Skip to main content

An Adaptive Max-Plus Eigenvector Method for Continuous Time Optimal Control Problems

  • Chapter
  • First Online:
Numerical Methods for Optimal Control Problems

Part of the book series: Springer INdAM Series ((SINDAMS,volume 29))

Abstract

An adaptive max-plus eigenvector method is proposed for approximating the solution of continuous time nonlinear optimal control problems. At each step of the method, given a set of quadratic basis functions, a standard max-plus eigenvector method is applied to yield an approximation of the value function of interest. Using this approximation, an approximate level set of the back substitution error defined by the Hamiltonian is tessellated according to where each basis function is active in approximating the value function. The polytopes obtained, and their vertices, are sorted according to this back substitution error, allowing “worst-case” basis functions to be identified. The locations of these basis functions are subsequently evolved to yield new basis functions that reduce this worst-case. Basis functions that are inactive in the value function approximation are pruned, and the aforementioned steps repeated. Underlying algebraic properties associated with max-plus linearity, dynamic programming, and semiconvex duality are provided as a foundation for the development, and the utility of the proposed method is illustrated by example.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. McEneaney, W.: Max-Plus Methods for Nonlinear Control and Estimation. Systems & Control: Foundations & Application. Birkhauser, Basel (2006)

    Google Scholar 

  2. McEneaney, W.: A curse-of-dimensionality-free numerical method for solution of certain HJB PDEs. SIAM J. Control Optim. 46(4), 1239–1276 (2007)

    Article  MathSciNet  Google Scholar 

  3. Akian, M., Gaubert, S., Lakhoua, A.: The max-plus finite element method for solving deterministic optimal control problems: basic properties and convergence analysis. SIAM J. Control Optim. 47(2), 817–848 (2008)

    Article  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

  5. McEneaney, W.: A new fundamental solution for differential Riccati equations arising in control. Automatica 44, 920–936 (2008)

    Article  MathSciNet  Google Scholar 

  6. Dower, P., McEneaney, W., Zhang, H.: Max-plus fundamental solution semigroups for optimal control problems. In: Proceedings of SIAM Conference on Control Theory and Its Applications (Paris), 2015, pp. 368–375 (2015)

    Google Scholar 

  7. Qu, Z.: A max-plus based randomized algorithm for solving a class of HJB PDEs. In: Proceedings of 53rd IEEE Conference on Decision and Control (Los Angeles, CA) (2014)

    Google Scholar 

  8. Dower, P.: An approximation arising in max-plus based optimal stopping. In: Proceedings of Australian Control Conference (Sydney), pp. 271–276 (2012)

    Google Scholar 

  9. Grune, L., Dower, P.: Hamiltonian based a posteriori error estimation for Hamilton-Jacobi-Bellman equations, Technical report. Universitat Bayreuth. https://epub.uni-bayreuth.de/id/eprint/3578 (2018)

  10. Rockafellar, R.: Conjugate Duality and Optimization. SIAM Regional Conference Series in Applied Mathematics, vol. 16. SIAM, Philadelphia (1974)

    Google Scholar 

  11. Baccelli, F., Cohen, G., Olsder, G., Quadrat, J.-P.: Synchronization and Linearity. Wiley, New York (1992)

    MATH  Google Scholar 

  12. Kolokoltsov, V., Maslov, V.: Idempotent Analysis and Applications. Kluwer Publishing House, Dordrecht (1997)

    Book  Google Scholar 

  13. Litvinov, G., Maslov, V., Shpiz, G.: Idempotent functional analysis: an algebraic approach. Math. Notes 69(5), 696–729 (2001)

    Article  MathSciNet  Google Scholar 

  14. Cohen, G., Gaubert, S., Quadrat, J.-P.: Duality and separation theorems in idempotent semimodules. Linear Algebra Appl. 379, 395–422 (2004)

    Article  MathSciNet  Google Scholar 

  15. Fleming, W., McEneaney, W.: A max-plus-based algorithm for a Hamilton-Jacobi-Bellman equation of nonlinear filtering. SIAM J. Control Optim. 38(3), 683–710 (2000)

    Article  MathSciNet  Google Scholar 

  16. McEneaney, W., Dower, P.: The principle of least action and fundamental solutions of mass-spring and n-body two-point boundary value problems. SIAM J. Control Optim. 53(5), 2898–2933 (2015)

    Article  MathSciNet  Google Scholar 

  17. Zhang, H., Dower, P.: Max-plus fundamental solution semigroups for a class of difference Riccati equations. Automatica 52, 103–110 (2015)

    Article  MathSciNet  Google Scholar 

  18. Dower, P., McEneaney, W.: A max-plus dual space fundamental solution for a class of operator differential Riccati equations. SIAM J. Control Optim. 53(2), 969–1002 (2015)

    Article  MathSciNet  Google Scholar 

  19. Dower, P., Zhang, H.: A max-plus primal space fundamental solution for a class of differential Riccati equations. Math. Control Signals Syst. 29(3), 1–33 (2017) [Online]. http://dx.doi.org/10.1007/s00498-017-0200-2

  20. Dower, P., McEneaney, W.: Solving two-point boundary value problems for a wave equation via the principle of stationary action and optimal control. SIAM J. Control Optim. 55(4), 2151–2205 (2017)

    Article  MathSciNet  Google Scholar 

  21. Dower, P.: Basis adaptation for a max-plus eigenvector method arising in optimal control. In: Proceedings of 23rd International Symposium on Mathematical Theory of Networks and Systems (Hong Kong), pp. 350–355 (2018)

    Google Scholar 

  22. Avis, D., Fukuda, K.: A pivoting algorithm for convex hulls and vertex enumeration of arrangements and polyhedra. Discret. Comput. Geom. 8, 295–313 (1992)

    Article  MathSciNet  Google Scholar 

  23. Bremner, D., Fukuda, K., Marzetta, A.: Primal-dual methods for vertex and facet enumeration. Discret. Comput. Geom. 20, 333–357 (1998)

    Article  MathSciNet  Google Scholar 

  24. de Berg, M., Cheong, O., van Kreveld, M., Overmars, M.: Computational Geometry: Algorithms and Applications, 3rd ed. Springer, Berlin (2008)

    Book  Google Scholar 

  25. Rockafellar, R., Wets, R.: Variational Analysis. Springer, Berlin (1997)

    MATH  Google Scholar 

Download references

Acknowledgements

This research was partially supported by AFOSR/AOARD grant FA2386-16-1-4066.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter M. Dower .

Editor information

Editors and Affiliations

Appendix

Appendix

In the statement of Theorem 3, the semiconvex transform D φ is as specified by (12), and its candidate inverse is as per (13). For convenience, formally define

(54)

in which \({{\mathcal {R}}_-^{{\mathsf {{K}}}}}\) is as per (14).

Lemma 16

WithK ∈ Σ M ∪{M} fixed, D φ and \({\mathsf {{D}}}_\varphi ^{\sharp }\) of (12) and (54) satisfy the following properties:

  1. 1)

    \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^\sharp ) {{ \ = {\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ) = {{{{\mathcal {R}}_-^{{\mathsf {{K}}}}}}} \subset {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}}}\);

  2. 2)

    \({\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^{\sharp }) {{ \ = {\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ) = {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}} \subset {{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}}}\);

  3. 3)

    \({\mathsf {{D}}}_\varphi \, {\mathsf {{D}}}_\varphi ^{\sharp } = {\mathsf {{I}}}\) on \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^{\sharp }) = {{\mathcal {R}}_-^{{\mathsf {{K}}}}}\);

  4. 4)

    \({\mathsf {{D}}}_\varphi ^{\sharp }\, {\mathsf {{D}}}_\varphi = {\mathsf {{I}}}\) on \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ) = {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\);

  5. 5)

    \({{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}} = {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\);

  6. 6)

    IfK > M then \({{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\setminus {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}} \ne \emptyset \) and \({{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\setminus {{\mathcal {R}}_-^{{\mathsf {{K}}}}} \ne \emptyset \).

The following observations are useful in establishing Lemma 16.

  1. 1.

    Given invertible M ∈ Σ,

    $$\displaystyle \begin{aligned} & \langle x,\, z \rangle + {\,{\frac{1}{2}}}\, \langle x,\, {\mathsf{{M}}}\, x\rangle = {\,{\frac{1}{2}}}\, \langle x - \zeta(z),\, {\mathsf{{M}}} \, (x - \zeta(z)) \rangle - {\,{\frac{1}{2}}}\, \langle \zeta(z),\, {\mathsf{{M}}}\, \zeta(z) \rangle {} \end{aligned} $$
    (55)

    for all \(x,z\in {\mathbb {R}}^n\), where .

  2. 2.

    Given O 1, O 2 ∈ Σ satisfying O 1 < O 2,

    $$\displaystyle \begin{aligned} {{{\mathcal{S}}_+^{{{{{\,\mathsf{{O}}}_1}}}}}} \subset {{{\mathcal{S}}_+^{{{{{\,\mathsf{{O}}}_2}}}}}}\,, \qquad {{{\mathcal{S}}_-^{{{{{\,\mathsf{{O}}}_1}}}}}} \subset {{{\mathcal{S}}_-^{{{{{\,\mathsf{{O}}}_2}}}}}}\,. {} \end{aligned} $$
    (56)

Proof of Lemma 16

Assertion 1) By definitions (14) and (54),

$$\displaystyle \begin{aligned} {\mathrm{ran}\,}({\mathsf{{D}}}_\varphi) = {{\mathcal{R}}_-^{{\mathsf{{K}}}}} = {\mathrm{dom}\,}({\mathsf{{D}}}_\varphi^\sharp). {} \end{aligned} $$

Fix an arbitrary \(a\in {{\mathcal {R}}_-^{{\mathsf {{K}}}}}\). Applying definition (14), there exists a \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) such that D φ ψ = a. Define \(\psi _+,\psi _\pm :{\mathbb {R}}^n{\rightarrow }{{{{\overline {{\mathbb {R}}}}}}}\) by

(57)

for all \(x\in {\mathbb {R}}^n\), and note that

$$\displaystyle \begin{aligned} \psi_\pm(x) & = \psi(x) - {\,{\frac{1}{2}}}\, \langle x,\, {\mathsf{{M}}}\, x\rangle {} \end{aligned} $$
(58)

for all \(x\in {\mathbb {R}}^n\). By inspection of (57), ψ + is convex and lower closed on \({\mathbb {R}}^n\), as \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\). Hence, ψ ± is also convex and lower closed on \({\mathbb {R}}^n\), as − (K + M) ≥ 0 by definition of −K ∈ Σ M ∪{M}. Hence, the convex conjugate \(\psi _\pm ^*:{\mathbb {R}}^n{\rightarrow }{{{{\overline {{\mathbb {R}}}}}}}\) of ψ ± is also convex and lower closed [25, Theorem 5, p. 16], with

(59)

for all \(z\in {\mathbb {R}}^n\). Define \(\widetilde a:{\mathbb {R}}^n{\rightarrow }{{{{\overline {{\mathbb {R}}}}}}}\) by

(60)

for all \(\zeta \in {\mathbb {R}}^n\). Note that \(\widetilde a\) is upper closed, as \(\psi _\pm ^*\) and \({\, {\frac {1}{2}}}\, \langle \cdot ,\, {\mathsf {{M}}}\, \cdot \rangle \) are lower closed. Furthermore, \(\zeta \mapsto \tilde a(\zeta ) - {\, {\frac {1}{2}}}\, \langle \zeta , \, -{\mathsf {{M}}}\, \zeta \rangle \) is concave, as \(\psi _\pm ^*\) is convex. Hence, \(\widetilde a\in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\). Recalling that \(\psi _\pm ^*\) is the convex conjugate of ψ ±, (58), (60) imply that

$$\displaystyle \begin{aligned} \widetilde a(\zeta) & {{\ = - \psi_\pm^*(-{\mathsf{{M}}}\, \zeta) - {\,{\frac{1}{2}}}\, \langle \zeta,\, {\mathsf{{M}}}\, \zeta \rangle}} = - \int_{{\mathbb{R}}^n}^\oplus \langle -{\mathsf{{M}}}\, \zeta,\, x \rangle \otimes (-\psi_\pm(x)) \, dx - {\,{\frac{1}{2}}} \, \langle \zeta,\, {\mathsf{{M}}}\, \zeta \rangle {}\\ & = -\int_{{\mathbb{R}}^n}^\oplus {\,{\frac{1}{2}}} \, \langle x,\, {\mathsf{{M}}}\, x \rangle - \langle {\mathsf{{M}}}\, \zeta,\, x \rangle + {\,{\frac{1}{2}}} \, \langle \zeta,\, {\mathsf{{M}}}\, \zeta \rangle -\psi(x) \, dx {}\\ & = ({\mathsf{{D}}}_\varphi\, \psi)(\zeta) = a(\zeta) {} \end{aligned} $$
(61)

for all \(\zeta \in {\mathbb {R}}^n\). That is, \(a = \widetilde a\in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\), and as \(a\in {{\mathcal {R}}_-^{{\mathsf {{K}}}}}\) is arbitrary, \({{\mathcal {R}}_-^{{\mathsf {{K}}}}} \subset {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\).

2) By definitions (12) and (54), \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ) = {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) and \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^\sharp ) = {{\mathcal {R}}_-^{{\mathsf {{K}}}}}\). Fix an arbitrary \(a\in {\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^\sharp ) = {{\mathcal {R}}_-^{{\mathsf {{K}}}}}\). Following the proof of Assertion 1), there exists a \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) such that D φ ψ = a, defining a convex and lower closed \(\psi _\pm :{\mathbb {R}}^n\rightarrow {{{{\overline {{\mathbb {R}}}}}}}\) as per (57), (58). Note further that \(a = \widetilde a\in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\), where \(\widetilde a\) is defined as per (60). As \(\psi _\pm = \psi _\pm ^{**}\), see [25, Theorem 5, p. 16], (54), (58) and (60) imply that

$$\displaystyle \begin{aligned} \psi(x) & = (\psi_\pm^*)^*(x) + {\,{\frac{1}{2}}}\, \langle x,\, {\mathsf{{M}}}\, x\rangle = \int_{{\mathbb{R}}^n}^\oplus \langle x,\, z\rangle + {\,{\frac{1}{2}}}\, \langle x,\, {\mathsf{{M}}}\, x\rangle - \psi_\pm^*(z)\, dz {}\\ & = \int_{{\mathbb{R}}^n}^\oplus {\,{\frac{1}{2}}}\, \langle x - \zeta,\, {\mathsf{{M}}} \, (x - \zeta) \rangle + \left[ - \psi_\pm^*(-{\mathsf{{M}}}\, \zeta) - {\,{\frac{1}{2}}}\, \langle \zeta,\, {\mathsf{{M}}}\, \zeta \rangle \right] \, d\zeta {}\\ & \,{{ = \int_{{\mathbb{R}}^n}^\oplus {\,{\frac{1}{2}}}\, \langle x - \zeta,\, {\mathsf{{M}}} \, (x - \zeta) \rangle \otimes \widetilde a(\zeta)\, d\zeta = ({\mathsf{{D}}}_\varphi^{\sharp}\, \widetilde a)(x) = ({\mathsf{{D}}}_\varphi^{\sharp}\, a)(x)}} {} \end{aligned} $$
(62)

where the third equality uses (55) and the change of variable . Hence, \({\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^\sharp ) \subset {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\), as \(\psi = {\mathsf {{D}}}_\varphi ^\sharp \, a\in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\), and \(a\in {\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^\sharp )\) is arbitrary.

Alternatively, fix an arbitrary \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\), and construct ψ ± directly using (57), (58). Subsequently define \(a = \widetilde a\in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\) via (60), and note that \(\psi = {\mathsf {{D}}}_\varphi ^\sharp \, a\) as per (62). That is, \({{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\subset {\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^\sharp )\), as \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) is arbitrary. Recalling the earlier conclusion \({\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^\sharp ) \subset {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) yields \({{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}={\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^\sharp )\). Finally, as −K ∈ Σ M ∪{M}, i.e. K ≤−M, (56) implies that \({{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\subset {{{\mathcal {S}}_+^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\) as required.

3), 4) Applying Assertions 1) and 2), the compositions \({\mathsf {{D}}}_\varphi \, {\mathsf {{D}}}_\varphi ^\sharp :{{\mathcal {R}}_-^{{\mathsf {{K}}}}}{\rightarrow }{{\mathcal {R}}_-^{{\mathsf {{K}}}}}\) and \({\mathsf {{D}}}_\varphi ^\sharp \, {\mathsf {{D}}}_\varphi :{{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}{\rightarrow }{{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) are well-defined. Applying (61) and (62) yields \(a = {\mathsf {{D}}}_\varphi \, \mathsf {D}_\varphi ^\sharp \, a\) and \(\psi = \mathsf {D}_\varphi ^\sharp \, \mathsf {D}_\varphi \, \psi \) for any \(a\in {{\mathcal {R}}_-^{\mathsf {K}}} = {\mathrm{dom}\,}(\mathsf {D}_\varphi ^\sharp )\), \(\psi \in {{\mathcal {S}}_+^{\,\mathsf {K}}} = {\mathrm{dom}\,}(\mathsf {D}_\varphi )\), i.e.,

$$\displaystyle \begin{aligned} & {\mathsf{{D}}}_\varphi\, {\mathsf{{D}}}_\varphi^\sharp = {\mathsf{{I}}}\,, \qquad && {\mathrm{dom}\,}({\mathsf{{D}}}_\varphi\, {\mathsf{{D}}}_\varphi^\sharp) = {{\mathcal{R}}_-^{{\mathsf{{K}}}}}\,, {}\\ & {\mathsf{{D}}}_\varphi^\sharp\, {\mathsf{{D}}}_\varphi = {\mathsf{{I}}}\,, \qquad && {\mathrm{dom}\,}({\mathsf{{D}}}_\varphi^\sharp\, {\mathsf{{D}}}_\varphi) = {{\mathcal{S}}_+^{{\,\mathsf{{K}}}}}\,. {} \end{aligned} $$

5) Fix any \(a\in {{{\mathcal {S}}_-^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\). Define \(a_\mp :{\mathbb {R}}^n{\rightarrow }{{{{\overline {{\mathbb {R}}}}}}}\) by

$$\displaystyle \begin{aligned} a_\mp(\zeta) & = - \left[ a(\zeta) - {\,{\frac{1}{2}}}\, \langle \zeta,\, -{\mathsf{{M}}}\, \zeta\rangle \right]\,, {} \end{aligned} $$

for all \(\zeta \in {\mathbb {R}}^n\). Observe that a is convex and lower closed by definition of a, so that \(a_\mp = a_\mp ^{**}\), see [10, Theorem 5, p. 16]. Hence,

$$\displaystyle \begin{aligned} a(y) & = -(a_\mp^*)^*(y) - {\,{\frac{1}{2}}}\, \langle y,\, {\mathsf{{M}}}\, y\rangle = - \int_{{\mathbb{R}}^n}^\oplus \langle y,\, z\rangle + {\,{\frac{1}{2}}}\, \langle y,\, {\mathsf{{M}}}\, y\rangle - a_\mp^*(z)\, dz {}\\ & = - \int_{{\mathbb{R}}^n}^\oplus {\,{\frac{1}{2}}}\, \langle y - \zeta, \, {\mathsf{{M}}}\, (y - \zeta) \rangle - \left[ {\,{\frac{1}{2}}}\, \langle \zeta,\, {\mathsf{{M}}}\, \zeta \rangle + a_\mp^*({\mathsf{{M}}}\, \zeta) \right] \, d\zeta {} \end{aligned} $$
(63)

where the final equality follows by application of (55). Define \(\psi _\mp :{\mathbb {R}}^n{\rightarrow }{{{{\overline {{\mathbb {R}}}}}}}\) by

(64)

for all \(\zeta \in {\mathbb {R}}^n\). As \(a_\mp ^*\) is also convex and lower closed [10], \(\psi _\mp \in {{{\mathcal {S}}_+^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\). Substituting (64) in (63), and recalling (12) with selected in the definition of dom (D φ),

$$\displaystyle \begin{aligned} a(y) & = - \int_{{\mathbb{R}}^n}^\oplus {\,{\frac{1}{2}}}\, \langle y - \zeta, \, {\mathsf{{M}}}\, (y - \zeta) \rangle \otimes (-\psi_\mp(\zeta))\, d\zeta = ({\mathsf{{D}}}_\varphi\, \psi_\mp)(y) {} \end{aligned} $$

for all \(y\in {\mathbb {R}}^n\). Hence, \(a\in {{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\) by inspection of (14). As \(a\in {{{\mathcal {S}}_-^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\) is arbitrary, it follows immediately that \({{{\mathcal {S}}_-^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\subset {{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\). However, applying Assertions 1) and 2) for K = −M yields that \({{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\subset {{{\mathcal {S}}_-^{{\,\mathsf {{-{\mathsf {{M}}}}}}}}}\), so that the claimed property holds.

6) Let \({\mathsf {{D}}}_\varphi ^{\mathsf {{M}}}\) and \({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}\sharp }\) denote the operators D φ and \({\mathsf {{D}}}_\varphi ^\sharp \) of (12) and (54) with respective domains given by \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^{\mathsf {{M}}}) = {{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\) and \({\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}\sharp }) = {{\mathcal {R}}_-^{{\mathsf {{-M}}}}}\). Fix any −K ∈ Σ M, i.e. so that −K > M. Observe that \({{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\setminus {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\ne \emptyset \), as it contains \({\, {\frac {1}{2}}}\, \langle \cdot ,\, {\mathsf {{N}}} \, \cdot \rangle \) for any N ∈ Σ satisfying M < N < −K. Fix any \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\setminus {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\). By Assertions 1) through 5) applied to \({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}}\) and \({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}\sharp }\), there exists an \(a\in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}} = {{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\supset {{{\mathcal {R}}}_-^{{{\mathsf {{K}}}}}}\) such that \({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}}\, \psi = a\) and \({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}\sharp }\, a = \psi \). Suppose that \(a\in {{{\mathcal {R}}}_-^{{{\mathsf {{K}}}}}} = {\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^\sharp )\subset {\mathrm{dom}\,}({\mathsf {{D}}}_\varphi ^{{\mathsf {{M}}}\sharp })\). By (12) and Assertion 4), applied to D φ and \({\mathsf {{D}}}_\varphi ^\sharp \),

$$\displaystyle \begin{aligned} & {\mathsf{{D}}}_\varphi^\sharp \, a = {\mathsf{{D}}}_\varphi^\sharp \, {\mathsf{{D}}}_\varphi^{{\mathsf{{M}}}}\, \psi = {\mathsf{{D}}}_\varphi^\sharp \, {\mathsf{{D}}}_\varphi\, \psi = \psi\,. {} \end{aligned} $$

Hence, \(\psi \in {\mathrm{ran}\,}({\mathsf {{D}}}_\varphi ^\sharp ) = {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\), by assertion 2), which contradicts \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\setminus {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\). That is, \(a\in {{\mathcal {R}}_-^{{\mathsf {{-M}}}}}\) and \(a\not \in {{{\mathcal {R}}}_-^{{{\mathsf {{K}}}}}}\), so that \(a\in {{{\mathcal {R}}}_-^{{-{\mathsf {{M}}}}}}\setminus {{{\mathcal {R}}}_-^{{{\mathsf {{K}}}}}}\ne \emptyset \). \(\square \)

Corollary 17

GivenK ∈ Σ M ∪{M}, and φ as per (10), the semiconvex transform \({\mathsf {{D}}}_\varphi :{{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}{\rightarrow }{{\mathcal {R}}_-^{{\mathsf {{K}}}}}\) of (12) has a well-defined inverse \({\mathsf {{D}}}_\varphi ^{-1}:{{\mathcal {R}}_-^{{\mathsf {{K}}}}} {\rightarrow } {{\mathcal {S}}_+^{{\,\mathsf {{K}}}}}\) given by

where \({\mathsf {{D}}}_\varphi ^\sharp \) is as per (54).

Proof of Corollary 17 and Theorem 3

Immediate by inspection of Lemma 16. \(\square \)

Lemma 18

The asserted forms (16), (17) of the semiconvex transform and its inverse (12), (13) hold.

Proof

Given any \(\psi \in {{\mathcal {S}}_+^{{\,\mathsf {{-M}}}}}\), Theorem 3 implies that \({\mathsf {{D}}}_\varphi \, \psi \in {{\mathcal {S}}_-^{{\,\mathsf {{-M}}}}}\). Define \(a_\mp :{\mathbb {R}}^n\rightarrow {{{{\overline {{\mathbb {R}}}}}}}\) by

(65)

for all \(z\in {\mathbb {R}}^n\), and observe that a is convex and lower closed, i.e. a  = cl co a , see [10]. Motivated by the form of the argument of the convex hull operation in the right-hand side of (16), define \(\widetilde a_\mp :{\mathbb {R}}^n\rightarrow {{{{\overline {{\mathbb {R}}}}}}}\) by for all \(z\in {\mathbb {R}}^n\). Note by definition (8) of \(\delta _{z_i}^-\) that

$$\displaystyle \begin{aligned} \widetilde a_\mp(z) & = \left\{ \begin{array}{cl} -[{\mathsf{{D}}}_\varphi\, \psi](z_i) - \varphi(0,z_i), & z = z_i\,, \ i\in{\mathbb{N}}\,, \\ +\infty\,, & \text{otherwise}, \end{array} \right. {} \end{aligned} $$
(66)

for all \(z\in {\mathbb {R}}^n\). As \(\{z_i\}_{i\in {\mathbb {N}}}\) is dense in \({\mathbb {R}}^n\), inspection of (65) and (66) yields that \(a_\mp = {\text{cl}}^-\, {\text{co}}\ \widetilde a_\mp \). Recalling the definition of \(\widetilde a_\mp \) subsequently yields (16).

The remaining assertion (17) is an immediate consequence of (13), as \(\{z_i\}_{i\in {\mathbb {N}}}\) is dense in \({\mathbb {R}}^n\). \(\square \)

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Dower, P.M. (2018). An Adaptive Max-Plus Eigenvector Method for Continuous Time Optimal Control Problems. In: Falcone, M., Ferretti, R., Grüne, L., McEneaney, W. (eds) Numerical Methods for Optimal Control Problems. Springer INdAM Series, vol 29. Springer, Cham. https://doi.org/10.1007/978-3-030-01959-4_10

Download citation

Publish with us

Policies and ethics