Skip to main content
Log in

Mean-Field Pontryagin Maximum Principle

  • Published:
Journal of Optimization Theory and Applications Aims and scope Submit manuscript

Abstract

We derive a maximum principle for optimal control problems with constraints given by the coupling of a system of ordinary differential equations and a partial differential equation of Vlasov type with smooth interaction kernel. Such problems arise naturally as Gamma-limits of optimal control problems constrained by ordinary differential equations, modeling, for instance, external interventions on crowd dynamics by means of leaders. We obtain these first-order optimality conditions in the form of Hamiltonian flows in the Wasserstein space of probability measures with forward–backward boundary conditions with respect to the first and second marginals, respectively. In particular, we recover the equations and their solutions by means of a constructive procedure, which can be seen as the mean-field limit of the Pontryagin Maximum Principle applied to the optimal control problem for the discretized density, under a suitable scaling of the adjoint variables.

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.

Similar content being viewed by others

Notes

  1. We follow the notation of [33].

  2. For simplicity of computation, we consider minimization of 4 times the variance.

References

  1. Helbing, D., Farkas, I., Vicsek, T.: Simulating dynamical features of escape panic. Nature 407(6803), 487–490 (2000)

    Article  Google Scholar 

  2. Hughes, R.L.: A continuum theory for the flow of pedestrians. Transp. Res. B Methodol. 36(6), 507–535 (2002)

    Article  Google Scholar 

  3. Reynolds, C.W.: Flocks, herds and schools: a distributed behavioral model. ACM SIGGRAPH Comput. Graph. 21(4), 25–34 (1987)

    Article  Google Scholar 

  4. Vicsek, T., Czirók, A., Ben-Jacob, E., Cohen, I., Shochet, O.: Novel type of phase transition in a system of self-driven particles. Phys. Rev. Lett. 75(6), 1226 (1995)

    Article  MathSciNet  Google Scholar 

  5. Cucker, F., Dong, J.G.: A general collision-avoiding flocking framework. IEEE Trans. Autom. Control 56(5), 1124–1129 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cucker, F., Smale, S.: Emergent behavior in flocks. IEEE Trans. Autom. Control 52(5), 852–862 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  7. D’Orsogna, M.R., Chuang, Y.L., Bertozzi, A.L., Chayes, L.S.: Self-propelled particles with soft-core interactions: patterns, stability, and collapse. Phys. Rev. Lett. 96(10), 104302 (2006)

    Article  Google Scholar 

  8. Motsch, S., Tadmor, E.: Heterophilious dynamics enhances consensus. SIAM Rev. 56(4), 577–621 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  9. Huang, M., Caines, P.E., Malhamé, R.P.: Individual and mass behaviour in large population stochastic wireless power control problems: centralized and Nash equilibrium solutions. In: 42nd IEEE Conference on Decision and Control 1, pp. 98–103. IEEE (2003)

  10. Lasry, J.M., Lions, P.L.: Mean field games. Jpn. J. Math. 2(1), 229–260 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  11. Bongini, M., Fornasier, M.: Sparse stabilization of dynamical systems driven by attraction and avoidance forces. Netw. Heterog. Media 9(1), 1–31 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  12. Bongini, M., Fornasier, M., Frölich, F., Hagverdi, L : Sparse control of force field dynamics. In: International Conference on NETwork Games, COntrol and OPtimization (2014)

  13. Bongini, M., Fornasier, M., Junge, O., Scharf, B.: Sparse control of alignment models in high dimension. Netw. Heterog. Media 10(3), 647–697 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  14. Bongini, M., Fornasier, M., Kalise, D.: (Un)conditional consensus emergence under feedback controls. Discrete Contin. Dyn. Syst. 35(9), 4071–4094 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  15. Caponigro, M., Fornasier, M., Piccoli, B., Trélat, E.: Sparse stabilization and control of alignment models. Math. Models Methods Appl. Sci. 25(03), 521–564 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  16. Bensoussan, A., Frehse, J., Yam, P.: Mean Field Games and Mean Field Type Control Theory. Springer, New York (2013)

    Book  MATH  Google Scholar 

  17. Fornasier, M., Solombrino, F.: Mean-field optimal control. ESAIM Control Optim. Calc. Var. 20(4), 1123–1152 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  18. Piccoli, B., Rossi, F., Trélat, E.: Control to flocking of the kinetic Cucker–Smale model. SIAM J. Math. Anal. 47(6), 4685–4719 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  19. Clason, C., Kunisch, K.: A duality-based approach to elliptic control problems in non-reflexive Banach spaces. ESAIM Control Optim. Calc. Var. 17(01), 243–266 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  20. Pieper, K., Vexler, B.: A priori error analysis for discretization of sparse elliptic optimal control problems in measure space. SIAM J. Control Optim. 51(4), 2788–2808 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  21. Privat, Y., Trélat, E., Zuazua, E.: Optimal location of controllers for the one-dimensional wave equation. Ann. Inst. H. Poincaré Anal. Non Linéaire 30(6), 1097–1126 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  22. Stadler, G.: Elliptic optimal control problems with \(L^1\)-control cost and applications for the placement of control devices. Comput. Optim. Appl. 44(2), 159–181 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  23. Fornasier, M., Piccoli, B., Rossi, F.: Mean-field sparse optimal control. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 372(2028), 20130400 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  24. Albi, G., Bongini, M., Cristiani, E., Kalise, D.: Invisible control of self-organizing agents leaving unknown environments. SIAM J. Appl. Math. 76(4), 1683–1710 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  25. Andersson, D., Djehiche, B.: A maximum principle for SDEs of mean-field type. Appl. Math. Comput. 63(3), 341–356 (2011)

    MathSciNet  MATH  Google Scholar 

  26. Burger, M., Di Francesco, M., Markowich, P., Wolfram, M.T.: Mean field games with nonlinear mobilities in pedestrian dynamics. Discrete Contin. Dyn. Syst. Ser. B 19, 1311–1333 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  27. Carmona, R., Delarue, F., Lachapelle, A.: Control of McKean–Vlasov dynamics versus mean field games. Math. Finance Econ. 7(2), 131–166 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  28. Raymond, J.P., Zidani, H.: Hamiltonian Pontryagin’s principles for control problems governed by semilinear parabolic equations. Appl. Math. Optim. 39(2), 143–177 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  29. Albi, G., Choi, Y.P., Fornasier, M., Kalise, D.: Mean field control hierarchy. arXiv:1608.01728

  30. Clarke, F.: Functional Analysis, Calculus of Variations and Optimal Control. Springer, New York (2013)

    Book  MATH  Google Scholar 

  31. Ambrosio, L., Gangbo, W.: Hamiltonian ODEs in the Wasserstein space of probability measures. Commun. Pure Appl. Math. 61(1), 18–53 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  32. Cullen, M., Gangbo, W., Pisante, G.: The semigeostrophic equations discretized in reference and dual variables. Arch. Ration. Mech. Anal. 185(2), 341–363 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  33. Ambrosio, L., Gigli, N., Savaré, G.: Gradient Flows in Metric Spaces and in the Space of Probability Measures. Lectures in Mathematics ETH Zürich, 2nd edn. Birkhäuser Verlag, Basel (2008)

    MATH  Google Scholar 

  34. Villani, C.: Topics in Optimal Transportation, Graduate Studies in Mathematics, vol. 58. American Mathematical Society, Providence (2003)

    MATH  Google Scholar 

  35. Hegselmann, R., Krause, U.: Opinion dynamics and bounded confidence models, analysis, and simulation. J. Artif. Soc. Soc. Simul. 5(3) (2002)

  36. Levine, H., Rappel, W.J., Cohen, I.: Self-organization in systems of self-propelled particles. Phys. Rev. E 63(1), 017101 (2000)

    Article  Google Scholar 

  37. Chuang, Y.L., D’Orsogna, M.R., Marthaler, D., Bertozzi, A.L., Chayes, L.S.: State transitions and the continuum limit for a 2D interacting, self-propelled particle system. Phys. D 232(1), 33–47 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  38. Wongkaew, S., Caponigro, M., Borzi, A.: On the control through leadership of the Hegselmann–Krause opinion formation model. Math. Models Methods Appl. Sci. 25(3), 565–585 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  39. Dobrušin, R.L.: Vlasov equations. Funktsional. Anal. i Prilozhen 13(2), 48–58, 96 (1979)

  40. Villani, C.: Landau damping. In: Numerical Models for Fusion. Panor. Synthèses, vol. 39/40, pp. 237–326. Soc. Math. France, Paris (2013)

  41. Pfaffelmoser, K.: Global classical solutions of the Vlasov–Poisson system in three dimensions for general initial data. J. Differ. Equ. 95(2), 281–303 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  42. Lions, P.L., Perthame, B.: Propagation of moments and regularity for the \(3\)-dimensional Vlasov–Poisson system. Invent. Math. 105(2), 415–430 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  43. Piccoli, B., Rossi, F.: Transport equation with nonlocal velocity in Wasserstein spaces: convergence of numerical schemes. Acta Appl. Math. 124(1), 73–105 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  44. Piccoli, B., Rossi, F.: Generalized Wasserstein distance and its application to transport equations with source. Arch. Ration. Mech. Anal. 211(1), 335–358 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  45. Kinderlehrer, D., Stampacchia, G.: An Introduction to Variational Inequalities and Their Applications. Academic Press, New York (1980)

    MATH  Google Scholar 

  46. Dal Maso, G.: An Introduction to \(\varGamma \)-Convergence. Progress in Nonlinear Differential Equations and Their Applications. Birkhäuser, Boston (1993)

    Google Scholar 

  47. Ambrosio, L., Fusco, N., Pallara, D.: Functions of Bounded Variation and Free Discontinuity Problems, vol. 254. Clarendon Press, Oxford (2000)

    MATH  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the support of the PHC-PROCOPE Project “Sparse Control of Multiscale Models of Collective Motion.” Mattia Bongini and Massimo Fornasier additionally acknowledge the support of the ERC-Starting Grant Project “High-Dimensional Sparse Optimal Control.” Francesco Rossi additionally acknowledges the support of the ANR project CroCo ANR-16-CE33-0008.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Francesco Rossi.

Additional information

Communicated by Aram Arutyunov.

Appendix: Semiconvexity Along Geodesics of \(\mathbb {H}_c\)

Appendix: Semiconvexity Along Geodesics of \(\mathbb {H}_c\)

Throughout the section, \({\mathcal {K}}\) shall denote a convex, compact subset of \({\mathbb R}^{2d}\). The following property shall be used to prove that the subdifferential of \(\mathbb {H}_c\) is nonempty.

Definition A.1

Let \(\psi : {\mathcal {P}}_2({\mathbb R}^n) \rightarrow ]-\infty ,+\infty ]\) be a proper, lower semicontinuous functional. We say that \(\psi \) is semiconvex along geodesics whenever, for every \(\nu _0, \nu _1 \in {\mathcal {P}}_2({\mathbb R}^n)\) and every optimal transport plan \(\rho \in \varPi _o(\nu _0, \nu _1)\) there exists \(C \in {\mathbb R}\) for which for every \(s \in [0,1]\) it holds

$$\begin{aligned}&\psi (((1-s)\pi _1 + s\pi _2)_{\#}\rho ) \\&\quad \le (1-s) \psi (\nu _0) + s \psi (\nu _1) + Cs(1-s) {\mathcal W}^2_2(\nu _0, \nu _1)). \end{aligned}$$

In order to prove the semiconvexity of \(\mathbb {H}_c\), we shall establish the semiconvexity of the following functionals:

where \(\hat{{\mathcal {F}}}\), \(\hat{{\mathcal {G}}}\), \(\hat{\ell }\), and \(\hat{\omega }\) are \({\mathcal {C}}^2\) functions. The desired result will then follow by noticing that \(\mathbb {H}_c(\nu ) = \hat{\mathbb {H}}_c^1(\nu ) + \hat{\mathbb {H}}_c^2(\nu )\) for \(\hat{{\mathcal {F}}} = {\mathcal {F}}\), \(\hat{{\mathcal {G}}} = {\mathcal {G}}\), \(\hat{\ell } = -\ell \circ (\pi _1, Id )\), \(\hat{\omega } = \omega \circ \pi _1\) and \({\mathcal {K}} = cl \!\left( {B(0,R_T)}\right) \).

The following simple property will be needed to prove semiconvexity of the above functionals.

Lemma A.1

Let \(\nu _0,\nu _1 \in {\mathcal P}_c({\mathbb R}^{2d})\) with support contained in \({\mathcal {K}}\). Let \(\rho \in \varPi (\nu _0,\nu _1)\) and set

$$\begin{aligned} \nu _s = ((1-s)\pi _1 + s\pi _2)_{\#} \rho , \end{aligned}$$
(42)

for every \(s \in [0,1]\). Then, it holds

$$\begin{aligned} \mathrm {supp}(\nu _s) \subseteq {\mathcal {K}} \quad \text { for all } s \in [0,1]. \end{aligned}$$

Proof

We first notice that for every \(\rho \in \varPi (\nu _0,\nu _1)\) it holds

$$\begin{aligned} \mathrm {supp}(\rho ) \subseteq {\mathcal {K}} \times {\mathcal {K}}\,. \end{aligned}$$
(43)

This follows from the equality

$$\begin{aligned} {\mathbb R}^{4d} \backslash ({\mathcal {K}} \times {\mathcal {K}}) = ({\mathbb R}^{2d} \times ({\mathbb R}^{2d} \backslash {\mathcal {K}})) \cup (({\mathbb R}^{2d} \backslash {\mathcal {K}}) \times {\mathbb R}^{2d}) \end{aligned}$$

and from the fact that both \({\mathbb R}^{2d} \times ({\mathbb R}^{2d} \backslash {\mathcal {K}}))\) and \(({\mathbb R}^{2d} \backslash {\mathcal {K}}) \times {\mathbb R}^{2d}\) are \(\rho \)-null sets by hypothesis.

To prove Lemma A.1, it suffices to show that for all \(f \in {\mathcal {C}}({\mathbb R}^{2d})\) satisfying \(f \equiv 0\) on \({\mathcal {K}}\) it holds

$$\begin{aligned} \int _{{\mathbb R}^{2d}} f \hbox {d}\nu _s = 0. \end{aligned}$$
(44)

Indeed,

$$\begin{aligned} \int _{{\mathbb R}^{2d}} f \hbox {d}\nu _s&= \int _{{\mathbb R}^{4d}} f d((1-s)\pi _1 + s\pi _2)_{\#} \rho (z_0,z_1) \\&=\int _{{\mathbb R}^{4d}} f((1-s)z_0 + s z_1) d \rho (z_0,z_1) \\&=\int _{{\mathcal {K}} \times {\mathcal {K}}} f((1-s)z_0 + s z_1) d \rho (z_0,z_1), \end{aligned}$$

since, by (43), \(\mathrm {supp}(\rho ) \subseteq {\mathcal {K}} \times {\mathcal {K}}\). The convexity of \({\mathcal {K}}\) implies \((1-s)z_0 + sz_1 \in {\mathcal {K}}\) for every \(s \in [0,1]\), which, together with the assumption \(f \equiv 0\) in \({\mathcal {K}}\), yield (44), as desired. \(\square \)

In what follows, we shall make use of the following, well-known result.

Remark A.1

Let \({\mathcal {K}}\) be a convex, compact subset of \({\mathbb R}^{2d}\) and let \(f \in {\mathcal {C}}^2({\mathbb R}^{2d};{\mathbb R})\). Then, there exists \(C_{{\mathcal {K}},f} \in {\mathbb R}\) depending only on \({\mathcal {K}}\) and f such that

$$\begin{aligned} f((1-s)x_0 + sx_1) \le (1-s)f(x_0) + s f(x_1) + C_{{\mathcal {K}},f} s(1-s) \Vert x_0 - x_1\Vert ^2, \end{aligned}$$
(45)

for every \(x_0,x_1 \in {\mathbb R}^{2d}\) and \(s \in [0,1]\).

We now prove the semiconvexity of \(\hat{\mathbb {H}}^1_c\).

Lemma A.2

Let \(\nu _0,\nu _1 \in {\mathcal P}_c({\mathbb R}^{2d})\) and let \(\rho \in \varPi (\nu _0,\nu _1)\). Then, there exists \(C \in {\mathbb R}\) independent of \(\nu _0\) and \(\nu _1\) for which

$$\begin{aligned} \hat{\mathbb {H}}^1_c(((1-s)\pi _1 + s\pi _2)_{\#}\rho ) \le (1-s) \hat{\mathbb {H}}^1_c(\nu _0) + s \hat{\mathbb {H}}^1_c(\nu _1) + Cs(1-s) {\mathcal W}^2_2(\nu _0, \nu _1) \end{aligned}$$

holds for every \(s \in [0,1]\).

Proof

We may assume \(\mathrm {supp}(\nu _0),\mathrm {supp}(\nu _1) \subseteq {\mathcal {K}}\) for some convex and compact set \({\mathcal {K}} \subset {\mathbb R}^{2d}\), otherwise the inequality is trivial. Hence, from Lemma A.1, it follows \(\mathrm {supp}(\nu _s) \subseteq {\mathcal {K}}\) for every \(s \in [0,1]\). But then, since \(\hat{{\mathcal {F}}}\) and \(\hat{{\mathcal {G}}}\) are both \({\mathcal {C}}^2\), the result follows as in [33, Proposition 9.3.2, Proposition 9.3.5]. \(\square \)

Corollary A.1

Let \(\hat{\omega } \in {\mathcal {C}}^2({\mathbb R}^{2d};{\mathbb R}^{d})\), \(\nu _0,\nu _1 \in {\mathcal P}_c({\mathbb R}^{2d})\), \(\rho \in \varPi (\nu _0,\nu _1)\) and define \(\nu _s\) as in (42) for \(s \in [0,1]\). If we set

$$\begin{aligned} \xi _s = \int _{{\mathbb R}^{2d}} \hat{\omega } d \nu _s, \end{aligned}$$
(46)

then

$$\begin{aligned} \Vert \xi _s - (1-s)\xi _0 - s\xi _1\Vert \le C s (1-s) {\mathcal W}^2_2(\nu _0,\nu _1), \end{aligned}$$

for all \(s\in [0,1]\), where C is independent of \(\nu _0\) and \(\nu _1\).

Proof

Follows from Lemma A.2 applied first to the functions \(\hat{{\mathcal {F}}} \equiv 0\) and \(\hat{{\mathcal {G}}} \equiv \hat{\omega }\), and then to \(\hat{{\mathcal {F}}} \equiv 0\) and \(\hat{{\mathcal {G}}} \equiv -\hat{\omega }\). \(\square \)

The semiconvexity of \(\hat{\mathbb {H}}^2_c\) will be deduced as a corollary of the following estimate.

Lemma A.3

Suppose that \(\hat{\ell } \in {\mathcal {C}}^2({\mathbb R}^{2d}\times {\mathbb R}^d;{\mathbb R})\), let \(z_0,z_1 \in {\mathcal {K}}\) and set \(z_s = (1-s)z_0 + sz_1\) for all \(s \in [0,1]\). Furthermore, let \(\nu _0,\nu _1 \in {\mathcal P}_c({\mathbb R}^{2d})\), \(\rho \in \varPi (\nu _0,\nu _1)\) and define \(\nu _s\) and \(\xi _s\) as in (42) and (46) for \(s \in [0,1]\). Then, for all \(s \in [0,1]\), it holds

$$\begin{aligned} \hat{\ell }(z_s,\xi _s)&\le (1-s) \hat{\ell }(z_0,\xi _0) + s \hat{\ell }(z_1,\xi _1) + C_{{\mathcal {K}},\hat{\ell },\hat{\omega }} s(1-s) {\mathcal W}^2_2(\nu _0,\nu _1) \\&\quad + C_{{\mathcal {K}},\hat{\ell },\hat{\omega }} s(1-s) \Vert z_0 - z_1\Vert ^2, \end{aligned}$$

for some constant \(C_{{\mathcal {K}},\hat{\ell },\hat{\omega }}\) depending only on \({\mathcal {K}},\hat{\ell }\) and \(\hat{\omega }\).

Proof

Since \({\mathcal {K}}\) is compact, \(z_s \in {\mathcal {K}}\) for all \(s \in [0,1]\). Moreover, \((1-s)\xi _0 + s\xi _1 \in {\mathcal {K}}'\) for all \(s \in [0,1]\), for some convex and compact set \({\mathcal {K}}' \subset {\mathbb R}^d\). Notice that from (45) follows

$$\begin{aligned} \begin{aligned} \hat{\ell }(z_s,(1-s)\xi _0 + s\xi _1)&\le (1-s) \hat{\ell }(z_0,\xi _0) + s \hat{\ell }(z_1,\xi _1)\\&\quad + C_{{\mathcal {K}},{\mathcal {K}}'}s(1-s)\left( \Vert z_0 - z_1\Vert ^2 + \Vert \xi _0 - \xi _1\Vert ^2 \right) , \end{aligned} \end{aligned}$$
(47)

and from the definition of \(\xi _s\) and Jensen’s inequality, we get

$$\begin{aligned} \Vert \xi _0 - \xi _1\Vert ^2&\le {\mathrm {Lip}}_{{\mathcal {K}}}(\omega ) {\mathcal W}^2_1(\nu _0,\nu _1) \le {\mathrm {Lip}}_{{\mathcal {K}}}(\omega ) {\mathcal W}^2_2(\nu _0,\nu _1). \end{aligned}$$
(48)

Moreover, for every \(s\in [0,1]\) it holds

$$\begin{aligned} \begin{aligned} \Vert \hat{\ell }(z_s,\xi _s) - \hat{\ell }(z_s,(1-s)\xi _0 + s\xi _1)\Vert&\le {\mathrm {Lip}}_{{\mathcal {K}} \times {\mathcal {K}}'} \Vert \xi _s - (1-s)\xi _0 - s\xi _1\Vert \\&\le {\mathrm {Lip}}_{{\mathcal {K}} \times {\mathcal {K}}'} s(1-s) C {\mathcal W}^2_2(\nu _0,\nu _1). \end{aligned} \end{aligned}$$
(49)

Hence, for every \(s\in [0,1]\), using (47), (48) and (49), we get

$$\begin{aligned} \hat{\ell }(z_s,\xi _s)&= \hat{\ell }(z_s,\xi _s) - \hat{\ell }(z_s,(1-s)\xi _0 + s\xi _1) + \hat{\ell }(z_s,(1-s)\xi _0 + s\xi _1) \\&\le (1-s) \hat{\ell }(z_0,\xi _0) + s \hat{\ell }(z_1,\xi _1) + C_{{\mathcal {K}},\hat{\ell },\hat{\omega }} s(1-s) {\mathcal W}^2_2(\nu _0,\nu _1) \\&\quad + C_{{\mathcal {K}},\hat{\ell },\hat{\omega }} s(1-s) \Vert z_0 - z_1\Vert ^2. \end{aligned}$$

This concludes the proof. \(\square \)

Corollary A.2

Let \(\nu _0,\nu _1 \in {\mathcal P}_c({\mathbb R}^{2d})\) and \(\rho \in \varPi _o(\nu _0,\nu _1)\). Then, there exists \(C \in {\mathbb R}\) independent of \(\nu _0\) and \(\nu _1\) for which

$$\begin{aligned} \hat{\mathbb {H}}^2_c(((1-s)\pi _1 + s\pi _2)_{\#}\rho ) \le (1-s) \hat{\mathbb {H}}^2_c(\nu _0) + s \hat{\mathbb {H}}^2_c(\nu _1) + Cs(1-s) {\mathcal W}^2_2(\nu _0, \nu _1) \end{aligned}$$

holds for every \(s \in [0,1]\).

Proof

Notice that, by Lemma A.1, \(\hat{H}^2_c(\nu _s)\) can be rewritten as

$$\begin{aligned} \hat{\mathbb {H}}^2_c(\nu _s) = \int _{{\mathcal {K}} \times {\mathcal {K}}} \hat{\ell }(z_s,\xi _s)\hbox {d}\rho (z_0,z_1), \end{aligned}$$

Furthermore, since \(\rho \in \varPi _o(\nu _0,\nu _1)\) it holds

$$\begin{aligned} \int _{{\mathcal {K}} \times {\mathcal {K}}} \Vert z_0 - z_1\Vert ^2 \hbox {d}\rho (z_0,z_1) = \int _{{\mathbb R}^{4d}} \Vert z_0 - z_1\Vert ^2 \hbox {d}\rho (z_0,z_1) = {\mathcal W}^2_2(\nu _0,\nu _1), \end{aligned}$$

the thesis follows from Lemma A.3. \(\square \)

Proposition A.1

The functional \(\mathbb {H}_c\) is semiconvex along geodesics.

Proof

Follows directly from Lemma A.2 and Corollary A.2, by noticing that \(\mathbb {H}_c(\nu ) = \hat{\mathbb {H}}_c^1(\nu ) + \hat{\mathbb {H}}_c^2(\nu )\) for \(\hat{{\mathcal {F}}} = {\mathcal {F}}\), \(\hat{{\mathcal {G}}} = {\mathcal {G}}\), \(\hat{\ell } = -\ell \circ (\pi _1, Id )\), \(\hat{\omega } = \omega \circ \pi _1\) and \({\mathcal {K}} = cl \!\left( {B(0,R_T)}\right) \). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bongini, M., Fornasier, M., Rossi, F. et al. Mean-Field Pontryagin Maximum Principle. J Optim Theory Appl 175, 1–38 (2017). https://doi.org/10.1007/s10957-017-1149-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-017-1149-5

Keywords

Mathematics Subject Classification

Navigation