Skip to main content

Advertisement

Log in

Ensuring successful introduction of Wolbachia in natural populations of Aedes aegypti by means of feedback control

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The control of the spread of dengue fever by introduction of the intracellular parasitic bacterium Wolbachia in populations of the vector Aedes aegypti, is presently one of the most promising tools for eliminating dengue, in the absence of an efficient vaccine. The success of this operation requires locally careful planning to determine the adequate number of individuals carrying the Wolbachia parasite that need to be introduced into the natural population. The introduced mosquitoes are expected to eventually replace the Wolbachia-free population and guarantee permanent protection against the transmission of dengue to human. In this study, we propose and analyze a model describing the fundamental aspects of the competition between mosquitoes carrying Wolbachia and mosquitoes free of the parasite. We then use feedback control techniques to devise an introduction protocol that is proved to guarantee that the population converges to a stable equilibrium where the totality of mosquitoes carry Wolbachia.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. See e.g. the page http://www.eliminatedengue.com/co/progress/article/690/ in the site of the project Eliminate Dengue (Hoffmann et al. 2012).

References

  • Alphey L (2014) Genetic control of mosquitoes. Annu Rev Entomol 59(1):205–224

    Article  Google Scholar 

  • Alphey L, Benedict M, Bellini R, Clark GG, Dame DA, Service MW, Dobson SL (2010) Sterile-insect methods for control of mosquito-borne diseases: an analysis. Vector-Borne Zoonotic Dis 10(3):295–311

  • Angeli D, Sontag ED (2003) Monotone control systems. IEEE Trans Autom Control 48:1684–1698

    Article  MathSciNet  MATH  Google Scholar 

  • Angeli D, De Leenheer P, Sontag ED (2004) A small-gain theorem for almost global convergence of monotone systems. Syst Control Lett 52(5):407–414

    Article  MathSciNet  MATH  Google Scholar 

  • Barton N, Turelli M (2011) Spatial waves of advance with bistable dynamics: cytoplasmic and genetic analogues of Allee effects. Am Nat 178(3):E48–E75

    Article  Google Scholar 

  • Bian G, Xu Y, Lu P, Xie Y, Xi Z (2010) The endosymbiotic bacterium Wolbachia induces resistance to dengue virus in Aedes aegypti. PLoS Pathog 6(4):e1000833

    Article  Google Scholar 

  • Bliman PA, Aronna MS, Coelho FC, da Silva MAHB (2015) Ensuring successful introduction of Wolbachia in natural populations of Aedes aegypti by means of feedback control. arxiv http://arxiv.org/abs/1503.05216

  • Brogdon WG, McAllister JC (1998) Insecticide resistance and vector control. Emerg Infect Dis 4(4):605–613

    Article  Google Scholar 

  • Brownstein JS, Hett E, O’Neill SL (2003) The potential of virulent Wolbachia to modulate disease transmission by insects. J Invertebr Pathol 84(1):24–29

    Article  Google Scholar 

  • Christophers S (1960) Aedes aegypti (L.) the yellow fever mosquito: its life history, bionomics and structure. Cambridge University Press, Cambridge

  • Coppel WA (1965) Stability and asymptotic behavior of differential equations. Heath, Heath mathematical monographs

  • de Freitas RM, Valle D (2014) Challenges encountered using standard vector control measures for dengue in Boa Vista. Brazil. Bull World Health Organ 92(9):685–689

    Article  Google Scholar 

  • de Freitas RM, Avendanho FC, Santos R, Sylvestre G, Araujo SC, Lima JBP, Martins AJ, Coelho GE, Valle D (2014) Undesirable consequences of insecticide resistance following Aedes aegypti control activities due to a dengue outbreak. PLoS ONE 9:1–9

    Google Scholar 

  • Dutra HLC, Santos LMB, Caragata EP, Silva JBL, Villela DAM, Maciel-de Freitas R, Moreira LA (2015) From lab to field: the influence of urban landscapes on the invasive potential of Wolbachia in Brazilian Aedes aegypti mosquitoes. PLOS Negl Trop Dis 9(4):e0003689

    Article  Google Scholar 

  • Enciso G (2014) Fixed points and convergence in monotone systems under positive or negative feedback. Int J Control 87(2):301–311

    Article  MathSciNet  MATH  Google Scholar 

  • Enciso G, Sontag ED (2006) Nonmonotone systems decomposable into monotone systems with negative feedback. J Differ Equ 224:205–227

    Article  MathSciNet  MATH  Google Scholar 

  • Farnesi LC, Martins AJ, Valle D, Rezende GL (2009) Embryonic development of Aedes aegypti (Diptera: Culicidae): influence of different constant temperatures. Mem Inst Oswaldo Cruz 104(1):124–6

    Article  Google Scholar 

  • Ferreira CP, Godoy WA (eds) (2014) Ecological modelling applied to entomology. Springer, Berlin

    Google Scholar 

  • Focks DA (2003) A review of entomological sampling methods and indicators for dengue vectors. WHO, Geneva

    Google Scholar 

  • Focks DA, Haile DG, Daniels E, Mount GA (1993) Dynamic life table model of Aedes aegypti (Diptera: Culicidae)—analysis of the literature and model development. J Med Entomol 30:1003–1017

    Article  Google Scholar 

  • Frentiu FD, Zakir T, Walker T, Popovici J, Pyke AT, van den Hurk A, McGraw EA, O’Neill SL (2014) Limited dengue virus replication in field-collected Aedes aegypti mosquitoes infected with Wolbachia. PLoS Negl Trop Dis 8(2):e2688

    Article  Google Scholar 

  • Gedeon T, Hines G (2009) Multi-valued characteristics and morse decompositions. J Differ Equ 247(4):1013–1042

    Article  MathSciNet  MATH  Google Scholar 

  • Gouzé JL (1988) A criterion of global convergence to equilibrium for differential systems with an application to Lotka–Volterra systems. Research report 0894, Inria, France

  • Hancock PA, Godfray HCJ (2012) Modelling the spread of Wolbachia in spatially heterogeneous environments. J R Soc Interface 9(76):3045–3054

    Article  Google Scholar 

  • Hancock PA, Sinkins SP, Godfray HCJ (2011a) Population dynamic models of the spread of Wolbachia. Am Nat 177(3):323–333

    Article  Google Scholar 

  • Hancock PA, Sinkins SP, Godfray HCJ (2011b) Strategies for introducing Wolbachia to reduce transmission of mosquito-borne diseases. PLoS Negl Trop Dis 5(4):e1024

    Article  Google Scholar 

  • Hirsch MW (1988) Stability and convergence in strongly monotone dynamical systems. J Reine Angew Math 383:1–53

    MathSciNet  MATH  Google Scholar 

  • Hoffmann AA (2014) Facilitating Wolbachia invasions. Aust Entomol 53(2):125–132

    Article  Google Scholar 

  • Hoffmann AA, Montgomery BL, Popovici J, Iturbe-Ormaetxe I, Johnson PH, Muzzi F, Greenfield M, Durkan M, Leong YS, Dong Y, Cook H, Axford J, Callahan AG, Kenny N, Omodei C, McGraw EA, Ryan PA, Ritchie SA, Turelli M, O’Neill SL (2011) Successful establishment of Wolbachia in Aedes populations to suppress dengue transmission. Nature 476(7361):454–457

    Article  Google Scholar 

  • Hoffmann A, Moreira L, O’Neill S, Osorio JE, Ritchie S, Simmons C, Turelli M (2012), Eliminate dengue. http://www.eliminatedengue.com/

  • Hu L, Huang M, Tang M, Yu J, Zheng B (2015) Wolbachia spread dynamics in stochastic environments. Theor Popul Biol 106:32–44

    Article  MATH  Google Scholar 

  • Huang M, Tang M, Yu J (2015) Wolbachia infection dynamics by reaction–diffusion equations. Sci China Math 58(1):77–96

    Article  MathSciNet  MATH  Google Scholar 

  • Hughes H, Britton NF (2013) Modelling the use of Wolbachia to control dengue fever transmission. Bull Math Biol 75(5):796–818

    Article  MathSciNet  MATH  Google Scholar 

  • Jeffery JAL, Yen NT, Nam VS, Nghia LT, Hoffmann AA, Kay BH, Ryan PA (2009) Characterizing the Aedes aegypti population in a Vietnamese village in preparation for a Wolbachia-based mosquito control strategy to eliminate dengue. PLOS Negl Trop Dis 3(11):e552

    Article  Google Scholar 

  • Keeling MJ, Jiggins FM, Read JM (2003) The invasion and coexistence of competing Wolbachia strains. Heredity 91(4):382–388

    Article  Google Scholar 

  • Koiller J, Da Silva M, Souza M, Codeço C, Iggidr A, Sallet G (2014) Aedes, Wolbachia and Dengue. Research report RR-8462, Inria, France

  • La Salle JP (1976) The stability of dynamical systems. SIAM, Philadelphia

    Book  Google Scholar 

  • Malisoff M, De Leenheer P (2006) A small-gain theorem for monotone systems with multi-valued input-state characteristics. IEEE Trans Autom Control 41(2):287–292

    Article  MATH  Google Scholar 

  • McMeniman CJ, Lane RV, Cass BN, Fong AW, Sidhu M, Wang YF, O’Neill SL (2009) Stable introduction of a life-shortening Wolbachia infection into the mosquito Aedes aegypti. Science 323(5910):141–4

    Article  Google Scholar 

  • Moreira LA, Iturbe-Ormaetxe I, Jeffery JA, Lu G, Pyke AT, Hedges LM, Rocha BC, Hall-Mendelin S, Day A, Riegler M, Hugo LE, Johnson KN, Kay BH, McGraw EA, van den Hurk AF, Ryan PA, O’Neill SL (2009a) A Wolbachia symbiont in Aedes aegypti limits infection with dengue, chikungunya, and plasmodium. Cell 139(7):1268–1278

    Article  Google Scholar 

  • Moreira LA, Saig E, Turley AP, Ribeiro JMC, O’Neill SL, McGraw AE (2009) Human probing behavior of Aedes aegypti when infected with a life-shortening strain of Wolbachia. PLOS Negl Trop Dis 12(3):1–6

    Google Scholar 

  • Murray JV, Jansen CC, De Barro P (2016) Risk associated with the release of Wolbachia-infected Aedes aegypti mosquitoes into the environment in an effort to control dengue. Front Public Health 4:43

    Article  Google Scholar 

  • Ndii MZ, Hickson R, Allingham D, Mercer G (2015) Modelling the transmission dynamics of dengue in the presence of Wolbachia. Math Biosci 262:157–166

    Article  MathSciNet  MATH  Google Scholar 

  • O’Neill S (2015) The dengue stopper. Sci Am 312(6):72–7

    Article  Google Scholar 

  • O’Neill SL, Hoffman AA, Werren JH (eds) (1998) Influential passengers: inherited microorganisms and arthropod reproduction. Oxford University Press, Oxford

    Google Scholar 

  • Ocampo CB, Salazar-Terreros MJ, Mina NJ, McAllister J, Brogdon W (2011) Insecticide resistance status of Aedes aegypti in 10 localities in Colombia. Acta Trop 118(1):37–44

    Article  Google Scholar 

  • Otero M, Solari HG, Schweigmann N (2006) A stochastic population dynamics model for Aedes aegypti: formulation and application to a city with temperate climate. Bull Math Biol 68:1945–1974

    Article  MathSciNet  MATH  Google Scholar 

  • Otero M, Schweigmann N, Solari HG (2008) A stochastic spatial dynamical model for Aedes aegypti. Bull Math Biol 70:1297–1325

    Article  MathSciNet  MATH  Google Scholar 

  • Rasgon JL, Scott TW (2004) An initial survey for Wolbachia (Rickettsiales: Rickettsiaceae) infections in selected California mosquitoes (Diptera: Culicidae). J Med Entomol 41(2):255–257

    Article  Google Scholar 

  • Ruang-Areerate T, Kittayapong P (2006) Wolbachia transinfection in Aedes aegypti: a potential gene driver of dengue vectors. Proc Natl Acad Sci 103(33):12534–9

    Article  Google Scholar 

  • Silver JB (2007) Mosquito ecology: field sampling methods. Springer, Berlin

    Google Scholar 

  • Smith HL (1995) Monotone dynamical systems: an introduction to the theory of competitive and cooperative systems. Mathematical surveys and monographs, vol 41. American Mathematical Society

  • Smith DL, Perkins TA, Tusting LS, Scott TW, Lindsay SW (2013) Mosquito population regulation and larval source management in heterogeneous environments. PLoS One 8(8):e71247

    Article  Google Scholar 

  • Southwood TR, Murdie G, Yasuno M, Tonn RJ, Reader PM (1972) Studies on the life budget of Aedes aegypti in Wat Samphaya, Bangkok, Thailand. Bull World Health Organ 46(2):211

    Google Scholar 

  • Turelli M (2010) Cytoplasmic incompatibility in populations with overlapping generations. Evolution 64(1):232–241

    Article  Google Scholar 

  • Walker T, Johnson P, Moreira L, Iturbe-Ormaetxe I, Frentiu F, McMeniman C, Leong YS, Dong Y, Axford J, Kriesner P et al (2011) The wMel Wolbachia strain blocks dengue and invades caged Aedes aegypti populations. Nature 476(7361):450–453

    Article  Google Scholar 

  • Xi Z, Khoo CC, Dobson SL (2005) Wolbachia establishment and invasion in an Aedes aegypti laboratory population. Science 310(5746):326–8

    Article  Google Scholar 

  • Yang HM, Macoris ML, Galvani KC, Andrighetti MT, Wanderley DM (2009) Assessing the effects of temperature on the population of Aedes aegypti, the vector of dengue. Epidemiol Infect 137(08):1188–1202

    Article  Google Scholar 

  • Yeap HL, Mee P, Walker T, Weeks AR, O’Neill SL, Johnson P, Ritchie SA, Richardson KM, Doig C, Endersby NM, Hoffmann AA (2011) Dynamics of the popcorn Wolbachia infection in outbred Aedes aegypti informs prospects for mosquito vector control. Genetic 187(2):583–95

    Article  Google Scholar 

  • Zheng B, Tang M, Yu J (2014) Modeling Wolbachia spread in mosquitoes through delay differential equations. SIAM J Appl Math 74(3):743–770

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We wish to thank the anonymous referees for their valuable advices that helped improve this manuscript. The first author is indebted to T. Gedeon for valuable discussions. The second author thanks H. Solari for useful bibliographical references. This work was done while the second author was a postdoctoral fellow at IMPA, Rio de Janerio, funded by CAPES-Brazil. The authors also acknowledge CAPES-Brazil for the STIC AmSud funding for the MOSTICAW Project, Process No. 99999.007551/2015-00, and FGV for the Pesquisa Aplicada funding of the project “Controle da Dengue através do uso da bactéria Wolbachia”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre-Alexandre Bliman.

Appendices

Appendix A: A sexual version of the infestation model

We here provide a sexual version of model (1). For simplicity, no control input is written. We denote respectively \({\mathbf {m}}_U, {\mathbf {m}}_W, {\mathbf {M}}_U, {\mathbf {M}}_W\), the numbers of uninfected, resp. Wolbachia-infected, males in early and adult phases; and similarly \({\mathbf {f}}_U, {\mathbf {f}}_W, {\mathbf {F}}_U, {\mathbf {F}}_W\) for the females. The model is:

$$\begin{aligned} \dot{{\mathbf {m}}}_U= & {} \lambda _U\alpha _U \frac{{\mathbf {M}}_U}{{\mathbf {M}}_U+{\mathbf {M}}_W} {\mathbf {F}}_U -\nu {\mathbf {m}}_U \nonumber \\&-\,\mu (1+k ({\mathbf {m}}_W+{\mathbf {m}}_U+{\mathbf {f}}_W+{\mathbf {f}}_U)){\mathbf {m}}_U \end{aligned}$$
(39a)
$$\begin{aligned} \dot{{\mathbf {M}}}_U= & {} \nu {\mathbf {m}}_U -\mu _U {\mathbf {M}}_U\end{aligned}$$
(39b)
$$\begin{aligned} \dot{{\mathbf {m}}}_W= & {} \lambda _W\alpha _W {\mathbf {F}}_W -\nu {\mathbf {m}}_W - \mu (1+k ({\mathbf {m}}_W+{\mathbf {m}}_U+{\mathbf {f}}_W+{\mathbf {f}}_U)){\mathbf {m}}_W\end{aligned}$$
(39c)
$$\begin{aligned} \dot{{\mathbf {M}}}_W= & {} \nu {\mathbf {m}}_W -\mu _W {\mathbf {M}}_W\end{aligned}$$
(39d)
$$\begin{aligned} \dot{{\mathbf {f}}}_U= & {} \alpha _U \frac{{\mathbf {M}}_U}{{\mathbf {M}}_U+{\mathbf {M}}_W}{\mathbf {F}}_U -\nu {\mathbf {f}}_U - \mu (1+k ({\mathbf {m}}_W+{\mathbf {m}}_U+{\mathbf {f}}_W+{\mathbf {f}}_U)){\mathbf {f}}_U \end{aligned}$$
(39e)
$$\begin{aligned} \dot{{\mathbf {F}}}_U= & {} \nu {\mathbf {f}}_U -\mu _U {\mathbf {F}}_U\end{aligned}$$
(39f)
$$\begin{aligned} \dot{{\mathbf {f}}}_W= & {} \alpha _W {\mathbf {F}}_W -\nu {\mathbf {f}}_W - \mu (1+k ({\mathbf {m}}_W+{\mathbf {m}}_U+{\mathbf {f}}_W+{\mathbf {f}}_U)){\mathbf {f}}_W\end{aligned}$$
(39g)
$$\begin{aligned} \dot{{\mathbf {F}}}_W= & {} \nu {\mathbf {f}}_W -\mu _W {\mathbf {F}}_W \end{aligned}$$
(39h)

Here \(\lambda _U, \lambda _W\) are the sex ratio (ratio of males to females) of the offspring for the uninfected and infected populations. The other parameters have the same meaning than for model (1) (see Table 2). Here they have been chosen identical for both sex, and in such conditions, it is straightforward to see that the variables defined by

$$\begin{aligned} {\mathbf {L}}_U:= {\mathbf {m}}_U+ {\mathbf {f}}_U,\ {\mathbf {L}}_W:= {\mathbf {m}}_W+ {\mathbf {f}}_W,\ {\mathbf {A}}_U:= {\mathbf {M}}_U+ {\mathbf {F}}_U,\ {\mathbf {A}}_W:= {\mathbf {M}}_W+ {\mathbf {F}}_W \end{aligned}$$
(40)

obey the following equations:

$$\begin{aligned} \dot{{\mathbf {L}}}_U= & {} (1+\lambda _U)\alpha _U \frac{{\mathbf {M}}_U}{{\mathbf {M}}_U+{\mathbf {M}}_W}{\mathbf {F}}_U -\nu {\mathbf {L}}_U - \mu (1+k ({\mathbf {L}}_W+{\mathbf {L}}_U)){\mathbf {L}}_U \end{aligned}$$
(41a)
$$\begin{aligned} \dot{{\mathbf {A}}}_U= & {} \nu {\mathbf {L}}_U -\mu _U {\mathbf {A}}_U \end{aligned}$$
(41b)
$$\begin{aligned} \dot{{\mathbf {L}}}_W= & {} (1+\lambda _W) \alpha _W {\mathbf {F}}_W -\nu {\mathbf {L}}_W - \mu (1+k ({\mathbf {L}}_W+{\mathbf {L}}_U)){\mathbf {L}}_W \end{aligned}$$
(41c)
$$\begin{aligned} \dot{{\mathbf {A}}}_W= & {} \nu {\mathbf {L}}_W -\mu _W {\mathbf {A}}_W \end{aligned}$$
(41d)

If moreover \(\lambda _U=\lambda _W\) and the sex ratio are initially equal to this common value, that is:

$$\begin{aligned} \frac{{\mathbf {m}}_U(0)}{{\mathbf {f}}_U(0)} = \frac{{\mathbf {M}}_U(0)}{{\mathbf {F}}_U(0)} = \lambda _U,\quad \frac{{\mathbf {m}}_W(0)}{{\mathbf {f}}_W(0)} = \frac{{\mathbf {M}}_W(0)}{{\mathbf {F}}_W(0)} = \lambda _W \end{aligned}$$
(42)

then the same proportions are conserved along the evolution, and it is possible to replace \(\frac{{\mathbf {M}}_U}{{\mathbf {M}}_U+{\mathbf {M}}_W}\) by \(\frac{{\mathbf {L}}_U}{{\mathbf {L}}_U+{\mathbf {L}}_W}, (1+\lambda _U) {\mathbf {F}}_U\) by \({\mathbf {L}}_U\) and \((1+\lambda _W) {\mathbf {F}}_W\) by \({\mathbf {L}}_W\) in (41a), (41c), showing that (41) boils down to the simpler model (1).

Appendix B: Proof of Theorem 7

1.1 Computation and ordering of the equilibrium points

One here computes the equilibrium points. The latter verify

$$\begin{aligned}&\gamma _U\mathcal{R}_U \frac{ A_U}{ A_U+ A_W} A_U - (1+ L_W+ L_U) L_U = 0 \end{aligned}$$
(43a)
$$\begin{aligned}&\gamma _W \mathcal{R}_W A_W - (1+ L_W+ L_U) L_W =0 \end{aligned}$$
(43b)
$$\begin{aligned}&L_U = \gamma _U A_U,\quad L_W =\gamma _W A_W \end{aligned}$$
(43c)

The point \(x_{0,0} := (0,0,0,0)\) is clearly an equilibrium. Let us look for an equilibrium \(x_{U,0} := (L_U^*,A_U^*,0,0)\). The quantities \(L_U^*,A_U^*\) then have to satisfy

$$\begin{aligned} \gamma _U\mathcal{R}_U A_U^* - \left( 1 + L_U^*\right) L_U^* =0,\quad L_U^* = \gamma _U A_U^*. \end{aligned}$$
(44)

Dividing by \(L_U^*\ne 0\) yields \(1+ L_U^* = \mathcal{R}_U\). One thus gets the unique solution of this form verifying

$$\begin{aligned} L_U^* = \mathcal{R}_U-1,\quad A_U^* = \frac{\mathcal{R}_U-1}{\gamma _U}, \end{aligned}$$

which is positive due to the sustainability hypothesis (6).

Similarly, one now looks for an equilibrium defined as \(x_{0,W} := (0,0,L_W^*,A_W^*)\). The values of \(L_W^*,A_W^*\) must verify

$$\begin{aligned} \gamma _W \mathcal{R}_W A_W^* - \left( 1+ L_W^*\right) L_W^* =0,\quad L_W^* = \gamma _W A_W^*. \end{aligned}$$

This is identical to (44), and as for the \(x_{U,0}\) case, one gets a unique, positive, solution, namely

$$\begin{aligned} L_W^* = \mathcal{R}_W-1,\quad A_W^* = \frac{\mathcal{R}_W-1}{\gamma _W}. \end{aligned}$$
(45)

We show now that system (7) also admits a unique coexistence equilibrium with positive components \(x_{U,W}= (L_U^{**}, A_U^{**}, L_W^{**}, A_W^{**})\). Coming back to (45) and expressing the value of the factor common to the first and second identity leads to

$$\begin{aligned} 1+L_U^{**}+L_W^{**}= & {} \gamma _W\mathcal{R}_W\frac{A_W^{**}}{L_W^{**}} = \mathcal{R}_W\\= & {} \gamma _U\mathcal{R}_U\frac{A_U^{**}}{A_U^{**}+A_W^{**}}\frac{A_U^{**}}{L_U^{**}}\\= & {} \mathcal{R}_U\frac{A_U^{**}}{A_U^{**}+A_W^{**}} \end{aligned}$$

One thus deduces

$$\begin{aligned} \frac{A_U^{**}}{A_U^{**}+A_W^{**}} = \frac{\mathcal{R}_W}{\mathcal{R}_U}, \end{aligned}$$

and one can express all three remaining unknowns in function of \(A_W^{**}\):

$$\begin{aligned} L_W^{**} = \gamma _W A_W^{**},\quad A_U^{**} = \frac{\mathcal{R}_W}{\mathcal{R}_U-\mathcal{R}_W} A_W^{**},\quad L_U^{**} = \gamma _U A_U^{**} = \gamma _U\frac{\mathcal{R}_W}{\mathcal{R}_U-\mathcal{R}_W} A_W^{**}. \end{aligned}$$

Using the value of \(L_U^{**}\) and \(L_W^{**}\) now yields the relation

$$\begin{aligned} \mathcal{R}_W -1 = L_U^{**}+L_W^{**} = \gamma _W\left( 1+ \frac{\gamma _U}{\gamma _W}\frac{\mathcal{R}_W}{\mathcal{R}_U-\mathcal{R}_W} \right) A_W^{**}, \end{aligned}$$

which has a unique, positive, solution when (6) holds. Hence, the fourth equilibrium is given by

$$\begin{aligned} L_U^{**} = \frac{\delta }{1+\delta }(\mathcal{R}_W-1),\quad A_U^{**} = \frac{\delta }{(1+\delta )\gamma _U}(\mathcal{R}_W-1)\\ L_W^{**} = \frac{1}{1+\delta }(\mathcal{R}_W-1),\quad A_W^{**} = \frac{1}{(1+\delta )\gamma _W}(\mathcal{R}_W-1) \end{aligned}$$

where \(\delta \) was given in Eq. (14d) in the statement of the theorem.

So far we have found all equilibrium points. Actually, it is easy to see that no equilibria is missing: for \(L_U= 0\) we necessarily have \(A_U=0,\) and this gives us \(x_{0,0}\) and \(x_{0,W},\) while for \(L_U \ne 0 \) we get \(A_U\ne 0,\) and this leads us to \(x_{U,0}\) and to \(x_{U,W}.\)

Notice that the last equilibrium can be expressed alternatively by use of the values of the equilibrium \(x_{0,W}\):

$$\begin{aligned} L_U^{**}= & {} \frac{\delta }{1+\delta }L_W^*,\quad A_U^{**} = \frac{\delta }{1+\delta }\frac{\gamma _W}{\gamma _U}A_W^* \end{aligned}$$
(46a)
$$\begin{aligned} L_W^{**}= & {} \frac{1}{1+\delta }L_W^*,\quad A_W^{**} = \frac{1}{1+\delta }A_W^* \end{aligned}$$
(46b)

and this provides straightforward comparison result:

$$\begin{aligned} L_U^{**}< L_W^*<L_U^* \quad \text { and } \quad L_W^{**}< L_W^* <L_U^* \end{aligned}$$
(47a)

and thus \(A_\eta ^{**} = \gamma _\eta L_\eta ^{**} < \gamma _\eta L_\eta ^* = A_\eta ^*\), for \(\eta \in \{U,W\}\) and, therefore,

$$\begin{aligned} A_U^{**}<A_U^*\quad \text { and }\quad A_W^{**} < A_W^*, \end{aligned}$$
(47b)

the second inequality being directly deduced from (46b). The relations (50) allow us to establish the inequalities (15).

1.2 Local stability analysis

The local stability analysis is conducted through analysis of the eigenvalues of the Jacobian matrices. Recall that the Jacobian has been computed in (13).

Stability of \(x_{0,0}\). The value of \({ Df}\) is not defined at \(x_{0,0}\). To show the instability of the equilibrium \(x_{0,0}\), let the function V be defined on \(\mathbb {R}_+^4\) by

$$\begin{aligned} V(x):= \rho (L_U+(1+\varepsilon ) A_U) + L_W+(1+\varepsilon )A_W, \end{aligned}$$

for values of \(\varepsilon , \rho \) still to be defined. Notice that V is positive definite for any positive \(\varepsilon \) and \(\rho .\) We will show that there exists \(\rho >0\) for which the derivative \({\dot{V}}\) of V along the trajectories is positive definite in a sufficiently small relative neighborhood of \(x_{0,0}\) in \(\mathbb {R}_+^4.\)

One can check that

$$\begin{aligned} \dot{V}(x)= & {} (\varepsilon - L_U-L_W) (\rho L_U+ L_W)\nonumber \\&+\,\rho \gamma _U\left( \mathcal{R}_U\frac{A_U}{A_U+A_W}-1-\varepsilon \right) A_U\nonumber \\&+\,\gamma _W(\mathcal{R}_W-1-\varepsilon )A_W. \end{aligned}$$
(48)

The first term of the last expression is positive for all values of \((L_U,L_W)\) in some relative neighborhood of the origin in \(\mathbb {R}_+^2\). Assuming from now on that \(\varepsilon \in (0,\mathcal{R}_W-1)\), one verifies easily that the sum of the two remaining terms in the right-hand side of (48) is positive when exactly one of the two numbers \(A_U, A_W\) is zero, due to (6). Assume now that e.g. \(A_U\ne 0\). Then the sum of the last two terms in (48) is equal to

$$\begin{aligned} \gamma _WA_U\left( b \left( \mathcal{R}_U \frac{1}{1+a}-1-\varepsilon \right) +(\mathcal{R}_W-1-\varepsilon ) a \right) ,\quad a:=\frac{A_W}{A_U},\ b:= \frac{\rho \gamma _U}{\gamma _W}. \end{aligned}$$

We will now prove that there exists \(b>0\) (and therefore \(\rho >0\)) such that the previous expression is positive for any nonnegative a (and therefore for any pair \((A_U,A_W)\) with \(A_U>0, A_W\ge 0\)).

The map

$$\begin{aligned} a \mapsto b \left( \mathcal{R}_U \frac{1}{1+a}-1-\varepsilon \right) +(\mathcal{R}_W-1-\varepsilon ) a \end{aligned}$$
(49)

is clearly convex. It has a positive value at the origin, where its derivative is equal to \(-b\mathcal{R}_U+\mathcal{R}_W-1-\varepsilon \). Taking now \(0<b<\frac{\mathcal{R}_W-1-\varepsilon }{\mathcal{R}_U}\), this expression is positive, which ensures that the map (49) takes on positive values on \(\mathbb {R}^+\). Set \(\rho :=\frac{b\gamma _W}{\gamma _U}.\)

Therefore, positive values of \(\varepsilon \) and \(\rho \) have been exhibited, for which \({\dot{V}}\) is positive definite in a relative neighborhood of \(x_{0,0}\) in \(\mathbb {R}_+^4.\) This demonstrates the instability of \(x_{0.0}.\)

Stability of \(x_{U,0}\). Using (13) and recalling the value of \(x_{U,0}\) given in (16), we see that \({ Df}(x_{U,0})\) is the upper block-triangular matrix

$$\begin{aligned} \begin{pmatrix} 1-2\mathcal{R}_U &{}\quad \gamma _U\mathcal{R}_U &{}\quad 1-\mathcal{R}_U &{}\quad -\gamma _U\mathcal{R}_U\\ 1 &{}\quad -\gamma _U &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad -\mathcal{R}_U &{}\quad \gamma _W\mathcal{R}_W\\ 0 &{}\quad 0 &{}\quad 1 &{}\quad -\gamma _W \end{pmatrix}. \end{aligned}$$

The eigenvalues of this block-triangular matrix have negative real parts if and only if

$$\begin{aligned} \frac{\mathcal{R}_U}{2\mathcal{R}_U-1}<1 \quad \text { and } \quad \mathcal{R}_W<\mathcal{R}_U. \end{aligned}$$

These conditions are satisfied since the sustainability condition (6) holds. In conclusion, the equilibrium \(x_{U,0}\) is locally asymptotically stable.

Stability of \(x_{0,W}\). From (13) and (16) we get that the Jacobian \({ Df}(x_{0,W})\) of f at \(x_{0,W}\) is the lower block-triangular matrix

$$\begin{aligned} \begin{pmatrix} -\mathcal{R}_W &{} 0 &{} 0 &{} 0\\ 1 &{} -\gamma _U &{} 0 &{} 0\\ 1-\mathcal{R}_W &{} 0 &{} 1-2\mathcal{R}_W &{} \gamma _W\mathcal{R}_W\\ 0 &{} 0 &{} 1 &{} -\gamma _W \end{pmatrix}. \end{aligned}$$
(50)

The left-upper \(2\times 2\)-block is a Hurwitz matrix, while asymptotic stability of the second one is equivalent to the condition

$$\begin{aligned} 2\mathcal{R}_W-1 > \mathcal{R}_W, \end{aligned}$$

that is \(\mathcal{R}_W>1\), which holds true, due to hypothesis (6). The equilibrium \(x_{0,W}\) is thus locally asymptotically stable.

Stability of \(x_{U,W}\). The instability of \(x_{U,W}\) can be proved by showing that the determinant of the Jacobian matrix \({ Df}(x_{U,W})\) is negative, which, together with the fact that the state space has even dimension 4, establishes the existence of a positive real root to the characteristic polynomial; and thus that the Jacobian is not a Hurwitz matrix. This argument yields lengthy computations.

It is more appropriate to use here the monotonicity properties of system (7), established in Theorem 5. As a matter of fact, bringing together the inequalities (15) (already proved in the end of the previous section, see (50)), the asymptotical stability of \(x_{U,0}\) and \(x_{0,W}\) and the strongly order-preserving property of the reference problem, Theorem 2.2 in Smith (1995) shows that the intermediate point \(x_{U,W}\) cannot be stable. This finally achieves the stability analysis, as well as the proof of Theorem 7.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bliman, PA., Aronna, M.S., Coelho, F.C. et al. Ensuring successful introduction of Wolbachia in natural populations of Aedes aegypti by means of feedback control. J. Math. Biol. 76, 1269–1300 (2018). https://doi.org/10.1007/s00285-017-1174-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-017-1174-x

Keywords

Mathematics Subject Classification

Navigation