Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Monotonicity properties arising in a simple model of Wolbachia invasion for wild mosquito populations

  • In this paper, we propose a simplified bidimensional Wolbachia infestation model in a population of Aedes aegypti mosquitoes, preserving the main features associated with the biology of this species that can be found in higher-dimensional models. Namely, our model represents the maternal transmission of the Wolbachia symbiont, expresses the reproductive phenotype of cytoplasmic incompatibility, accounts for different fecundities and mortalities of infected and wild insects, and exhibits the bistable nature leading to the so-called principle of competitive exclusion. Using tools borrowed from monotone dynamical system theory, in the proposed model, we prove the existence of an invariant threshold manifold that allows us to provide practical recommendations for performing single and periodic releases of Wolbachia-carrying mosquitoes, seeking the eventual elimination of wild insects that are capable of transmitting infections to humans. We illustrate these findings with numerical simulations using parameter values corresponding to the wMelPop strain of Wolbachia that is considered the best virus blocker but induces fitness loss in its carriers. In these tests, we considered multiple scenarios contrasting a periodic release strategy against a strategy with a single inundative release, comparing their effectiveness. Our study is presented as an expository and mathematically accessible tool to study the use of Wolbachia-based biocontrol versus more complex models.

    Citation: Diego Vicencio, Olga Vasilieva, Pedro Gajardo. Monotonicity properties arising in a simple model of Wolbachia invasion for wild mosquito populations[J]. Mathematical Biosciences and Engineering, 2023, 20(1): 1148-1175. doi: 10.3934/mbe.2023053

    Related Papers:

    [1] Yunfeng Liu, Guowei Sun, Lin Wang, Zhiming Guo . Establishing Wolbachia in the wild mosquito population: The effects of wind and critical patch size. Mathematical Biosciences and Engineering, 2019, 16(5): 4399-4414. doi: 10.3934/mbe.2019219
    [2] Luis Almeida, Michel Duprez, Yannick Privat, Nicolas Vauchelet . Mosquito population control strategies for fighting against arboviruses. Mathematical Biosciences and Engineering, 2019, 16(6): 6274-6297. doi: 10.3934/mbe.2019313
    [3] Bo Zheng, Wenliang Guo, Linchao Hu, Mugen Huang, Jianshe Yu . Complex wolbachia infection dynamics in mosquitoes with imperfect maternal transmission. Mathematical Biosciences and Engineering, 2018, 15(2): 523-541. doi: 10.3934/mbe.2018024
    [4] Bo Zheng, Lihong Chen, Qiwen Sun . Analyzing the control of dengue by releasing Wolbachia-infected male mosquitoes through a delay differential equation model. Mathematical Biosciences and Engineering, 2019, 16(5): 5531-5550. doi: 10.3934/mbe.2019275
    [5] Martin Strugarek, Nicolas Vauchelet, Jorge P. Zubelli . Quantifying the survival uncertainty of Wolbachia-infected mosquitoes in a spatial model. Mathematical Biosciences and Engineering, 2018, 15(4): 961-991. doi: 10.3934/mbe.2018043
    [6] Mugen Huang, Moxun Tang, Jianshe Yu, Bo Zheng . The impact of mating competitiveness and incomplete cytoplasmic incompatibility on Wolbachia-driven mosquito population suppressio. Mathematical Biosciences and Engineering, 2019, 16(5): 4741-4757. doi: 10.3934/mbe.2019238
    [7] Yufei Wang, Huidong Cheng, Qingjian Li . Dynamic analysis of wild and sterile mosquito release model with Poincaré map. Mathematical Biosciences and Engineering, 2019, 16(6): 7688-7706. doi: 10.3934/mbe.2019385
    [8] Daiver Cardona-Salgado, Doris Elena Campo-Duarte, Lilian Sofia Sepulveda-Salcedo, Olga Vasilieva, Mikhail Svinin . Optimal release programs for dengue prevention using Aedes aegypti mosquitoes transinfected with wMel or wMelPop Wolbachia strains. Mathematical Biosciences and Engineering, 2021, 18(3): 2952-2990. doi: 10.3934/mbe.2021149
    [9] Rajivganthi Chinnathambi, Fathalla A. Rihan . Analysis and control of Aedes Aegypti mosquitoes using sterile-insect techniques with Wolbachia. Mathematical Biosciences and Engineering, 2022, 19(11): 11154-11171. doi: 10.3934/mbe.2022520
    [10] Yun Li, Hongyong Zhao, Kai Wang . Dynamics of an impulsive reaction-diffusion mosquitoes model with multiple control measures. Mathematical Biosciences and Engineering, 2023, 20(1): 775-806. doi: 10.3934/mbe.2023036
  • In this paper, we propose a simplified bidimensional Wolbachia infestation model in a population of Aedes aegypti mosquitoes, preserving the main features associated with the biology of this species that can be found in higher-dimensional models. Namely, our model represents the maternal transmission of the Wolbachia symbiont, expresses the reproductive phenotype of cytoplasmic incompatibility, accounts for different fecundities and mortalities of infected and wild insects, and exhibits the bistable nature leading to the so-called principle of competitive exclusion. Using tools borrowed from monotone dynamical system theory, in the proposed model, we prove the existence of an invariant threshold manifold that allows us to provide practical recommendations for performing single and periodic releases of Wolbachia-carrying mosquitoes, seeking the eventual elimination of wild insects that are capable of transmitting infections to humans. We illustrate these findings with numerical simulations using parameter values corresponding to the wMelPop strain of Wolbachia that is considered the best virus blocker but induces fitness loss in its carriers. In these tests, we considered multiple scenarios contrasting a periodic release strategy against a strategy with a single inundative release, comparing their effectiveness. Our study is presented as an expository and mathematically accessible tool to study the use of Wolbachia-based biocontrol versus more complex models.



    It is widely known that female mosquitoes of the species Aedes aegypti are major transmitters of dengue and other vector-borne infections. When deliberately infected with Wolbachia, they lose their vector competence by becoming far less capable of developing a viral load sufficient for transmission of the virus to humans. Due to this remarkable feature, Wolbachia-based biocontrol of mosquito populations has recently emerged as a novel method for the prevention and control of vector-borne infections and is accepted as an ecologically friendly and potentially cost-effective method [1,2,3,4,5].

    The goal of Wolbachia-based biocontrol is the eventual elimination of wild insects (capable of transmitting the virus to humans) by performing single or periodic releases of Wolbachia-carrying mosquitoes in some determined localities initially populated by wild mosquitoes. The practical implementation of this method requires the mass-rearing of a large quantity of Wolbachia-infected insects for posterior releases, and the desired result depends on the progressive Wolbachia invasion and its durable establishment in wild mosquito populations. The final outcome of this process is usually referred to as "population replacement" [6,7,8].

    In the literature from the last decade, one can find a variety of Wolbachia infestation models created with the purpose, among others, of evaluating biocontrol strategies that seek suppression of wild mosquito populations. Previous studies include some frequency-based models formulated as a single equation [9,10,11], models considering only female insects [6,12,13,14,15], models accounting for sex structure [16,17,18,19] or stage structure [20,21,22,23], and more sophisticated models that take into account the mosquito maturation delay or their age structure [18,19,24].

    The main features of all these models describing the natural dynamics of Wolbachia are related to mimicking vertical transmission and the interference of the reproductive outcomes induced by cytoplasmic incompatibility (or CI reproductive phenotype). In this context, it is meaningful to note that CI occurs when a female, uninfected by Wolbachia, is inseminated by an infected male, producing inviable eggs. Thus, the CI reproductive phenotype suppresses the growth of the uninfected population and facilitates Wolbachia spread.

    Different strains of Wolbachia may induce either perfect (100%) or imperfect (less than 100 %) maternal transmission and CI reproductive phenotype. Notably, some of the existent models are designed to address imperfect maternal transmission and/or imperfect CI [6,9,17,18,20,21,23,24]. Nevertheless, laboratory trials evince that Aedes aegypti mosquitoes that were deliberately infected by the wMelPop Wolbachia strain (regarded as the best blocker of arboviral infections [3,5,25,26,27,28]) exhibit almost perfect CI and maternal transmission [25,29]. Table 1 summarizes the results of matings between infected and uninfected mosquitoes when the maternal transmission and CI are perfect (100%). On the other hand, many scholars also indicate that the wMelPop strain is associated with high "fitness costs" since it reduces female fecundity, the viability of eggs, and the lifespan of infected mosquitoes [7,8,10,25]. The latter makes the spread of wMelPop Wolbachia infection a rather challenging task.

    Table 1.  Illustration of the CI reproductive phenotype and maternal transmission of Wolbachia.
    Mosquito offspring
    Adults Wolbachia-infected female Uninfected female
    Wolbachia-infected male Infected Inviable eggs
    Uninfected male Infected Uninfected

     | Show Table
    DownLoad: CSV

    The group of models assuming perfect CI and maternal transmission of Wolbachia exhibit a bistable nature that makes them fully compliant with the so-called principle of competitive exclusion [30]. This phenomenon implies the existence of a certain (dynamic) threshold in the current level (or frequency) of infection above which Wolbachia is capable of invading and persisting in the uninfected population and below which the Wolbachia-carrying population is driven toward extinction.

    For each infection frequency, this threshold can be expressed in terms of the so-called minimal viable sizes of each population (infected and uninfected) that are tightly related to the frequency-dependent Allee effect. Anticipated knowledge of the minimal viable size of the Wolbachia-infected population corresponding to the current size of the wild population is the key issue for determining the appropriate size of release(s) for implementation of Wolbachia-based biocontrol, and one of the main goals of the present work consists of assessing the release size(s) through the use of a simplified bidimensional model of Wolbachia invasion.

    In effect, many scholars have intuitively detected the presence of the aforementioned threshold in the infection frequency (see, e.g., [6,10,11,12,14,15,16,17,18,19,22,23]). However, none of these works were focused on explicit identification of the minimal viable sizes of infected and uninfected populations nor their useful significance for the practical implementation of Wolbachia-based biocontrol. The present paper intends to contribute to this strand of research and corroborate the result obtained in [31] based on a different model.

    All models describing Wolbachia invasion are competitive; however, only a few of them rigorously exhibit the property of monotonicity, namely:

    the four-dimensional stage-structured model proposed in [22];

    the bidimensional model that is further reduced to a single equation in terms of the infection frequency [6];

    the reduced bidimensional version [31] of the sex-structured model initially proposed in [16]*;

    *Notably, the four-dimensional sex-structured model developed in [16] is not monotone, while its reduced bidimensional version studied in [31] clearly exhibits the property of monotonicity.

    the simplified version without maturation delay considered in [19].

    This important property makes the theoretical analysis of these models simpler by making use of the variety of research results developed for monotone dynamical systems and assembled in [32]. In particular, the application of the theory of monotone systems to competitive dynamics allows identifying a partial order under which a competitive monotone system exhibits the so-called saddle-point behavior [33], and the latter bears a strong relationship to the principle of competitive exclusion [30]. In general terms, a system exhibiting the saddle-point behavior possesses two locally stable equilibria on the boundary and one unstable (saddle-point) equilibrium in the interior of the state domain. Moreover, the state domain is divided into three disjointed and invariant parts: two attraction basins of boundary equilibria and the so-called "threshold manifold" (or separatrix) containing the unstable equilibrium that separates the attraction basins. One of the goals of the present work is to identify the threshold manifold and study its properties in light of the practical implementation of Wolbachia-based biocontrol.

    For that purpose, we have developed a simplified bidimensional variant of the four-dimensional model presented in [22] that retains all the key properties of the original four-dimensional model, including the property of monotonicity. Moreover, the model studied in [22] does not require the anticipated knowledge of the carrying capacity parameter (the maximum number of individuals supported by the environment), in contrast to the model proposed in [6].

    Notably, our reduced model bears a certain degree of similarity to the models studied in [19] and [31], but they are not the same. The bidimensional model obtained in [19] by omitting the maturation delay assumes only the density-dependent mortality of both mosquito populations and ignores their natural mortalities, which differ for infected and uninfected insects. In contrast, our bidimensional model accounts for both mortality types (natural and density-dependent). Furthermore, the authors of [19] have, in effect, identified the threshold manifold for their bidimensional model, but they did not discuss its underlying properties in light of the practical implementation of Wolbachia-based biocontrol.

    In its turn, the bidimensional model studied in [31] is the fruit of the reduction of a four-dimensional sex-structured Ricker-type model encompassing the evolution of only adult insects. Therefore, the Ricker-type model presented in [31] is structurally different from our model and the one presented in [19]. Namely, the density-dependence in [31] is included in the recruitment term (which is natural for Ricker-type models) and expresses the insect survival to the adult stage. In contrast, our model and the one studied in [19] exhibit the density-dependent mortality of the insect population attributed to inter- and intra-species competition at the aquatic stage.

    Our main motivation in proposing a bidimensional (or planar) dynamical system is that, in general, such dynamical systems have several important advantages compared to higher-dimensional systems. First, they concede a comprehensive visualization of the system's behavior in the phase plane that not only facilitates the conceptual theoretical analysis of the model but also provides meaningful interpretations of the potential outcomes of the model. Second, for planar dynamical systems, there are numerous optimal control tools [34,35,36,37] that can be applied to the model proposed in this paper to obtain an analytically optimal synthesis of various optimization problems related to biocontrol purposes. Such analytical solutions can be later tested in more complex models. Although we do not analyze optimal control problems in this work, the proposed model has been designed to serve these purposes in the future and to provide major insights regarding the evolution of both mosquito populations under the action of biocontrol. In addition, as we discuss in the final section, the properties of the proposed model enable the impulsive releases of Wolbachia-infected mosquitoes in order to achieve the population replacement. Notably, in [38] the author proposes to employ feedback controls for reaching the steady state representing the population replacement. Motivated by that work and our exercise carried out with impulsive releases (Section 5) we believe that our proposed model is well-suited for determining optimal feedback in analytical form.

    In this context, it is worth pointing out that planar systems are less precise in describing the evolution of populations than higher-dimensional models.

    The paper is organized as follows. In Section 2, we introduce a bidimensional Wolbachia infestation model for populations of Aedes aegypti mosquitoes. In Section 3, we establish some basic properties of the introduced model. In Section 4, we prove that the reduced model retains the property of monotonicity inherent from the original four-dimensional model developed in [22]. Using this remarkable property and other tools from the theory of monotone dynamical systems, we establish the existence of an invariant threshold manifold for the proposed bidimensional model. Finally, in Section 5, we discuss the core properties of the points located on the threshold manifold and propose their practical interpretations for performing single and periodic releases of Wolbachia-infected mosquitoes to reach an eventual elimination of wild insects and thus achieve the population replacement.

    Let us consider two populations of mosquitoes, PN(t):=FN(t)+MN(t) and PW(t):=FW(t)+MW(t) present at the day t0 in some locality, where FN(t) and MN(t) stand, respectively, for the density of female and male insects that are free of Wolbachia symbiotic bacterium while FW(t) and MW(t) denote, respectively, the density of female and male insects infected with Wolbachia bacterium.

    Scientific evidence [39,54] suggests that wild female and male mosquitoes are often evenly distributed; therefore, let us suppose that N(t):=FN(t)=MN(t) for all t0. A similar assumption can be introduced, for the sake of simplicity, regarding Wolbachia-carrying mosquitoes, that is, W(t):=FW(t)=MW(t) for all t0. Then, the frequency of Wolbachia infection in the total mosquito population, PW(t)PN(t)+PW(t) can be determined by N(t) and W(t) as W(t)N(t)+W(t).

    To propose a simplified model of Wolbachia invasion, we have chosen as a starting point the stage-structured model of Wolbachia infestation with four state variables developed by Bliman et al [22], and our final goal is to design a bidimensional (reduced) version of this model with similar characteristics. The original four-dimensional model [22] accounts for the larval (or aquatic) stage; therefore, it is worth recalling that immature mosquitoes undergo inter- and intra-species competition and exhibit density-dependent mortality at this stage. Thus, our reduced bidimensional version of the model should bear density-dependence in the mortality term, in contrast to another bidimensional model [31] where density-dependence is included in the recruitment term.

    Let us err on the side of caution while reducing the model's dimension and recall that a "good" model for describing Wolbachia invasion with perfect maternal transmission and CI must necessarily account for the following features:

    (ⅰ) Maternal transmission of the bacterium Wolbachia to the next generation. This feature implies that Wolbachia-infected mosquitoes are progenies of Wolbachia-carrying females. A wild female cannot produce Wolbachia-infected offspring (see Table 1).

    (ⅱ) The reproductive phenotype of cytoplasmic incompatibility (CI). This feature implies that Wolbachia-carrying females are capable of producing viable and Wolbachia-infected offspring after mating with either wild or Wolbachia-carrying males. On the other hand, wild females produce inviable offspring after mating with Wolbachia-carrying males.

    (ⅲ) Positive invariance, well-posedness, and the existence of the absorbing set that attracts all the system trajectories. Any model is "biologically viable" if its state variables are nonnegative for all t0 and their underlying trajectories are bounded.

    (ⅳ) Bistable nature. This feature expresses the so-called principle of competitive exclusion [30] according to which only one mosquito population (either with or without Wolbachia) should ultimately survive and persist.

    In mathematical terms, feature (ⅳ) implies the presence of two local attractors (boundary equilibria) and an unstable strictly positive coexistence equilibrium in between. From the biological standpoint, the last feature (ⅳ) is also directly related to the so-called frequency-dependent Allee effect and implies the existence of a certain threshold in the frequency of Wolbachia infection above which the wild population is eventually driven toward extinction, and Wolbachia successfully invade the mosquito population. Such a threshold is usually referred to as the "minimal viable population size" of wild mosquitoes.

    In addition to key features given by (ⅰ)–(ⅳ), the four-dimensional model developed in [22] possesses another important property related to its monotonicity. Namely, the semiflow associated with that four-dimensional model is monotone and strongly order-preserving for the partial order induced by the cone K:=R×R×R+×R+. Notably, there are several models of Wolbachia invasion with perfect maternal transmission and CI, both four-dimensional [14,16,17] and bidimensional [12,13], that do not exhibit such an important property. Therefore, it is highly desirable to conserve this property in the reduced bidimensional version of the original model developed in [22].

    To reduce the four-dimensional model developed in [22] to a bidimensional scenario, the aquatic and aerial stages of each mosquito class (with and without Wolbachia) can be "merged" into one state variable denoted by N(t) and W(t) for noninfected and Wolbachia-infected insects, respectively. As a result, we obtain the following model without the stage structure:

    {dNdt=F(N,W):=ρNN(NN+W)αNNβNN(N+W)(2.1a)dWdt=G(N,W):=ρWWαWWβWW(N+W)(2.1b)

    that is bidimensional and describes the time evolution of adult mosquito populations, N(t) and W(t).

    Let us provide a brief description of positive constant parameters included in the model (2.1):

    ● The parameters ρN and ρW represent the fecundity rates of uninfected and Wolbachia-infected insects, respectively, in the absence of competition (i.e., a mean number of adult mosquitoes produced by one female on average per day during her lifespan).

    ● The parameters αN and αW refer to a natural mortality rate of uninfected and Wolbachia-infected insects, respectively (note that 1/αN and 1/αW express the average lifespan of noninfected and Wolbachia-infected mosquitoes).

    Notably, system (2.1) bears a certain degree of similarity to the one studied in [19]. However, system (2.1) explicitly accounts for the natural mortalities, αN and αW, of wild and Wolbachia-infected mosquitoes while the model in [19] ignores them and only considers the density-dependent mortalities (expressed by the last terms in both equations of (2.1)).

    ● The parameters βN and βW are associated with the competition between two mosquito populations for food resources, breeding sites, and mating opportunities, including larvae development into adults under density dependence.

    To guarantee survival and persistence of each population in the absence of another, the following conditions are introduced:

    ρN>αN,ρW>αW. (2.2)

    It is worth pointing out that conditions (2.2) assert that each mosquito population is viable and persistent in the absence of another. In other words, the recruitment or birth rate ρN (resp. ρW) is greater than the natural density-independent mortality rate αN (resp. αW). Therefore, conditions (2.2) are very realistic, whereas their violation would imply the inviability of each mosquito population.

    It is easy to verify that condition (ⅰ) on p. 1152 is fulfilled by Eq (2.1b) according to which Wolbachia-carriers are progenies of W only (cf. the positive term in the right-hand side of (2.1b) that expresses the recruitment of W). Additionally, Eq (2.1a) captures the condition (ⅱ) on p. 1152 that refers to the CI-phenotype in the sense that recruitment of wild mosquitoes (cf. the positive term in the right-hand side of (2.1b)) is proportional to the number of matings between wild males and females. These outcomes agree with the description of maternal transmission and the reproductive phenotype of cytoplasmic incompatibility presented in Table 1.

    To illustrate the remaining features, a deeper analysis of the model (2.1) is required, and the latter is presented in the next section.

    This section is focused on establishing the key features (ⅲ) and (ⅳ) on p. 1152 for the reduced bidimensional model (2.1). The well-posedness of the model is attested in subsection 3.1, and the stability analysis of the system (2.1) is carried out in subsection 3.2.

    To verify the condition (ⅲ) on p. 1152, we establish and prove the following result.

    Proposition 1. With condition (2.2) in force, dynamical system (2.1) is invariant in the positive cone R2+. Additionally, there exists a compact set XR2+ such that the trajectories of (2.1) engendered by initial conditions (N(0),W(0))X remain in X for all t0.

    Proof. It is immediately noted that

    dNdt|N=0=0anddWdt|W=0=0

    which plainly indicates that system (2.1) is invariant in the positive cone R2+, i.e., its trajectories N(t) and W(t) engendered by (N(0),W(0))R2+ remain in R2+ for all t0, and thus

    N(t)0andW(t)0forallt0.

    Furthermore, let us observe that for F(N,W) and G(N,W) the following hold:

    F(N,W)βNN(ρNαNβNN) and  G(N,W)βWW(ρWαWβWW).

    Thus, one has

    dNdt|N=ρNαNβN0anddWdt|W=ρWαWβW0,

    where these quantities are strictly negative when N>ρNαNβN and W>ρWαWβW. Hence, it follows that

    N(t)max{N,N(0)}andW(t)max{W,W(0)}

    where

    N:=ρNαNβNandW:=ρWαWβW. (3.1)

    In other words, the compact set

    X:={(N,W)R2+: 0NN, 0WW} (3.2)

    is invariant in the sense that all trajectories (N(t),W(t)) of (2.1) engendered by the initial conditions (N(0),W(0))X remain in X for all t0.

    It stems from the proof of Proposition 1 that the reduced dynamical system (2.1) is dissipative [32], and X is referred to as an absorbing set. This means that all trajectories N(t),W(t) of (2.1) engendered by (N(0),W(0))R2+X are attracted to X and there is a finite ˉt>0 such that

    (N(ˉt),W(ˉt))Xforall tˉt.

    Thus, we can conclude that all solutions of (2.1) engendered by nonnegative initial conditions (N(0),W(0))R2+ are uniformly ultimately bounded.

    Remark 1. It is easy to check that R2+ contains three subsets that are invariant with respect to solutions of the system (2.1):

    1. Set Ω0:={(0,0)} containing the origin is invariant since for the initial condition (N(0),W(0))Ω0 we have that (N(t),W(t))Ω0 for all t0.

    2. Set ΩN:={(N,W)R2+: N>0,W=0} that contains the N-axis is invariant. In effect, W(0)=0 implies the absence of Wolbachia-carriers for all t0 and the dynamical system (2.1) turns into the logistic equation for the wild population:

    dNdt=N(ρNαNβNN), (3.3)

    whose solutions engendered by N(0)>0 tend to the carrying capacity N as t. Therefore, if (N(0),W(0))ΩN, we have that (N(t),W(t))ΩN for all t0.

    3. Set ΩW:={(N,W)R2+: N=0,W>0} that contains the W-axis is also invariant because N(0)=0 implies the absence of wild mosquitoes for all t0 and the dynamical system (2.1) turns into the logistic equation for the Wolbachia-carrying population

    dWdt=W(ρWαWβWW), (3.4)

    whose solutions engendered by W(0)>0 tend to the carrying capacity W as t. Therefore, if (N(0),W(0))ΩW, we have that (N(t),W(t))ΩW for all t0.

    To verify the condition (ⅳ) on p. 1152, it is necessary to determine all possible equilibria of the system (2.1) that are solutions of the algebraic system

    {0=F(N,W)=ρNN(NN+W)αNNβNN(N+W)(3.5a)0=G(N,W)=ρWWαWWβWW(N+W).(3.5b)

    After some manipulations, we determine that the dynamical system (2.1) has four steady states (or equilibria), all of them are located in XR2+, defined in (3.2), and their coordinates can be explicitly expressed in terms of the model's parameters (the detailed derivation of the model's equilibria is provided in Appendix A).

    ● Trivial steady state E0=(0,0) that corresponds to the extinction of both mosquito populations.

    ● Boundary steady state EN=(N,0) that corresponds to the survival and persistence of the wild mosquito population and eventual extinction of the Wolbachia-carrying population.

    ● Boundary steady state EW=(0,W) that corresponds to the survival and persistence of the Wolbachia-carrying population and eventual extinction of the wild mosquito population.

    ● Strictly positive steady state Ec=(Nc,Wc) that corresponds to the coexistence of both mosquito populations and where

    Nc=W(1βNρN(NW)),Wc=WβNρN(NW) (3.6)

    and Nc+Wc=W.

    Figure 1 displays the abovementioned equilibria on the phase portrait of the system (2.1). This figure also shows two nullclines or zero-growth isoclines of the system: the N-nullcline (i.e., the curve F(N,W)=0) is plotted in blue color, and the W-nullcline (i.e., the curve G(N,W)=0) is plotted in red color.

    Figure 1.  Phase portrait of the dynamical system (2.1) with four equilibria E0,EN,EW, and Ec, joined by N-nullcline (blue-colored curve) and W-nullcline (red-colored curve) and the attraction regions of boundary equilibria EN,EW defined by the two order intervals in Proposition 4.

    Recall that the violation of conditions (2.2) (i.e., a situation with ρN<αN or ρW<αW) represents the case in which, for each population, the mortality rate is higher than the recruitment rate. The only outcome in such a case would be the extinction of both populations, and the trivial steady state E0=(0,0) would be globally asymptotically stable because the other (nontrivial) equilibria EN,EW, and Ec would become unfeasible (that is, with negative coordinates, cf. (3.1) and (3.6)). In other words, the absorbing set X defined by (3.2) would contain only the trivial equilibrium, and all the system trajectories would be attracted to E0. We will not take into consideration this case, and from now on, we will assume that both conditions in (2.2) are always in force.

    On the other hand, the existence of a strictly positive equilibrium Ec=(Nc,Wc) requires that

    NW>0. (3.7)

    From the biological standpoint, condition (3.7) is rather credible. Let us recall that Wolbachia infection negatively affects the individual fitness of its carriers by reducing the females' fecundity and increasing the natural mortality of mosquitoes (see the exhaustive review by Dorigatti et al., 2018 [25] and more precise references therein). Therefore, it holds that

    ρW<ρNandαW>αN.

    Furthermore, several recent studies have determined that, at high levels of intraspecific competition, Wolbachia-infected larvae experience reduced survival [45,46,55]. In other words, the Wolbachia-carrying insects exhibit higher mortality due to intraspecific competition than wild insects, meaning that βNβW. From the above rationale, it follows that

    ρNαNβN=N>W=ρWαWβW

    which is equivalent to (3.7).

    The following result is claimed in order to complete the proof of the condition (ⅳ) on p. 1152.

    Proposition 2. With conditions (2.2) and (3.7) in force, the steady states EN=(N,0) and EW=(0,W) are asymptotically stable while the coexistence equilibrium Ec=(Nc,Wc) is unstable. Furthermore, the trivial steady state E0 is a source (nodal repeller).

    Proof. Local stability of each equilibrium can be determined by the signs of eigenvalues of the Jacobian matrix associated with the system (2.1):

    J(N,W):=[ρN(1W2(N+W)2)αNβN(W+2N)N(βN+ρNN(N+W)2)βWWρWαWβW(N+2W)]. (3.8)

    Direct evaluation of (3.8) in the boundary steady state EN=(N,0) renders that

    J(N,0)=[ρNαN2βNNNβNρN0ρWαWβWN],

    which is an upper-triangular matrix and its eigenvalues λNi,i=1,2 are located on the main diagonal. In effect,

    λN1=ρNαN2βNN=(ρNαN)<0,λN2=ρWαWβWN=βW(NW)<0,

    are both negative due to the conditions (2.2), (3.7). Therefore, EN=(N,0) is locally asymptotically stable (nodal attractor).

    Furthermore, by direct substitution of the boundary steady state EW=(0,W) in (3.8), we obtain

    J(0,W)=[αNβNW0βWWρWαW2βWW]

    which is a lower-triangular matrix and its eigenvalues λWi,i=1,2 are located on the main diagonal. Consequently,

    λW1=αNβNW<0,λW2=ρWαW2βWW=(ρWαW)<0

    are both negative by virtue of (2.2). Therefore, EW=(0,W) is locally asymptotically stable (nodal attractor).

    On the other hand, direct evaluation of the Jacobian matrix (3.8) in the coexistence steady state Ec=(Nc,Wc) defined by (3.6) results in a rather cumbersome approach. However, let us recall that det(J(Nc,Wc))=λc1λc2 where λci,i=1,2 denote two eigenvalues of J(Nc,Wc). Therefore, to prove that the coexistence equilibrium Ec=(Nc,Wc) is unstable, it is sufficient to show that det(J(Nc,Wc)) is strictly negative. In effect,

    det(J(Nc,Wc))=[ρNαNρNW2c(Nc+Wc)2βN(Wc+2Nc)][ρWαWβW(Nc+2Wc)]βWWcNc[βN+ρNNc(Nc+Wc)2]{Nc+Wc=WρNαN=βNNρWαW=βWW}=βWWc[βN(NWNc)ρN(WcW)2]βWWc[βNNc+ρN(NcW)2]=βWWc[βN(NW)+ρNW2(N2cW2c)]{ρN=βNWWc(NW)}=βWβN[Wc(NW)+(NW)(NcWc)]=βWβN(NW)Nc<0.

    Thus, we have det(J(Nc,Wc))=λc1λc2<0 meaning that λc1 and λc2 have opposite signs. Therefore, the coexistence equilibrium Ec=(Nc,Wc) is unstable (saddle point).

    To show that E0=(0,0) is repelling, we cannot merely substitute its coordinates in the Jacobian matrix (3.8) since J(0,0) cannot be directly computed at this steady state. Let us now recall (see Remark 1) that the origin is a self-contained invariant set and that the system trajectories engendered by W(0)=0 and N(0)>0 (which can be arbitrarily small) are attracted to EN, while the trajectories engendered by N(0)=0 and W(0)>0 (which can also be arbitrarily small) are attracted to EW. Thus, every vicinity of E0 contains initial conditions from which the system trajectories move away from the origin E0=(0,0) in the direction of either EN or EW. This means that E0 is a source. This completes the proof of Proposition 2 and also establishes the validity of the condition (ⅳ) presented on p. 1152.

    In the next section, we explore monotonicity and other important properties of the reduced system (2.1).

    To show that a certain dynamical system is monotone, it should be recalled that a dynamical system is called monotone when the flow generated by this system preserves some partial order [32]. Therefore, first, we define the partial order induced by a convex cone KR2 which is typically a quadrant of R2 for bidimensional dynamical systems [30,53]. Given two elements X=(x1,x2)K and Y=(y1,y2)K, we write:

    XKY if YXK;

    X<KY if YXK0;

    XKY if YXIntK.

    Partial order defined by the last two items is also referred to as "strict order" and "strong order", respectively.

    The sets

    ]]X,Y[[K:={ZR2+: X<KZ<KY},[[X,Y]]K:={ZR2+: XKZKY}

    are referred to as order intervals (open and closed, respectively) induced by the cone K.

    Let XR2+ such that X=(x1,x2):=(N,W) defines the state vector of the system (2.1). Denote as F:R2+R2+,F:=(F(X),G(X)) the vector field whose components F and G are defined in (2.1) as scalar functions of X=(N,W). Using these notations, our system (2.1) can be written as

    dXdt=F(X) (4.1)

    and its solution, engendered by an initial condition X0=(N(0),W(0))R2+, can be denoted as Φ(t,X0).

    The positive semiflow of the system (2.1) or (4.1) generated by the vector field F(X) is then the continuous mapping defined by Φ: R+×R2+×R2+, where Φ(t,X) denotes the solution of (4.1) that satisfies X(0)=X. Here, we consider only the positive semiflow of the system (4.1), that is, with t[0,), since we are interested in the system's behavior in forward time.

    The semiflow Φ(t,) is called monotone (resp. strongly monotone) on a subset AR2+, with respect to partial order induced by the cone K if

    Φ(t,X)KΦ(t,Y)(resp. Φ(t,X)KΦ(t,Y) )wheneverX,YA;XKY(resp. X<KY )forall t0.

    Furthermore, the semiflow Φ(t,) is called strongly order-preserving on AR2+ (or SOP, for briefness) if Φ(t,) is monotone on A and, whenever X<KY, there exist open neighborhoods VX of X and VY of Y such that

    Φ(t,VX)KΦ(t,VY)for t>0.

    In other words, the partial order induced by the cone K is preserved for every ordered pair ˆXKˆY, where ˆXVX,ˆYVY within these open vicinities VX and VY for all t>0.

    Proposition 3. System (2.1), (4.1) is monotone in R2+ and strongly order-preserving (SOP) in the interior of R2+ with respect to the partial order induced by the cone K:=R+×R.

    Proof. It is easy to see that the partial order induced by K:=R+×R is related to the "standard" order (induced by R2+) in the following sense. For any two elements X=(x1,x2)R2+ and Y=(y1,y2)R2+, it is fulfilled that

    XKYifandonlyifx1y1andx2y2.

    A similar statement also holds with "K" replacing "K" and "<" replacing "".

    Using the idea of order isomorphism (for more details, see [32,Section 3.5]) induced by the diagonal matrix

    P:=[1001],P=P1=PT

    it is immediate to conclude that

    XKYifandonlyifPXPY.

    According to [32,40], a simple change of variables

    Z:=PX=P[NW]=[NW]andPF(X)=PF(PZ)

    leads to the cooperative dynamical system

    dZdt=PF(PZ) (4.2)

    which is monotone with respect to the "standard" order (induced by R2+). Furthermore, the Jacobian of this system J(Z) is linked to the Jacobian J(N,W) of (2.1), (4.1) by the relationship

    J(Z)=PJ(X)P

    and is a Metzler matrix

    PJ(N,W)P=[ρN(1W2(N+W)2)αNβN(W+2N)N(βN+ρNN(N+W)2)βWWρWαWβW(N+2W)] (4.3)

    for all (N,W)R2+. The latter implies that the semiflow generated by (4.2) preserves the "standard" order (induced by R2+), while the semiflow Φ(t,) generated by the system (2.1), (4.1) preserves the partial order induced by the cone K:=R+×R for all (N,W)R2+. Moreover, the Jacobian matrix (4.3) is irreducible whenever N0,W0. Therefore, the system (2.1), (4.1) is strongly order-preserving in the interior of R2+ in accordance with the results from [32,Chapter 4].

    Let us recall the basin of attraction of a locally asymptotically stable equilibrium X for the dynamical system (4.1) is the set of initial conditions X0 such that the solutions Φ(t;X0) engendered by X0 converge to X as t. According to Proposition 2, our system (2.1) possesses two local attractors, EN=(N,0) and EW=(0,W), whose respective basins of attraction can be written as

    BEN:={(N,W)R2+: limt(N(t),W(t))=(N,0)} (4.4a)
    BEW:={(N,W)R2+: limt(N(t),W(t))=(0,W)} (4.4b)

    where the limits are understood in the "componentwise" sense.

    To determine sets included in the above basins of attraction, we introduce the following lemma obtained from Proposition 3.

    Lemma 1. Systems (2.1) and (4.1) are strongly order-preserving (SOP) on [[EN,EW]]K with respect to the partial order induced by the cone K:=R+×R.

    Proof. We follow the proof scheme used in [22,Theorem 5], using the fact that strong monotonicity implies the SOP property, as indicated in [32,Proposition 1.1.1]. First, from Proposition 1, it can be easily deduced that systems (2.1) and (4.1) are positively invariant in the interior of R2+. Furthermore, from Remark 1, the sets R+×{0}, {0}×R+ and {0}×{0} are invariant as well. Proposition 3 proves that the flow is SOP in the interior of R2+ which is invariant. In each invariant set R+×{0} and {0}×R+, systems (2.1) and (4.1) are reduced to a quadratic equation, equivalent to a logistic equation growth in each case. In each of these sets, in the order relation restricted to each invariant set, the flow is strongly monotone, and thus, the flow is also SOP in these sets. The condition holds trivially in {0}×{0} given that this point is a steady state.

    Finally, having proven that the flow is SOP for initial conditions in each of the invariant sets, from [32,Remark 5.1.1], we can deduce that given an initial condition in one of these sets (i.e., in the border of R2+) and another initial condition in the interior of R2+, strong monotonicity is also preserved, which concludes the proof that the flow is SOP on [[EN,EW]]K because this interval is a subset of R2+.

    The combination of results established by Propositions 2 and 3 and Lemma 1 together provide an essential basis for applying the fundamental Order Interval Trichotomy Theorem (see, e.g., [32,Theorem 2.2.2]) to the order interval [[EN,EW]]K and two subintervals it contains. Let us recall the formulation of this theorem.

    Theorem 1 ([32], Theorem 2.2.2). Let the semiflow Φ(t,) of the dynamical system given in general form (4.1) be SOP with respect to the partial order induced by some cone C on the order interval [[X,Y]]C where X<CY and X,Y are equilibria of (4.1). If ¯Φ(t,[[X,Y]]C) is compact for each t>0, then one of the following holds:

    (ⅰ) There exists another equilibrium Z]]X,Y[[C of the system (4.1).

    (ⅱ) For any X0[[X,Y]]C{Y}, all solutions Φ(t;X0) are attracted to X, that is, limtΦ(t;X0)=X.

    (ⅲ) For any X0[[X,Y]]C{X}, all solutions Φ(t;X0) are attracted to Y, that is, limtΦ(t;X0)=Y.

    To adapt the hypotheses of this theorem to the system (2.1) with C=K, X=EN, and Y=EW, it must be shown that Φ(t,[[EN,EW]]K) has compact closure in R2+. In this context, it is instructive to note that any orbit

    O(N,W):={(N(t),W(t))R2+: t0} (4.5)

    of (2.1) has a compact closure due to the existence of the absorbing set XR2+ (given by (3.2)) which, in effect, coincides with the order interval [[EN,EW]]K:=[0,N]×[0,W].

    It is clear that direct application of Theorem 1 to the order interval [[EN,EW]]K reaffirms, via item (ⅰ), the existence of Ec=(Nc,Wc) such that EN<KEc<KEW. On the other hand, the following proposition relates the two order subintervals of [[EN,EW]]K with the two basins of attraction BEN and BEW of the boundary equilibria EN and EW defined by (4.4).

    Proposition 4. For the system (2.1), it is fulfilled that

    [[EN,Ec[[KBEN,]]Ec,EW]]KBEW,

    where

    [[EN,Ec[[K:=[[EN,Ec]]K{Ec},]]Ec,EW]]K:=[[Ec,EW]]K{Ec}.

    Proof. Dynamical system (2.1) fulfills the hypotheses of Theorem 1, which will be applied to the order intervals [[EN,Ec]]K and [[Ec,EW]]K, sets included in [[EN,EW]]K where the flow is SOP (see Lemma 1). It is immediate to check that there is no equilibrium point inside the order interval [[EN,Ec]]K, and EN is an attractor. By virtue of item (ii) of Theorem 1, all orbits O(N,W) started in [[EN,Ec[[K are attracted to EN. Therefore, [[EN,Ec[[K belongs to BEN. A similar rationale applies to the order interval ]]Ec,EW]]K making use of item (iii) of Theorem 1.

    Figure 1 displays order intervals [[EN,Ec]]K and [[Ec,EW]]K as pink and green rectangles, respectively. The pink rectangle belongs to BENX while the green one lies inside BEWX, that is, each order subinterval is included in the intersection of the underlying basin of attraction and the absorbing set X. The unmarked areas of X contain points X=(N,W) that may belong to either BEN or BEW.

    The following result establishes further properties of the reduced system (2.1) related to its bistability.

    Proposition 5. Dynamical system (2.1) exhibits the saddle-point behavior and there exists an invariant threshold manifold that passes through the positive steady state Ec separating the attraction basins BEN and BEW.

    Proof. Notably, according to [33], a dynamical system is said to admit a "saddle-point behavior" if it possesses two locally stable equilibria on the boundary of the state domain and one unstable (saddle-point) equilibrium in the interior of the state domain. Furthermore, the state domain of the system can be divided into three disjoint and invariant parts: two attraction basins (each containing one stable equilibrium on the boundary) and the so-called "threshold" manifold containing the unstable equilibrium that separates the attraction basins of two locally stable equilibria. Such a manifold is also referred to as the separatrix of two attraction basins.

    The saddle-point behavior of the system (2.1) will be shown by applying the result summarized by H. Smith [53,Theorem 3.2]. For that purpose, we establish the cogency of four necessary hypotheses:

    (H1) The semiflow Φ(t,) generated by the system (2.1) is strictly order-preserving on R2+ with respect to <K and order-compact for each t>0.

    (H2) The origin E0=(0,0) is a repelling equilibrium.

    (H3) All orbits originated on the boundaries of R2+ are confined to these boundaries.

    (H4) If X,YR2+ satisfy X<KY and either X or Y belongs to Int R2+, then Φ(t,X)KΦ(t,Y) for t>0. If X=(X1,X2)R2+ satisfies Xi0,i=1,2, then Φ(t,X) Int R2+ for t>0.

    To show the validity of (H1), we recall the statement of Proposition 3 according to which the semiflow Φ(t,) generated by the system (2.1) is SOP in the interior of R2+ and, therefore, it is also strictly order-preserving in Int R2+. Strict monotonicity of Φ(t,) on the borders of R2+ (which are, in fact, invariant sets ΩN and ΩW, see Remark 1) follows from strict monotonicity of solutions of logistic equations (3.3) and (3.4). Furthermore, the semiflow Φ(t,) generated by the system (2.1) is order-compact since any orbit (4.5) of (2.1) has a compact closure due to the existence of the absorbing set (3.2).

    The hypothesis (H2) is cogent by virtue of Proposition 2, and the hypothesis (H3) is justified by invariance of the boundaries (sets ΩN and ΩW, see Remark 1).

    Finally, the hypothesis (H4) is corroborated by the positiveness of the system trajectories engendered by positive initial conditions (see Proposition 1) along with the SOP property of the semiflow Φ(t,) on Int R2+ (see Proposition 3).

    With the hypotheses (H1)-(H4) in force, we can now apply the result recaptured by H. Smith [53,Theorem 3.2] which basically affirms the following. If there is a unique equilibrium in Ec]]EN,EW[[KIntR2+ and it is a saddle point, then there exists an unordered positively invariant set

    SEc:=R2+(BENBEW)

    that contains the unstable equilibria E0,Ec and consists of points XsIntR2+ such that

    limtΦ(t,Xs)=EcforallXsSEc.

    This completes the proof of Proposition 5 and establishes the saddle-point behavior of the system (2.1).

    The thorough analysis of the reduced model (2.1) performed in Sections 3 and 4 provides very useful insights for practical applications. By recalling the Stable Manifold Theorem (see, e.g., [50,p. 107]), it can be concluded that the invariant "threshold" manifold SEc is, in effect, the stable invariant manifold of the saddle point Ec that is tangent to the eigenvector generated by the negative eigenvalue of J(Ec). Furthermore, there also exists the unstable invariant manifold of the saddle point Ec that is tangent to the eigenvector generated by the positive eigenvalue of J(Ec). The unstable manifold contains two monotone heteroclinic orbits [32] that connect the saddle point Ec with two local attractors EN and EW.

    Figure 2 displays the plots of the stable and unstable manifolds of Ec (green and purple curves, respectively). It is clearly shown that the green curve plays the role of separatrix SEc that divides R2+ into two attraction basins BEN (unshaded area) and BEW (green-shaded region). The situation displayed in Figure 2 fully agrees with previous findings obtained for the case of perfect maternal transmission and CI [19,24,31,43].

    Figure 2.  Separatrix SEc or the stable manifold of Ec (green curve) that separates the attraction basins BEN and BEW (green-shaded region) and the unstable manifold of Ec (purple curve) that connects three equilibria ENEcEW.

    From the biological standpoint, points in SEc indicate minimal viable population sizes of wild and Wolbachia-carrying mosquito populations that are related to the frequency-dependent Allee effect. In fact, if the initial wild mosquito population size is N00, one can compute the minimal viable population size ˆW0=ˆW0(N0), such that (N0,ˆW0)SEc. Thus, if the initial Wolbachia-carrying mosquito population W0 is lower than the minimal viable population size ˆW0, then (N0,W0)BEN and the underlying solution (N(t),W(t)) of the system (2.1) is attracted to EN=(N,0), meaning the ultimate persistence of the wild population and progressive extinction of the Wolbachia-infected population.

    Similarly, if the initial condition (N0,W0) assigned to the system (2.1) lies above SEc, that is, W0>ˆW0, it implies the initial size of the Wolbachia-infected population W0 exceeds its minimal viable population size ˆW0. In such a case, we have that (N0,W0)BEW and the underlying solution (N(t),W(t)) of the system (2.1) is attracted to the desired equilibrium EW=(0,W), meaning the ultimate persistence of the Wolbachia-infected population and progressive extinction of the wild mosquito population.

    In this context, it is worthwhile to highlight that while staying in line with the results presented in [8], our work goes further. In particular, the authors of [22] have identified only the "secured parts" of the attraction basins BEN and BEW in the form of four-dimensional order intervals§. In contrast, our model has allowed us not only to identify the entire basins of attraction BEN and BEW corresponding to both boundary equilibria but also to estimate the minimum viable population sizes. The latter was not addressed in the work [22].

    §Note that their bidimensional analogs corresponding to our model are displayed in Figure 1 as pink and green rectangles, respectively.

    To plot all figures presented in this paper, we have used the parameter values given in Table 2. These values correspond to the wMelPop strain of Wolbachia, which is regarded as the best one for controlling dengue infections among human individuals since it confers the most profound resistance to the replication of dengue virus in mosquitoes [25,28]. However, many scholars point out that the wMelPop strain is associated with high "fitness cost" since it reduces female fecundity, the viability of eggs, and the lifespan of infected mosquitoes [7,8,25]. Therefore, the successful invasion of mosquitoes carrying the wMelPop strain of Wolbachia and their durable persistence even in small detached localities appears to be a challenging task.

    Table 2.  Parameter values for the reduced model (2.1) corresponding to Aedes aegypti mosquitoes and the wMelPop strain of Wolbachia.
    Description Units Assumed value References
    Fecundity rate of uninfected insects time1 ρN=4.55 [16,54]
    Fecundity rate of infected insects time1 ρW=0.5×ρN=2.27 [2,7,16,25]
    Natural mortality rate of uninfected insects time1 αN=0.03333 [16,54]
    Natural mortality rate of infected insects time1 αW=2×αN=0.06666 [2,7,16,25]
    Competition parameter of uninfected insects (mosquito × time)1 βN=2.61258×103 fitted using data from
    Competition parameter of infected insects (mosquito × time)1 βW=3.12792×103 [45,46,55]

     | Show Table
    DownLoad: CSV

    It should be recalled that the ultimate goal of Wolbachia-based biocontrol consists of seeking the eventual elimination of wild insects (capable of transmitting the virus to human individuals) by performing periodic releases of Wolbachia-carrying mosquitoes in some determined localities initially populated by wild mosquitoes. The practical implementation of this method requires mass-rear a massive quantity of Wolbachia-infected insects for posterior releases, and the desired result is propelled by the progressive Wolbachia invasion and its durable establishment in wild mosquito populations. The final outcome of this process is usually referred to as "population replacement".

    Let us now provide some useful insights and practical interpretations derived from Figure 2 while keeping in mind the primary goal of Wolbachia-based biocontrol. First, we recall that the attraction basin BEW contains the initial conditions (N(0),W(0)) starting from which the trajectories of the system (2.1) converge to the desired boundary equilibrium EW=(0,W) and the population replacement will be eventually achieved.

    Suppose now that the population replacement is sought to be achieved with a single (or inundative) initial release of Wolbachia-carrying insects. To determine the size of such an abundant release, the current size of the wild mosquito population should be first assessed by some known technique [44,49]. Once the abundance of wild mosquitoes N0 is fairly estimated, the information regarding the minimal viable sizes ˆW0=ˆW0(N0) such that (N0,ˆW0)SEc of Wolbachia-carrying mosquito populations will be of the utmost importance, and this information is explicitly supplied by the model (2.1) and its underlying parameters.

    Upon closer inspection of Figure 2, we observe that the area inside BEW surrounding W (portion of the green-shaded region around W) is much smaller than the area inside BEN surrounding N (portion of the unshaded region around N). Therefore, when the size of the wild mosquito population is close to saturation or its carrying capacity N, an extreme amount of Wolbachia-carriers will be necessary for a single inundative release.

    On the other hand, it is instructive to recall that wild mosquito populations may exhibit seasonal size variations [47,49]. In this context, the best timing for Wolbachia-based biocontrol by a single inundative release will be the period of relatively low mosquito abundance. Such periods usually correlate with cooler and windier seasons in tropical and subtropical regions [49] or arise after carefully planned and thoroughly implemented vector control measures [41,52].

    For different initial sizes N0=λN of wild mosquito population expressed as fractions λ{0.25,0.5,0.75,1} of N, one can estimate the corresponding minimal release sizes of Wolbachia-carrying insects ˆW0=ˆW0(λN)=ˆλN (also expressed as the multiplicatives of N) that ensure the population replacement by a single inundative release. The corresponding values of λ and ˆλ are presented in Table 3 (considering the parameters of Table 2), and the points (λN,ˆλN)SEc are displayed in Figure 3 in red color.

    Table 3.  Minimal Wolbachia-infected population size ˆW0=ˆW0(λN) for different initial sizes N0=λN,λ{0.25,0.5,0.75,1} of the wild mosquito population, see Figure 3.
    λ (Initial wild population N0=λN) ˆλ (Minimum viable population ˆW0=ˆW0(λN)=ˆλN
    0.25 0.38
    0.5 0.83
    0.75 1.32
    1 1.85

     | Show Table
    DownLoad: CSV
    Figure 3.  Minimal Wolbachia-infected population sizes ˆW0=ˆW0(λN) for different initial sizes N0=λN,λ{0.25,0.5,0.75,1} of the wild mosquito population.

    From Table 3 and Figure 3, we observe that the minimal Wolbachia-infected population sizes needed for single inundative releases always exceed the initial sizes of the wild mosquito population. When the initial size of the wild population is higher than 0.5N, the population replacement becomes substantially more difficult to reach with a single inundative release, as it requires mass-rearing vast numbers of Wolbachia-carriers.

    As an alternative to a single inundative release, one may perform several periodic (or inoculative) releases. This strategy may seem reasonable if the mass-rearing facility cannot produce the vast quantity ˆλN of Wolbachia-carriers en masse but is capable of producing a smaller quantity of Wolbachia-carrying insects every τ days. However, under such a setting, the total amount of Wolbachia-infected mosquitoes necessary to reach the population replacement will be larger than in the case of a single inundative release.

    From the mathematical standpoint, periodic inoculative releases of Wolbachia-carrying insects can be modeled by the following impulsive dynamical system:

    {dNdt=ρNN(NN+W)αNNβNN(N+W),(5.1a)dWdt=ρWWαWWβWW(N+W),(5.1b)
    N(0)=N0,W(0)=Λ,W(iτ+)=W(iτ)+Λ, i=1,2,,n1, (5.1c)

    where τ denotes the period of releases, Λ=const stands the release size (i.e., the number of Wolbachia-carriers to be released), and n defines the number of releases. Notably, W(iτ±) denote the right and left limits of the function W(t) at t=iτ. Formal analysis of the impulsive system (5.1) is a challenging task and may be proposed as an object for further studies. Therefore, in the context of this paper, we limit ourselves to revising its numerical solutions in order to assess the practical value of periodic releases and to compare their overall performance with an outcome of a single inundative release.

    By performing a series of numerical simulations, we have estimated the minimal release sizes Λ:=ˆλN (also expressed as the multiplicatives of N) for τ=1 day and τ=3 days and taking different initial sizes of wild mosquito population N0, expressed as fractions λ{0.25,0.5,0.75,1} of N.

    In Table 4, we present minimum release sizes Λ that ensure the population replacement by n periodic releases even when the initial condition (N0,Λ) lies inside the attraction basin BEN of the boundary equilibrium EN=(N,0) (that is, strictly below the separatrix SEc). Revising the entries of Table 4, it is easy to detect several patterns or "tradeoffs" between the frequency of releases τ, constant release size Λ, and the overall number of releases n needed to ensure the population replacement. Namely, more frequent releases (τ=1) require smaller release sizes Λ and shorter overall time of the release program but a greater number of releases n. The latter is quite logical, and not only aligns with common sense, but also bears similarities with other works dealing with periodic releases of mosquitoes [42]. In this context, the anticipated knowledge of the production costs related to the mass-rearing of Wolbachia-carrying insects and the logistics costs for performing field releases are important for choosing the release frequency. Although we have no reliable information regarding such costs, the impulsive system (5.1) may serve to be of potential utility in the future when healthcare entities eventually decide to evaluate this method of biological vector control. In particular, the model (5.1) may help adjust the frequency and the sizes of releases to the production capacity of artificial mass-rearing facilities of the Wolbachia-carrying insects.

    Table 4.  For different initial sizes of the wild mosquito population (column 1), the release size of the Wolbachia-carrying population (column 2) necessary to ensure the population replacement by inoculative releases each τ days (column 3) and the number of releases (column 4) are indicated.
    λ (N0=λN) ˆλ (Λ=ˆλN, release size) Period of releases (τ days) Number of releases, n
    0.25 0.25 1 5
    0.25 0.3615 3 3
    0.5 0.39 1 9
    0.5 0.773 3 3
    0.75 0.43 1 11
    0.75 1.178 3 4
    1 0.43 1 12
    1 1.39 3 8

     | Show Table
    DownLoad: CSV

    On the other hand, the outcomes of numerical simulations performed on the original model (2.1) can be also compared with those obtained for the impulsive system (5.1). Contrasting the values of ˆλ from Tables 3 and 4 for the same values of N0=λN, we observe that they bear a more striking difference for τ=1 than for τ=3. Moreover, the mentioned difference is smaller for the smaller values of N0 (such as 0.25N and 0.5N) and becomes more noticeable for the larger values of N0 (such as 0.75N and N). Thus, the release programs based on periodic inoculative releases seem more practicable when the initial size of the wild mosquito population is close to its saturation level N.

    Figure 4 displays simulation results for the impulsive system (5.1) when the wild mosquito population is at saturation N0=N. The left column of Figure 4 corresponds to daily releases (τ=1 day) and the right one corresponds to inoculative releases performed every three days (τ=3). The upper charts of Figure 4 present the system's trajectories N(t) and W(t) drawn by blue and red-colored curves, respectively, and also bear two dashed lines marking the coordinates of the coexistence equilibrium Ec=(Nc,Wc). The lower charts exhibit the underlying parts of orbits O(N,W)={(N(t),W(t))R2+: t0} in the phase space that start in (N0,W0)BEN (red-colored point below the separatrix SEc) and move the system states to the attraction basin BEW of the desired boundary equilibrium EW=(0,W) (red-colored point above the separatrix SEc).

    Figure 4.  Trajectories N(t),W(t) (upper row) and the orbits in the phase space (lower row) of the impulsive system (5.1) engendered by the initial conditions N0=N,W0=Λ=ˆλN corresponding to periodic releases with τ=1 (left column) and τ=3 (right column) for constant release sizes Λ=ˆλN given in the last two rows of Table 4.

    The periodic inoculative releases are suspended when the orbit O(N,W) of (5.1) crosses the separatrix SEc and enters the attraction basin BEW (green-shaded region in the lower charts of Figure 4). The latter is also clearly visible in the upper charts of Figure 4: the system trajectory N(t) decays and crosses the blue-colored dashed line, while the trajectory W(t) remains strictly above the red-colored dashed line.

    Browsing once again the simulation results given in Tables 3 and 4 and contrasting them for each particular value of N0=λN,λ{0.25,0.5,0.75,1}, we may conclude that, from the practical standpoint, an implementation of a single inundative release seems more rational and operative than several periodic inoculative releases. Effectively, under the "worst scenario", i.e., when N0=N (this situation is illustrated in Figure 4) it is necessary to mass-rear at least 5.16N of Wolbachia-carriers during 12 days (with τ=1) or at least 11.12N of Wolbachia-carriers during 18 days (when τ=3), while a single inundative release only requires to mass-rear 1.85N of Wolbachia-infected insects, albeit all at once.

    Thus, the reduced bidimensional model (2.1) has resulted in a quite handy and easily interpretable tool for determining the appropriate size of a single inundative release or periodic releases of Wolbachia-carrying insects since it explicitly yields the dependence between minimal viable population sizes of wild and Wolbachia-infected mosquito populations.

    To conclude this study, let us mention that the proposed bidimensional model (2.1) is adaptable to other Wolbachia strains (e.g., wMEl, wAlb) that are being tested for vector control. Furthermore, this model can be adjusted to other insect species in which Wolbachia induces the virus inhibition (e.g., rice planthoppers and fruit flies [48,51]). Therefore, an anticipated knowledge of the separatrix SEc may constitute the key element of Wolbachia-based biocontrol for human and crop diseases.

    Diego Vicencio was supported by the program CONICYT PFCHA/ Doctorado Becas Chile/2017-21171813 and FONDECYT grant N 1200355 ANID-Chile program. Pedro Gajardo was partially supported by FONDECYT grant N 1200355 ANID-Chile program, and ANID/BASAL FB210005. Olga Vasilieva acknowledges financial support from the National Fund for Science, Technology, and Innovation (Autonomous Heritage Fund Francisco José de Caldas) by way of the Research Program No. 1106-852-69523 (Principal Investigator: Hector J. Martinez), Contract: CT FP 80740-439-2020 (Colombian Ministry of Science, Technology, and Innovation – Minciencias), Grant ID: CI-71241 (Universidad del Valle, Colombia). Olga Vasilieva also appreciates the endorsement obtained from the STIC AmSud Program for regional cooperation (20-STIC-05 NEMBICA project, international coordinator: Pierre-Alexandre Bliman, INRIA – France).

    The authors declare that they have no conflict of interest.

    Let us recall that to find the steady states for our model, we must solve the following algebraic system:

    {0=F(N,W)=ρNN(NN+W)αNNβNN(N+W)(A1a)0=G(N,W)=ρWWαWWβWW(N+W).(A1b)

    Note that equation (A.1b) has a solution in W=0. By replacing that value of W in (A.1a), we can easily obtain the first possible steady state, E0=(0,0). If we assume that N0, we obtain the following equation:

    ρNNαNNβNN2=N(ρNαNβNN)=0.

    Taking this expression, it is evident that the value of N must be N=ρNαNβN, thus obtaining the second steady state EN=(N,0).

    Let us now assume that N=0, but W0. By replacing these values in (A.1b), we obtain the following equation.

    ρWWαWWβWW2=W(ρWαWβWW)=0.

    Taking this expression, it is evident that the value of W must be W=ρWαWβW, thus obtaining the third steady state EW=(0,W).

    For the final steady state, we must assume both N0 and W0. By equation (A.1b), we obtain:

    ρWWcαWWcβWW2c+βWWcN=Wc(ρWαWβW(Wc+Nc))=0.

    From this equation, we obtain that Wc+Nc=ρWαWβW=W. Replacing this value in (A.1b), we get:

    ρNN2cWαNNcβNNcW=Nc(ρNNcWαNβNW)=0.

    Hence,

    Nc=W(αNρN+βNρNW)=W(1ρNρN´+αNρN+βNρNW)=W(1βNρN(ρNαNβN)+βNρNW)=W(1βNρN(NW)).

    Finally, from this last expression, and using that Nc+Wc=W we obtain:

    Wc=W(βNρN(NW)).

    Thus, the last steady state is Ec=(Nc,Wc), and its coordinates match the formula (3.6).



    [1] A. Hoffmann, B. Montgomery, J. Popovici, I. Iturbe-Ormaetxe, P. Johnson, F. Muzzi, et al., Successful establishment of Wolbachia in Aedes populations to suppress dengue transmission, Nature, 476 (2011), 454–457. https://doi.org/10.1038/nature10356 doi: 10.1038/nature10356
    [2] C. McMeniman, R. Lane, B. Cass, A. Fong, M. Sidhu, Y.-F. Wang, et al., Stable introduction of a life-shortening Wolbachia infection into the mosquito Aedes aegypti, Science, 323 (2009), 141–144. https://doi.org/10.1126/science.1165326 doi: 10.1126/science.1165326
    [3] L. Moreira, I. Iturbe-Ormaetxe, J. Jeffery, G. Lu, A. Pyke, L. Hedges, et al., A Wolbachia symbiont in Aedes aegypti limits infection with dengue, chikungunya, and plasmodium, Cell, 139 (2009), 1268–1278. https://doi.org/10.1016/j.cell.2009.11.042 doi: 10.1016/j.cell.2009.11.042
    [4] T. Ruang-Areerate, P. Kittayapong, Wolbachia transinfection in Aedes aegypti: a potential gene driver of dengue vectors, Proc. Natl. Acad. Sci. U.S.A., 103 (2006), 12534–12539. https://doi.org/10.1073/pnas.0508879103 doi: 10.1073/pnas.0508879103
    [5] T. Walker, P. Johnson, L. Moreira, I. Iturbe-Ormaetxe, F. Frentiu, C. McMeniman, et al., The wMel Wolbachia strain blocks dengue and invades caged Aedes aegypti populations, Nature, 476 (2011), 450–453. https://doi.org/10.1038/nature10355 doi: 10.1038/nature10355
    [6] L. Almeida, Y. Privat, M. Strugarek, N. Vauchelet, Optimal releases for population replacement strategies: application to Wolbachia, SIAM J. Math. Anal., 51 (2019), 3170–3194. https://doi.org/10.1137/18M1189841 doi: 10.1137/18M1189841
    [7] C. McMeniman, S. O'Neill, A virulent Wolbachia infection decreases the viability of the dengue vector Aedes aegypti during periods of embryonic quiescence, PLoS Negl. Trop. Dis., 4 (2010), e748. https://doi.org/10.1371/journal.pntd.0000748 doi: 10.1371/journal.pntd.0000748
    [8] S. Ritchie, M. Townsend, C. Paton, A. Callahan, A. Hoffmann, Application of wMelPop Wolbachia strain to crash local populations of Aedes aegypti, PLoS Negl. Trop. Dis., 9 (2015), e0003930. https://doi.org/10.1371/journal.pntd.0003930 doi: 10.1371/journal.pntd.0003930
    [9] L. Almeida, A. Haddon, C. Kermorvant, A. Léculier, Y. Privat, M. Strugarek, et al., Optimal release of mosquitoes to control dengue transmission, ESAIM: Proc. Surv., 67 (2020), 16–29. https://doi.org/10.1051/proc/202067002 doi: 10.1051/proc/202067002
    [10] J. Schraiber, A. Kaczmarczyk, R. Kwok, M. Park, R. Silverstein, F. Rutaganira, et al., Constraints on the use of lifespan-shortening Wolbachia to control dengue fever, J. Theor. Biol., 297 (2012), 26–32. https://doi.org/10.1016/j.jtbi.2011.12.006 doi: 10.1016/j.jtbi.2011.12.006
    [11] M. Turelli, Cytoplasmic incompatibility in populations with overlapping generations, Evolution, 64 (2010), 232–241. https://doi.org/10.1111/j.1558-5646.2009.00822.x doi: 10.1111/j.1558-5646.2009.00822.x
    [12] D. E. Campo-Duarte, D. Cardona-Salgado, O. Vasilieva, Establishing wMelPop Wolbachia infection among wild Aedes aegypti females by optimal control approach, Appl. Math. Inf. Sci., 11 (2017), 1011–1027. https://doi.org/10.18576/amis/110408 doi: 10.18576/amis/110408
    [13] D. E. Campo-Duarte, O. Vasilieva, D. Cardona-Salgado, Optimal control for enhancement of Wolbachia frequency among Aedes aegypti females, Int. J. Pure Appl. Math., 112 (2017), 219–238.
    [14] D. Contreras-Julio, P. Aguirre, J. Mujica, O. Vasilieva, Finding strategies to regulate propagation and containment of dengue via invariant manifold analysis, SIAM J. Appl. Dyn. Syst., 19 (2020), 1392–1437. https://doi.org/10.1137/20M131299X doi: 10.1137/20M131299X
    [15] A. Fenton, K. Johnson, J. Brownlie, G. Hurst, Solving the Wolbachia paradox: modeling the tripartite interaction between host, Wolbachia, and a natural enemy, Am. Nat., 178 (2011), 333–342. https://doi.org/10.1086/661247 doi: 10.1086/661247
    [16] D. E. Campo-Duarte, O. Vasilieva, D. Cardona-Salgado, M. Svinin, Optimal control approach for establishing wMelPop Wolbachia infection among wild Aedes aegypti populations, J. Math. Biol., 76 (2018), 1907–1950. https://doi.org/10.1007/s00285-018-1213-2 doi: 10.1007/s00285-018-1213-2
    [17] J. Farkas, S. Gourley, R. Liu, A.-A. Yakubu, Modelling Wolbachia infection in a sex-structured mosquito population carrying West Nile virus, J. Math. Biol., 75 (2017), 621–647. https://doi.org/10.1007/s00285-017-1096-7 doi: 10.1007/s00285-017-1096-7
    [18] C. Ferreira, Aedes aegypti and Wolbachia interaction: population persistence in an environment changing, Theor. Ecol., 13 (2020), 137–148. https://doi.org/10.1007/s12080-019-00435-9 doi: 10.1007/s12080-019-00435-9
    [19] B. Zheng, M. Tang, J. Yu, Modeling Wolbachia spread in mosquitoes through delay differential equations, SIAM J. Appl. Math., 74 (2014), 743–770. https://doi.org/10.1137/13093354X doi: 10.1137/13093354X
    [20] A. Adekunle, M. Meehan, E. McBryde, Mathematical analysis of a Wolbachia invasive model with imperfect maternal transmission and loss of Wolbachia infection, Infect. Dis. Model., 4 (2019), 265–285. https://doi.org/10.1016/j.idm.2019.10.001 doi: 10.1016/j.idm.2019.10.001
    [21] L. Almeida, M. Duprez, Y. Privat, N. Vauchelet, Mosquito population control strategies for fighting against arboviruses, Math. Biosci. Eng., 16 (2019), 6274–6297. https://doi.org/10.3934/mbe.2019313 doi: 10.3934/mbe.2019313
    [22] P.-A. Bliman, M. S. Aronna, F. Coelho, M. da Silva, Ensuring successful introduction of Wolbachia in natural populations of Aedes aegypti by means of feedback control, J. Math. Biol., 76 (2018), 1269–1300. https://doi.org/10.1007/s00285-017-1174-x doi: 10.1007/s00285-017-1174-x
    [23] L. Xue, C. Manore, P. Thongsripong, J. Hyman, Two-sex mosquito model for the persistence of Wolbachia, J. Biol. Dyn., 11 (2017), 216–237. https://doi.org/10.1080/17513758.2016.1229051 doi: 10.1080/17513758.2016.1229051
    [24] J. Farkas, P. Hinow, Structured and unstructured continuous models for Wolbachia infections, Bull. Math. Biol., 72 (2010), 2067–2088. https://doi.org/10.1007/s11538-010-9528-1 doi: 10.1007/s11538-010-9528-1
    [25] I. Dorigatti, C. McCormack, G. Nedjati-Gilani, N. Ferguson, Using Wolbachia for dengue control: insights from modelling, Trends Parasitol., 34 (2018), 102–113. https://doi.org/10.1016/j.pt.2017.11.002 doi: 10.1016/j.pt.2017.11.002
    [26] H. Dutra, M. Rocha, F. Dias, S. Mansur, E. Caragata, L. Moreira, Wolbachia blocks currently circulating Zika virus isolates in Brazilian Aedes aegypti mosquitoes, Cell Host Microbe, 19 (2016), 771–774. https://doi.org/10.1016/j.chom.2016.04.021 doi: 10.1016/j.chom.2016.04.021
    [27] N. Ferguson, D. Kien, H. Clapham, R. Aguas, V. Trung, T. Chau, et al., Modeling the impact on virus transmission of Wolbachia-mediated blocking of dengue virus infection of Aedes aegypti, Sci. Transl. Med., 7 (2015), 279ra37–279ra37.
    [28] M. Woolfit, I. Iturbe-Ormaetxe, J. Brownlie, T. Walker, M. Riegler, A. Seleznev, et al., Genomic evolution of the pathogenic Wolbachia strain, wMelPop, Genome Biol. Evol., 5 (2013), 2189–2204. https://doi.org/10.1093/gbe/evt169 doi: 10.1093/gbe/evt169
    [29] H. Yeap, P. Mee, T. Walker, A. Weeks, S. O'Neill, P. Johnson, et al., Dynamics of the "popcorn" Wolbachia infection in outbred Aedes aegypti informs prospects for mosquito vector control, Genetics, 187 (2011), 583–595. https://doi.org/10.1534/genetics.110.122390 doi: 10.1534/genetics.110.122390
    [30] S.-B. Hsu, H. Smith, P. Waltman, Competitive exclusion and coexistence for competitive systems on ordered Banach spaces, Trans. Am. Math. Soc., 348 (1996), 4083–4094. https://doi.org/10.1090/S0002-9947-96-01724-2 doi: 10.1090/S0002-9947-96-01724-2
    [31] O. E. Escobar-Lasso, O. Vasilieva, A simplified monotone model of Wolbachia invasion encompassing Aedes aegypti mosquitoes, Stud. Appl. Math., 146 (2021), 565–585.
    [32] H. Smith, Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems, vol. 41 of Mathematical Surveys and Monographs, American Mathematical Society, Providence RI, USA, 1995.
    [33] J. Jiang, X. Liang, X.-Q. Zhao, Saddle-point behavior for monotone semiflows and reaction–diffusion models, J. Differ. Equ., 203 (2004), 313–330. https://doi.org/10.1016/j.jde.2004.05.002 doi: 10.1016/j.jde.2004.05.002
    [34] U. Boscain, B. Piccoli, Optimal syntheses for control systems on 2-D manifolds, vol. 43 of Mathématiques & Applications, Springer-Verlag, Berlin, Germany, 2004.
    [35] H. Sussmann, Regular synthesis for time-optimal control of single-input real analytic systems in the plane, SIAM J. Control Optim., 25 (1987), 1145–1162. https://doi.org/10.1137/0325062 doi: 10.1137/0325062
    [36] H. Sussmann, The structure of time-optimal trajectories for single-input systems in the plane: the c nonsingular case, SIAM J. Control Optim., 25 (1987), 433–465. https://doi.org/10.1137/0325025 doi: 10.1137/0325025
    [37] H. Sussmann, The structure of time-optimal trajectories for single-input systems in the plane: the general real analytic case, SIAM J. Control Optim., 25 (1987), 868–904. https://doi.org/10.1137/0325048 doi: 10.1137/0325048
    [38] P.-A. Bliman, A feedback control perspective on biological control of dengue vectors by Wolbachia infection, Eur. J. Contr., 59 (2021), 188–206. https://doi.org/10.1016/j.ejcon.2020.09.006 doi: 10.1016/j.ejcon.2020.09.006
    [39] H. Aida, H. Dieng, T. Satho, A. Nurita, M. Salmah, F. Miake, et al., The biology and demographic parameters of Aedes albopictus in northern peninsular Malaysia, Asian Pac. J. Trop. Biomed., 1 (2011), 472–477. https://doi.org/10.1016/S2221-1691(11)60103-2 doi: 10.1016/S2221-1691(11)60103-2
    [40] D. Angeli, E. Sontag, Monotone control systems, IEEE Trans. Automat. Contr., 48 (2003), 1684–1698. https://doi.org/10.1109/TAC.2003.817920 doi: 10.1109/TAC.2003.817920
    [41] J. H. Arias-Castro, H. J. Martinez-Romero, O. Vasilieva, Biological and chemical control of mosquito population by optimal control approach, Games, 11 (2020), 62. https://doi.org/10.3390/g11040062 doi: 10.3390/g11040062
    [42] P.-A. Bliman, D. Cardona-Salgado, Y. Dumont, O. Vasilieva, Implementation of control strategies for sterile insect techniques, Math. Biosci., 314 (2019), 43–60. https://doi.org/10.1016/j.mbs.2019.06.002 doi: 10.1016/j.mbs.2019.06.002
    [43] M. H. Chan, P. S. Kim, Modelling a Wolbachia invasion using a slow–fast dispersal reaction–diffusion approach, Bull. Math. Biol., 75 (2013), 1501–1523. https://doi.org/10.1007/s11538-013-9857-y doi: 10.1007/s11538-013-9857-y
    [44] D. Cianci, J. Van Den Broek, B. Caputo, F. Marini, D. Torre, H. Heesterbeek, et al., Estimating mosquito population size from mark–release–recapture data, J. Med. Entomol., 50 (2013), 533–542. https://doi.org/10.1603/ME12126 doi: 10.1603/ME12126
    [45] P. Crain, J. Mains, E. Suh, Y. Huang, P. Crowley, S. Dobson, Wolbachia infections that reduce immature insect survival: Predicted impacts on population replacement, BMC Evol. Biol., 11 (2011), 290. https://doi.org/10.1186/1471-2148-11-290 doi: 10.1186/1471-2148-11-290
    [46] S. De Oliveira, D. Villela, F. Dias, L. Moreira, R. de Freitas, How does competition among wild type mosquitoes influence the performance of Aedes aegypti and dissemination of Wolbachia pipientis?, PLoS Negl. Trop. Dis., 11 (2017), e0005947.
    [47] H. Delatte, G. Gimonneau, A. Triboire, D. Fontenille, Influence of temperature on immature development, survival, longevity, fecundity, and gonotrophic cycles of Aedes albopictus, vector of chikungunya and dengue in the Indian Ocean, J. Med. Entomol., 46 (2009), 33–41. https://doi.org/10.1603/033.046.0105 doi: 10.1603/033.046.0105
    [48] J.-T. Gong, Y. Li, T.-P. Li, Y. Liang, L. Hu, D. Zhang, et al., Stable introduction of plant-virus-inhibiting Wolbachia into plant hoppers for rice protection, Curr. Biol., 30 (2020), 4837–4845. https://doi.org/10.1016/j.cub.2020.09.033 doi: 10.1016/j.cub.2020.09.033
    [49] L. Gouagna, J.-S. Dehecq, D. Fontenille, Y. Dumont, S. Boyer, Seasonal variation in size estimates of Aedes albopictus population based on standard mark-release-recapture experiments in an urban area on Reunion Island, Acta Trop., 143 (2015), 89–96. https://doi.org/10.1016/j.actatropica.2014.12.011 doi: 10.1016/j.actatropica.2014.12.011
    [50] L. Perko, Differential Equations and Dynamical Systems, Texts in Applied Mathematics, Springer, New York, USA, 2013.
    [51] A. C. Pimentel, C. S. Cesar, M. Martins, R. Cogni, The antiviral effects of the symbiont bacteria Wolbachia in insects, Front. Immunol., 11 (2021), 3690. https://doi.org/10.3389/fimmu.2020.626329 doi: 10.3389/fimmu.2020.626329
    [52] E. Pliego-Pliego, O. Vasilieva, J. Velázquez-Castro, A. Fraguela-Collar, Control strategies for a population dynamics model of Aedes aegypti with seasonal variability and their effects on dengue incidence, Appl. Math. Model., 81 (2020), 296–319.
    [53] H. Smith, Monotone dynamical systems: Reflections on new advances & applications, Discrete Contin. Dyn. Syst. Ser. A, 37 (2017), 485–504.
    [54] L. Styer, S. Minnick, A. Sun, T. Scott, Mortality and reproductive dynamics of Aedes aegypti (Diptera: Culicidae) fed human blood, Vector Borne Zoonotic Dis., 7 (2007), 86–98. https://doi.org/10.1089/vbz.2007.0216 doi: 10.1089/vbz.2007.0216
    [55] E. Suh, S. Dobson, Reduced competitiveness of Wolbachia infected Aedes aegypti larvae in intra-and inter-specific immature interactions, J. Invertebr. Pathol., 114 (2013), 173–177.
  • This article has been cited by:

    1. Jose L. Orozco-Gonzales, Antone dos Santos Benedito, Daiver Cardona-Salgado, Claudia Pio Ferreira, Helenice de Oliveira Florentino, Lilian S. Sepulveda-Salcedo, Olga Vasilieva, Comparing the long-term persistence of different Wolbachia strains after the release of bacteria-carrying mosquitoes, 2024, 372, 00255564, 109190, 10.1016/j.mbs.2024.109190
    2. Jose Luis Orozco Gonzales, Antone dos Santos Benedito, Helenice de Oliveira Florentino, Claudia Pio Ferreira, Daiver Cardona-Salgado, Lilian S. Sepulveda-Salcedo, Olga Vasilieva, Optimization approaches to Wolbachia-based biocontrol, 2025, 137, 0307904X, 115663, 10.1016/j.apm.2024.115663
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2376) PDF downloads(104) Cited by(2)

Figures and Tables

Figures(4)  /  Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog