Skip to main content

Advertisement

Log in

Optimal control of vaccination in a vector-borne reaction–diffusion model applied to Zika virus

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

Abstract

Zika virus has acquired worldwide concern after a recent outbreak in Latin America that started in Brazil, with associated neurological conditions such as microcephaly in newborns from infected mothers. The virus is transmitted mainly by Aedes aegypti mosquitoes, but direct (sexual) transmission has been documented. We formulate a reaction diffusion model that considers spatial movement of humans and vectors, with local contact transmission of Zika virus. Vaccination is introduced as a control variable, giving immunity to susceptible humans, in order to characterize an optimal vaccination strategy that minimizes the costs associated with infections and vaccines. The optimal control characterization is obtained in terms of state and adjoint equations. Parameter estimation and numerical simulations are carried out using data for the initial 2015 Zika outbreak in the state of Rio Grande do Norte in Brazil. Several scenarios are considered and analyzed in terms of number of new infections and costs, showing that the optimal control application is successful, significantly reducing these quantities.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  • Abboubakar H, Kamgang JC, Nkamba LN, Tieudjo D (2017) Bifurcation thresholds and optimal control in transmission dynamics of arboviral diseases. J Math Biol 76:1–49

    MathSciNet  MATH  Google Scholar 

  • Agusto F, Bewick S, Fagan W (2017) Mathematical model for Zika virus dynamics with sexual transmission route. Ecol Complex 29:61–81

    Article  Google Scholar 

  • Althaus CL, Low N (2016) How relevant is sexual transmission of Zika virus? PLoS Med. https://doi.org/10.1371/journal.pmed.1002157

    Google Scholar 

  • Anguelov R, Dufourd C, Dumont Y (2017) Simulations and parameter estimation of a trap-insect model using a finite element approach. Math Comput Simul 133:47–75

    Article  MathSciNet  Google Scholar 

  • Asano E, Gross LJ, Lenhart S, Real LA (2008) Optimal control of vaccine distribution in a rabies metapopulation model. Math Biosci Eng 5(2):219–238

    Article  MathSciNet  MATH  Google Scholar 

  • Besnard M, Lastere S, Teissier A, Cao-Lormeau V, Musso D (2014) Evidence of perinatal transmission of Zika virus, French Polynesia, December 2013 and February 2014. Eurosurveillance 19(13):20,751

    Article  Google Scholar 

  • Betts JT (2010) Practical methods for optimal control and estimation using nonlinear programming, vol 19. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  • Bonyah E, Khan MA, Okosun K, Islam S (2017) A theoretical model for Zika virus transmission. PLoS ONE. https://doi.org/10.1371/journal.pone.0185540

    Google Scholar 

  • Brockmann D, Hufnagel L, Geisel T (2006) The scaling laws of human travel. Nature 439(7075):462

    Article  Google Scholar 

  • De Araujo AL, Boldrini JL, Calsavara BM (2016) An analysis of a mathematical model describing the geographic spread of dengue disease. J Math Anal Appl 444(1):298–325

    Article  MathSciNet  MATH  Google Scholar 

  • De Silva KR, Phan TV, Lenhart S (2017) Advection control in parabolic PDE systems for competitive populations. Discrete Contin Dyn Syst Ser B 22(3):1049–1072

    MathSciNet  MATH  Google Scholar 

  • Douglas J Jr, Dupont T, Ewing RE (1979) Incomplete iteration for time-stepping a Galerkin method for a quasilinear parabolic problem. SIAM J Numer Anal 16(3):503–522

    Article  MathSciNet  MATH  Google Scholar 

  • D’Ortenzio E, Matheron S, de Lamballerie X, Hubert B, Piorkowski G, Maquart M, Descamps D, Damond F, Yazdanpanah Y, Leparc-Goffart I (2016) Evidence of sexual transmission of Zika virus. N Engl J Med 374(22):2195–2198

    Article  Google Scholar 

  • Evans LC (1998) Partial differential equations. American Mathematical Society, Providence

    MATH  Google Scholar 

  • Fister R (1997) Optimal control of harvesting in a predator-prey parabolic system. Houst J Math 23(2):341–355

    MathSciNet  MATH  Google Scholar 

  • Fitzgibbon W, Morgan J, Webb G (2017) An outbreak vector-host epidemic model with spatial structure: the 2015–2016 zika outbreak in rio de janeiro. Theoretical Biology and Medical Modelling 14(1):7

    Article  Google Scholar 

  • Gao D, Lou Y, He D, Porco TC, Kuang Y, Chowell G, Ruan S (2016) Prevention and control of zika as a mosquito-borne and sexually transmitted disease: a mathematical modeling analysis. Sci Rep. https://doi.org/10.1038/srep28070

    Google Scholar 

  • Gonzalez MC, Hidalgo CA, Barabasi AL (2008) Understanding individual human mobility patterns. Nature 453(7196):779

    Article  Google Scholar 

  • Hackbusch W (1978) A numerical method for solving parabolic equations with opposite orientations. Computing 20(3):229–240

    Article  MathSciNet  MATH  Google Scholar 

  • Health Office (2017) Rio Grande do Norte epidemiological bulletin, 2016. http://www.saude.rn.gov.br. Accessed 16 August 2017

  • Hinze M, Schulz V (2010) Preface of the guest-editors. GAMM-Mitteilungen 33(2):131–132

    Article  MathSciNet  Google Scholar 

  • Hiriart-Urruty JB, Korytowski A, Maurer H, Szymkat M (2016) Advances in mathematical modeling. Optimization and optimal control, vol 109. Springer, Berlin

    Book  MATH  Google Scholar 

  • Honório NA, Castro MG, Barros FSMd, Magalhães MdAFM, Sabroza PC (2009) The spatial distribution of Aedes aegypti and Aedes albopictus in a transition zone, Rio de Janeiro, Brazil. Cadernos de Saúde Pública 25(6):1203–1214

    Article  Google Scholar 

  • Hughes TJ (2012) The finite element method: linear static and dynamic finite element analysis. Courier Corporation, North Chelmsford

    Google Scholar 

  • Keeling MJ, Rohani P (2008) Modeling infectious diseases in humans and animals. Princeton University Press, Princeton

    Book  MATH  Google Scholar 

  • Kon CML, Labadin J, Tiga J (2016) Generic reaction-diffusion model for transmission of mosquito-borne diseases: results of simulation with actual cases. In: ECMS, pp 93–99

  • Lenhart S, Workman JT (2007) Optimal control applied to biological models. Crc Press, New York

    MATH  Google Scholar 

  • LeVeque RJ (2007) Finite difference methods for ordinary and partial differential equations: steady-state and time-dependent problems, vol 98. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  • Li X, Yong J (2012) Optimal control theory for infinite dimensional systems. Springer, Berlin

    Google Scholar 

  • Lions JL (1971) Optimal control of systems governed by partial differential equations. Springer, Berlin

    Book  MATH  Google Scholar 

  • Lou Y, Zhao XQ (2011) A reaction-diffusion malaria model with incubation period in the vector population. J Math Biol 62(4):543–568

    Article  MathSciNet  MATH  Google Scholar 

  • Malone RW, Homan J, Callahan MV, Glasspool-Malone J, Damodaran L, Schneider ADB, Zimler R, Talton J, Cobb RR, Ruzic I, Smith-Gagen J, Janies D, Wilson J (2016) Zika virus: medical countermeasure development challenges. PLoS Negl Trop Dis. https://doi.org/10.1371/journal.pntd.0004530

    Google Scholar 

  • Marcondes CB, Ximenes MdFFd (2016) Zika virus in Brazil and the danger of infestation by Aedes (Stegomyia) mosquitoes. Revista da Sociedade Brasileira de Medicina Tropical 49(1):4–10

    Article  Google Scholar 

  • Marino S, Hogue IB, Ray CJ, Kirschner DE (2008) A methodology for performing global uncertainty and sensitivity analysis in systems biology. J Theor Biol 254(1):178–196

    Article  MathSciNet  MATH  Google Scholar 

  • Miyaoka TY, Meyer JFCA, Souza JMR (2017) A general boundary condition with linear flux for advection–diffusion models. Trends Appl Comput Math 18(2):253–272

    MathSciNet  Google Scholar 

  • Mlakar J, Korva M, Tul N, Popović M, Poljšak-Prijatelj M, Mraz J, Kolenc M, Resman Rus K, Vesnaver Vipotnik T, Fabjan Vodušek V, Vizjak A, Jože P, Petrovec M (2016) Avšič Županc T (2016) Zika virus associated with microcephaly. New England Journal of Medicine 374:951–958

    Article  Google Scholar 

  • Mocenni C, Madeo D, Sparacino E (2011) Linear least squares parameter estimation of nonlinear reaction diffusion equations. Math Comput Simul 81(10):2244–2257

    Article  MathSciNet  MATH  Google Scholar 

  • Motta IJ, Spencer BR, Cordeiro da Silva SG, Arruda MB, Dobbin JA, Gonzaga YB, Arcuri IP, Tavares RC, Atta EH, Fernandes RF, Costa DA, Ribeiro LJ, Limonte F, Higa LM, Voloch CM, Brindeiro RM, Tanuri A, Ferreira OC (2016) Evidence for transmission of Zika virus by platelet transfusion. N Engl J Med 375(11):1101–1103

    Article  Google Scholar 

  • Murray JD (2002) Mathematical biology. Vol. 2: spatial models and biomedical applications. Springer, New York

    Book  MATH  Google Scholar 

  • Neilan RM, Lenhart S (2011) Optimal vaccine distribution in a spatiotemporal epidemic model with an application to rabies and raccoons. J Math Anal Appl 378(2):603–619

    Article  MathSciNet  MATH  Google Scholar 

  • Neitzel I, Tröltzsch F (2008) On convergence of regularization methods for nonlinear parabolic optimal control problems with control and state constraints. Control Cybernet 37(4):1013–1043

    MathSciNet  MATH  Google Scholar 

  • NIAID (2017) National Institute of Allergy and Infectious Diseases. https://www.niaid.nih.gov/news-events/nih-begins-testing-investigational-zika-vaccine-humans. Accessed 26 Oct 2017

  • Oehler E, Watrin L, Larre P, Leparc-Goffart I, Lastere S, Valour F, Baudouin L, Mallet H, Musso D, Ghawche F (2014) Zika virus infection complicated by Guillain-Barre syndrome–case report, French Polynesia, December 2013. Eurosurveillance. https://doi.org/10.2807/1560-7917.ES2014.19.9.20720

    Google Scholar 

  • Otero M, Schweigmann N, Solari HG (2008) A stochastic spatial dynamical model for Aedes aegypti. Bull Math Biol 70(5):1297

    Article  MathSciNet  MATH  Google Scholar 

  • PAHO (2017) Pan American Health Organization. https://www.paho.org/zika. Accessed 26 Oct 2017

  • Powell MJ (2009) The BOBYQA algorithm for bound constrained optimization without derivatives. Cambridge NA Report NA2009/06, University of Cambridge, Cambridge

  • Raimundo SM, Yang HM, Massad E (2016) Modeling vaccine preventable vector-borne infections: yellow fever as a case study. J Biol Syst 24(02n03):193–216

    Article  MathSciNet  MATH  Google Scholar 

  • Rodrigues HS, Monteiro MTT, Torres DF (2014) Vaccination models and optimal control strategies to dengue. Math Biosci 247:1–12

    Article  MathSciNet  MATH  Google Scholar 

  • Soubeyrand S, Roques L (2014) Parameter estimation for reaction–diffusion models of biological invasions. Popul Ecol 56(2):427–434

    Article  Google Scholar 

  • Valega-Mackenzie W, Ríos-Soto KR (2018) Can vaccination save a Zika virus epidemic? Bull Math Biol 80(3):598–625

    Article  MathSciNet  MATH  Google Scholar 

  • Wang L, Zhao H, Oliva SM, Zhu H (2017) Modeling the transmission and control of Zika in brazil. Sci Rep. https://doi.org/10.1038/s41598-017-07264-y

    Google Scholar 

  • WHO (2017) World Health Organization. https://www.who.int/topics/zika. Accessed 26 Oct 2017

  • Zheng B, Tang M, Yu J, Qiu J (2017) Wolbachia spreading dynamics in mosquitoes with imperfect maternal transmission. J Math Biol 76:1–29

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tiago Yuzo Miyaoka.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This work was supported by CAPES (Process Number 88881.134100/2016-01) and by the National Institute for Mathematical and Biological Synthesis, an Institute sponsored by the National Science Foundation through NSF Award #DBI-1300426, with additional support from The University of Tennessee, Knoxville.

Appendix: Derivation of optimal control

Appendix: Derivation of optimal control

We seek to find an optimal control \(u^* \in U\) satisfying (4), where U is the set of admissible controls defined in (3). This optimization problem is subject to the state system (1), with state functions in \(L^2(0,T;H^1(\varOmega ))\) and corresponding time derivatives in \(L^2(0,T;H^1(\varOmega )^*)\). In order to find \(u^*\), we have to solve an optimality system, which consists of the state PDEs, adjoint PDEs, and the optimal control characterization. To obtain the sensitivity system, we must differentiate the control-to-state map \(u \rightarrow \left( S,I,R,S_v,I_v \right) (u)\). We use that system and the objective functional to find the adjoint system. Then we use the sensitivity and adjoint system to simplify the derivative of the map \(u \rightarrow J(u)\), finally obtaining the desired optimal control characterization. The derivatives of both of these maps are computed as Fréchet derivatives (derivatives in dircetion \(\ell \)). In the following, we provide an overview of the derivation of the sensitivity and adjoint equations, and the control characterization. Details about optimal control theory in PDEs can be found in De Silva et al. (2017), Fister (1997), Lenhart and Workman (2007), Li and Yong (2012), Lions (1971).

1.1 Sensitivity system

From the control-to-state map, we define \((S^\epsilon ,I^\epsilon ,R^\epsilon ,S_v^\epsilon ,I_v^\epsilon ) = (S,I,R,S_v,I_v)(u + \epsilon \ell )\). We then define the sensitivity functions, which are Gateaux derivatives of the control-to-state map:

$$\begin{aligned}&\dfrac{S^\epsilon - S}{\epsilon } \rightharpoonup \psi _1, \ \dfrac{I^\epsilon - I}{\epsilon } \rightharpoonup \psi _2, \ \dfrac{R^\epsilon - R}{\epsilon } \rightharpoonup \psi _3, \ \dfrac{S_v^\epsilon - S_v}{\epsilon } \rightharpoonup \psi _4, \\&\quad \dfrac{I_v^\epsilon - I_v}{\epsilon } \rightharpoonup \psi _5, \text { as } \epsilon \rightarrow 0. \end{aligned}$$

We also have:

$$\begin{aligned} S^\epsilon \rightarrow S, \ I^\epsilon \rightarrow I, \ R^\epsilon \rightarrow R, \ S_v^\epsilon \rightarrow S_v, \ I_v^\epsilon \rightarrow I_v, \text { as } \epsilon \rightarrow 0. \end{aligned}$$
(13)

The sensitivity functions \(\varvec{\psi }=(\psi _1,\psi _2,\psi _3,\psi _4,\psi _5)\) satisfy a system corresponding to a linearized version of the state equations. See works by De Silva et al. (2017), Li and Yong (2012), and Lions (1971) to show justification of such convergence of these quotients and the existence of the sensitivity functions and their PDEs. These equations are used to derive PDEs for the adjoint \(\varvec{\lambda }=(\lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5)\), where each adjoint variable is associated with a state variable.

We now calculate equations for the sensitivity functions. We calculate an equation for the quotient \(\dfrac{S^\epsilon - S}{\epsilon }\):

$$\begin{aligned} \begin{aligned}&\left( \dfrac{S^\epsilon - S}{\epsilon } \right) _t - \dfrac{\nabla \cdot \left( \alpha \nabla S^\epsilon \right) - \nabla \cdot \left( \alpha \nabla S \right) }{\epsilon } = - \beta \left( \dfrac{ S^\epsilon I_v^\epsilon - S^\epsilon I_v + S^\epsilon I_v - S I_v}{\epsilon } \right) \\&\quad - \beta _d \left( \dfrac{ S^\epsilon I^\epsilon - S^\epsilon I + S^\epsilon I - S I}{\epsilon } \right) - u \left( \dfrac{S^\epsilon - S}{\epsilon }\right) - \ell S^\epsilon . \end{aligned} \end{aligned}$$

Letting \(\epsilon \rightarrow 0\), passing to the limit gives:

$$\begin{aligned} \begin{aligned} (\psi _1)_t - \nabla \cdot \left( \alpha \psi _1 \right) + \beta \left( S \psi _5 + \psi _1 I_v \right) + \beta _d \left( S \psi _2 + \psi _1 I \right) + u \psi _1 = - \ell S. \end{aligned} \end{aligned}$$

In the same way, letting \(\epsilon \rightarrow 0\) in

$$\begin{aligned} \begin{aligned}&\left( \dfrac{I^\epsilon - I}{\epsilon } \right) _t - \dfrac{\nabla \cdot \left( \alpha _I \nabla I^\epsilon \right) - \nabla \cdot \left( \alpha \nabla I \right) }{\epsilon } = \beta \left( \dfrac{ S^\epsilon I_v^\epsilon - S^\epsilon I_v + S^\epsilon I_v - S I_v}{\epsilon } \right) \\&\quad +\beta _d \left( \dfrac{ S^\epsilon I^\epsilon - S^\epsilon I + S^\epsilon I - S I}{\epsilon } \right) - \delta \left( \dfrac{I^\epsilon - I}{\epsilon }\right) + \left( \dfrac{f - f}{\epsilon } \right) \end{aligned} \end{aligned}$$

gives:

$$\begin{aligned} \begin{aligned} (\psi _2)_t - \nabla \cdot \left( \alpha _I \psi _2 \right) - \beta \left( S \psi _5 + \psi _1 I_v \right) - \beta _d \left( S \psi _2 + \psi _1 I \right) + \delta \psi _2 = 0. \end{aligned} \end{aligned}$$

Note that the term with external source f vanishes because it depends only on (xt). Similarly:

$$\begin{aligned} \begin{aligned}&\left( \dfrac{R^\epsilon - R}{\epsilon } \right) _t - \dfrac{\nabla \cdot \left( \alpha \nabla R^\epsilon \right) - \nabla \cdot \left( \alpha \nabla R \right) }{\epsilon } = u \left( \dfrac{S^\epsilon - S}{\epsilon }\right) + \ell S^\epsilon + \delta \left( \dfrac{I^\epsilon - I}{\epsilon } \right) , \end{aligned} \end{aligned}$$

gives:

$$\begin{aligned} \begin{aligned} (\psi _3)_t - \nabla \cdot \left( \alpha \psi _3 \right) - u \psi _1 - \delta \psi _2 = \ell S. \end{aligned} \end{aligned}$$

For \(S_v\), we obtain:

$$\begin{aligned} \begin{aligned}&\left( \dfrac{S_v^\epsilon - S_v}{\epsilon } \right) _t - \dfrac{\nabla \cdot \left( \alpha _v \nabla S_v^\epsilon \right) - \nabla \cdot \left( \alpha _v \nabla S_v \right) }{\epsilon } = - \beta _v \left( \dfrac{ S_v^\epsilon I^\epsilon - S_v^\epsilon I + S_v^\epsilon I - S_v I}{\epsilon } \right) \\&\quad + r_v \left( \dfrac{S_v^\epsilon - S_v}{\epsilon }\right) + r_v \left( \dfrac{I_v^\epsilon - I_v}{\epsilon }\right) - \dfrac{r_v}{\kappa _v} \left( \dfrac{ S_v^\epsilon S_v^\epsilon - S_v^\epsilon S_v + S_v^\epsilon S_v - S_v S_v}{\epsilon } \right) \\&\quad - \dfrac{2 r_v}{\kappa _v} \left( \dfrac{ S_v^\epsilon I_v^\epsilon - S_v^\epsilon I_v + S_v^\epsilon I_v - S_v I_v}{\epsilon } \right) - \dfrac{r_v}{\kappa _v} \left( \dfrac{ I_v^\epsilon I_v^\epsilon - I_v^\epsilon I_v + I_v^\epsilon I_v - I_v I_v}{\epsilon } \right) , \end{aligned} \end{aligned}$$

and:

$$\begin{aligned} \begin{aligned}&(\psi _4)_t - \nabla \cdot \left( \alpha _v \psi _4 \right) + \beta _v \left( S_v \psi _2 + \psi _4 I \right) - r_v \left( \psi _4 + \psi _5 \right) \\&\qquad + \dfrac{2 r_v}{\kappa _v} \left( S_v + I_v \right) \left( \psi _4 + \psi _5 \right) = 0. \end{aligned} \end{aligned}$$

Similarly, for \(I_v\) we have:

$$\begin{aligned} \begin{aligned}&\left( \dfrac{I_v^\epsilon - I_v}{\epsilon } \right) _t - \dfrac{\nabla \cdot \left( \alpha _v \nabla I_v^\epsilon \right) - \nabla \cdot \left( \alpha _v \nabla I_v \right) }{\epsilon }\\&\quad = + \beta _v \left( \dfrac{ S_v^\epsilon I^\epsilon - S_v^\epsilon I + S_v^\epsilon I - S_v I}{\epsilon } \right) - \mu _v \left( \dfrac{S_v^\epsilon - S_v}{\epsilon }\right) , \end{aligned} \end{aligned}$$

and:

$$\begin{aligned} \begin{aligned} (\psi _5)_t - \nabla \cdot \left( \alpha _v \psi _5 \right) - \beta _v \left( S_v \psi _2 + \psi _4 I \right) + \mu _v \psi _5 = 0. \end{aligned} \end{aligned}$$

In a similar way, we obtain initial conditions:

$$\begin{aligned} \psi _1(\varvec{x},0) = \psi _2(\varvec{x},0) = \psi _3(\varvec{x},0) = \psi _4(\varvec{x},0) = \psi _5(\varvec{x},0) = 0, \ \text {in} \ \varOmega , \end{aligned}$$

and boundary conditions:

$$\begin{aligned} \dfrac{\partial \psi _1}{\partial \varvec{n}} = \dfrac{\partial \psi _2}{\partial \varvec{n}} = \dfrac{\partial \psi _3}{\partial \varvec{n}} = \dfrac{\partial \psi _4}{\partial \varvec{n}} = \dfrac{\partial \psi _5}{\partial \varvec{n}} = 0 \ \text {in} \ \partial \varOmega \times (0,T). \end{aligned}$$

We can write the sensitivity system in vector form as following:

$$\begin{aligned} \mathscr {L} \begin{pmatrix} \psi _1 \\ \psi _2 \\ \psi _3 \\ \psi _4 \\ \psi _5 \\ \end{pmatrix} = \begin{pmatrix} -\ell S \\ 0 \\ \ell S \\ 0 \\ 0 \\ \end{pmatrix}, \end{aligned}$$

where:

$$\begin{aligned} \mathscr {L} \begin{pmatrix} \psi _1 \\ \psi _2 \\ \psi _3 \\ \psi _4 \\ \psi _5 \\ \end{pmatrix} = \begin{pmatrix} (\psi _1)_t - \nabla \cdot (\alpha \nabla \psi _1) \\ (\psi _2)_t - \nabla \cdot (\alpha _I \nabla \psi _2) \\ (\psi _3)_t - \nabla \cdot (\alpha \nabla \psi _3) \\ (\psi _4)_t - \nabla \cdot (\alpha _v \nabla \psi _4) \\ (\psi _5)_t - \nabla \cdot (\alpha _v \nabla \psi _5) \\ \end{pmatrix} + \varvec{M} \begin{pmatrix} \psi _1 \\ \psi _2 \\ \psi _3 \\ \psi _4 \\ \psi _5 \\ \end{pmatrix}, \end{aligned}$$

and the matrix \(\varvec{M}^T\) is given by (6).

1.2 Adjoint system

For the adjoint variables \(\varvec{\lambda }=(\lambda _1,\ldots , \lambda _5)\), the adjoint operator is obtained from the relation:

$$\begin{aligned} \langle \varvec{\lambda } , \mathscr {L} \varvec{\psi } \rangle = \langle \mathscr {L}^* \varvec{\lambda } , \varvec{\psi } \rangle \end{aligned}$$
(14)

in the appropriate weak \(L^2(Q)\) sense. The right hand side of the adjoint system is given by the derivatives of the integrand of the objective functional, with respect to the states:

$$\begin{aligned} \begin{aligned} \dfrac{\partial }{\partial S} \left( c_1 I + c_2 u S + c_3 u^2 \right)&= c_2 u, \\ \dfrac{\partial }{\partial I} \left( c_1 I + c_2 u S + c_3 u^2 \right)&= c_1, \\ \dfrac{\partial }{\partial R} \left( c_1 I + c_2 u S + c_3 u^2 \right)&= 0, \\ \dfrac{\partial }{\partial S_v} \left( c_1 I + c_2 u S + c_3 u^2 \right)&= 0, \\ \dfrac{\partial }{\partial I_v} \left( c_1 I + c_2 u S + c_3 u^2 \right)&= 0. \end{aligned} \end{aligned}$$

We use the expression (14) in order to derive the operators in the adjoint system. This is done using integration by parts on \(\langle \varvec{\lambda } , \mathscr {L} \varvec{\psi } \rangle \) and the boundary conditions of the sensitivity equations. The adjoint system obtained this way is:

$$\begin{aligned} \mathscr {L^*} \begin{pmatrix} \lambda _1 \\ \lambda _2 \\ \lambda _3 \\ \lambda _4 \\ \lambda _5 \\ \end{pmatrix} = \begin{pmatrix} c_2 u \\ c_1 \\ 0 \\ 0 \\ 0 \\ \end{pmatrix}, \end{aligned}$$

where:

$$\begin{aligned} \mathscr {L}^* \begin{pmatrix} \lambda _1 \\ \lambda _2 \\ \lambda _3 \\ \lambda _4 \\ \lambda _5 \\ \end{pmatrix} = \begin{pmatrix} \mathscr {L}_1^* \lambda _1 \\ \mathscr {L}_2^* \lambda _2 \\ \mathscr {L}_3^* \lambda _3 \\ \mathscr {L}_4^* \lambda _4 \\ \mathscr {L}_5^* \lambda _5 \\ \end{pmatrix} + \varvec{M}^T \begin{pmatrix} \lambda _1 \\ \lambda _2 \\ \lambda _3 \\ \lambda _4 \\ \lambda _5 \\ \end{pmatrix}, \end{aligned}$$

and:

$$\begin{aligned} \begin{pmatrix} \mathscr {L}_1^* \lambda _1 \\ \mathscr {L}_2^* \lambda _2 \\ \mathscr {L}_3^* \lambda _3 \\ \mathscr {L}_4^* \lambda _4 \\ \mathscr {L}_5^* \lambda _5 \\ \end{pmatrix} = \begin{pmatrix} -(\lambda _1)_t - \nabla \cdot (\alpha \nabla \lambda _1) \\ -(\lambda _2)_t - \nabla \cdot (\alpha _I \nabla \lambda _2) \\ -(\lambda _3)_t - \nabla \cdot (\alpha \nabla \lambda _3) \\ -(\lambda _4)_t - \nabla \cdot (\alpha _v \nabla \lambda _4) \\ -(\lambda _5)_t - \nabla \cdot (\alpha _v \nabla \lambda _5) \\ \end{pmatrix}, \end{aligned}$$

with final time conditions given by (7) and boundary conditions by (8).

1.3 Optimal control characterization

We now derive an expression for an optimal control \(u^*\) as a function of the state and adjoint variables. The derivative of the map \(u \rightarrow J(u)\) at \(u^*\) in the direction \(\ell \) is:

$$\begin{aligned} \lim _{\epsilon \rightarrow 0^+} \dfrac{J(u^*+\epsilon \ell ) - J(u^*)}{\epsilon } \ge 0, \end{aligned}$$

with:

$$\begin{aligned} \begin{aligned} J(u^*+\epsilon \ell )&= \int _Q \left[ c_1 I^\epsilon + c_2 (u^* + \epsilon \ell ) S^\epsilon + c_3 ( u^* + \epsilon \ell )^2 \right] \mathrm {d}\varvec{x} \mathrm {d}t \\&= \int _Q \left[ c_1 I^\epsilon + c_2 u^* S^\epsilon + \epsilon c_2 \ell S^\epsilon + c_3 (u^*)^2 + 2 c_3 u^* \epsilon \ell + c_3 \epsilon ^2 \ell ^2 \right] \mathrm {d}\varvec{x} \mathrm {d}t. \\ J(u^*)&= \int _Q \left[ c_1 I^\epsilon + c_2 u^* + c_3 ( u^*)^2 \right] \mathrm {d}\varvec{x} \mathrm {d}t. \end{aligned} \end{aligned}$$

Using the sensitivity and adjoint systems we find a characterization for \(u^*\):

$$\begin{aligned} \begin{aligned} 0&\le \lim _{\epsilon \rightarrow 0^+} \dfrac{J(u^*+\epsilon \ell ) - J(u^*)}{\epsilon } \\ 0&\le \lim _{\epsilon \rightarrow 0^+} \int _Q \left[ c_1 \dfrac{I^\epsilon -I}{\epsilon } + c_2 u^* \dfrac{S^\epsilon -S}{\epsilon } + c_2 \ell S^\epsilon + 2 c_3 u^* \ell + c_3 \epsilon \ell ^2 \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ c_1 \psi _2 + c_2 u^* \psi _1 + c_2 \ell S^* + 2 c_3 u^* \ell \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ \left( \psi _1, \psi _2, \psi _3, \psi _4, \psi _5 \right) \left( c_2 u^*, c_1, 0, 0, 0 \right) ^T + \ell \left( c_2 S^* + 2 c_3 u^* \right) \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ \left( \psi _1, \psi _2, \psi _3, \psi _4, \psi _5 \right) \mathscr {L}^*\left( \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5 \right) ^T + \ell \left( c_2 S^* + 2 c_3 u^* \right) \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ \left( \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5 \right) \mathscr {L}\left( \psi _1, \psi _2, \psi _3, \psi _4, \psi _5 \right) ^T + \ell \left( c_2 S^* + 2 c_3 u^* \right) \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ \left( \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5 \right) \left( - \ell S^*, 0, \ell S^*,0,0 \right) ^T + \ell \left( c_2 S^* + 2 c_3 u^* \right) \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \left[ -\ell \lambda _1 S^* + \ell \lambda _3 S^* + \ell \left( c_2 S^* + 2 c_3 u^* \right) \right] \mathrm {d}\varvec{x} \mathrm {d}t \\ 0&\le \int _Q \ell \left[ \left( -\lambda _1 + \lambda _3 + c_2 \right) S^* + 2 c_3 u^* \right] \mathrm {d}\varvec{x} \mathrm {d}t. \\ \end{aligned} \end{aligned}$$

On the set where \(u_{\min }< u^*(x, t) < u_{\max }\), the variation \(\ell \) can be an \(L^{\infty }\) function with arbitrary sign, and thus on this set, we obtain:

$$\begin{aligned} u^* = \dfrac{(\lambda _1 - \lambda _3 - c_2)S^*}{2c_3}. \end{aligned}$$

Due to the boundedness \(u_{\min } \le u \le u_{\max }\), and taking variations \(\ell \) with appropriate signs when \(u^*\) is at the lower or upper bounds, we can obtain the following optimal control characterization:

$$\begin{aligned} u^* = \min \left( u_{\max }, \max \left( \dfrac{(\lambda _1 - \lambda _3 - c_2)S^*}{2c_3}, u_{\min } \right) \right) . \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Miyaoka, T.Y., Lenhart, S. & Meyer, J.F.C.A. Optimal control of vaccination in a vector-borne reaction–diffusion model applied to Zika virus. J. Math. Biol. 79, 1077–1104 (2019). https://doi.org/10.1007/s00285-019-01390-z

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01390-z

Keywords

Mathematics Subject Classification

Navigation