r+K)>0. Set K1=2α/b0, r1=(b0K2α)/3. Then x2>x1>0 if and only if (15) K>K1andr<r1.(15) As we only care about the positive solutions of g(x,r)=0, we assume (Equation15) holds in this Part. The discriminant of D(r,K)=0 is r=32(b0+1)[(b0+2)Kα]K. It is easy to verify that r>0 if and only if K>K2:=α/(b0+2). Note that K2=α/(b0+2)<2α/b0=K1. If K>K1, then K>K2 and D(r,K)=0 has two roots r=(3b0+4)K2α22(b0+1)[(b0+2)Kα]K,r2=(3b0+4)K2α+22(b0+1)[(b0+2)Kα]K. Since r<r2 and r+r2=2[(3b0+4)K2α]>2[(3b0+4)K12α]=4α(b0+2)/b0>0, we obtain r2>0. To find out whether r is greater than 0, we define (16) F(K)=rr2=(b0K2α)28αK=b02K24α(b0+2)K+4α2.(16) Then r>0 if and only if F(K)>0. As F(K)=0 has two positive zeros (17) K3=2αb02b0+112andK=2αb02b0+1+12,(17) we have r>0 when K>K or 0<K<K3, r=0 when K=K or K=K3, and r<0 when K3<K<K. By using KK1=2αb02b0+1+122αb0=4αb021+b0+1>0 and K1K3=2αb02αb02b0+112=4αb02b0+11>0, we obtain K3<K1<K. Thus r>0 only when K>K.

Next we consider the relation among r, r1 and r2. As K>K1, we have (2b0+3)Kα>(2b0+3)K1α=3α(b0+2)b0>0, and so r2r1=43[(2b0+3)Kα]+22(b0+1)[(b0+2)Kα]K>0. Since rr1=43[(2b0+3)Kα]22(b0+1)[(b0+2)Kα]K, we have r<r1 if and only if 2[(2b0+3)Kα]<32(b0+1)[(b0+2)Kα]K, i.e. (18) h(K):=b0(b0+3)K2α(b03)K2α2>0.(18) Note that the unique positive zero K4 of h(K) satisfying K4=2α/(b0+3)<2α/b0=K1. Then K>K1 implies K>K4 and so h(K)>0. Thus r<r1<r2. As r<r1 from condition (Equation15), we derive that x1,2>0 only when 0r<r. Then for the case K>K, g(x,r)=0 has two different positive roots when 0r<r, a unique positive root x~=(b0K2α3r)/2 when r=r, and no positive root when r>r1. The proof of Part (1) is completed.

2.2. The bistability and semistability under insufficient releasing

In this section, we discuss the suppression dynamics based on the releasing number R(t) and the density-dependent competitive constant K according to Lemma 2.1. We first introduce Fluctuation Lemma which is important throughout this section.

Lemma 2.2

Fluctuation Lemma [Citation36]: Appendix A.5

Let f:[b,)R be bounded and differentiable. Then there exist sequences sk, tk such that f(sk)f_,f(sk)0,andf(tk)f¯,f(tk)0as k, where f¯:=lim suptf(t) and f_:=lim inftf(t).

Now we start to discuss the stabilities of the equilibria of system (Equation4). We first show that the mosquito-free equilibrium A0 is locally stable for small initial size when K>K.

Lemma 2.3

Let K>K, r¯0=inft[t0τ,)R(t) and r0=min{r¯0,r}. If the initial function satisfies φ(t)<min{A1(r0),A~1(r0)}, where A~1(r0) and A1(r0) are equilibria of (Equation4) defined in (Equation8) and (Equation9), respectively, then the solution A(t) of (Equation4) satisfies (19) limtA(t)=0.(19)

Proof.

We first prove the positivity and boundedness of the solutions of (Equation4). If A(t)>0 does not hold for all t0, we let t¯>t0 be the least time such that A(t¯)=0. Hence A(t¯)0, and so 0<bA2(t¯τ)2α+A(t¯τ)+2R(t¯τ)0. This contradiction verifies the positivity of A(t). If A(t) is an unbounded solution of (Equation4), then there is an increasing sequence {tn} with tn as n, such that {A(tn)} increases monotonically to infinity as n, A(tnτ)<A(tn), and A(tn)0. Thus m1+A(tn)+R(tn)K<bA(tn)2α+A(tn)+2R(t¯τ), which is impossible for large n, and therefore all the solutions of (Equation4) are bounded.

The nonnegativity of R(t) indicates that r0=inft[t0τ,)R(t)0. By Lemma 2.1, A1(r0) and A~1(r0) are well defined and positive when r0<r and r0=r, respectively. For any a with 0<a<min{A1(r0),A~1(r0)}, we claim that (20) 0<A(t)<afor tt0if 0<φ(t)<aon [t0τ,t0].(20) If not, then there is some t¯>t0, such that A(t¯)=a, and A(t)<a for t[t0,t¯), then A(t¯)0. Substituting t=t¯ into (Equation4) leads to m1+a+r0Kam1+a+R(t¯)Ka<ba22α+a+2R(t¯τ)ba22α+a+2r0. Hence m1+a+r0K<ba2α+a+2r0. Setting x = a and r=r0 in (Equation13) leads to g(a,r0)<0. Note that g(x,r0)>0 when 0x<A1(r0) for the case r0<r, or 0x<A~1(r0) for the case r0=r. This contradiction guarantees (Equation20), which verifies the locally uniform stability of A0.

Assume that 0φ(t)<min{A1(r0),A~1(r0)} for t[t0τ,t0]. By the claim (Equation20), if φ(t)<a<min{A1(r0),A~1(r0)} on [t0τ,t0], then A(t)<a and so the upper limit A¯=lim suptA(t)a. As the solution is bounded, there is a sequence {tn} such that (21) A(tn)A¯=lim suptA(t),A(tn)0andtn,as n,(21) by Fluctuation Lemma (Lemma 2.2). Taking the limit of (Equation4) along this sequence gives (22) m1+A¯+r0KA¯m1+A¯+R(tn)KA¯=lim supnbA2(tnτ)2α+A(tnτ)+2R(tnτ)bA¯22α+A¯+2R(tnτ)bA¯22α+A¯+2r0.(22) Hence A¯=0 or m1+A¯+r0KbA¯2α+A¯+2r0, which leads to g(A¯,r0)0. The facts g(x,r0)>0 on [0,min{A1(r0),A~1(r0)}) and 0A¯a<min{A1(r0),A~1(r0)} indicate g(A¯,0)>0, which contradicts the second case. Thus A¯=0. By the positivity of A(t), we obtain that the lower limit of the solution A_=lim inftA(t)0. Hence A_=A¯=0, which verifies the asymptotic stability of A0.

When K>K and the releasing number maintains at a constant level such that R(t)r(0,r), Lemma 2.1 indicates that (Equation4) has a mosquito-free equilibrium A0=0 and two positive equilibria A1 and A2 defined in (Equation9). We next show that (Equation4) exhibits bistable dynamics with an unstable equilibrium A1 being sandwiched by two asymptotically stable equilibria A0 and A2.

Theorem 2.1

Let K>K and R(t)r(0,r) for t0. Let A(t) be a solution of (Equation4). Then we have:

  1. A1 is unstable;

  2. limtA(t)=0 when φ(t)A1 on [t0τ,t0] and φ(t)A1;

  3. limtA(t)=A2 when φ(t)A1 on [t0τ,t0] and φ(t)A1.

Proof.

(1) The instability of A1 follows from Lemma 2.3 directly.

(2) Let A0(t) be an arbitrary solution of (Equation4) with φ(t)A1 on [t0τ,t0] and φ(t)A1. By setting φ(t)A1 for t[t0τ,t0] we obtain a constant solution A1(t)A1. We claim that A0(t)A1(t) for all t0. Let Aε(t) be a solution of (Equation4) with Aε(t)=A0(t)ε for tτtt0. For sufficiently small ε>0, it is easy to verify Aε(t)<A1(t) for t0. Thus Limε0+Aε(t)A1(t), implying that A0(t)A1(t) for t0. Since φ(t)A1, then A0(t)A1 when t[t0,t0+τ] and we can find t1 and t2 such that A0(t)<A1 when t[t1,t2][t0,t0+τ]. Let A2(t) be a solution of (Equation4) with the continuous initial function φ¯(t)=A1for t[t2τ,t1],A0(t)φ¯(t)A1for t(t1,t2),φ¯(t2)=A0(t2). Since A0(t)φ¯(t) when t[t2τ,t1], we obtain A0(t)A2(t) when t[t2,t1+τ] by the similar method as the proof for A0(t)A1(t). Note that dA1(t)dt=bA122α+A1+2rm1+A1(t)+rKA1(t),dA2(t)dt=bA122α+A1+2rm1+A2(t)+rKA2(t) when t[t2,t1+τ], which are two ordinary differential equations. Then by the fact A2(t2)<A1(t2) we get A2(t)<A1(t) when t[t2,t1+τ]. Thus we have A0(t)A2(t)<A1(t)=A1 for t[t2,t1+τ], then we derive limtA(t)=0 by Lemma 2.3.

(3) By the proof of Part (1), we only need to prove the case φ(t)>A1 on [t0τ,t0]. Similar to the claim (Equation20) in Lemma 2.3, we claim that, for any positive constant a<A2A1, (23) |A(t)A2|<afor tt0if |φ(t)A2|<aon [t0τ,t0].(23) If not, we let t¯>t0 again be the least time at which A=A2a or A=A2+a. If A(t¯)=A2a, then A(t)>A2a for t[t0,t¯) and A(t¯)0. From (Equation4) we have m1+A2a+rK(A2a)bA2(t¯τ)2α+A(t¯τ)+2r>b(A2a)22α+A2a+2r, which implies g(A2a,r)>0. However, A1<A2a<A2 for 0<a<A2A1 leads to g(A2a,r)<0 as g(x,r)<0 for A1<x<A2 and 0<r<r.

If A(t¯)=A2+a, then A(t)<A2+a for t[t0,t¯) and A(t¯)0. Hence m1+A2+a+rK(A2a)bA2(t¯τ)2α+A(t¯τ)+2r<b(A2+a)22α+A2+a+2r, which leads to g(A2+a,r)<0 and therefore contradicts to the fact g(x,r)>0 for x>A2 and 0<r<r. Then (Equation23) holds, which verifies the uniform stability of A2.

Now we discuss the attractiveness of A2. Let a(0,A2A1). By a similar argument as the proof for (Equation23), it can be shown that (24) A(t)>A1+a,for tt0,when φ(t)>A1+aon [t0τ,t0].(24) In fact, if there is some t¯>t0 such that A(t¯)=A1+a and A(t)>A1+a in [t0,t¯), then A(t¯)=bA2(t¯τ)2α+A(t¯τ)+2rm1+A1+a+rK(A1+a)0, which implies m1+A1+a+rK>b(A1+a)2α+A1+a+2r. Therefore, g(A1+a,r)>0. As 0<a<A2A1, we have A1<A1+a<A2, and so g(A1+a,r)<0. This contradiction verifies (Equation24). Thus the lower limit of the solution satisfies A_=lim inftA(t)A1+a. Let {sn} be a divergent sequence along which (25) A(sn)A_=lim inftA(t),A(sn)0,andsnas n.(25) Taking the limit in (Equation4) along this sequence implies (26) m1+A_+rKA_=lim infnbA2(snτ)2α+A(snτ)+2rbA_22α+A_+2r.(26) Hence m1+A_+rKbA_2α+A_+2r, which implies g(A_,r)0. By using the fact A_A1+a and the property of g(x,r) we have A_A2, and so A¯A_A2.

Let {tn} again be a sequence along which (Equation21) holds. By the Fluctuation Lemma again and taking the limit in (Equation4) along {tn} yields a similar inequality in (Equation22): (27) m1+A¯+rKA¯=lim supnbA2(tnτ)2α+A(tnτ)+2rbA¯22α+A¯+2r.(27) Hence g(A¯,r)0. Since g(x,r)0 when A1xA2 and g(x,r)>0 when x>A2 for 0<r<r, we get A¯A2, and so A_=A¯=A2, which verifies the attractiveness of A2.

When R(t)0, (Equation4) reduces to (Equation6) which describes the field mosquito population without the interference of Wolbachia. In this case, the positive equilibria A1 and A2 of (Equation4) reduce to A1 and A2, respectively. By the similar argument as in Theorem 2.1, we find that (Equation6) also exhibits bistable dynamics when K>K and we omit the detailed proof.

Theorem 2.2

Let K>K and R(t)0. Then the solution A(t) of (Equation6) satisfies the following statements.

  1. A1 is unstable;

  2. limtA(t)=0 when φ(t)A1 on [t0τ,t0] and φ(t)A1;

  3. limtA(t)=A2 when φ(t)A1 on [t0τ,t0] and φ(t)A1.

For the special case K>K and R(t)r, the positive equilibria A1 coincides with A2 to A~1. We show that (Equation4) displays semistable dynamics with A~1 being asymptotically stable from right-side.

Theorem 2.3

If K>K and R(t)r, then the solution A(t) of (Equation4) satisfies:

  1. limtA(t)=0 when φ(t)A~1 on [t0τ,t0] and φ(t)A~1;

  2. limtA(t)=A~1 when φ(t)A~1 on [t0τ,t0] and φ(t)A~1.

Proof.

The asymptotic stability of A0=0 follows directly from Lemma 2.3 and Theorem 2.1. To prove the stability of A~1 from right-side, we claim (28) A~1<A(t)<A~1+afor tt0(28) when A~1<φ(t)<A~1+a on [t0τ,t0] for any a>0 as in (Equation23). In fact, if there is some t¯>t0, such that A(t¯)=A~1 and A(t)>A~1 for t[t0,t¯), then A(t¯)=bA2(t¯τ)2α+A(t¯τ)+2rm1+A~1+rKA~10, which implies m1+A~1+rKA~1bA2(t¯τ)2α+A(t¯τ)+2r>b(A~1)22α+A~1+2r. Hence g(A~1,r)>0, which contradicts to the fact g(A~1,r)=0. Furthermore, if there is some t~>t0 such that A(t~)=A~1+a and A(t)<A~1+a for t[t0,t~), then A(t~)0. Substituting t=t~ into (Equation4) leads to m1+A~1+a+rK(A~1+a)bA2(t~τ)2α+A(t~τ)+2r<b(A~1+a)22α+A~1+a+2r. Then we have g(A~1+a,r)<0, contradicting to the fact that g(x,r)>0 when x>A~1. Thus (Equation28) holds, which verifies that A~1 is stable from right-side.

The claim in (Equation28) shows that the upper limit A¯ and the lower limit A_ satisfy A¯A_A~1 when the initial function φ(t)>A~1 on [t0τ,t0]. By Fluctuation Lemma, we can find a sequence {tn} such that (Equation21) holds. Taking the limit of (Equation4) along {tn} implies m1+A¯+rKA¯=lim supnbA2(tnτ)2α+A(tnτ)+2rbA¯22α+A¯+2r, which gives g(A¯,r)0. By employing the proof of Lemma 2.1 we obtain g(x,r)0 for all x0 except g(A~1,r)=0. Hence A¯=A~1, and so A_=A¯=A~1, which verifies the asymptotical stability of A~1 from right-side when φ(t)>A~1 on [t0τ,t0]. If φ(t)=A~1 for some t[t0τ,t0] but φ(t)A~1 in this time interval, we can prove limtA(t)=A~1 directly from the above result.

2.3. A threshold releasing level

In this part, we show that r defines a threshold releasing level above which the wild mosquitoes could be eradicated.

Theorem 2.4

Assume that K>K and r0=inft[t0τ,)R(t)>r. Then the mosquito-free equilibrium A0 of (Equation4) is globally asymptotically stable.

Proof.

By the definition of g(x,r) in (Equation13), we have g(x,r0)>0 for any x0 when K>K and r0>r. By an exactly similar proof, the claim (Equation20) holds for any positive constant a, which guarantees the globally stability of A0=0.

To obtain the global attractiveness of A0, it is sufficient to prove that the upper limit A¯=0. The positivity of the solution A(t) shows that A¯A_0. Let {tn} be a sequence such that (Equation21) holds. Taking the limit of (Equation4) along {tn} implies that the inequality in (Equation22) holds. Thus A¯=0 or g(A¯,r0)0. As the latter case contradicts to the fact g(x,r0)>0 for x0, we have A¯=0, and so A¯=A_=0, which verifies the global asymptotical stability of A0.

2.4. A threshold mortality rate level

Lemma 2.1 confirms that r vanishes when K=K. If R(t)0, then the unique positive equilibrium of (Equation4) or (Equation6) A~1=(b0K2α3r)/2=b0K/2α=A~2. By the same argument in the proof of Theorem 2.3, we obtain that (Equation4) or (Equation6) exhibits semistable dynamics directly.

Theorem 2.5

Let K=K and R(t)0. Then the solution A(t) of (Equation4) or (Equation6) satisfies the following statements:

  1. limtA(t)=0 when φ(t)A~2 on [t0τ,t0] and φ(t)A~2;

  2. limtA(t)=A~2 when φ(t)>A~2 on [t0τ,t0].

From Lemma 2.1 we see that A0 is the unique equilibrium of (Equation4) when K<K, or K=K and r0=inft[t0τ,)R(t)>0. We next show that A0 is globally asymptotically stable in both cases.

Theorem 2.6

The mosquito-free equilibrium A0=0 in (Equation4) is globally asymptotically stable if one of the following conditions holds:

  1. K<K;

  2. K=K and R(tn)>0, where {tn} is an increasing sequence with tn as n.

Proof.

(1) Similar to the proof of Theorem 2.4, for any a>0, the claim (Equation20) holds when K<K. In fact, if there is some t¯>t0, such that A(t¯)=a and A(t)<a for t[t0τ,t¯), then A(t¯)0, which implies m1+aKam1+a+R(t¯)KabA2(t¯τ)2α+A(t¯τ)+2R(t¯τ)<ba22α+a. Hence g(a,0)<0, contradicting to the fact g(x,0)>0 for any x0 when K<K. The claim (Equation20) verifies the global stability of A0.

To obtain the global attractiveness of A0, it is sufficient to prove A¯=0. The positivity of the solution A(t) shows that A¯A_0. Let {tn} be a sequence such that (Equation21) holds. Taking the limit of (Equation4) along {tn} implies m1+A¯KA¯m1+A¯+R(tn)KA¯lim supnbA2(tnτ)2α+A(tnτ)+2R(tnτ)bA¯22α+A¯. Thus A¯=0 or g(A¯,0)0. The second case contradicts to the fact g(x,0)>0 for all x0. Hence A_=A¯=0, which proves the global asymptotical stability of A0.

(2) The proof follows from Theorem 2.5 directly.

Let δA=m/K be the additional mortality rate of adults, which is mainly caused by intraspecific and interspecific competition. K and the natural mortality rate m identify (29) δA(α,b,m):=mK=m2αbm12.(29) By fixing m, we have δA>δA if and only if K<K. Theorem 2.6 shows that δA defines the threshold mortality rate level of adults, over which all mosquitoes go to extinction naturally.

3. Applications of Aedes albopictus population control

An efficient strategy to control dengue is to reduce the vector Aedes mosquitoes in wild area to a substantially lower level within the peak season of dengue [Citation51]. To assess and optimize the mosquito control strategies, we need to clarify the key factors by studying the interactive dynamics of wild and Wolbachia-infected mosquitoes. In this section, we assess quantitatively the effects of suppression by numerical simulations.

3.1. Threshold dynamics of wild mosquito population

Our theoretical analysis in the previous part has identified an important parameter K=2αb02b0+1+12=2αb/m12, which is an important factor of δA(α,b,m). Obviously, δA(α,b,m) decreases in α. In addition, it follows from δAb=12α1mb>0andδAm=12αbm1<0 that δA(α,b,m) increases in b, and decreases in m.

Theorems 2.1–2.6 show the rich dynamics of wild mosquito population, including bistability and semistability. When δA>δA, or equivalently, K<K, A0=0 is globally asymptotically stable, which guarantees the extinction of wild mosquito population. When δA=δA, besides the locally stable equilibrium A0, (Equation6) has a semistable positive equilibrium A~2=b0K/2α, with all the solutions converging to A~2 uniformly when the initial function φ(t)>A~2 on [t0τ,t0]. When δA<δA, (Equation6) displays bistable dynamics with two locally asymptotically stable equilibria A0 and A2=b0K2α+(b0K2α)28αK2, which gives mosquito population carrying capacity in the inhabiting area. Hence δA defines the threshold mortality rate of adults, above which the wild mosquitoes would be eliminated ultimately naturally. Interestingly, A0 is always locally asymptotically stable even when δAδA, which indicates that the wild mosquito population will be wiped off when the population size is small.

Integrated with experimental and field data of Aedes albopictus mosquito population in Guangzhou, we use numerical examples to display the theoretical results more clearly. We first list the meaning of each parameter and their ranges in Table . By experimental findings, some of the life-table parameters of mosquitoes are strongly sensitive to temperature (T for short) [Citation29,Citation31,Citation38,Citation53]. For example, the mean hatching rate of eggs μE, mortality rate of immature stage (including egg, larva and pupa) δI, and natural mortality rate of adults m. The average number of larvae β from a female can be estimated by β=NμE/τA [Citation19].

Table 1. Definitions of the parameters used in the model.

It is a challenging task to estimate the mating inhomogeneity constant α and the density-dependent competition constant K in the field. In order to estimate them and demonstrate the dependence of the threshold δA and the carrying capacity A2 on the parameters, we assume that all the parameters are kept constant as follows unless otherwise specified [Citation21,Citation44], (30) b=3,m=0.05,τ=17.(30) The basic assumption (Equation5) holds obviously. As shown in Figure , we note that A2 increases almost linearly in K, and decreases almost linearly in α. Note that the magnitude of A2 depends on the size of the studying area. Without loss of generality, we specify (31) α=100,K=1000.(31) It is noticed that K and A2 are dependent on the size of the release area. From (Equation31) we have K=4.3948,δA=0.0114,A1=3.4,A2=58797. As the unstable equilibrium A1 is small, the migrating mosquitoes are easy to reach this threshold and then grow to the stable equilibrium A2, which can explain the successful invading of Aedes albopictus in most tropical and subtropical areas.

Figure 1. The dependence of the carrying capacity A2 of wild mosquitoes on α and K. With the parameters specified in (Equation30) and (Equation31), A2 increases almost linearly in K, and decreases almost linearly in α.

Figure 1. The dependence of the carrying capacity A2∗ of wild mosquitoes on α and K. With the parameters specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ), A2∗ increases almost linearly in K, and decreases almost linearly in α.

As shown in Figure  (A), with the increasing of α, δA decreases steeply for small α, and decreases steadily for large α. By fixing all the parameter values as in (Equation30) and (Equation31) except α, the simulation shows that δA decreases from 1.138 to 0.02231 when α increases from 1 to 50, and decreases from 0.02231 to 0.00291 when α increases from 50 to 400. However, δA decreases almost linearly with m, and increases almost linearly with b. For example, by fixing all the parameter values as in (Equation30) and (Equation31) except m, δA decreases from 0.01319 to 0.005986 as m increases from α=0.01 to α=0.4.

Figure 2. The dependence of the threshold mortality rate δA on α, b and m. The figure is simulated by (Equation29) with the parameters specified in (Equation30) and (Equation31).

Figure 2. The dependence of the threshold mortality rate δA∗ on α, b and m. The figure is simulated by (Equation29(29) δA∗(α,b,m):=mK∗=m2αbm−12.(29) ) with the parameters specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ).

The threshold dynamics of the wild mosquito population remind us a feasible way to reduce the number of wild mosquitoes by raising its mortality rate. Applying insecticides is the traditional and most frequently used method of mosquito eradication in most countries in the world. Since insecticide spraying influences mainly the natural mortality rate of mosquitoes, we model the effect of insecticide spraying by raising the natural mortality rate m of adults. Consider the peak season of Aedes albopictus mosquitoes in which the population stabilizes at the steady-state level that coincides with the high-incidence of dengue in Guangzhou. We define the suppression rate γ as the ratio of the carrying capacity A2(m) of the inhabiting area with applying insecticides to that of without applying insecticides. We estimate the mortality rate m which is sufficient for γ5%.

The carrying capacity A2(m) decreases gradually in m. Let m = 0.05 be the natural mortality rate of adult mosquitoes without interference of insecticides. Then A2(0.05)=58797. If we double m to m = 0.1, then A2(0.1)=28793, and so γ=28793/58797=48.97%. Similarly, as shown in Figure , we have γ=14.93% for A2(0.3)=8777 with m = 0.3, and γ=4.93% for A2(0.72)=2898 with m = 0.72. Hence, to reduce more than 95% of wild mosquitoes in the peak season, we need to spray insecticides so that the mortality rate m of adults above the level of 0.72. This indirect mosquito control method strongly depends on the effect of the insecticide.

Figure 3. The effect of insecticides spraying on wild mosquito population. Let φ(t)=500 on [17,0], and the parameters, except m, be specified in (Equation30) and (Equation31). The curves simulated by (Equation6) show that the steady-state level of wild mosquitoes decreases with m.

Figure 3. The effect of insecticides spraying on wild mosquito population. Let φ(t)=500 on [−17,0], and the parameters, except m, be specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ). The curves simulated by (Equation6(6) dA(t)dt=bA2(t−τ)2α+A(t−τ)−m1+A(t)KA(t),(6) ) show that the steady-state level of wild mosquitoes decreases with m.

3.2. Threshold dynamics of compensation suppression policy

Although having an instant effect and being wildly used in the world, insecticide spraying is not long-term effective and sustainable method to control mosquitoes due to mosquito resistance to insecticides [Citation3]. Recently, by using some special strains of Wolbachia pipientis, Xi et al. reduced more than 90% of the Aedes albopictus population in a field trial in Guangzhou [Citation52]. The released Wolbachia-infected males keep their crossability in about three days in the field. Compensating the loss of crossable males by new releasing in every three days, we can maintain the number of crossable infected males in wild almost at a constant level. This compensation suppression policy can be modelled by setting R(t)r0 in (Equation4). For the case KK, Theorem 2.6 shows that the wild mosquitoes can be eliminated naturally when r>0. When K>K, Theorems 2.1–2.4 shows that r defines the threshold releasing level above which all wild mosquitoes will be suppressed ultimately. As shown in Figure  (A), when m0.05, r increases steeply as m decreases from 0.05 to 0, or b increases from 1 to 5. However, r varies steadily when m>0.05. Unlike b and m, as shown in Figure  (B), r increases almost linearly with K, and decreases with α.

Figure 4. The dependence of the threshold releasing number r on m, b, K and α. Let the parameters be specified in (Equation30) and (Equation31). (A) The nonlinear dependence of r on m and b, (B) r increases almost linearly in K, and decreases almost linearly in α.

Figure 4. The dependence of the threshold releasing number r∗ on m, b, K and α. Let the parameters be specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ). (A) The nonlinear dependence of r∗ on m and b, (B) r∗ increases almost linearly in K, and decreases almost linearly in α.

With the parameters specified in (Equation30) and (Equation31), we have K=4.3948 and r=9826. Hence K>K. According to Theorems 2.1–2.4, (Equation4) displays similar dynamics to the wild mosquito population in this case, including bisability and semistability. Let the constant releasing number r=5000<r, as shown in Figure  (A), wild mosquitoes go to extinction when the initial mosquito population size φ(t)<A1=1445 on [17,0], while the mosquito suppression fails when φ(t)>A1 on [17,0] and the size of mosquito population tends to A2=42355. Similarly, as shown in Figure (B) for the case r=r, the semistable dynamics shows that a successful population suppression is also dependent on the initial mosquito population size. It remains a rather subtle question to determine the ultimate fate of mosquito suppression when the initial functions cross over the unstable equilibrium A1 or semistable equilibrium A1.

Figure 5. The bistability and semistability of (Equation4). The parameters are specified in (Equation30) and (Equation31). (A) Let R(t)5000<r=9826, (Equation4) displays bistability: A(t)0 as t when φ(t)<A1=1445 on [17,0], and A(t)A2=42355 as t when φ(t)>A1=1445 on [17,0]. (B) With R(t)r, the unique positive equilibrium A1=14661 is semistable, with A(t)A1 as t when φ(t)>A1 on [17,0].

Figure 5. The bistability and semistability of (Equation4(4) dA(t)dt=bA2(t−τ)2α+A(t−τ)+2R(t−τ)−m1+A(t)+R(t)KA(t).(4) ). The parameters are specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ). (A) Let R(t)≡5000<r∗=9826, (Equation4(4) dA(t)dt=bA2(t−τ)2α+A(t−τ)+2R(t−τ)−m1+A(t)+R(t)KA(t).(4) ) displays bistability: A(t)→0 as t→∞ when φ(t)<A1=1445 on [−17,0], and A(t)→A2=42355 as t→∞ when φ(t)>A1=1445 on [−17,0]. (B) With R(t)≡r∗, the unique positive equilibrium A1∗=14661 is semistable, with A(t)→A1∗ as t→∞ when φ(t)>A1∗ on [−17,0].

3.3. The impact of temperature on suppression

As typical subtropical monsoon climate, Guangzhou has long hot and rainy summers, and warm and dry winters. As shown in Figure  (A), the daily mean temperatures of Guangzhou present strongly periodicity. A large number of experimental findings indicate that most of the life-table parameters of mosquitoes are strongly sensitive to climatic conditions, in particular, temperature and precipitations [Citation10,Citation26,Citation38]. Deep understanding the mosquito-environment relationship is essential for the control of mosquito-borne diseases. To integrate temperature into our model, we introduce the temperature-dependent hatching rate of eggs μE [Citation10,Citation26,Citation38] as μE(T)=0.24273.15+T298exp107981.98712981273.15+T1+exp1000001.9871141841273.15+T. According to [Citation8,Citation9,Citation38], the mean mortality rate in immature stage is estimated by δI(T)=0.0000866T20.00368T+0.09,T12.5C,0.5,else. Hence the parameter b is given by (32) b(T)=NμE(T)2τAeδI(T)τ,(32) where N, τA and τ are defined in Table . Although the delays τA and τ are sensitive to temperature [Citation10,Citation26,Citation38], our previous studies show that mosquito suppression does not sensitive to the delays, and so we assume them keep constants [Citation19,Citation22]. Furthermore, we estimate the natural mortality rate of adults by [Citation8,Citation9,Citation38] (33) m(T)=0.000114T20.00427T+0.1278,T15C,0.5,else.(33) We take January 1st as the initial time t0=0 to display the significant seasonal variation of Aedes albopictus population. As shown in Figure  (B), the simulated curves generated by (Equation4) with the temperature-dependent parameters in (Equation32)–(Equation33) show a strong periodicity annually as the temperature, and capture several critical features of natural Aedes albopictus mosquito population in Guangzhou [Citation29–31]. The eggs cannot hatch until the weekly average temperature rises to 1011C, and the daily photoperiod reaches 11-11.5 h, which results in most of eggs laid after October accumulating into egg bank [Citation30,Citation37]. As the warm and rainy season starts in early March, the hatching of a large amount of diapause eggs leads to a rapid growth of Aedes albopictus population from early spring. The suitable temperature in the fall helps the population reaches its peak in late September and early October, after which the population size decays sharply and nearly vanishes from December to February in next year [Citation30,Citation31].

Figure 6. The Aedes albopictus population in Guangzhou shows strong periodicity annually as the temperature. (A) The daily mean temperature of Guangzhou from 2011 to 2017; Data source: China meteorological data sharing service system: http://www.cma.gov.cn/. (B) The simulated curve of wild Aedes albopictus population over a seven-year period in Guangzhou (2011–2017). With the parameters specified in (Equation30) and (Equation31), except b(T) in (Equation32) and m(T) in (Equation33) being temperature-dependent functions, the curve is generated by (Equation4) with φ(t)=1000 for t[17,0].

Figure 6. The Aedes albopictus population in Guangzhou shows strong periodicity annually as the temperature. (A) The daily mean temperature of Guangzhou from 2011 to 2017; Data source: China meteorological data sharing service system: http://www.cma.gov.cn/. (B) The simulated curve of wild Aedes albopictus population over a seven-year period in Guangzhou (2011–2017). With the parameters specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ), except b(T) in (Equation32(32) b(T)=NμE(T)2τAe−δI(T)τ,(32) ) and m(T) in (Equation33(33) m(T)=0.000114T2−0.00427T+0.1278,T≥15∘C,0.5,else.(33) ) being temperature-dependent functions, the curve is generated by (Equation4(4) dA(t)dt=bA2(t−τ)2α+A(t−τ)+2R(t−τ)−m1+A(t)+R(t)KA(t).(4) ) with φ(t)=1000 for t∈[−17,0].

By the definitions, both of the threshold mortality rate δA and threshold releasing level r are temperature-dependent, and so they vary in time. As shown in Figure , δA and r display quasi-periodicity annually as temperature, and peaks in the period from July to October. Hence this period is the critical timing for the control of Aedes albopictus population by both chemical and biological measures.

Figure 7. The temperature dependence of the threshold mortality rate δA and threshold releasing level r. The curves are simulated by δA in (Equation29) and r in (Equation7) with the parameters specified in (Equation30) and (Equation31), except b(T) in (Equation32) and m(T) in (Equation33). Both of δA and r display quasi-periodicity annually as temperature, and take large values in the period from July to October.

Figure 7. The temperature dependence of the threshold mortality rate δA∗ and threshold releasing level r∗. The curves are simulated by δA∗ in (Equation29(29) δA∗(α,b,m):=mK∗=m2αbm−12.(29) ) and r∗ in (Equation7(7) K∗=2αb02(b0+1+1)2,r∗=(3b0+4)K−2α−22(b0+1)[(b0+2)K−α]K.(7) ) with the parameters specified in (Equation30(30) b=3,m=0.05,τ=17.(30) ) and (Equation31(31) α=100,K=1000.(31) ), except b(T) in (Equation32(32) b(T)=NμE(T)2τAe−δI(T)τ,(32) ) and m(T) in (Equation33(33) m(T)=0.000114T2−0.00427T+0.1278,T≥15∘C,0.5,else.(33) ). Both of δA∗ and r∗ display quasi-periodicity annually as temperature, and take large values in the period from July to October.

The coincidence of the high-incidence period of dengue fever with the peak season of Aedes albopictus population in Guangzhou indicates the close relation between mosquitoes size and dengue cases, so it is essential to reduce the wild mosquito population to a substantially lower level during the peak season of dengue. As shown in Figure  (B), the highest Aedes albopictus density in 2012 in Guangzhou was observed around September 12. We use our model to design suppression policy, aiming for the elimination of more than 95% mosquitoes in the peak season and keeping the effect until the end of November. We divide the suppression into three phases, a strong suppression phase with large releasing number (phase I) before September 12, followed by a transitional phase (phase II) after September 12 in about 15 days and a maintenance phase (phase III) after September 28 with small releasing number. In the field trial of mosquito suppression in Guangzhou, the infected males were released three times a week with constant amount [Citation52]. In the simulation, we assume that R(t)ri>0, i = 1, 2, 3 during the three phases, respectively.

To suppress the wild mosquito population during the peak season in the most cost-efficient way, we estimate the starting date of releasing and the least releasing amount. We initiate the simulation in January 1, 2011 to avoid the impact of initial mosquito population size on the estimation. The temporal profiles of adult mosquitoes shown in Figure (A) indicate that the number of wild mosquitoes decreases sharply after releasing, and more than 95% has been reduced at the peak. As shown in Figure  (B), although the values of r1 reduce sharply with the increase of releasing time, the total releasing numbers of infected males in phase I varies in a small interval. When the highest mosquito density in September 12 has been suppressed by more than 95%, it needs much fewer releasing number in the transitional phase II and maintenance phase III to keep the suppression level until the end of November. The variations of the total releasing numbers in phase II and III are insignificant as we expect. If the releasing starts five weeks ahead of the peak of mosquitoes instead of four weeks, the total releasing amount in all three phases reduces sharply from 184,430 to 152,630. However, the total releasing number has little variation when the length of phase I is shifted to 6, 7, 8, and 9 weeks, respectively, and keeps the lengths of the last two phases the same. The variation of the total releasing amount in three phases and its durations indicate that the most cost-efficient control should be started no less than 5 weeks before the start of the peak season of dengue.

Figure 8. Suppression of wild Aedes albopictus population in the high-incidence season of dengue fever. The curves are generated by (Equation4) with the same parameters as in Figure (B), except R(t)r1 in phase I, R(t)r2 in phase II, and R(t)r3 in phase III. (A) A three phases releasing policy achieves the target of suppressing 95% of wild mosquitoes on September 12, 2012, and keeps the low density until the end of November. (B) The minimum constant releasing abundance in each phase and the total releasing number required to reach the target of suppression.

Figure 8. Suppression of wild Aedes albopictus population in the high-incidence season of dengue fever. The curves are generated by (Equation4(4) dA(t)dt=bA2(t−τ)2α+A(t−τ)+2R(t−τ)−m1+A(t)+R(t)KA(t).(4) ) with the same parameters as in Figure 6(B), except R(t)≡r1 in phase I, R(t)≡r2 in phase II, and R(t)≡r3 in phase III. (A) A three phases releasing policy achieves the target of suppressing 95% of wild mosquitoes on September 12, 2012, and keeps the low density until the end of November. (B) The minimum constant releasing abundance in each phase and the total releasing number required to reach the target of suppression.

Acknowledgments

Thank the two anonymous reviewers very much for their valuable comments and suggestions which help us greatly improve the writing of this manuscript. This work was supported by National Natural Science Foundation of China (11471085, 11631005), Program for Changjiang Scholars and Innovative Research Team in University (IRT_16R16), and Natural Science Foundation of Guangdong Province (2017A030310597).

Disclosure statement

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

Additional information

Funding

Thank the two anonymous reviewers very much for their valuable comments and suggestions which help us greatly improve the writing of this manuscript. This work was supported by National Natural Science Foundation of China [grant number 11471085], [grant number 11631005], and Natural Science Foundation of Guangdong Province [grant number 2017A030310597].

References

  • F. Arkin, Dengue researcher faces charges in vaccine fiasco. Science 364(6438) (2019), pp. 320–320.
  • M.P. Atkinson, Z. Su, N. Alphey, L.S. Alphey, P.G. Coleman, and L.M. Wein, Analyzing the control of mosquito-borne diseases by a dominant lethal genetic system, PNAS 104(22) (2007), pp. 9540–9545.
  • F. Baldacchino, F. Caputo, A. Drago, T.A. Della, F. Montarsi, and A. Rizzoli, Control methods against invasive Aedes mosquitoes in Europe: a review, Pest. Manag. Sci. 71 (2015), pp. 1471–1485.
  • S. Bhatt, P.W. Gething, O.J. Brady, J.P. Messina, A.W. Farlow, C.L. Moyes, J.M. Drake, J.S. Brownstein, A.G. Hoen, O. Sankoh, M.F. Myers, D.B. George, T. Jaenisch, G.R.W. Wint, C.P.Simmons, T.W. Scott, J.J. Farrar, and S.I. Hay, The global distribution and burden of dengue, Nature 496 (2013), pp. 504–507.
  • M.S. Blagrove, C. Arias-Goata, A.B. Failloux, and S.P. Sinkins, Wolbachia strain wMel induces cytoplasmic incompatibility and blocks dengue transmission in Aedes albopictus, Proc. Natl. Acad. Sci. USA 109 (2012), pp. 255–260.
  • L. Cai, S. Ai, and G. Fan, Dynamics of delayed mosquitoes populations models with two different strategies of releasing sterile mosquitoes, Math. Biosci. Eng. 15 (2018), pp. 1181–1202.
  • L. Cai, S. Ai, and J. Li, Dynamics of mosquitoes populations with different strategies for releasing sterile mosquitoes, SIAM J. Appl. Math. 74(6) (2014), pp. 1786–1809.
  • P. Cailly, A. Tran, T. Balenghien, G. L'Ambert, C. Toty, and P. Ezanno, A climate driven abundance model to assess mosquito control strategies, Ecol. Model. 227 (2012), pp. 7–17.
  • Q. Cheng, Q. Jing, R.C. Spear, J.M. Marshall, Z. Yang, P. Gong, and S.V. Scarpino, Climate and the timing of imported cases as determinants of the dengue outbreak in Guangzhou, 2014: Evidence from a mathematical model, PLoS Negl. Trop. Dis. 10(2) (2016), p. e0004417.
  • D.A. Focks, D.G. Haile, E. Daniels, and G.A. Mount, Dynamic life table model for Aedes aegypti (Diptera: Culicidae): simulation results and validation, J. Med. Entomol. 30(6) (1993), pp. 1018–1028.
  • P.E. Fine, On the dynamics of symbiote-dependent cytoplasmic incompatibility in Culicine mosquitoes, J. Invertebr. Pathol 31(1) (1978), pp. 10–18.
  • J.E. Gentile, S. Rund, and G.R. Madey, Modelling sterile insect technique to control the population of Anopheles gambiae, Malaria J. 14(1) (2015), p. 92.
  • M. Hertig and S.B. Wolbach, Studies on rickettsia-like micro-organisms in insects, J. Med. Res. 44(3) (1924), pp. 329–374.
  • 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, and Z. Wuet al, The threshold infection level for Wolbachia invasion in random environment. J. Diff. Equ. 266(7) (2019), pp. 4377–4393. Available at https://doi.org/10.1016/j.jde.2018.09.035.
  • L. Hu, M. Huang, M. Tang, J. Yu, and B. Zheng, Wolbachia spread dynamics in multi-regimes of environmental conditions, J. Theor. Biol. 462 (2019), pp. 247–258.
  • M. Huang, M. Tang, and J. Yu, Wolbachia infection dynamics by reaction-diffusion equations, Sci. China Math. 58 (2015), pp. 77–96.
  • M. Huang, J. Yu, L. Hu, and B. Zheng, Qualitative analysis for a Wolbachia infection model with diffusion, Sci. China Math. 59 (2016), pp. 1249–1266.
  • M. Huang, J. Lou, L. Hu, B. Zheng, and J. Yu, Assessing the efficiency of Wolbachia driven Aedes mosquito suppression by delay differential equations, J. Theor. Biol. 440 (2018), pp. 1–11.
  • 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. 1–20.
  • 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. Bios. 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 (2020). Available at https://doi.org/10.3934/dcdsb.2020118.
  • P. Jia, X. Chen, J. Chen, L. Guo, X. Yu, and Q. Liu, A climate-driven mechanistic population model of Aedes albopictus with diapause, Para. Vect 9 (2016), p. 175.
  • J.L. Kyle and E. Harris, Global spread and persistence of dengue, Annu. Rev. Microbiol. 62 (2008), pp. 71–92.
  • R.M. Lana, M.M. Morais, T.F.M. de Lima, T.G.S. Carneiro, L.M. Stolerman, J.P.C. dos Santos, J.J.C. Corts, A.E. Eiras, C.T. Codeo, and L.A. Moreira, Assessment of a trap based Aedes aegypti surveillance program using mathematical modelling, PLoS One 13(1) (2018), p. e0190673.
  • H. Laven, Cytoplasmic inheritance in culex, Nature 177 (1956), pp. 141–142.
  • J. Li and S. Ai, Impusive releases of sterile mosquitoes and interactive dynamics with time delay, J. Biol. Dynam. 14(1) (2020), pp. 289–307.
  • Y. Li, F. Kamara, G. Zhou, S. Puthiyakunnon, C. Li, Y. Liu, Y. Zhou, L. Yao, G. Yan, X. Chen, and P. Kittayapong, Urbanization increases Aedes albopictus larval habitats and accelerates mosquito development and survivorship, PLoS Negl. Trop. Dis. 8(11) (2014), p. e3301.
  • F. Liu, C. Zhou, and P. Lin, Studies on the population ecology of Aedes albopictus 5. The seasonal abundance of natural population of Aedes albopictus in Guangzhou, Acta Sci. Natur. Universitatis Sunyatseni 29(2) (1990), pp. 118–122.
  • F. Liu, C. Yao, P. Lin, and C. Zhou, Studies on life table of the natural population of Aedes albopictus, Acta Sci. Natur. Universitatis Sunyatseni 31(3) (1992), pp. 84–93.
  • Y. Liu, Z. Guo, M. Smaily, and L. Wang, A Wolbachia infection model with free boundary, J. Biol. Dynam. 14(1) (2020), pp. 515–542.
  • Z. Liu, Y. Zhang, and Y. Yang, Population dynamics of Aedes (Stegomyia) albopictus (Skuse) under laboratory conditions, Acta Entomol. Sin. 28(3) (1985), pp. 274–280.
  • E.E. Ooi, K.T. Goh, and D.J. Gubler, Dengue prevention and 35 years of vector control in Singapore, Emerg. Infect. Diseases 12 (2006), pp. 887–893.
  • P.A. Ross, M. Turelli, and A.A. Hoffmann, Evolutionary ecology of Wolbachia releases for disease control, Annu. Rev. Genet. 53 (2019), pp. 93–116.
  • H.L. Smith, An Introduction to Delay Differential Equations with Applications to the Life Sciences, Springer, New York, Vol. 57, 2011.
  • L. Toma, F. Severini, M. Luca, A. Bella, and R. Romi, Seasonal patterns of oviposition and egg hatching rate of Aedes albopiczus in Rome, J. Am. Mosq. Control. Assoc. 19(1) (2003), pp. 19–22.
  • J. Waldock, N.L. Chandra, J. Lelieveld, Y. Proestos, E. Michael, G. Christophides, and P.E. Parham, The role of environment variables on Aedes albopictus biology and Chikungunya epidemiology, Pathogens Global Health. 107 (2013), pp. 224–240.
  • T. Walker, P.H. Johnson, L.A. Moreika, I. Iturbe-Ormaetxe, F.D. Frentiu, C.J. Mcmeniman, Y.S. Leong, Y. Dong, A.L. Lloyd, S.A. Ritchie, S.L. O'Neill, and A.A. Hoffmann, The wMel Wolbachia strain blocks dengue and invades caged Aedes aegypti populations, Nature 476 (2011), pp. 450–453.
  • E. Waltz, US reviews plan to infect mosquitoes with bacteria to stop disease, Nature 89(1) (2016), pp. 450–451.
  • Y. Wang, X. Liu, C. Li, T. Su, J. Jin, Y. Guo, D. Ren, Z. Yang, and F Meng, A survey of insecticide resistance in Aedes albopictus (Diptera: Culicidae) during a 2014 dengue fever outbreak in Guangzhou, China, China. J. Econ. Entomol. 110(1) (2017), pp. 239–244.
  • WHO, Global strategy for dengue prevention and control 2012–2020. Geneva: World Health Organization, 2012.
  • Z. Xi, C.C. Khoo, and S.L. Dobson, Wolbachia establishment and invasion in an Aedes aegypti laboratory population, Science 310 (2005), pp. 326–328.
  • J. Yu, Modeling mosquito population suppression based on delay differential equations, SIAM J. Appl. Math. 78(6) (2018), pp. 3168–3187.
  • J. Yu and J. Li, Dynamics of interactive wild and sterile mosquitoes with time delay, J. Biol. Dynam.13(1) (2019), pp. 606–620.
  • J. Yu and J. Li, Global asymptotic stability in an interactive wild and sterile mosquito model, J. Diff. Equ. 269(7) (2020), pp. 6193–6215.
  • J. Yu and B. Zheng, Modeling Wolbachia infection in mosquito population via discrete dynamical models, J. Diff. Equ. Appl. 25 (2019), pp. 1549–1567.
  • L. Zhang, L. Tan, H. Ai, C. Lei, and F. Yuan, Laboratory and field studies on the oviposition pattern of Aedes albopictus, Acta Parasitol. Et. Med. Entomol. Sin. 16 (2009), pp. 219–223.
  • B. Zheng, M. Tang, and J. Yu, Modeling Wolbachia spread in mosquitoes through delay differential equation, SIAM J. Appl. Math. 74 (2014), pp. 743–770.
  • B. Zheng and J. Yu, Characterization of Wolbachia enhancing domain in mosquitoes with imperfect maternal transmission, J. Biol. Dynam. 12(1) (2018), pp. 596–610.
  • 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 (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.G. Parker, J.R.L. Gilles, 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.A. Baton, A.A. Hoffmann, and Z. Xi, Incompatible and sterile insect techniques combined eliminate mosquitoes, Nature 572 (2019), pp. 56–61.
  • Z. Zhong and G. He, The life table of laboratory Aedes albopictus under various temperatures, Academic J. Sun Yat-sen University Medical Sci. 9(3) (1988), pp. 35–39.
  • Z. Zhong and G. He, The life and fertility table of Aedes albopictus under different temperatures, Acta Entom. Sinica 33(1) (1990), pp. 64–70.