Skip to main content
Log in

Convergence Rate for a Gauss Collocation Method Applied to Unconstrained Optimal Control

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

Abstract

A local convergence rate is established for an orthogonal collocation method based on Gauss quadrature applied to an unconstrained optimal control problem. If the continuous problem has a smooth solution and the Hamiltonian satisfies a strong convexity condition, then the discrete problem possesses a local minimizer in a neighborhood of the continuous solution, and as the number of collocation points increases, the discrete solution converges to the continuous solution at the collocation points, exponentially fast in the sup-norm. Numerical examples illustrating the convergence theory are provided.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  1. Benson, D.A., Huntington, G.T., Thorvaldsen, T.P., Rao, A.V.: Direct trajectory optimization and costate estimation via an orthogonal collocation method. J. Guid. Control Dyn. 29(6), 1435–1440 (2006)

    Article  Google Scholar 

  2. Garg, D., Patterson, M.A., Hager, W.W., Rao, A.V., Benson, D.A., Huntington, G.T.: A unified framework for the numerical solution of optimal control problems using pseudospectral methods. Automatica 46, 1843–1851 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  3. Elnagar, G., Kazemi, M., Razzaghi, M.: The pseudospectral Legendre method for discretizing optimal control problems. IEEE Trans. Automat. Control 40(10), 1793–1796 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  4. Fahroo, F., Ross, I.M.: Costate estimation by a Legendre pseudospectral method. J. Guid. Control Dyn. 24(2), 270–277 (2001)

    Article  Google Scholar 

  5. Elnagar, G.N., Kazemi, M.A.: Pseudospectral Chebyshev optimal control of constrained nonlinear dynamical systems. Comput. Optim. Appl. 11(2), 195–217 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  6. Fahroo, F., Ross, I.M.: Direct trajectory optimization by a Chebyshev pseudospectral method. J. Guid. Control Dyn. 25, 160–166 (2002)

    Article  Google Scholar 

  7. Fahroo, F., Ross, I.M.: Pseudospectral methods for infinite-horizon nonlinear optimal control problems. J. Guid. Control Dyn. 31(4), 927–936 (2008)

    Article  Google Scholar 

  8. Garg, D., Patterson, M.A., Darby, C.L., Françolin, C., Huntington, G.T., Hager, W.W., Rao, A.V.: Direct trajectory optimization and costate estimation of finite-horizon and infinite-horizon optimal control problems using a Radau pseudospectral method. Comput. Optim. Appl. 49(2), 335–358 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  9. Liu, F., Hager, W.W., Rao, A.V.: Adaptive mesh refinement method for optimal control using nonsmoothness detection and mesh size reduction. J. Frankl. Inst. 352, 4081–4106 (2015)

    Article  MathSciNet  Google Scholar 

  10. Patterson, M.A., Hager, W.W., Rao, A.V.: A \(ph\) mesh refinement method for optimal control. Optim. Control Appl. Meth. 36, 398–421 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Williams, P.: Jacobi pseudospectral method for solving optimal control problems. J. Guid. Control Dyn. 27(2), 293–297 (2004)

    Article  Google Scholar 

  12. Kang, W.: The rate of convergence for a pseudospectral optimal control method. In: Proceeding of the 47th IEEE Conference on Decision and Control, IEEE, pp. 521–527 (2008)

  13. Kang, W.: Rate of convergence for the Legendre pseudospectral optimal control of feedback linearizable systems. J. Control Theory Appl. 8, 391–405 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gong, Q., Ross, I.M., Kang, W., Fahroo, F.: Connections between the covector mapping theorem and convergence of pseudospectral methods for optimal control. Comput. Optim. Appl. 41(3), 307–335 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  15. Dontchev, A.L., Hager, W.W.: Lipschitzian stability in nonlinear control and optimization. SIAM J. Control Optim. 31, 569–603 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  16. Dontchev, A.L., Hager, W.W.: The Euler approximation in state constrained optimal control. Math. Comput. 70, 173–203 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  17. Dontchev, A.L., Hager, W.W., Veliov, V.M.: Second-order Runge–Kutta approximations in constrained optimal control. SIAM J. Numer. Anal. 38, 202–226 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  18. Dontchev, A.L., Hager, W.W., Malanowski, K.: Error bounds for Euler approximation of a state and control constrained optimal control problem. Numer. Funct. Anal. Optim. 21, 653–682 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  19. Hager, W.W.: Runge-Kutta methods in optimal control and the transformed adjoint system. Numer. Math. 87, 247–282 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  20. Kameswaran, S., Biegler, L.T.: Convergence rates for direct transcription of optimal control problems using collocation at Radau points. Comput. Optim. Appl. 41(1), 81–126 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  21. Reddien, G.W.: Collocation at Gauss points as a discretization in optimal control. SIAM J. Control Optim. 17(2), 298–306 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  22. Nocedal, J., Wright, S.J.: Numerical Optimization, 2nd edn. Springer, New York (2006)

    MATH  Google Scholar 

  23. Hager, W.W.: Numerical analysis in optimal control. In: Hoffmann, K.H., Lasiecka, I., Leugering, G., Sprekels, J., Tröltzsch, F. (eds.) International Series of Numerical Mathematics, vol. 139, pp. 83–93. Birkhauser Verlag, Basel (2001)

    Google Scholar 

  24. Hager, W.W., Hou, H., Rao, A.V.: Lebesgue constants arising in a class of collocation methods (2015). arXiv:1507.08316

  25. Elschner, J.: The h-p-version of spline approximation methods for Mellin convolution equations. J. Integral Equ. Appl. 5(1), 47–73 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  26. Horn, R.A., Johnson, C.R.: Matrix Analysis. Cambridge University Press, Cambridge (2013)

    MATH  Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge support by the Office of Naval Research under Grants N00014-11-1-0068 and N00014-15-1-2048 and by the National Science Foundation under Grants DMS-1522629 and CBET-1404767. Comments and suggestions from the reviewers are gratefully acknowledged. In particular, in the initial draft of this paper, it was assumed that \(\nabla ^2 C(\mathbf{{x}}^*(1))\) was positive definite, while one of the reviewers correctly pointed out that this assumption could be relaxed to positive semidefinite without effecting the convergence results for a stationary point of the discrete problem.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to William W. Hager.

Appendix

Appendix

In (15), we define a new matrix \(\mathbf{{D}}^\dagger \) in terms of the differentiation matrix \(\mathbf{{D}}\). The following proposition shows that the elements of \(\mathbf{{D}}^\dagger \) can be gotten by rearranging the elements of \(\mathbf{{D}}\).

Proposition 10.1

The entries of the matrices \(\mathbf{{D}}\) and \(\mathbf{{D}}^\dagger \) satisfy

$$\begin{aligned} D_{ij} = -D_{N+1-i, N+1-j}^\dagger , \quad 1 \le i \le N, \quad 1 \le j \le N. \end{aligned}$$
(60)

In other words, \(\mathbf{{D}}_{1:N} = -\mathbf{{J}} \mathbf{{D}}_{1:N}^\dagger \mathbf{{J}}\) where \(\mathbf{{J}}\) is the exchange matrix with ones on its counterdiagonal and zeros elsewhere. Equivalently, \(\mathbf{{D}}_{1:N}^\dagger = -\mathbf{{J}} \mathbf{{D}}_{1:N} \mathbf{{J}}\).

Proof

By (9) the elements of \(\mathbf{{D}}\) can be expressed in terms of the derivatives of a set of Lagrange basis functions evaluated at the collocation points:

$$\begin{aligned} D_{ij} = \dot{L}_j(\tau _i) \quad \text{ where } L_j \in \mathcal{{P}}_N, \quad L_j (\tau _k) = \left\{ \begin{array}{ll} 1 &{} \text{ if } k = j, \\ 0 &{} \text{ if } 0 \le k \le N, \; k \ne j. \end{array} \right. \end{aligned}$$

In (9) we give an explicit formula for the Lagrange basis functions, while here we express the basis function in terms of polynomials \(L_j\) that equal one at \(\tau _j\) and vanish at \(\tau _k\) where \(0 \le k \le N\), \(k \ne j\). These \(N +1\) conditions uniquely define \(L_j \in \mathcal{{P}}_N\). Similarly, by Garg et al. [2, Thm. 1], the entries of \(\mathbf{{D}}_{1:N}^\dagger \) are given by

$$\begin{aligned} D_{ij}^\dagger = \dot{M}_j(\tau _i) \quad \text{ where } M_j \in \mathcal{{P}}_N, \quad M_j (\tau _k) = \left\{ \begin{array}{ll} 1 &{} \text{ if } k = j, \\ 0 &{} \text{ if } 1 \le k \le N+1, \; k \ne j . \end{array} \right. \end{aligned}$$

Observe that \(M_{N+1-j}(t) = L_j (-t)\) due to the symmetry of the quadrature points around \(t = 0\):

  1. (a)

    Since \(-\tau _{N+1-j} = \tau _j\), we have \(L_j (-\tau _{N+1-j}) = L_j (\tau _j) = 1\).

  2. (b)

    Since \(\tau _{N+1} = 1\) and \(\tau _0 = -1\), we have \(L_j (-\tau _{N+1}) = L_j (\tau _0) = 0\).

  3. (c)

    Since \(-\tau _i = \tau _{N+1-i}\), we have \(L_j (-\tau _{i}) = L_j (\tau _{N+1-i}) = 0\) if \(i \ne N+1-j\).

Since \(M_{N+1-j}(t)\) is equal to \(L_j (-t)\) at \(N+1\) distinct points, the two polynomials are equal everywhere. Replacing \(M_{N+1-j}(t)\) by \(L_j (-t)\), we have

$$\begin{aligned} D_{N+1-i, N+1-j}^\dagger = -\dot{L}_j (-\tau _{N+1-i}) = - \dot{L}_j (\tau _i) =-D_{ij}. \end{aligned}$$

\(\square \)

Tables 1 and 2 illustrate properties (P1) and (P2) for the differentiation matrix \(\mathbf{{D}}\). In Table 1, we observe that \(\Vert \mathbf{{D}}_{1:N}^{-1}\Vert _\infty \) monotonically approaches the upper limit 2. More precisely, it is found that \(\Vert \mathbf{{D}}_{1:N}^{-1}\Vert _\infty = 1 + \tau _N\), where the final collocation point \(\tau _N\) approaches one as N tends to infinity. In Table 2, we give the maximum 2-norm of the rows of \([\mathbf{{W}}^{1/2} \mathbf{{D}}_{1:N}]^{-1}\). It is found that the maximum is attained by the last row of \([\mathbf{{W}}^{1/2} \mathbf{{D}}_{1:N}]^{-1}\), and the maximum monotonically approaches \(\sqrt{2}\).

Table 1 \(\Vert \mathbf{D}_{1:N}^{-1}\Vert _\infty \)
Table 2 Maximum Euclidean norm for the rows of \([\mathbf{{W}}^{1/2}{} \mathbf{D}_{1:N}]^{-1}\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hager, W.W., Hou, H. & Rao, A.V. Convergence Rate for a Gauss Collocation Method Applied to Unconstrained Optimal Control. J Optim Theory Appl 169, 801–824 (2016). https://doi.org/10.1007/s10957-016-0929-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-016-0929-7

Keywords

Mathematics Subject Classification

Navigation