Skip to main content
Log in

Branch-and-Lift Algorithm for Deterministic Global Optimization in Nonlinear Optimal Control

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

Abstract

This paper presents a branch-and-lift algorithm for solving optimal control problems with smooth nonlinear dynamics and potentially nonconvex objective and constraint functionals to guaranteed global optimality. This algorithm features a direct sequential method and builds upon a generic, spatial branch-and-bound algorithm. A new operation, called lifting, is introduced, which refines the control parameterization via a Gram–Schmidt orthogonalization process, while simultaneously eliminating control subregions that are either infeasible or that provably cannot contain any global optima. Conditions are given under which the image of the control parameterization error in the state space contracts exponentially as the parameterization order is increased, thereby making the lifting operation efficient. A computational technique based on ellipsoidal calculus is also developed that satisfies these conditions. The practical applicability of branch-and-lift is illustrated in a numerical example.

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.

Algorithm 1
Fig. 1
Algorithm 2
Fig. 2
Algorithm 3
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. In a more general implementation of the branching operation, the set \(\mathbb{A}\) can be subdivided into more than two subsets. This is useful, for instance, when running the algorithm on a multiprocessor computer, thereby enabling multiple branches to be analyzed in parallel.

References

  1. Mitsos, A., Chachuat, B., Barton, P.I.: Towards global bilevel dynamic optimization. J. Glob. Optim. 45(1), 63–93 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  2. Betts, J.T.: Practical Methods for Optimal Control Using Nonlinear Programming, 2nd edn. Advances in Design and Control Series. SIAM, Philadelphia (2010)

    Book  MATH  Google Scholar 

  3. Biegler, L.T.: Nonlinear Programming: Concepts, Algorithms, and Applications to Chemical Processes. MOS-SIAM Series on Optimization. SIAM, Philadelphia (2010)

    Book  Google Scholar 

  4. Teo, K.L., Goh, C.J., Wong, K.H.: A Unified Computational Approach to Optimal Control Problems. Longman, New York (1991)

    MATH  Google Scholar 

  5. Banga, J.R., Moles, C.G., Alonso, A.A.: Global optimization of bioprocesses using stochastic and hybrid methods. In: Floudas, C.A., Pardalos, P.M. (eds.) Frontiers in Global Optimization. Nonconvex Optimization and Its Applications, vol. 74, pp. 45–70. Kluwer, Dordrecht (2003)

    Chapter  Google Scholar 

  6. Egea, J.A., Vazquez, E., Banga, J.R., Marti, R.M.: Improved scatter search for the global optimization of computationally expensive dynamic models. J. Glob. Optim. 43(2–3), 175–190 (2009)

    Article  MATH  Google Scholar 

  7. Luss, R.: Iterative Dynamic Programming. Chapman & Hall/CRC, Boca Raton (2000)

    Book  Google Scholar 

  8. Bellman, R.E.: Dynamic Programming. Princeton University Press, Princeton (1957)

    MATH  Google Scholar 

  9. Björnberg, J., Diehl, M.: Approximate robust dynamic programming and robustly stable MPC. Automatica 42(5), 777–782 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  10. Grüne, L., Semmler, W.: Using dynamic programming with adaptive grid scheme to solve nonlinear dynamic models in economics. Comput. Econ. Finance 2002(99) (2002)

  11. Luss, R.: Optimal control by dynamic programming using systematic reduction in grid size. Int. J. Control 51(5), 995–1013 (1990)

    Article  Google Scholar 

  12. Bryson, A.E., Ho, Y.: Applied Optimal Control. Hemisphere, Washington (1975)

    Google Scholar 

  13. von Stryk, O., Bulirsch, R.: Direct and indirect methods for trajectory optimization. Ann. Oper. Res. 37, 357–373 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  14. Pontryagin, L.S., Boltyanskii, V.G., Gamkrelidze, R.V., Mishchenko, E.F.: The Mathematical Theory of Optimal Processes. Wiley, New York (1962)

    MATH  Google Scholar 

  15. Brusch, R., Schappelle, R.: Solution of highly constrained optimal control problems using nonlinear programming. AIAA J. 11(2), 135–136 (1973)

    Article  Google Scholar 

  16. Sargent, R.W.H., Sullivan, G.R.: The development of an efficient optimal control package. In: Stoer, J. (ed.) Proc. 8th IFIP Conf. Optimization Techniques, 1977, Heidelberg (1978)

    Google Scholar 

  17. Chachuat, B., Singer, A.B., Barton, P.I.: Global methods for dynamic optimization and mixed-integer dynamic optimization. Ind. Eng. Chem. Res. 45(25), 8373–8392 (2006)

    Article  Google Scholar 

  18. Esposito, W.R., Floudas, C.A.: Deterministic global optimization in nonlinear optimal control problems. J. Glob. Optim. 17, 96–126 (2000)

    Article  MathSciNet  Google Scholar 

  19. Esposito, W.R., Floudas, C.A.: Global optimization for the parameter estimation of differential-algebraic systems. Ind. Eng. Chem. Res. 39(5), 1291–1310 (2000)

    Article  Google Scholar 

  20. Lin, Y., Stadtherr, M.A.: Validated solutions of initial value problems for parametric ODEs. Appl. Numer. Math. 57(10), 1145–1162 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  21. Papamichail, I., Adjiman, C.S.: A rigorous global optimization algorithm for problems with ordinary differential equations. J. Glob. Optim. 24, 1–33 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  22. Singer, A.B., Barton, P.I.: Global optimization with nonlinear ordinary differential equations. J. Glob. Optim. 34(2), 159–190 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  23. Sahlodin, A.M.: Global optimization of dynamic process systems using complete search methods. Ph.D. thesis, McMaster University, Ontario, Canada (2012)

  24. Horst, R., Tuy, H.: Global Optimization: Deterministic Approaches, 3rd edn. Springer, Berlin (1996)

    Book  MATH  Google Scholar 

  25. Bock, H.G., Plitt, K.J.: A multiple shooting algorithm for direct solution of optimal control problems. In: Proc. IFAC 9th World Congress, Budapest, Hungary, pp. 242–247 (1984)

    Google Scholar 

  26. Houska, B., Ferreau, H.J., Diehl, M.: ACADO Toolkit—an open source framework for automatic control and dynamic optimization. Optim. Control Appl. Methods 32(3), 298–312 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  27. Leineweber, D.B., Bauer, I., Bock, H.G., Schlöder, J.P.: An efficient multiple shooting based reduced SQP strategy for large-scale dynamic process optimization. Part I: Theoretical aspects. Comput. Chem. Eng. 27, 157–166 (2003)

    Article  Google Scholar 

  28. Neuman, C.P., Sen, A.: A suboptimal control algorithm for constrained problems using cubic splines. Automatica 9, 601–603 (1973)

    Article  MATH  Google Scholar 

  29. Tsang, T.H., Himmelblau, D.M., Edgar, T.F.: Optimal control via collocation and nonlinear programming. Int. J. Control 21, 763–768 (1975)

    Article  MATH  Google Scholar 

  30. Tawarmalani, M., Sahinidis, N.V.: A polyhedral branch-and-cut approach to global optimization. Math. Program. 103(2), 225–249 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  31. Sahinidis, N.V.B.: A general purpose global optimization software package. J. Glob. Optim. 8(2), 201–205 (1996)

    Article  MATH  MathSciNet  Google Scholar 

  32. Flores Tlacuahuac, A., Terrazas, S., Biegler, L.T.: Global optimization of highly nonlinear dynamic systems. Ind. Eng. Chem. Res. 47(8), 2643–2655 (2008)

    Article  Google Scholar 

  33. Galperin, E.F., Zheng, Q.: Variation-free iterative method for global optimal control. Int. J. Control 50(5), 1731–1743 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  34. Scott, J.K., Barton, P.I.: Convex relaxations for nonconvex optimal control problems. In: Proc. 50th IEEE Conf. Decision Control, Orlando, Florida, pp. 1042–1047 (2011)

    Google Scholar 

  35. Verfaillie, G., Lemaître, M., Schiex, T.: Russian doll search for solving constraint optimization problems. In: Proc. 13th Nat. Conf. Artificial Intelligence, Portland, Oregon, vol. 1, pp. 181–187 (1996)

    Google Scholar 

  36. Floudas, C.A.: Deterministic Global Optimization: Theory, Methods, and Applications. Kluwer, Dordrecht (1999)

    Google Scholar 

  37. Neumaier, A.: Complete search in continuous global optimization and constraint satisfaction. Acta Numer. 13, 271–369 (2004)

    Article  MathSciNet  Google Scholar 

  38. Achterberg, T., Koch, T., Martin, A.: Branching rules revisited. Oper. Res. Lett. 33, 42–54 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  39. Belotti, P., Lee, J., Liberti, L., Margot, F., Wächter, A.: Branching and bounds tightening techniques for non-convex MINLP. Oper. Res. Lett. 33, 42–54 (2005)

    Article  MathSciNet  Google Scholar 

  40. Lasserre, J.B.: Moments, Positive Polynomials and Their Applications. Imperial College Press, London (2009)

    Book  Google Scholar 

  41. Walter, W.: Differential and Integral Inequalities. Springer, Berlin (2009)

    Google Scholar 

  42. Sahlodin, A.M., Chachuat, B.: Discretize-then-relax approach for convex/concave relaxations of the solutions of parametric ODEs. Appl. Numer. Math. 61, 803–820 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  43. Scott, J.K., Chachuat, B., Barton, P.I.: Nonlinear convex and concave relaxations for the solutions of parametric ODEs. Optim. Control Appl. Methods 34(2), 145–163 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  44. Scott, J.K., Barton, P.I.: Improved relaxations for the parametric solutions of ODEs using differential inequalities. J. Glob. Optim. 57(1), 143–176 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  45. Singer, A.B., Barton, P.I.: Bounding the solutions of parameter dependent nonlinear ordinary differential equations. SIAM J. Sci. Comput. 27(6), 2167–2182 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  46. McCormick, G.P.: Computability of global solutions to factorable nonconvex programs: Part I—Convex underestimating problems. Math. Program. 10, 147–175 (1976)

    Article  MATH  MathSciNet  Google Scholar 

  47. Berz, M., Makino, K.: Performance of Taylor model methods for validated integration of ODEs. In: Lecture Notes Comput. Sci., vol. 3732, pp. 65–74. Springer, Berlin (2006)

    Google Scholar 

  48. Neher, M., Jackson, K.R., Nedialkov, N.S.: On Taylor model based integration of ODEs. SIAM J. Numer. Anal. 45, 236–262 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  49. Sahlodin, A.M., Chachuat, B.: Convex/concave relaxations of parametric ODEs using Taylor models. Comput. Chem. Eng. 35, 844–857 (2011)

    Article  Google Scholar 

  50. Bompadre, A., Mitsos, A., Chachuat, B.: Convergence analysis of Taylor and McCormick–Taylor models. J. Glob. Optim. 57(1), 75–114 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  51. Houska, B., Logist, F., van Impe, Y., Diehl, M.: Robust optimization of nonlinear dynamic systems with application to a jacketed tubular reactor. J. Process Control 22(6), 1152–1160 (2012)

    Article  Google Scholar 

  52. Kurzhanski, A.B., Valyi, P.: Ellipsoidal Calculus for Estimation and Control. Birkhäuser, Boston (1997)

    Book  MATH  Google Scholar 

  53. Kurzhanski, A.B., Varaiya, P.: Reachability analysis for uncertain systems—the ellipsoidal technique. Dyn. Contin. Discrete Impuls. Syst., Ser. B, Appl. Algorithms 9, 347–367 (2002)

    MATH  MathSciNet  Google Scholar 

  54. Lin, Y., Stadtherr, M.A.: Deterministic global optimization of nonlinear dynamic systems. AIChE J. 53(4), 866–875 (2007)

    Article  Google Scholar 

  55. Zhao, Y., Stadtherr, M.A.: Rigorous global optimization for dynamic systems subject to inequality path constraints. Ind. Eng. Chem. Res. 50(22), 12678–12693 (2011)

    Article  Google Scholar 

  56. Adjiman, C.S., Dallwig, S., Floudas, C.A., Neumaier, A.: A global optimization method, αBB, for general twice-differentiable constrained NLPs—I. Theoretical advances. Comput. Chem. Eng. 22(9), 1137–1158 (1998)

    Article  Google Scholar 

  57. Chachuat, B., Latifi, M.A.: A new approach in deterministic global optimization of problems with ordinary differential equations. In: Floudas, C.A., Pardalos, P.M. (eds.) Frontiers in Global Optimization, Nonconvex Optimization and Its Applications, vol. 74, pp. 83–108. Kluwer, Dordrecht (2003)

    Chapter  Google Scholar 

  58. Mitsos, A., Chachuat, B., Barton, P.I.: McCormick-based relaxations of algorithms. SIAM J. Optim. 20(2), 573–601 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  59. Horst, R.: Deterministic global optimization with partition sets whose feasibility is not known: application to concave minimization, reverse convex constraints, DC-programming, and Lipschitzian optimization. J. Optim. Theory Appl. 58(1), 11–37 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  60. Kearfott, R.B.: Rigorous Global Search: Continuous Problems, Nonconvex Optimization and Its Applications vol. 13. Kluwer, Dordrecht (1996)

    Book  Google Scholar 

  61. Kearfott, R.B.: On proving existence of feasible points in equality constrained optimization problems. Math. Program. 83, 89–100 (1998)

    MATH  MathSciNet  Google Scholar 

  62. Filippov, A.F.: On certain questions in the theory of optimal control. J. SIAM Control Ser. A 1(1), 76–84 (1962)

    MATH  MathSciNet  Google Scholar 

  63. Saff, E.B., Totik, V.: Polynomial approximation of piecewise analytic functions. J. Lond. Math. Soc. 39(2), 487–498 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  64. Gottlieb, D., Shu, C.W.: On the Gibbs phenomenon and its resolution. SIAM Rev. 39(4), 644–668 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  65. Bhattacharjee, B., Lemonidis, P., Green, W.H., Barton, P.I.: Global solution of semi-infinite programs. Math. Program., Ser. B 103(2), 283–307 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  66. Nedialkov, N.S., Jackson, K.R., Corliss, G.F.: Validated solutions of initial value problems for ordinary differential equations. Appl. Math. Comput. 105(1), 21–68 (1999)

    Article  MATH  MathSciNet  Google Scholar 

  67. Parulekar, S.J.: Analysis of forced periodic operations of continuous bioprocesses—single input variations. Chem. Eng. Sci. 53(14), 2481–2502 (1998)

    Article  Google Scholar 

  68. Ruan, L., Chen, X.D.: Comparison of several periodic operations of a continuous fermentation process. Biotechnol. Prog. 12, 286–288 (1996)

    Article  Google Scholar 

  69. Agarwal, P., Koshy, G., Ramirez, M.: An algorithm for operating a fed-batch fermentor at optimum specific growth rate. Biotechnol. Bioeng. 33, 115–125 (1989)

    Article  Google Scholar 

  70. Houska, B., Logist, F., van Impe, Y., Diehl, M.: Approximate robust optimization of time-periodic stationary states with application to biochemical processes. In: Proc. 50th IEEE Conf. Decision Control, Shanghai, China, pp. 6280–6285 (2009)

    Google Scholar 

  71. Azhmyakov, V., Raisch, J.: Convex control systems and convex optimal control. IEEE Trans. Autom. Control 53(4), 993–998 (2008)

    Article  MathSciNet  Google Scholar 

  72. Boyd, S., Vandenberghe, L.: Convex Optimization. University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

This paper is based upon work supported by the Engineering and Physical Sciences Research Council (EPSRC) under Grant EP/J006572/1. Financial support from the Centre of Process Systems Engineering (CPSE) of Imperial College is gratefully acknowledged. The authors are grateful to the reviewers for the thoughtful comments that led to substantial improvement of the article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Benoît Chachuat.

Appendices

Appendix A: Proof of Theorem 4.1

This section aims at constructing convergent bounds on the set \({\mathbb{E}}_{M}(t,a)\), thereby providing a proof of Theorem 4.1. Recall the definition of the set \({\mathbb{E}}_{M}(t,a)\) in Sect. 4 as

$${\mathbb{E}}_M(t,a) := \bigl\{ x(t,u) - y(t,a) \mid u \in\mathbb{U}_M(a) \bigr\} . $$

The blanket Assumption A2 guarantees that the solution trajectories x and y are well defined, so that a sufficiently large and compact a priori enclosure \(\widetilde{\mathbb{E}} \subseteq\mathbb{R}^{n_{x}}\) can always be found such that

$$\forall t \in[0,T] , \quad{\mathbb{E}}_M(t,a) \subseteq\widetilde{ \mathbb{E}} . $$

The focus here is on tightening the a priori enclosure \(\widetilde {\mathbb{E}}\), which is crucial for analyzing the convergence of the set \({\mathbb{E}}_{M}(t,a)\). We start by defining the response defect

$$d(t,u,a) = x(t,u) - y(t,a) , $$

which by construction satisfies an ODE of the form

$$\begin{aligned} & \quad\forall t \in[0,T] , \\ & \dot{d}(t,u,a) \overset{(\scriptsize3,9)}{=} f\bigl(x(t,u)\bigr) - f\bigl(y(t,a)\bigr) + G\bigl(x(t,u)\bigr) u(t) - G\bigl(y(t,a)\bigr) \Biggl( \sum _{i=0}^{M} a_i \varPhi_i(t) \Biggr) \\ & \quad\text{with} \quad d(0,u,a) = 0 . \end{aligned}$$
(25)

In order to bound the solution trajectory of (25), consider a Taylor expansion of the term f(x(t,u))−f(y(t,a)) in the form

$$f\bigl(x(t,u)\bigr) - f\bigl(y(t,a)\bigr) = \frac{\partial f(y(t,a))}{\partial x} d(t,u,a) + R_f(t,u,a) d(t,u,a) , $$

where the matrix \(R_{f}(t,u,a) \in\mathbb{R}^{n_{x} \times n_{x}}\) denotes a nonlinear remainder term. By the mean-value theorem, R f is bounded as

$$\begin{aligned} \forall u \in\mathbb{U}_M(a) , \quad\big\Vert R_f(t,u,a) \big\Vert\leq\max_{d',d'' \in\widetilde{\mathbb{E}}} \biggl \Vert \frac{1}{2} \frac{\partial^2 f(y(t,a) + d')}{\partial x^2} d'' \biggr \Vert . \end{aligned}$$
(26)

Similarly, the mean-value theorem can be applied to the term [G(x(t,u))−G(y(t,a))]u(t), giving

$$\bigl[ G\bigl(x(t,u)\bigr) - G\bigl(y(t,a)\bigr) \bigr] u(t) = R_G(t,u,a) d(t,u,a) , $$

where the nonlinear remainder term R G (t,u,a) is bounded by

$$\begin{aligned} \forall u \in\mathbb{U}_M(a) , \quad\big\Vert R_G(t,u,a)\big \Vert\leq\max_{u' \in\mathbb{F}_u(t)} \max _{d' \in\widetilde{\mathbb{E}}} \biggl \Vert \frac{\partial G( y(t,a) + d')}{\partial x} u' \biggr \Vert . \end{aligned}$$
(27)

In the following, we introduce the shorthand notation R(t,u,a):=R f (t,u,a)+R G (t,u,a), so that (25) can be written in the form

$$\begin{aligned} &\quad \forall t \in[0,T] , \\ & \dot{d}(t,u,a) = \frac{\partial f(y(t,a))}{\partial x} d(t,u,a) + R(t,u,a) d(t,u,a) \\ &\phantom{\dot{d}(t,u,a)=}{}+ G \bigl(y(t,a)\bigr) \Biggl[ u(t) - \sum_{i=0}^{M} a_i \varPhi_i(t) \Biggr] \quad\text{with} \ d(0,u,a) = 0 . \end{aligned}$$
(28)

A few remarks are in order regarding the previous ODE (28):

  1. 1.

    The Jacobian of f and the function G are evaluated at the point y(t,a). Therefore, these matrices are independent of the control input u.

  2. 2.

    The matrix-valued remainder term R(t,u,a) depends on u in general, but standard tools such as interval arithmetics can be applied to construct a bound \(\overline{R}(t,a)<\infty\) independent of u such that

    $$\begin{aligned} & \quad \forall t\in[0,T] , \\ & \max_{d',d'' \in\widetilde{\mathbb{E}}} \biggl \Vert \frac {\partial^2 f(y(t,a) + d')}{\partial x^2} \frac{d''}{2} \biggr \Vert _2 + \max_{u' \in \mathbb{F}_u(t)} \max_{d' \in\widetilde{\mathbb{E}}} \biggl \Vert \frac {\partial G( y(t,a) + d')}{\partial x} u' \biggr \Vert _2 \leq\overline{R}(t,a) . \end{aligned}$$
    (29)

    The validity of this bound follows from the inequalities (26) and (27):

    $$\forall u \in\mathbb{U}_M(a) , \quad\bigl \Vert R(t,u,a) \bigr \Vert _2 \leq\overline{R}(t,a) . $$

    However, the tightness of the bound \(\overline{R}(t,a)\) depends strongly on the tightness of the a priori enclosure \(\widetilde {\mathbb{E}}\), and a tighter \(\widetilde{\mathbb{E}}\) will typically lead to a tighter bound on the norm of the remainder term R.

  3. 3.

    The right-hand side of the ODE (28) also depends on the control parameterization defect \((u - \sum_{i=0}^{M} a_{i} \varPhi_{i})\). It follows by the orthogonality of the Φ i s that the first M+1 Gram–Schmidt coefficients of this defect are all equal to zero. Consequently, an enclosure \(\mathbb{W}_{M}(a)\) of the control parameterization defect, which is independent of u and convex (Assumption A4), is obtained as:

    $$\begin{aligned} & \quad\forall u \in\mathbb{U}_M(a) , \\ & u - \sum_{i=0}^{M} a_i \varPhi_i \in\mathbb{W}_M(a) := \left \{ w \in L^2[0,T]^{n_u} \left|\begin{array}{l} \forall i \in\{ 0, \ldots, M \}, \forall t \in[0,t], \\ 0 = \langle w , \varPhi_i \rangle_\mu\\ \displaystyle w(t) + \sum_{i=0}^{M} a_i \varPhi_i(t) \in \mathbb{F}_u(t) \end{array} \right. \right \}. \end{aligned}$$
    (30)

Summarizing the above considerations, the main idea is to regard the solution trajectory of the differential equation (28) as a function of the remainder term and of the control parameterization defect rather than as a function of the control function u. In order to formalize this change of variable, let e(t,Δ,w) denote the solution of the ODE

$$\begin{aligned} \dot{e}(t,\Delta,w) =& \bigl[ A(t) + \Delta(t) \bigr] e(t,\Delta,w) + B(t) w(t) , \quad\text{with}\ e(0,\Delta,w) = 0, \end{aligned}$$
(31)

for any functions \(w \in L^{2}[0,T]^{n_{u}}\) and \(\Delta\in L^{2}[0,T]^{n_{x} \times n_{x}}\), where we have introduced the shorthand notations \(A(t) := \frac {\partial f(y(t,a))}{\partial x}\) and B(t):=G(y(t,a)).

Proposition A.1

Let \(\overline{R}(\cdot,a)\) be a remainder bound satisfying (29), and let \(\mathbb{W}_{M}(a)\) be defined by (30). Then,

$$\begin{aligned} &\quad\forall(t,a) \in[0,T]\times\mathbb{D}_M , \\ &{\mathbb{E}}_M(t,a) \subseteq \left \{ e(t,\Delta,w) \left| \begin{array}{l} w \in\mathbb{W}_M(a) \\ \forall\tau\in[0,T] \quad\Vert\Delta(\tau) \Vert_2 \leq \overline{R}(\tau,a) \end{array} \right.\right \} , \end{aligned}$$

where e(⋅,Δ,w) denotes the solution of (31).

Proof

From the definition of the defect function d we have

$${\mathbb{E}}_M(t,a) = \bigl\{ d(t,u,a) \mid u \in\mathbb{U}_M(a) \bigr\}. $$

Moreover, the function e is defined such that

$$d(t,u,a) = e \Biggl(t, R(\cdot,u,a), u - \sum_{i=0}^{M} a_i \varPhi_i \Biggr) . $$

Therefore, the result of the proposition follows by applying the change of variables

$$R(\cdot,u,a) \rightarrow\Delta\quad\text{and} \quad u - \sum _{i=0}^{M} a_i \varPhi_i \rightarrow w $$

and using that \(\overline{R}(\cdot,a)\) and \(\mathbb{W}_{M}(a)\) are bounds on the norm of the remainder term and on the control parameterization defect, respectively. □

In order to understand the motivation behind Proposition A.1, it is helpful to interpret the differential equation (31) as a low-pass filter with uncertain but bounded gain Δ, which would filter high-frequency modes of the “noise” w. This interpretation is useful, as the control parameterization defect \(w = u - \sum_{i=0}^{M} a_{i} \varPhi_{i}\) is bounded by the set W M (a). For example, if Φ 0,Φ 1,… denote the basis functions in a standard trigonometric Fourier expansion, the set \(\mathbb{W}_{M}(a)\) contains for large M only highly oscillatory functions, as the first M+1 Fourier coefficients of the control parameterization are zero in this case. The differential equation (31) can be expected to filter out these highly oscillating modes such that e(t,Δ,w)≈0. Having this interpretation in mind, intuition suggests that we should be able to compute tight bounds on the function e(t,Δ,w), which converge for M→∞, even when the uncertain gain Δ is not known exactly.

The aim of the following considerations is to formalize this intuition by translating it into a rigorous algorithm that computes convergent enclosures of the sets \({\mathbb{E}}_{M}(\cdot,a)\) on [0,T]. To do so, the function e(t,Δ,w) is decomposed into the sum of two functions e L(t,w) and e N(t,Δ,v) such that

$$e(t,\Delta,w) = e_\mathrm{L}(t,w) + e_\mathrm{N}\bigl(t, \Delta,e_\mathrm{L}(\cdot,w)\bigr). $$

Specifically, e L(⋅,w) and e N(⋅,Δ,v) are the solutions of the following auxiliary ODEs:

$$\begin{aligned} & \quad\forall t\in[0,T] , \\ & \dot{e}_\mathrm{L}(t,w) = A(t) e_\mathrm{L}(t,w) + B(t) w(t) \quad\text{with}\ \ e_\mathrm{L}(0,w) = 0 , \end{aligned}$$
(32)
$$\begin{aligned} & \dot{e}_\mathrm{N}(t, \Delta,v) = \bigl(A(t) + \Delta(t) \bigr) e_\mathrm{N}(t,\Delta ,v) + \Delta(t) v(t) \quad\text{with} \ \ e_\mathrm{N}(t,\Delta,v) = 0 . \end{aligned}$$
(33)

Note that the output \(e_{\rm L}\) of the first ODE (32) becomes an input in the second ODE (33). Moreover, we define the sets

$$\begin{aligned} {\mathbb{E}}_M^\mathrm{L}(t,a) :=& \bigl\{ e_\mathrm{L}(t,w) \mid w \in\mathbb{W}_M(a) \bigr\} \quad\text{and} \\ {\mathbb{E}}_M^\mathrm{N}(t,a) :=& \left \{ e_\mathrm{N}(t,\Delta,v) \left|\forall\tau\in[0,t] , \quad \begin{array}{l} v(\tau) \in{\mathbb{E}}_M^\mathrm{L}(\tau,a) \\ \Vert\Delta(\tau) \Vert_2 \leq{\overline{R}}(\tau,a) \end{array} \right. \right \} . \end{aligned}$$

The following proposition is a direct consequence of Proposition A.1 and of the foregoing decomposition.

Proposition A.2

The image set \({\mathbb{E}}_{M}(t,a)\) is contained in the Minkowski sum of the sets \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) and \({\mathbb{E}}_{M}^{\mathrm{N}}(t,a)\),

$$ \forall(t,a) \in[0,T]\times\mathbb{D}_M , \quad{\mathbb{E}}_M(t,a) \subseteq{\mathbb{E}}_M^\mathrm{L}(t,a) \oplus{\mathbb{E}}_M^\mathrm{N}(t,a) . $$
(34)

The following lemma establishes that, in order to find a convergent bound on the diameter of the set \({\mathbb{E}}_{M}(t,a)\), it is sufficient to find a convergent bound on the diameter of the set \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\).

Lemma A.1

There exists a constant (t)<∞ such that

$$\forall t \in[0,T] , \quad\operatorname{diam}\bigl({\mathbb{E}}_M(t,a)\bigr) \leq\ell(t) \max_{\tau\in[0,T]} \operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{L}( \tau,a)\bigr) . $$

Proof

Since the uncertain gain Δ is bounded, there exist constants 1 and 2 such that

$$\bigl \Vert A(t) + \Delta(t) \bigr \Vert _2 \leq \ell_1 \quad\text{and} \quad\bigl \Vert \Delta(t)v(t) \bigr \Vert _2 \leq\ell_2 \frac{\operatorname{diam}({\mathbb{E}}_M^\mathrm{L}(t,a))}{2} $$

for all v with \(v(t) \in{\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\), all Δ with \(\Vert\Delta(t) \Vert_{2} \leq{\overline{R}}(t,a)\), and all t∈[0,T]. Applying Gronwall’s lemma for bounding the norm of the solution of the differential equation (33) yields the following bound on the diameter of \({\mathbb{E}}_{M}^{\mathrm{N}}(t,a)\):

$$\operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{N}(t,a)\bigr) \leq\exp( \ell_1 t ) \ell_2 \max_{\tau\in[0,T]} \operatorname{diam}\bigl({ \mathbb{E}}_M^\mathrm{L}(\tau,a)\bigr) . $$

Then from Proposition A.2 it follows that

$$\begin{aligned} \operatorname{diam}\bigl({\mathbb{E}}_M(t,a)\bigr) \leq& \operatorname {diam}\bigl({\mathbb{E}}_M^\mathrm{L}(t,a)\bigr) + \operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{N}(t,a)\bigr) \\ \leq& \bigl[ 1 + \ell_2 \exp( \ell_1 t ) \bigr] \max _{\tau\in[0,T]} \operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{L}( \tau ,a)\bigr) , \end{aligned}$$
(35)

and the result follows by defining (t):=1+ 2exp( 1 t). □

In the remainder of this appendix, the focus is on bounding the set \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\). We start by noting that the image sets \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) are both compact and convex. The compactness of \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) follows from Filippov’s theorem [62]. Moreover, the convexity of \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) follows from the fact that linear transformations of convex sets are convex, recalling that the set \(\mathbb{W}_{M}(a)\) is convex and that the functional e L(t,⋅) is linear since the ODE (32) is itself linear; see also [71]. Because any compact and convex set is uniquely characterized by its support function [72], we define the support function \(V(t,\cdot): \mathbb{R}^{n_{x}} \to\mathbb{R}\) associated with the image set \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) by

$$ \forall c \in\mathbb{R}^{n_x} , \quad V(t,c) := \max _{w \in \mathbb{W}_M(a)} c^\mathsf {T}e_\mathrm{L}(t,w) . $$
(36)

It is not hard to see from (36) that

$$\sup_{c} \frac{V(t,c)}{\Vert c \Vert_2} \geq\frac{1}{2} \operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{L} ( t,a ) \bigr) , $$

thereby providing a means for bounding the diameter of \({\mathbb{E}}_{M}^{\mathrm{L}} ( t,a )\). In expanded form, V(t,c) reads

$$\begin{aligned} V(t,c) :=& \displaystyle\max_{w\in L^2[0,T]^{n_u}}\ c^\mathsf {T}\int _0^t Z(t,\tau) B(\tau) w(\tau) \,\mathrm{d}\tau \\ &\text{s.t.} \quad \left \{ \begin{array}{l} \forall i \in\{ 0, \ldots, M \}\quad \forall t \in[0,t] ,\\ 0 = \langle w , \varPhi_i \rangle_\mu, \\ \displaystyle w(t) + \sum_{i=0}^{M} a_i \varPhi_i(t) \in \mathbb{F}_u(t), \end{array} \right . \end{aligned}$$
(37)

with \(Z: [0,T] \times[0,T] \to\mathbb{R}^{n_{x} \times n_{x}}\) the fundamental solution of the parametric linear ODE (32), which satisfies

$$\forall\tau,t \in[0,T] , \quad\frac{\partial}{\partial t} Z(t,\tau) = A(t) Z(t,\tau) \quad\text{with} \ Z(\tau,\tau) = I. $$

Since the sets \(\mathbb{F}_{u}(t)\) are compact, one can always scale the dynamic system, and we shall therefore restrict the analysis to the L 2-integrable functions w with \(w(t)+\sum_{i=0}^{M} a_{i} \varPhi _{i}(t)\in\mathbb{F}_{u}(t)\) for all t∈[0,T] and such that

$$\Vert w \Vert_{\mu,2} := \int_0^T w(t)^\mathsf {T}w(t) \mu(t) \,\mathrm{d}t \leq1 . $$

Let the functions \(H_{t}: [0,T] \to\mathbb{R}^{n_{x} \times n_{x}}\) and \(\theta: [0,T] \times[0,T] \to\mathbb{R}\) be defined by

$$H_{t}(\tau) := \frac{1}{\mu(\tau)} Z(t,\tau) B(\tau) \theta (t,\tau) \quad\text{and} \quad\theta(t,\tau) := \left \{ \begin{array}{l@{\quad}l} 1 & \text{if $\tau\leq t$} , \\ 0 & \text{otherwise} , \end{array} \right . $$

for all t,τ∈[0,T]. The objective function in (38) can be written in the form

$$\begin{aligned} c^\mathsf {T}\int_0^t Z(t,\tau) B(\tau) w( \tau) \,\mathrm{d}\tau =& c^\mathsf {T}\int_0^T \frac{1}{\mu(\tau)} Z(t,\tau) B(\tau) \theta(t,\tau) w(\tau) \mu(\tau) \,\mathrm{d} \tau\\ =& \bigl\langle c^\mathsf {T}H_{t} , w \bigr \rangle_\mu, \end{aligned}$$

and an upper bound on the support function V(t,c) is given by

$$\begin{aligned} V(t,c) \leq& \max_{w} \bigl\langle c^\mathsf {T}H_{t} , w \bigr\rangle_\mu\quad \text{s.t.} \quad \left \{ \begin{array}{l@{\quad}l} 0 = \langle w , \varPhi_i \rangle_\mu& \forall i \in\{ 0, \ldots, M \} , \\ \Vert w \Vert_{\mu,2} \leq 1. \end{array} \right . \end{aligned}$$
(38)

Associating with each equality constraint 0=〈w,Φ i μ in (38) the scaled multipliers \(c^{\mathsf {T}}D_{i} \in\mathbb{R}^{n_{u}}\), the dual of the convex problem (38) gives

$$\begin{aligned} V(t,c) \leq& \inf_{D} \max_w \bigl \langle c^\mathsf {T}H_{t} , w \bigr\rangle_\mu- \sum _{i=0}^M c^\mathsf {T}D_i \langle\varPhi_i , w \rangle_\mu\quad\text{s.t.} \quad \Vert w \Vert_{\mu,2} \leq1 \\ = & \inf_{D} \Biggl \Vert c^\mathsf {T}\Biggl( H_{t} - \sum_{i=0}^M D_i \varPhi_i \Biggr) \Biggr \Vert _{\mu,2} = \Biggl \Vert c^\mathsf {T}\Biggl( H_{t} - \sum _{i=0}^M \frac{\langle H_{t} , \varPhi_i \rangle_\mu}{\sigma_i} \varPhi_i \Biggr) \Biggr \Vert _{\mu,2} , \end{aligned}$$
(39)

where the tight version of the Cauchy–Schwarz inequality for L 2-scalar products has been used in the last equality. At this point, it becomes clear that the derived bound on the support function V(t,c) depends crucially on how accurately the function H t can be approximated with the function \(\sum_{i=0}^{M} D_{i} \varPhi_{i}\). This observation is formalized in the following theorem.

Theorem A.1

Let Assumption 4.1 and blanket Assumptions A2 and A4 be satisfied. Then, there exist constants \(C_{E}^{L,0} \in\mathbb{R}\) and \(C_{E}^{L,1} \in\mathbb{R}_{++}\) such that the condition

$$\forall(t,a) \in[0,T]\times\mathbb{D}_M , \quad\log\bigl( \operatorname{diam}\bigl({ \mathbb{E}}_M^\mathrm{L} ( t,a ) \bigr) \bigr) \leq C_E^{L,0} - C_E^{L,1} M $$

is satisfied for all \(M \in\mathbb{N}\).

Proof

Since the function H t is piecewise smooth and, by Assumption 4.1, there exist a sequence \(D_{0},D_{1},\ldots\in\mathbb{R}^{n_{x} \times n_{u}}\) and constants \(C_{H}^{0} \in\mathbb{R}\) and \(C_{H}^{1} \in \mathbb{R}_{++}\) such that

$$\begin{aligned} \log\Biggl( \Biggl \Vert H_{t}(\tau) - \sum _{i=0}^M D_i \varPhi_i(\tau) \Biggr \Vert \Biggr) \leq& C_H^0 - C_H^1 M \end{aligned}$$
(40)

for almost all τ∈[0,T] and all \(M \in\mathbb{N}\). By combining (38) and (40), it follows that there exists a constant \(C_{\mu}\in\mathbb{R}_{++}\) such that

$$\begin{aligned} \log\bigl( \operatorname{diam}\bigl({\mathbb{E}}_M^\mathrm{L} ( t,a ) \bigr) \bigr) \leq\log\biggl( 2 \sup_{c} \frac{V(t,c)}{\Vert c \Vert_2} \biggr) \leq& 2 C_\mu\bigl[ C_H^0 - C_H^1 M \bigr] \end{aligned}$$
(41)

for almost all τ∈[0,T] and all \(M \in\mathbb{N}\). In the last step, we have used the Lebesgue dominated convergence theorem (or alternatively Fatou’s lemma), which guarantees that the Lebesgue zero measure of points τ∈[0,T] at which inequality (40) may be violated does not contribute to the bound on the L 2-norm of the function \(H_{t} - \sum_{i=0}^{M} D_{i} \varPhi_{i}\). The statement of the theorem follows by taking \(C_{E}^{L,0} := 2 C_{\mu}C_{H}^{0}\) and \(C_{E}^{L,1} := 2 C_{\mu}C_{H}^{1}\). □

Finally, a proof of Theorem 4.1 is obtained by combining the results in Theorem A.1 and Lemma A.1.

Appendix B: Computation of Convergent Enclosures \(\overline{\mathbb{E}}_{M}(t,\mathbb{A})\)

Since the proof in Appendix A is essentially constructive, a line-by-line transcription of this proof into a numerical bounding procedure that constructs an enclosure \(\overline {\mathbb{E}}_{M}(t,\mathbb{A})\) satisfying Assumption 5.2 is in principle possible, assuming that suitable interval tools are available. For the implementation in this paper, we have used the bounding and relaxation techniques available through the library MC++ [50, 58], and we have refined the procedure based on the following observations:

  1. 1.

    From inequality (39) we know that the set \({\mathbb{E}}_{M}^{\mathrm{L}}(t,a)\) is enclosed by an ellipsoid of the form \(\mathcal{E}(Q(t,a))\) with

    $$ Q(t,a) := \int_0^T \Biggl( H_{t}( \tau) - \sum_{i=0}^M \frac {\langle H_{t}, \varPhi_i \rangle_\mu}{\sigma_i} \varPhi_i(\tau) \Biggr) \Biggl( H_{t}(\tau) - \sum _{i=0}^M \frac{\langle H_{t}, \varPhi_i \rangle_\mu }{\sigma_i} \varPhi_i(\tau) \Biggr)^\mathsf {T}\,\mathrm{d}\tau. $$

    By using standard tools from interval analysis for differential equations, a convergent bounding matrix \(\mathbb{Q}(t,\mathbb{A}) \in \mathbb{R}^{n_{x} \times n_{x}}\) can be constructed such that \(Q(t,a) \preceq \mathbb{Q}(t,\mathbb{A})\) for all \(a \in\mathbb{A}\) and all t∈[0,T]. In particular, if there exists a constant C Q such that

    $$\bigl \Vert \mathbb{Q}(t,\mathbb{A}) - \mathbb{Q}\bigl(t,\{ \operatorname {mid}(\mathbb{A}) \} \bigr) \bigr \Vert \leq C_Q \operatorname{diam}(\mathbb{A}) , $$

    then the enclosure \(\overline{\mathbb{E}}_{M}^{\mathrm{L}}(t,\mathbb{A}) := \mathcal{E}( \mathbb{Q}(t,\mathbb{A}) )\) will satisfy the convergence conditions

    $$\begin{aligned} \log\bigl(\operatorname{diam}\bigl(\overline {\mathbb{E}}_M^\text{L}\bigl(t, \{ \operatorname {mid}(\mathbb{A}) \} \bigr)\bigr) \bigr ) \leq& C_{\mathbb{E}}^{\text{L},0} - C_{\mathbb{E}}^{\text {L},1} M , \end{aligned}$$
    (42)
    $$\begin{aligned} \operatorname{diam}\bigl(\overline{\mathbb{E}}_M^\text{L} ( t, \mathbb{A} ) \bigr) - \operatorname{diam}\bigl(\overline {\mathbb{E}}_M^\text{L}\bigl(t, \{ \operatorname {mid}(\mathbb{A}) \}\bigr)\bigr) \leq& C_{\mathbb{E}}^{\text{L},2} \operatorname{diam}(\mathbb{A}) , \end{aligned}$$
    (43)

    for all compact sets \(\mathbb{A}\), for all \(M \in\mathbb{N}\), and for some constants \(C_{\mathbb{E}}^{\text{L},0} \in\mathbb{R}\), \(C_{\mathbb{E}}^{\text{L},1} \in\mathbb{R}_{++}\), and \(C_{\mathbb{E}}^{\text{L},2} \in\mathbb{R}_{+}\).

  2. 2.

    The proof of Lemma A.1 is based on Gronwall’s lemma, which is known to provide mathematically valid, yet typically conservative, bounds for practical purposes. Instead, our implementation considers a modified version of state-of-the-art reachable set enclosure algorithms for bounding the solution of the differential equation (33), which prove to be much less conservative than with Gronwall’s lemma.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Houska, B., Chachuat, B. Branch-and-Lift Algorithm for Deterministic Global Optimization in Nonlinear Optimal Control. J Optim Theory Appl 162, 208–248 (2014). https://doi.org/10.1007/s10957-013-0426-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-013-0426-1

Keywords

Navigation