Skip to main content
Log in

Legendre’s Necessary Condition for Fractional Bolza Functionals with Mixed Initial/Final Constraints

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

Abstract

The present work was primarily motivated by our findings in the literature of some flaws within the proof of the second-order Legendre necessary optimality condition for fractional calculus of variations problems. Therefore, we were eager to elaborate a correct proof and it turns out that this goal is highly nontrivial, especially when considering final constraints. This paper is the result of our reflections on this subject. Precisely, we consider here a constrained minimization problem of a general Bolza functional that depends on a Caputo fractional derivative of order \(0 < \alpha \le 1\) and on a Riemann–Liouville fractional integral of order \(\beta > 0\), the constraint set describing general mixed initial/final constraints. The main contribution of our work is to derive corresponding first- and second-order necessary optimality conditions, namely the Euler–Lagrange equation, the transversality conditions and, of course, the Legendre condition. A detailed discussion is provided on the obstructions encountered with the classical strategy, while the new proof that we propose here is based on the Ekeland variational principle. Furthermore, we underline that some subsidiary contributions are provided all along the paper. In particular, we prove an independent and intrinsic result of fractional calculus stating that it does not exist a nontrivial function which is, together with its Caputo fractional derivative of order \(0< \alpha <1\), compactly supported. Moreover, we also discuss some evidences claiming that Riemann–Liouville fractional integrals should be considered in the formulation of fractional calculus of variations problems in order to preserve the existence of solutions.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. As a first step toward the open challenge of proving a fractional version of the second-order Legendre necessary optimality condition, the cost functional considered in [47] is the most basic one, containing (only) a classical Lagrange cost depending on a fractional derivative of order \(0 < \alpha \le 1\), under fixed initial/final conditions.

References

  1. Agrachev, A.A., Sachkov, Y.L.: Control Theory from the Geometric Viewpoint, Encyclopaedia of Mathematical Sciences, vol. 87. Springer, Berlin (2004)

    Book  MATH  Google Scholar 

  2. Agrawal, O.P.: Formulation of Euler-Lagrange equations for fractional variational problems. J. Math. Anal. Appl. 272(1), 368–379 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  3. Agrawal, O.P.: A general formulation and solution scheme for fractional optimal control problems. Nonlinear Dyn. 38(1–4), 323–337 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  4. Agrawal, O.P., Defterli, O., Baleanu, D.: Fractional optimal control problems with several state and control variables. J. Vib. Control 16(13), 1967–1976 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  5. Almeida, R.: Variational problems involving a Caputo-type fractional derivative. J. Optim. Theory Appl. 174(1), 276–294 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  6. Almeida, R.: Optimality conditions for fractional variational problems with free terminal time. Discrete Contin. Dyn. Syst. Ser. S 11(1), 1–19 (2018)

    MathSciNet  MATH  Google Scholar 

  7. Almeida, R., Malinowska, A.B., Torres, D.F.M.: A fractional calculus of variations for multiple integrals with application to vibrating string. J. Math. Phys. 51(3), 033503 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  8. Almeida, R., Morgado, M.L.: The Euler-Lagrange and Legendre equations for functionals involving distributed-order fractional derivatives. Appl. Math. Comput. 331, 394–403 (2018)

    MathSciNet  MATH  Google Scholar 

  9. Almeida, R., Torres, D.F.M.: Necessary and sufficient conditions for the fractional calculus of variations with Caputo derivatives. Commun. Nonlinear Sci. Numer. Simul. 16(3), 1490–1500 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  10. Almeida, R., Torres, D.F.M.: A discrete method to solve fractional optimal control problems. Nonlinear Dyn. 80(4), 1811–1816 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Atanacković, T.M., et al.: Variational problems with fractional derivatives: invariance conditions and Noether’s theorem. Nonlinear Anal. 71(5–6), 1504–1517 (2009)

  12. Bagley, R.L., Torvik, P.J.: A theoretical basis for the application of fractional calculus in viscoelasticity. J. Rheol. 27, 201–210 (1983)

    Article  MATH  Google Scholar 

  13. Bagley, R.L., Torvik, P.J.: On the fractional calculus model of viscoelasticity behavior. J. Rheol. 30, 133–155 (1986)

    Article  MATH  Google Scholar 

  14. Bastos, N.R.O., Ferreira, R.A.C., Torres, D.F.M.: Necessary optimality conditions for fractional difference problems of the calculus of variations. Discrete Contin. Dyn. Syst. 29(2), 417–437 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  15. Bergounioux, M., Bourdin, L.: Pontryagin maximum principle for general Caputo fractional optimal control problems with Bolza cost and terminal constraints. ESAIM Control Optim. Calc. Var. 26, 38 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  16. Bourdin, L.: Existence of a weak solution for fractional Euler-Lagrange equations. J. Math. Anal. Appl. 399(1), 239–251 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  17. Bourdin, L., Cresson, J., Greff, I.: A continuous/discrete fractional Noether’s theorem. Commun. Nonlinear Sci. Numer. Simul. 18(4), 878–887 (2013)

  18. Bourdin, L., Idczak, D.: A fractional fundamental lemma and a fractional integration by parts formula–Applications to critical points of Bolza functionals and to linear boundary value problems. Adv. Differ. Equ. 20(3–4), 213–232 (2015)

    MathSciNet  MATH  Google Scholar 

  19. Bourdin, L., Odzijewicz, T., Torres, D.F.M.: Existence of minimizers for fractional variational problems containing Caputo derivatives. Adv. Dyn. Syst. Appl. 8(1), 3–12 (2013)

    MathSciNet  Google Scholar 

  20. Bourdin, L., Odzijewicz, T., Torres, D.F.M.: Existence of minimizers for generalized Lagrangian functionals and a necessary optimality condition–application to fractional variational problems. Differ. Integral Equ. 27(7–8), 743–766 (2014)

    MathSciNet  MATH  Google Scholar 

  21. Bressan, A., Piccoli, B.: Introduction to the mathematical theory of control. AIMS Series on Applied Mathematics, 2, American Institute of Mathematical Sciences. AIMS), Springfield, MO (2007)

  22. Brezis, H.: Functional analysis. Universitext, Springer, New York, Sobolev spaces and partial differential equations (2011)

  23. Bullo, F., Lewis, A.D.: Geometric Control of Mechanical Systems, Texts in Applied Mathematics, vol. 49. Springer, New York (2005)

    Book  Google Scholar 

  24. Comte, F.: Opérateurs fractionnaires en économétrie et en finance. Prépublication MAP5 (2001)

  25. Cong, N.D., Tuan, H.T.: Generation of nonlocal fractional dynamical systems by fractional differential equations. J. Integral Equ. Appl. 29(4), 585–608 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  26. Cresson, J.: Fractional embedding of differential operators and Lagrangian systems. J. Math. Phys. 48(3), 033504 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  27. Cresson, J., Inizan, P.: Variational formulations of differential equations and asymmetric fractional embedding. J. Math. Anal. Appl. 385(2), 975–997 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  28. Diethelm, K., Ford, N.J.: Volterra integral equations and fractional calculus: do neighboring solutions intersect? J. Integral Equ. Appl. 24(1), 25–37 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  29. Ekeland, I.: On the variational principle. J. Math. Anal. Appl. 47, 324–353 (1974)

    Article  MathSciNet  MATH  Google Scholar 

  30. Ferreira, R.A.C.: Fractional calculus of variations: a novel way to look at it. Fract. Calc. Appl. Anal. 22(4), 1133–1144 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  31. Ferreira, R.A.C., Malinowska, A.B.: A counterexample to a Frederico–Torres fractional Noether-type theorem. J. Math. Anal. Appl. 429(2), 1370–1373 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  32. Ferreira, R.A.C., Torres, D.F.M.: Fractional \(h\)-difference equations arising from the calculus of variations. Appl. Anal. Discrete Math. 5(1), 110–121 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  33. Frederico, G.S.F., Torres, D.F.M.: Fractional Noether’s theorem in the Riesz-Caputo sense. Appl. Math. Comput. 217(3), 1023–1033 (2010)

  34. Frederico, G.S.F., Torres, D.F.M.: Fractional isoperimetric Noether’s theorem in the Riemann-Liouville sense. Rep. Math. Phys. 71(3), 291–304 (2013)

  35. Gelfand, I.M., Fomin, S.V.: Calculus of variations, revised English edition translated and edited by Richard A. Silverman, Prentice-Hall Inc, Englewood Cliffs, NJ (1963)

  36. Gerolymatou, E., Vardoulakis, I., Hilfer, R.: Modelling infiltration by means of a nonlinear fractional diffusion model. J. Phys. D Appl. Phys. 39, 4104–4110 (2006)

    Article  Google Scholar 

  37. Glöckle, W.G., Nonnenmacher, T.F.: A fractional calculus approach to self-similar protein dynamics. Biophys. J . 68, 46–53 (1995)

    Article  Google Scholar 

  38. Guo, T.L.: The necessary conditions of fractional optimal control in the sense of Caputo. J. Optim. Theory Appl. 156(1), 115–126 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  39. Hélie, T., Matignon, D.: Diffusive representations for the analysis and simulation of flared acoustic pipes with visco-thermal losses. Math. Models Methods Appl. Sci. 16(4), 503–536 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  40. Herzallah, M.A.E., Baleanu, D.: Fractional Euler–Lagrange equations revisited. Nonlinear Dyn. 69(3), 977–982 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  41. Hestenes, M.R.: Calculus of variations and optimal control theory. Robert E. Krieger Publishing Co., Inc., Huntington, N.Y. (1980). Corrected reprint of the 1966 original

  42. Hilfer, R.: Applications of Fractional Calculus in Physics. World Scientific, River Edge (2000)

    Book  MATH  Google Scholar 

  43. Hilfer, R.: Fractional calculus and regular variation in thermodynamics. In: Applications of fractional calculus in physics, pp. 429–463. World Sci. Publ., River Edge, NJ (2000)

  44. Jelicic, Z.D., Petrovacki, N.: Optimality conditions and a solution scheme for fractional optimal control problems. Struct. Multidiscip. Optim. 38(6), 571–581 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  45. Kilbas, A.A., Srivastava, H.M., Trujillo, J.J.: Theory and Applications of Fractional Differential Equations, North-Holland Mathematics Studies, vol. 204. Elsevier Science B.V, Amsterdam (2006)

    Google Scholar 

  46. Klimek, M.: Solutions of Euler–Lagrange equations in fractional mechanics. In: XXVI Workshop on Geometrical Methods in Physics, 73–78, AIP Conf. Proc., 956, Amer. Inst. Phys., Melville, NY

  47. Lazo, M.J., Torres, D.F.M.: The Legendre condition of the fractional calculus of variations. Optimization 63(8), 1157–1165 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  48. Lévy, P.: L’addition des variables aléatoires définies sur un circonférence. Bull. Soc. Math. France 67, 1–41 (1939)

  49. Liberzon, D.: Calculus of Variations and Optimal Control Theory. Princeton University Press, Princeton, NJ (2012)

    Book  MATH  Google Scholar 

  50. Magin, R.L.: Fractional calculus models of complex dynamics in biological tissues. Comput. Math. Appl. 59(5), 1586–1593 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  51. Metzler, R., Klafter, J.: The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep. 339, no. 1, 77 pp (2000)

  52. Oldham, K.B., Spanier, J.: The replacement of Fick’s laws by a formulation involving semidifferentiation. J. Electroanal. Chem. 26, 331–341 (1970)

  53. Pfitzenreiter, T.: A physical basis for fractional derivatives in constitutive equations. Z. Angew. Math. Mech. 84(4), 284–287 (2004)

    Article  MATH  Google Scholar 

  54. Pontryagin, L. S. et al.: The mathematical theory of optimal processes, translated from the Russian by K. N. Trirogoff; edited by L. W. Neustadt. Interscience Publishers John Wiley & Sons, Inc. New York (1962)

  55. Pooseh, S., Almeida, R., Torres, D.F.M.: Fractional order optimal control problems with free terminal time. J. Ind. Manag. Optim. 10(2), 363–381 (2014)

    MathSciNet  MATH  Google Scholar 

  56. Riewe, F.: Nonconservative Lagrangian and Hamiltonian mechanics. Phys. Rev. E 53(2), 1890–1899 (1996)

    Article  MathSciNet  Google Scholar 

  57. Sabatier, J., Agrawal, O.P., Tenreiro Machado, J.A. (eds.): Advances in fractional calculus. Springer, Dordrecht (2007)

  58. Samko, S.G., Kilbas, A.A., Marichev, O.I.: Fractional integrals and derivatives, translated from the 1987 Russian original. Gordon and Breach Science Publishers, Yverdon (1993)

  59. Schättler, H., Ledzewicz, U.: Geometric Optimal Control, Interdisciplinary Applied Mathematics, vol. 38. Springer, New York (2012)

    Book  MATH  Google Scholar 

  60. Song, C.-J., Zhang, Y.: Noether symmetry and conserved quantity for fractional Birkhoffian mechanics and its applications. Fract. Calc. Appl. Anal. 21(2), 509–526 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  61. van Brunt, B.: The Calculus of Variations. Universitext. Springer, New York (2004)

    Book  MATH  Google Scholar 

  62. Zaslavsky, G.M.: Chaos, fractional kinetics, and anomalous transport. Phys. Rep. 371(6), 461–580 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  63. Zaslavsky, G.M.: Hamiltonian chaos and fractional dynamics, reprint of the 2005 original. Oxford University Press, Oxford (2008)

  64. Zelikin,. M. I.: Control theory and optimization. I, a translation of Homogeneous spaces and the Riccati equation in the calculus of variations (Russian), “Faktorial”, Moscow, 1998, translation by S. A. Vakhrameev, Encyclopaedia of Mathematical Sciences, 86, Springer-Verlag, Berlin (2000)

  65. Zoia, A., Néell, M.-C., Cortis, A.: Continuous-time random-walk model of transport in variably saturated heterogeneous porous media. Phys. Rev. E, 81(3) (2010)

  66. Zoia, A., Néell, M.-C., Joelson, M.: Mass transport subject to time-dependent flow with nonuniform sorption in porous media. Phys. Rev. E 80,(2009)

Download references

Acknowledgements

Rui A. C. Ferreira was supported by the “Fundação para a Ciência e a Tecnologia (FCT)” through the program “Stimulus of Scientific Employment, Individual Support-2017 Call” with reference CEECIND/00640/2017.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rui A. C. Ferreira.

Additional information

Communicated by Nikolai Osmolovskii.

Publisher's Note

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

Appendices

A Proof of Proposition 3.3

In this section we provide a detailed proof of Proposition 3.3. In order to accomplish it, recall that for all \(\rho \in {\mathbb {R}}\), there exists a real sequence \((\rho _k)_{k \in {\mathbb {N}}}\) such that

$$\begin{aligned} (1-\xi )^\rho = \sum _{k \in {\mathbb {N}}} \rho _k \xi ^k, \end{aligned}$$
(9)

for all \(\xi \in (-1,1)\). We also recall that the series convergence is uniform on all compact subsets included in \((-1,1)\). Moreover, if \(\rho \in {\mathbb {R}}\backslash {\mathbb {N}}\), the terms \(\rho _k\) are all different from zero.

Lemma A.1

Let \(c > a\) be a real number and \(u \in \mathrm {L}^1([a,c],{\mathbb {R}})\). If

$$\begin{aligned} \int _a^c (s-a)^k u(s) \, ds = 0, \end{aligned}$$

for all \(k \in {\mathbb {N}}\), then \(u=0\).

Proof

This result easily follows from the density of polynomial functions in \(\mathrm {C}([a,c],{\mathbb {R}})\) and from [22, Corollary 4.24 p.110]. \(\square \)

Lemma A.2

Let \(c > a\) be a real number and \(u \in \mathrm {L}^1([a,c],{\mathbb {R}})\). Let us consider the function

$$\begin{aligned}{}\begin{array}[t]{lrcl}\varPsi :&{}(c,+\infty ) &{}\longrightarrow &{}{\mathbb {R}}\\ {} &{}t&{} \longmapsto &{}\varPsi (t) := \displaystyle \int _a^c (t-s)^\mu u(s) \, ds , \end{array} \end{aligned}$$

where \(\mu \in {\mathbb {R}}\backslash {\mathbb {N}}\). If the function \(\varPsi \) is polynomial over a subinterval \(I \subset (c,+\infty )\) with a nonempty interior, then \(u = 0\).

Proof

Without loss of generality, we can assume that \(I = [c_1,c_2]\) is compact with \(c< c_1 < c_2\). From the LDC theorem, one can easily see that \(\varPsi \) is of class \(\mathrm {C}^\infty \) with

$$\begin{aligned} \varPsi ^{(r)}(t) = \mu (\mu - 1) \ldots (\mu - r+1) \displaystyle \int _a^c (t-s)^{\mu -r} u(s) \, ds, \end{aligned}$$

for all \(t > c\) and all \(r \ge 1\). In the sequel we fix some \(r \ge 1\) larger than the degree of \(\varPsi \) (polynomial over I) plus one. Since \(\mu \in {\mathbb {R}}\backslash {\mathbb {N}}\), we get that

$$\begin{aligned} \displaystyle \int _a^c (t-s)^{\mu -r} u(s) \, ds = 0, \end{aligned}$$

and thus

$$\begin{aligned} \displaystyle \int _a^c \left( 1 - \dfrac{s-a}{t-a} \right) ^{\mu -r} u(s) \, ds = 0, \end{aligned}$$

for all \(t \in I\). From Equality (9) (with \(\rho := \mu -r \in {\mathbb {R}}\backslash {\mathbb {N}}\)) and the uniform convergence of the power series (since \(0 \le \frac{s-a}{t-a} \le \frac{c-a}{c_1 - c } <1\) for all \((t,s) \in I \times [a,c]\)), we get that

$$\begin{aligned} \displaystyle \sum _{k \in {\mathbb {N}}} \dfrac{\rho _k}{(t-a)^k} \int _a^c (s-a)^k u(s) \, ds = 0, \end{aligned}$$

for all \(t \in I\). Finally, using the change of variable \(T=\frac{1}{t-a}\), we can write that \( \sum _{k \in {\mathbb {N}}} \lambda _k T^k = 0\) for all \(T \in [\frac{1}{c_2 - a},\frac{1}{c_1-a}]\), where \( \lambda _k := \rho _k \int _a^c (s-a)^k u(s) \, ds \) for all \(k \in {\mathbb {N}}\). Since the zeros of a nonzero power series are isolated, we deduce that \(\lambda _k = 0\) for all \(k \in {\mathbb {N}}\). Since all \(\rho _k\) are different from zero, Lemma A.1 concludes the proof. \(\square \)

We are now in a position to prove Proposition 3.3. Actually we can even prove the more general following statement.

Proposition A.1

Let \(0< \alpha <1\) and \(x \in {}_{\mathrm {c}} \mathrm {AC}^\alpha _{a+}\). If there exist two real numbers \(a \le c < d \le b\) such that:

  1. (i)

    x is polynomial over [cd];

  2. (ii)

    \({}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x](t) = 0_{{\mathbb {R}}^n}\) for almost every \(t \in [c,d]\);

then x is constant over [ad].

Proof

Without loss of generality, we assume in this proof that \(n=1\). From Proposition 2.5, it holds that \( x(t) = x(a) + \mathrm {I}^\alpha _{a+}[ {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x] ](t)\) for all \(t \in [a,b]\), and thus

$$\begin{aligned} x(t) = x(a) + \displaystyle \int _a^c \dfrac{(t-s)^{\alpha -1}}{\varGamma (\alpha )} {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x](s) \, ds, \end{aligned}$$

for all \(t \in [c,d]\). Let us denote by \(u \in \mathrm {L}^1([a,c],{\mathbb {R}})\) the restriction of \({}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x]\) over [ac]. From the hypothesis, one can easily deduce that the function

$$\begin{aligned}{}\begin{array}[t]{lrcl}\varPsi :&{}(c,+\infty ) &{}\longrightarrow &{}{\mathbb {R}}\\ {} &{}t&{} \longmapsto &{}\varPsi (t) := \displaystyle \int _{a}^{c} (t-s)^{\alpha -1} u(s) \, ds , \end{array} \end{aligned}$$

is polynomial over (cd]. From Lemma A.2, we deduce that \(u=0\) and thus \({}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x](t) = 0\) for almost every \(t \in [a,d]\). We deduce that \(x(t)=x(a)\) for all \(t \in [a,d]\) which completes the proof. \(\square \)

B Proof of Theorem 3.2

Our strategy of proof is based on the Ekeland variational principle [29]. Let us enunciate hereafter a simplified version (but sufficient for our purposes).

Proposition B.1

(Ekeland variational principle) Let \((\mathrm {E},\mathrm {d}_\mathrm {E})\) be a complete metric space and \(\mathrm {J}: \mathrm {E}\rightarrow {\mathbb {R}}^+\) be a continuous nonnegative map. Let \(\varepsilon >0\) and \(\lambda \in \mathrm {E}\) such that \(\mathrm {J}(\lambda ) = \varepsilon \). Then there exists \(\lambda _\varepsilon \in \mathrm {E}\) such that \(\mathrm {d}_\mathrm {E}( \lambda _\varepsilon , \lambda ) \le \sqrt{\varepsilon }\) and \(-\sqrt{\varepsilon } \; \mathrm {d}_\mathrm {E}( \lambda ' , \lambda _\varepsilon ) \le \mathrm {J}(\lambda ')-\mathrm {J}(\lambda _\varepsilon )\) for all \(\lambda ' \in \mathrm {E}\).

In Section B.1 we give some recalls about convex analysis. In Section B.2 we investigate the sensitivity analysis of the Bolza functional \({\mathcal {L}}\). Finally the proof of Theorem 3.2 is detailed in Section B.3 by applying the Ekeland variational principle on a penalized functional.

1.1 B.1 Basics of Convex Analysis

Let \(\mathrm {d}_\mathrm {S}: {\mathbb {R}}^j \rightarrow {\mathbb {R}}_{+}\) denote the standard distance function to the nonempty closed convex subset \(\mathrm {S}\subset {\mathbb {R}}^j\) defined by \(\mathrm {d}_\mathrm {S}(z):= \inf _{z' \in \mathrm {S}} \Vert z-z' \Vert _{{\mathbb {R}}^j}\) for all \(z \in {\mathbb {R}}^j\). We recall that for all \(z \in {\mathbb {R}}^j\), there exists a unique element \(\mathrm {P}_\mathrm {S}(z) \in \mathrm {S}\) (called the projection of z onto \(\mathrm {S}\)) such that \(\mathrm {d}_\mathrm {S}(z)=\Vert z-\mathrm {P}_\mathrm {S}(z) \Vert _{{\mathbb {R}}^j}\). It can easily be shown that the map \(\mathrm {P}_\mathrm {S}:{\mathbb {R}}^j \rightarrow \mathrm {S}\) is 1-Lipschitz continuous. Moreover it holds that \(( z- \mathrm {P}_\mathrm {S}(z)) \cdot (z'-\mathrm {P}_\mathrm {S}(z)) \le 0\) for all \(z'\in \mathrm {S}\), that is, \(z-\mathrm {P}_\mathrm {S}(z)\in \mathrm {N}_\mathrm {S}[\mathrm {P}_\mathrm {S}(z)]\) for all \(z \in {\mathbb {R}}^j\). Let us recall the two following required lemmas, whose proofs are detailed for the reader’s convenience.

Lemma B.1

Let \((z_k)_{k\in {\mathbb {N}}}\) be a sequence in \({\mathbb {R}}^j\) converging to some point \(z\in \mathrm {S}\) and let \((\zeta _k)_{k\in {\mathbb {N}}}\) be a positive real sequence. If \(\zeta _k(z_k-\mathrm {P}_\mathrm {S}(z_k))\) converges to some \({\overline{z}} \in {\mathbb {R}}^j\), then \({\overline{z}} \in \mathrm {N}_\mathrm {S}[z]\).

Proof

Since \(z_k - \mathrm {P}_\mathrm {S}(z_k) \in \mathrm {N}_\mathrm {S}[\mathrm {P}_\mathrm {S}(z_k)]\) and \(\zeta _k > 0\) for all \(k \in {\mathbb {N}}\), we obtain that \( \zeta _k (z_k - \mathrm {P}_\mathrm {S}(z_k)) \cdot (z' - \mathrm {P}_\mathrm {S}(z_k)) \le 0\) for all \(z' \in \mathrm {S}\) and all \(k \in {\mathbb {N}}\). Passing to the limit \(k \rightarrow \infty \), and since \(z \in \mathrm {S}\), we obtain that \( {\overline{z}} \cdot (z' - z) \le 0\) for all \(z' \in \mathrm {S}\) which exactly means that \({\overline{z}} \in \mathrm {N}_\mathrm {S}[z]\). \(\square \)

Lemma B.2

The map

$$\begin{aligned}{}\begin{array}[t]{lrcl}\mathrm {d}^2_\mathrm {S} :&{}{\mathbb {R}}^j &{}\longrightarrow &{}{\mathbb {R}}_+\\ {} &{}z&{} \longmapsto &{}\mathrm {d}^2_\mathrm {S}(z) := \mathrm {d}_\mathrm {S}(z)^2 , \end{array} \end{aligned}$$

is Fréchet-differentiable on \({\mathbb {R}}^j\), and its differential \({\mathcal {D}}\mathrm {d}^2_\mathrm {S}(z)\) at every \(z \in {\mathbb {R}}^j\) can be expressed as

$$\begin{aligned} {\mathcal {D}}\mathrm {d}^2_\mathrm {S}(z)(z') = 2 ( z-\mathrm {P}_\mathrm {S}(z) ) \cdot z' , \end{aligned}$$

for all \(z' \in {\mathbb {R}}^j\).

Proof

Let \(z \in {\mathbb {R}}^j\) and let us prove that \(\mathrm {d}^2_\mathrm {S}\) is Fréchet-differentiable at z with \( {\mathcal {D}}\mathrm {d}^2_\mathrm {S}(z)(z') = 2 ( z-\mathrm {P}_\mathrm {S}(z) ) \cdot z'\). One has

$$\begin{aligned} \mathrm {d}^2_\mathrm {S}(z+z') - \mathrm {d}^2_\mathrm {S}(z) \le \Vert z+z' - \mathrm {P}_\mathrm {S}(z) \Vert ^2 - \Vert z - \mathrm {P}_\mathrm {S}(z) \Vert ^2 = 2 ( z-\mathrm {P}_\mathrm {S}(z) ) \cdot z' + \Vert z' \Vert ^2, \end{aligned}$$

and, from Cauchy–Schwarz inequality and 1-Lipschitz continuity of \(\mathrm {P}_\mathrm {S}\), one gets

$$\begin{aligned}&\mathrm {d}^2_\mathrm {S}(z) - \mathrm {d}^2_\mathrm {S}(z+z') \le \Vert z - \mathrm {P}_\mathrm {S}(z+z') \Vert ^2 - \Vert z+z' - \mathrm {P}_\mathrm {S}(z+z') \Vert ^2 \\&\quad = -2 ( z-\mathrm {P}_\mathrm {S}(z+z') ) \cdot z' - \Vert z' \Vert ^2 \\&\quad = -2 ( z-\mathrm {P}_\mathrm {S}(z) ) \cdot z' + 2 ( \mathrm {P}_\mathrm {S}(z+z')-\mathrm {P}_\mathrm {S}(z)) \cdot z' - \Vert z' \Vert ^2 \\&\quad \le -2 ( z-\mathrm {P}_\mathrm {S}(z) ) \cdot z' + \Vert z' \Vert ^2 , \end{aligned}$$

for all \(z' \in {\mathbb {R}}^j\). Using both inequalities, the proof is complete. \(\square \)

1.2 B.2 Sensitivity Analysis of the Bolza Functional

In the proof of Theorem 3.2 (see Section B.3), we denote:

  • by \(r_\alpha \) some real number satisfying \(r_\alpha > \frac{1}{\alpha }\) and by \(r'_\alpha := \frac{r_\alpha }{r_\alpha - 1}\) the classical conjugate of \(r_\alpha \) satisfying \(\frac{1}{r_\alpha } + \frac{1}{r'_\alpha } = 1\);

  • and, for all \((u,y) \in \mathrm {L}^\infty \times {\mathbb {R}}^n\), by \(x(\cdot ,u,y) \in {}_\mathrm {c}\mathrm {AC}^{\alpha ,\infty }_{a+}\) the function defined by

    $$\begin{aligned} x(t,u,y) := y + \mathrm {I}^\alpha _{a+}[u](t), \end{aligned}$$

    for all \(t \in [a,b]\);

  • and, for all \((u,y) \in \mathrm {L}^\infty \times {\mathbb {R}}^n\), by \({\mathcal {P}}(u,y)\) the set of Lebesgue points \(\tau \in (a,b)\) of both the functions u and \(L(x(\cdot ,u,y),u,\cdot )\).

Remark B.1

Note that \(r'_\alpha (\alpha -1) +1 > 0\).

Remark B.2

Let \(x \in \mathrm {L}^1\). From Proposition 2.5 and Remark 2.5, note that \(x \in {}_\mathrm {c}\mathrm {AC}^{\alpha ,\infty }_{a+}\) if and only if there exists \((u,y) \in \mathrm {L}^\infty \times {\mathbb {R}}^n\) such that \(x=x(\cdot ,u,y)\). In that case, the couple (uy) is unique and is given by \(u = {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x]\) and \(y=x(a)\).

Remark B.3

For all \((u,y) \in \mathrm {L}^\infty \times {\mathbb {R}}^n\), note that the set \({\mathcal {P}}(u,y)\) is of full measure in [ab].

We introduce the set

$$\begin{aligned} \mathrm {L}^\infty _R := \mathrm {L}^\infty \Big ( [a,b] , {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R ) \Big ), \end{aligned}$$

for all \(R \ge 0\), where \({\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R )\) denotes the standard closed ball of \({\mathbb {R}}^n\) centered at the origin \(0_{{\mathbb {R}}^n}\) with radius \(R \ge 0\). We endow the set \(\mathrm {L}^\infty _R \times {\mathbb {R}}^n\) with the distance

$$\begin{aligned} \mathrm {d}_{ \mathrm {L}^\infty _R \times {\mathbb {R}}^n } \Big ( (u_2,y_2) , (u_1 , y_1) \Big ) := \Vert u_2 - u_1 \Vert _{\mathrm {L}^1} + \Vert y_2 - y_1 \Vert _{{\mathbb {R}}^n}, \end{aligned}$$

for all \((u_1,y_1)\), \((u_2,y_2) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\).

Lemma B.3

Let \(R \ge 0\). The following assertions are true:

  1. (i)

    The metric space \(( \mathrm {L}^\infty _R \times {\mathbb {R}}^n , \mathrm {d}_{ \mathrm {L}^\infty _R \times {\mathbb {R}}^n } )\) is complete;

  2. (ii)

    The map \((u,y) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n \longmapsto x(\cdot ,u,y) \in \mathrm {C}\) is continuous;

  3. (iii)

    The map

    $$\begin{aligned}{}\begin{array}[t]{lrcl}\varPhi _R :&{}\mathrm {L}^\infty _R \times {\mathbb {R}}^n &{}\longrightarrow &{}{\mathbb {R}}\\ {} &{}(u,y)&{} \longmapsto &{}\varPhi _R (u,y) := {\mathcal {L}}(x(\cdot ,u,y)), \end{array} \end{aligned}$$

    is continuous.

Proof

The first item can be easily derived from the PCLDC theorem. Secondly, it can be proved from the classical Hölder inequality that

$$\begin{aligned}&\Vert x(t,u_2,y_2) - x(t,u_1,y_1) \Vert _{{\mathbb {R}}^n} \\&\quad \le \Vert y_2 - y_1 \Vert _{{\mathbb {R}}^n} +\dfrac{1}{\varGamma (\alpha )} \left( \dfrac{ R (b-a)^{r'_\alpha (\alpha -1) +1}}{r'_\alpha (\alpha -1) +1 } \right) ^{1/r'_\alpha } \Vert u_2 - u_1 \Vert ^{1/r_\alpha }_{\mathrm {L}^1} , \end{aligned}$$

for all \(t \in [a,b]\) and all \((u_1,y_1)\), \((u_2,y_2) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\), which concludes the proof of the second item. The third item can be derived by contradiction and by using the LDC and PCLDC theorems. \(\square \)

The rest of this section is devoted to the sensitivity analysis of the Bolza functional \(\varPhi _R\) under perturbations of the couple (uy) (see Propositions B.2 and B.3). Before coming to these points, we first introduce the following notion of needle-perturbation of u.

Definition B.1

(Needle-perturbation of u) Let \(R \ge 0\) and let \((u,y) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\). A needle-perturbation of u associated with \((\tau ,v) \in {\mathcal {P}}(u,y) \times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\) and \(0 < h \le b-\tau \) is the function \(u^{(\tau ,v)}(\cdot ,h) \in \mathrm {L}^\infty _R\) defined by

$$\begin{aligned} u^{(\tau ,v)}(t,h) := \left\{ \begin{array}{lcl} v &{} \text {if} &{} t \in [\tau ,\tau +h) , \\ u(t) &{} \text {if} &{} t \notin [\tau ,\tau +h) , \end{array} \right. \end{aligned}$$

for almost every \(t \in [a,b]\).

Lemma B.4

Let \(R \ge 0\) and let \((u,y) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\). It holds that

$$\begin{aligned} \left\| x \Big ( t ,u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y) \right\| _{{\mathbb {R}}^n} \le \dfrac{2R}{\varGamma (\alpha +1)} h^\alpha , \end{aligned}$$
(10)

for all \(t \in [a,b]\) and

$$\begin{aligned}&\left\| \dfrac{ x \Big ( t , u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y)}{h} - \dfrac{(t-\tau )^{\alpha -1}}{\varGamma (\alpha )} (v-u(\tau )) \right\| _{{\mathbb {R}}^n} \nonumber \\&\quad \le \dfrac{( t - \tau )^{\alpha -1}}{\varGamma (\alpha )} \left\| \dfrac{1}{h} \int _\tau ^{\tau +h} u(s) \; ds - u(\tau ) \right\| _{{\mathbb {R}}^n} \nonumber \\&\qquad +\dfrac{2R}{\varGamma (\alpha )} \Bigg ( \Big ( t - (\tau +h) \Big )^{\alpha -1} - (t-\tau )^{\alpha -1} \Bigg ) , \end{aligned}$$
(11)

for all \(t \in (\tau +h,b]\), all \((\tau ,v) \in {\mathcal {P}}(u,y) \times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\) and all \(0 < h \le b-\tau \).

Proof

Let \((\tau ,v) \in {\mathcal {P}}(u,y) \times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\) and \(0 < h \le b-\tau \) being fixed for the whole proof. It holds that

$$\begin{aligned}&x \Big ( t ,u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y) = \mathrm {I}^\alpha _{a+} [ u^{(\tau ,v)}(\cdot ,h) -u ](t) \\&\quad = \int _a^t \dfrac{(t-s)^{\alpha - 1}}{\varGamma (\alpha )} \Big ( u^{(\tau ,v)}(s,h) -u(s) \Big ) \; ds \\&\quad = \left\{ \begin{array}{lcl} 0_{{\mathbb {R}}^n} &{} \text {if} &{} a \le t < \tau , \\ \displaystyle \int _{\tau }^t \dfrac{(t-s)^{\alpha - 1}}{\varGamma (\alpha )} ( v -u(s) ) \; ds &{} \text {if} &{} \tau \le t \le \tau + h , \\ \displaystyle \int _{\tau }^{\tau + h} \dfrac{(t-s)^{\alpha - 1}}{\varGamma (\alpha )} ( v -u(s) ) \; ds &{} \text {if} &{} t > \tau + h , \end{array} \right. \end{aligned}$$

for all \(t \in [a,b]\). Since \(u \in \mathrm {L}^\infty _R\) and \(v \in {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\), we obtain that

$$\begin{aligned}&\left\| x \Big ( t ,u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y) \right\| _{{\mathbb {R}}^n} \le \left\{ \begin{array}{lcl} 0 &{} \text {if} &{} a \le t< \tau , \\ 2 R \displaystyle \int _{\tau }^t \dfrac{(t-s)^{\alpha - 1}}{\varGamma (\alpha )} \; ds &{} \text {if} &{} \tau \le t \le \tau + h , \\ 2R \displaystyle \int _{\tau }^{\tau + h} \dfrac{(t-s)^{\alpha - 1}}{\varGamma (\alpha )} \; ds &{} \text {if} &{} t> \tau + h , \end{array} \right. \\&\quad = \left\{ \begin{array}{lcl} 0 &{} \text {if} &{} a \le t < \tau , \\ 2 R \dfrac{(t-\tau )^\alpha }{\varGamma (\alpha +1)} &{} \text {if} &{} \tau \le t \le \tau + h , \\ 2R \dfrac{(t-\tau )^\alpha - (t-(\tau +h))^\alpha }{\varGamma (\alpha +1)} &{} \text {if} &{} t > \tau + h , \end{array} \right. \end{aligned}$$

for all \(t \in [a,b]\). To prove Inequality (10), one has just to see that in all of the three above cases, the right-hand side term is less than \(\frac{2R}{\varGamma (\alpha +1)}h^\alpha \). For the last case, one has just to invoke the basic inequality \(\chi _2^\alpha - \chi _1^\alpha \le (\chi _2 - \chi _1 )^\alpha \) which is satisfied for all \(0 \le \chi _1 \le \chi _2\) (see the proof of Lemma 3.3 for some details). Now let us prove Inequality (11). Using similar arguments, one has

$$\begin{aligned}&\dfrac{ x \Big ( t , u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y)}{h} - \dfrac{(t-\tau )^{\alpha -1}}{\varGamma (\alpha )} (v-u(\tau )) \\&\quad = \dfrac{1}{h} \int _\tau ^{\tau +h} \dfrac{(t-s)^{\alpha -1}}{\varGamma (\alpha )} (v-u(s)) \; ds - \dfrac{(t-\tau )^{\alpha -1}}{\varGamma (\alpha )} (v-u(\tau )) \\&\quad = \dfrac{1}{h} \int _\tau ^{\tau +h} \dfrac{(t-s)^{\alpha -1} -(t-\tau )^{\alpha -1} }{\varGamma (\alpha )} (v-u(s)) \; ds + \dfrac{(t-\tau )^{\alpha -1}}{\varGamma (\alpha )} \left( u(\tau ) - \dfrac{1}{h} \int _\tau ^{\tau +h} u(s) \; ds \right) , \end{aligned}$$

for all \(t \in (\tau +h,b]\). Since \(u \in \mathrm {L}^\infty _R\) and \(v \in {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\), we obtain that

$$\begin{aligned}&\left\| \dfrac{ x \Big ( t , u^{(\tau ,v)}(\cdot ,h),y \Big )- x(t,u,y)}{h} - \dfrac{(t-\tau )^{\alpha -1}}{\varGamma (\alpha )} (v-u(\tau )) \right\| _{{\mathbb {R}}^n} \\&\quad \le \dfrac{2R}{\varGamma (\alpha )} \dfrac{1}{h} \int _\tau ^{\tau +h} (t-s)^{\alpha -1} -(t-\tau )^{\alpha -1} \; ds + \dfrac{( t - \tau )^{\alpha -1}}{\varGamma (\alpha )} \left\| \dfrac{1}{h} \int _\tau ^{\tau +h} u(s) \; ds - u(\tau ) \right\| _{{\mathbb {R}}^n} , \end{aligned}$$

for all \(t \in (\tau +h,b]\). Noting that \((t-s)^{\alpha -1} \le (t-(\tau +h))^{\alpha -1}\) for all \(s \in [\tau ,\tau +h]\) and all \(t \in (\tau +h,b]\), the proof of Inequality (11) is complete. \(\square \)

Proposition B.2

(Sensitivity analysis under needle-perturbation of u) Let \(R \ge 0\) and let \((u,y) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\). It holds that

$$\begin{aligned}&\lim \limits _{h \rightarrow 0^+} \dfrac{\varPhi _R \Big ( u^{(\tau ,v)}(\cdot ,h),y \Big )-\varPhi _R(u,y)}{h} = \dfrac{(b-\tau )^{\beta -1}}{\varGamma (\beta )} \Big ( L(x(\tau ),v,\tau )-L(x(\tau ),u(\tau ),\tau ) \Big ) \\&\quad +\, \left( \dfrac{ (b-\tau )^{\alpha -1}}{\varGamma (\alpha )} \partial _2 \varphi (x(a),x(b)) + \mathrm {I}^\alpha _{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (\tau ) \right) \cdot (v-u(\tau )) , \end{aligned}$$

for all \((\tau ,v) \in {\mathcal {P}}(u,y) \times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\), where \(x = x(\cdot ,u,y) \in {}_\mathrm {c}\mathrm {AC}^{\alpha ,\infty }_{a+}\).

Proof

Let \((\tau ,v) \in {\mathcal {P}}(u,y) \times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\). For simplicity of notations in this proof, we denote by

$$\begin{aligned} u_h := u^{(\tau ,v)}(\cdot ,h) \quad \text {and} \quad x_h := x \Big ( \cdot , u^{(\tau ,v)}(\cdot ,h),y \Big ), \end{aligned}$$

for all \(0 < h \le b-\tau \). From Inequality (11) and since \(\tau \) is a Lebesgue point of u, it is clear that

$$\begin{aligned} \lim \limits _{h \rightarrow 0^+} \dfrac{\varphi (x_h(a),x_h(b))-\varphi (x(a),x(b))}{h} = \dfrac{ (b-\tau )^{\alpha -1}}{\varGamma (\alpha )} \partial _2 \varphi (x(a),x(b)) \cdot (v-u(\tau )). \end{aligned}$$

Moreover the term

$$\begin{aligned} \dfrac{\mathrm {I}^{\beta }_{a+} [ L(x_h,u_h,\cdot )](b) - \mathrm {I}^{\beta }_{a+} [ L(x,u,\cdot ) ](b) }{h}, \end{aligned}$$

can be decomposed as:

$$\begin{aligned}&\dfrac{1}{h} \int _\tau ^{\tau +h} \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \Big ( L(x_h(s),v,s) - L(x(s),v,s) \Big ) \; ds \\&\quad +\, \dfrac{1}{h} \int _\tau ^{\tau +h} \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \Big ( L(x(s),v,s) - L(x(s),u(s),s) \Big ) \; ds \\&\quad +\, \dfrac{1}{h} \int _{\tau +h}^b \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \Big ( L(x_h(s),u(s),s) - L(x(s),u(s),s) \Big ) \; ds, \end{aligned}$$

for all \(0 < h \le b-\tau \). From the uniform convergence obtained in Inequality (10), the first term tends to zero when \(h \rightarrow 0^+\). Since \(\tau \) is a Lebesgue point of the function \(L(x,u,\cdot )\), it is clear that the second term tends to

$$\begin{aligned} \dfrac{(b-\tau )^{\beta -1}}{\varGamma (\beta )} \Big ( L(x(\tau ),v,\tau )-L(x(\tau ),u(\tau ),\tau ) \Big ), \end{aligned}$$

when \(h \rightarrow 0^+\). From a Taylor expansion with integral rest, the last term can be decomposed as:

$$\begin{aligned}&\int _{\tau +h}^b \int _0^1 \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \partial _1 L (\star _\theta ) \cdot \left( \dfrac{x_h(s)-x(s)}{h} - \dfrac{(s-\tau )^{\alpha -1}}{\varGamma (\alpha )} (v-u(\tau )) \right) \; d\theta ds \\&\quad +\, \int _{\tau +h}^b \int _0^1 \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \dfrac{(s-\tau )^{\alpha -1}}{\varGamma (\alpha )} ( \partial _1 L (\star _\theta ) - \partial _1 L (\star _0) ) \cdot (v-u(\tau )) \; d\theta ds \\&\quad +\, \int _{\tau +h}^b \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \dfrac{(s-\tau )^{\alpha -1}}{\varGamma (\alpha )} \partial _1 L (\star _0) \cdot (v-u(\tau )) \; ds \end{aligned}$$

where

$$\begin{aligned} \star _\theta := ( x(s)+\theta (x_h(s)-x(s)) , u(s) , s), \end{aligned}$$

for all \(\theta \in [0,1]\) and all \(0 < h \le b-\tau \). The last above term clearly tends to

$$\begin{aligned} \mathrm {I}^\alpha _{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (\tau ) \cdot (v-u(\tau )), \end{aligned}$$

when \(h \rightarrow 0^+\). From the LDC theorem, the second above term clearly tends to zero when \(h \rightarrow 0^+\). Finally one can prove from Inequality (11) that the norm of the first above term can be bounded by

$$\begin{aligned}&\dfrac{\Big ( b-(\tau +h) \Big )^{\alpha +\beta -1}}{\varGamma (\alpha +\beta )} \left\| \dfrac{1}{h} \int _\tau ^{\tau +h} u(s) \; ds - u(\tau ) \right\| _{{\mathbb {R}}^n} \\&\quad +\, \dfrac{2R}{\varGamma (\alpha +\beta )} \left( \Big ( b-(\tau +h) \Big )^{\alpha +\beta -1} - ( b-\tau )^{\alpha +\beta -1} \right) \\&\quad +\, 2R \int _\tau ^{\tau +h} \dfrac{(b-s)^{\beta -1}}{\varGamma (\beta )} \dfrac{(s-\tau )^{\alpha -1}}{\varGamma (\alpha )} \; ds, \end{aligned}$$

which tends to zero when \(h \rightarrow 0^+\) (in particular since \(\tau \) is a Lebesgue point of the function u). The proof is complete. \(\square \)

Proposition B.3

(Sensitivity analysis under perturbation of y) Let \(R \ge 0\) and let \((u,y) \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\). It holds that

$$\begin{aligned}&\lim \limits _{h \rightarrow 0^+} \dfrac{\varPhi _R(u,y+hy')-\varPhi _R(u,y)}{h} \\&\quad = \left( \partial _1 \varphi (x(a),x(b)) + \partial _2 \varphi (x(a),x(b)) + \mathrm {I}^\beta _{a+} \Big [ \partial _1 L (x, u, \cdot ) \Big ](b) \right) \cdot y', \end{aligned}$$

for all \(y' \in {\mathbb {R}}^n\), where \(x = x(\cdot ,u,y) \in {}_\mathrm {c}\mathrm {AC}^{\alpha ,\infty }_{a+}\).

Proof

We apply Proposition 3.1 with the constant variation \(\eta = y' \in {}_{\mathrm {c}} \mathrm {AC}^{\alpha ,\infty }_{a+}\). \(\square \)

1.3 B.3 Proof of Theorem 3.2 by Applying the Ekeland Variational Principle

Let \(x \in \mathrm {K}\) be a solution to Problem (P). Using the notations introduced in Section B.2, it holds that \(x = x(\cdot ,u,y)\) where \(u = {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x]\) and \(y=x(a)\). In particular note that \(g(y,x(b,u,y))=g(x(a),x(b)) \in \mathrm {S}\). Let \(R := \Vert u \Vert _{\mathrm {L}^\infty } + 1 \ge 0\) and let us consider a positive sequence \((\varepsilon _k)_{k \in {\mathbb {N}}}\) which tends to zero when \(k \rightarrow \infty \). We introduce the penalized functional

$$\begin{aligned}{}\begin{array}[t]{lrcl}\mathrm {J}_k :&{}\mathrm {L}^\infty _R \times {\mathbb {R}}^n &{}\longrightarrow &{}{\mathbb {R}}_+\\ {} &{}(u',y')&{} \longmapsto &{} \sqrt{ \Bigg ( \Big ( \varPhi _R (u',y') - \varPhi _R (u,y) + \varepsilon _k \Big )^+ \Bigg )^2 + \mathrm {d}^2_\mathrm {S}\Big ( g(y',x(b,u',y')) \Big ) } , \end{array} \end{aligned}$$

for all \(k \in {\mathbb {N}}\). From Lemma B.3 and the continuities of g and \(\mathrm {d}^2_\mathrm {S}\), it is clear that \(J_k\) is a continuous nonnegative map defined on a complete metric space for all \(k \in {\mathbb {N}}\) . Since \(\mathrm {J}_k (u,y) = \varepsilon _k\) for all \(k \in {\mathbb {N}}\), we deduce from the Ekeland variational principle (see Proposition B.1) that there exists a sequence \((u_k,y_k)_{k \in {\mathbb {N}}} \subset \mathrm {L}^\infty _R \times {\mathbb {R}}^n\) such that

$$\begin{aligned} \mathrm {d}_{\mathrm {L}^\infty _R \times {\mathbb {R}}^n} \Big ( (u_k,y_k),(u,y) \Big ) \le \sqrt{\varepsilon _k}, \end{aligned}$$

and

$$\begin{aligned} - \sqrt{\varepsilon _k} \; \mathrm {d}_{\mathrm {L}^\infty _R \times {\mathbb {R}}^n} \Big ( (u',y') , (u_k,y_k) \Big ) \le \mathrm {J}_k (u',y') - \mathrm {J}_k(u,y), \end{aligned}$$
(12)

for all \((u',y') \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\) and all \(k \in {\mathbb {N}}\). In the sequel, we denote by \(x_k := x(\cdot ,u_k,y_k)\) for all \(k \in {\mathbb {N}}\). Note that the sequence \((u_k,y_k)_{k \in {\mathbb {N}}}\) converges to (uy) in \(\mathrm {L}^\infty _R \times {\mathbb {R}}^n\), and thus the sequence \((x_k)_{k \in {\mathbb {N}}}\) converges to x in \(\mathrm {C}\) (see Lemma B.3).

From the optimality of x, one can easily see that \(\mathrm {J}_k (u',y') > 0\) for all \((u',y') \in \mathrm {L}^\infty _R \times {\mathbb {R}}^n\) and all \(k \in {\mathbb {N}}\). As a consequence, we can correctly define

$$\begin{aligned} \psi ^0_k := - \dfrac{1}{\mathrm {J}_k (u_k,y_k)} \Big ( \varPhi _R (u_k,y_k) - \varPhi _R (u,y) + \varepsilon _k \Big )^+ \le 0, \end{aligned}$$

and

$$\begin{aligned} \psi _k := - \dfrac{1}{\mathrm {J}_k (u_k,y_k)} \Bigg ( g(x_k(a),x_k(b)) - \mathrm {P}_\mathrm {S}\Big ( g(x_k(a),x_k(b)) \Big ) \Bigg ) \in {\mathbb {R}}^j, \end{aligned}$$

which satisfy \(\vert \psi ^0_k \vert ^2 + \Vert \psi _k \Vert _{{\mathbb {R}}^j}^2 = 1\) for all \(k \in {\mathbb {N}}\). From a standard compactness argument and from the PCLDC theorem, we can extract subsequences (that we do not relabel) such that \((\psi ^0_k)_{k \in {\mathbb {N}}}\) converges to some \(\psi ^0 \le 0\), \((\psi _k)_{k \in {\mathbb {N}}}\) converges to some \(\psi \in {\mathbb {R}}^j \) satisfying \(-\psi \in \mathrm {N}_\mathrm {S}[ g(x(a),x(b)) ]\) (see Lemma B.1) and \((u_k)_{k \in {\mathbb {N}}}\) converges to u pointwisely almost everywhere on [ab]. Moreover, note that \(\vert \psi ^0 \vert ^2 + \Vert \psi \Vert _{{\mathbb {R}}^j}^2 = 1\) and thus the couple \((\psi ^0,\psi )\) is not trivial.

Perturbation of \(y_k\). Let \(y' \in {\mathbb {R}}^n\) and let us fix some \(k \in {\mathbb {N}}\). From Inequality (12), it holds that

$$\begin{aligned} - \sqrt{\varepsilon _k} \; \Vert y' \Vert _{{\mathbb {R}}^n} \le \dfrac{1}{\mathrm {J}_k (u_k,y_k+hy') + \mathrm {J}_k (u_k,y_k)} \times \dfrac{\mathrm {J}^2_k ( u_k,y_k+hy' ) - \mathrm {J}^2_k (u_k,y_k)}{h}, \end{aligned}$$

for all \(h > 0\). Letting \(h \rightarrow 0^+\), we exactly get from Proposition B.3 that

$$\begin{aligned}&\Bigg ( \psi ^0_k \left( \partial _1 \varphi (x_k(a),x_k(b)) + \partial _2 \varphi (x_k(a),x_k(b)) + \mathrm {I}^\beta _{a+} \Big [ \partial _1 L (x_k, u_k, \cdot ) \Big ](b) \right) \\&\quad +\, \Big ( \partial _1 g ( x_k(a) , x_k(b) )^\top + \partial _2 g ( x_k(a) , x_k(b) )^\top \Big ) \times \psi _k \Bigg ) \cdot y' \le \sqrt{\varepsilon _k} \Vert y' \Vert _{{\mathbb {R}}^n}. \end{aligned}$$

Finally, letting \(k \rightarrow \infty \), we obtain (using in particular the LDC theorem) that

$$\begin{aligned}&\Bigg ( \psi ^0 \Big ( \partial _1 \varphi (x(a),x(b)) + \partial _2 \varphi (x(a),x(b)) + \mathrm {I}^\beta _{a+} \Big [ \partial _1 L (x, u, \cdot ) \Big ](b) \Big ) \\&\quad + \Big ( \partial _1 g ( x(a) , x(b) )^\top + \partial _2 g ( x(a) , x(b) )^\top \Big ) \times \psi \Bigg ) \cdot y' \le 0. \end{aligned}$$

Since the above inequality is satisfied for all \(y' \in {\mathbb {R}}^n\) and we can write

$$\begin{aligned} \mathrm {I}^\beta _{a+} \Big [ \partial _1 L (x, u, \cdot ) \Big ](b) = \mathrm {I}^1_{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (a), \end{aligned}$$

we deduce the crucial equality given by

$$\begin{aligned}&\Big ( \psi ^0 \partial _2 \varphi (x(a),x(b)) + \partial _2 g ( x(a) , x(b) )^\top \times \psi \Big ) + \psi ^0 \mathrm {I}^1_{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (a) \nonumber \\&\quad = - \psi ^0 \partial _1 \varphi (x(a),x(b))- \partial _1 g ( x(a) , x(b) )^\top \times \psi . \end{aligned}$$
(13)

Needle-perturbation of \(u_k\). Let \((\tau ,v) \in {\mathcal {P}}\times {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\) where

$$\begin{aligned} {\mathcal {P}}:= \Big \lbrace \tau \in (a,b) \mid (u_k(\tau ))_{k \in {\mathbb {N}}} \text { converges to } u(\tau ) \Big \rbrace \cap \bigcap _{k \in {\mathbb {N}}} {\mathcal {P}}(u_k,y_k). \end{aligned}$$

Note that \({\mathcal {P}}\) is of full measure in [ab]. Let us fix some \(k \in {\mathbb {N}}\). From Inequality (12), it holds that

$$\begin{aligned} - 2R \sqrt{\varepsilon _k} \le \dfrac{1}{\mathrm {J}_k \Big ( u_k^{(\tau ,v)}(\cdot ,h),y_k \Big ) + \mathrm {J}_k (u_k,y_k)} \times \dfrac{\mathrm {J}^2_k \Big ( u_k^{(\tau ,v)}(\cdot ,h),y_k \Big ) - \mathrm {J}^2_k (u_k,y_k)}{h}, \end{aligned}$$

for all \(0 < h \le b - \tau \). Letting \(h \rightarrow 0^+\), we exactly get from Lemma B.4 and Proposition B.2 that

$$\begin{aligned}&\Bigg ( \dfrac{(b-\tau )^{\alpha -1}}{\varGamma (\alpha )} \Big ( \psi ^0_k \partial _2 \varphi (x_k(a),x_k(b)) + \partial _2 g ( x_k(a) , x_k(b) )^\top \times \psi _k \Big ) \\&\quad +\, \psi ^0_k \mathrm {I}^\alpha _{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x_k,u_k,\cdot ) \right] (\tau ) \Bigg ) \cdot (v-u_k(\tau )) \\&\quad +\, \psi ^0_k \dfrac{(b-\tau )^{\beta -1}}{\varGamma (\beta )} \Big ( L(x_k(\tau ),v,\tau )-L(x_k(\tau ),u_k(\tau ),\tau ) \Big ) \le 2R \sqrt{\varepsilon _k}. \end{aligned}$$

Finally, letting \(k \rightarrow \infty \), we obtain (using in particular the LDC theorem and the fact that \((u_k(\tau ))_{k \in {\mathbb {N}}}\) converges to \(u(\tau )\)) that

$$\begin{aligned}&\Bigg ( \dfrac{(b-\tau )^{\alpha -1}}{\varGamma (\alpha )} \Big ( \psi ^0 \partial _2 \varphi (x(a),x(b)) + \partial _2 g ( x(a) , x(b) )^\top \times \psi \Big ) \nonumber \\&\quad +\, \psi ^0 \mathrm {I}^\alpha _{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (\tau ) \Bigg ) \cdot (v-u(\tau )) \nonumber \\&\quad +\, \psi ^0 \dfrac{(b-\tau )^{\beta -1}}{\varGamma (\beta )} \Big ( L(x(\tau ),v,\tau )-L(x(\tau ),u(\tau ),\tau ) \Big ) \le 0. \end{aligned}$$
(14)

Note that the above crucial inequality is satisfied for almost all \(\tau \in [a,b]\) and all \(v \in {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R)\).

Introduction of an adjoint vector. Let us introduce the adjoint vector p defined by

$$\begin{aligned}&p(t) := \dfrac{(b-t)^{\alpha -1} }{\varGamma (\alpha )} \Big ( \psi ^0 \partial _2 \varphi (x(a),x(b)) + \partial _2 g ( x(a) , x(b) )^\top \times \psi \Big ) \\&\quad +\, \psi ^0 \mathrm {I}^\alpha _{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (t), \end{aligned}$$

for almost every \(t \in [a,b]\). In particular it holds that

$$\begin{aligned}&\mathrm {I}^{1-\alpha }_{b-}[p](t) = \Big ( \psi ^0 \partial _2 \varphi (x(a),x(b)) + \partial _2 g ( x(a) , x(b) )^\top \times \psi \Big ) \\&\quad +\, \psi ^0 \mathrm {I}^1_{b-} \left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x,u,\cdot ) \right] (t), \end{aligned}$$

for all \(t \in [a,b]\). We deduce that \(p \in \mathrm {AC}^\alpha _{b-}\) with

$$\begin{aligned} \mathrm {D}^\alpha _{b-}[p] (t) = \psi ^0 \dfrac{(b-t)^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x(t),u(t),t), \end{aligned}$$

for almost every \(t \in [a,b]\). Moreover, in particular from Equality (13), it holds that

$$\begin{aligned} - \mathrm {I}^{1-\alpha }_{b-}[p](a)= & {} \psi ^0 \partial _1 \varphi (x(a),x(b)) + \partial _1 g ( x(a) , x(b) )^\top \times \psi , \\ \mathrm {I}^{1-\alpha }_{b-}[p](b)= & {} \psi ^0 \partial _2 \varphi (x(a),x(b)) + \partial _2 g ( x(a) , x(b) )^\top \times \psi . \end{aligned}$$

From Inequality (14), it is clear that

$$\begin{aligned} u(t) \in \mathop {\mathrm{arg max}}\limits _{v \in {\overline{\mathrm {B}}}_{{\mathbb {R}}^n}(0_{{\mathbb {R}}^n},R) } \Big \lbrace p(t) \cdot v +\psi ^0 \dfrac{(b-t)^{\beta -1}}{\varGamma (\beta )} L(x(t),v,t) \Big \rbrace , \end{aligned}$$
(15)

for almost every \(t \in [a,b]\). One can easily deduce that

$$\begin{aligned} p(t) = - \psi ^0 \dfrac{ (b-t)^{\beta -1}}{\varGamma (\beta )} \partial _2 L (x(t),u(t),t), \end{aligned}$$

for almost every \(t \in [a,b]\).

Normalization. By contradiction, let us assume that \(\psi ^0 = 0\). In that case, one can easily deduce from the above equalities that \( \partial _1 g(x(a),x(b))^\top \times \psi = \partial _2 g(x(a),x(b)) )^\top \times \psi = 0_{{\mathbb {R}}^n}\). Since g is assumed to be regular at (x(a), x(b)), we deduce that \(\psi = 0_{{\mathbb {R}}^j}\) which raises a contradiction with the nontriviality of the couple \((\psi ^0,\psi )\). We deduce that \(\psi ^0 < 0\). Moreover, since the couple \((\psi ^0,\psi )\) is defined up to a positive multiplicative constant, we now normalize the couple \((\psi ^0,\psi )\) such that \(\psi ^0 = -1\).

End of the proof. We deduce from the previous paragraphs that \(p = \frac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _2 L (x,u,\cdot ) \in \mathrm {AC}^\alpha _{b-}\) with

$$\begin{aligned} \mathrm {D}^\alpha _{b-}\left[ \dfrac{(b-\cdot )^{\beta -1}}{\varGamma (\beta )} \partial _2 L (x,u,\cdot ) \right] (t) = - \dfrac{(b-t)^{\beta -1}}{\varGamma (\beta )} \partial _1 L (x(t),u(t),t), \end{aligned}$$

for almost every \(t \in [a,b]\), which exactly corresponds to the Euler–Lagrange equation stated in Theorem 3.2 since \(u = {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x]\). We also deduce from the equalities on \(\mathrm {I}^{1-\alpha }_{b-} [p](a)\) and \(\mathrm {I}^{1-\alpha }_{b-} [p](b)\) in a previous paragraph that the transversality conditions given in Theorem 3.2 are satisfied. Finally, from the maximization condition (15), it is clear that the matrix \(\frac{(b-t)^{\beta -1}}{\varGamma (\beta )} \partial ^2_{22} L(x(t),u(t),t)\) is positive semi-definite for almost all \(t \in [a,b]\), which exactly corresponds to the Legendre condition given in Theorem 3.2 since \(u = {}_{\mathrm {c}} \mathrm {D}^\alpha _{a+}[x]\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bourdin, L., Ferreira, R.A.C. Legendre’s Necessary Condition for Fractional Bolza Functionals with Mixed Initial/Final Constraints. J Optim Theory Appl 190, 672–708 (2021). https://doi.org/10.1007/s10957-021-01908-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-021-01908-w

Keywords

Mathematics Subject Classification

Navigation