Skip to main content
Log in

Linear turnpike theorem

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

The turnpike phenomenon stipulates that the solution of an optimal control problem in large time remains essentially close to a steady-state of the dynamics, itself being the optimal solution of an associated static optimal control problem. Under general assumptions, it is known that not only the optimal state and the optimal control, but also the adjoint state coming from the application of the Pontryagin maximum principle, are exponentially close to that optimal steady-state, except at the beginning and at the end of the time frame. In such a result, the turnpike set is a singleton, which is a steady-state. In this paper, we establish a turnpike result for finite-dimensional optimal control problems in which some of the coordinates evolve in a monotone way, and some others are partial steady-states of the dynamics. We prove that the discrepancy between the optimal trajectory and the turnpike set is linear, but not exponential: we thus speak of a linear turnpike theorem.

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

Notes

  1. Equivalently, the pair \((A_1,{\bar{B}}_1)\) satisfies the Kalman condition.

  2. Indeed, let \((Y_1,Y_2)\in \textrm{IR}^{n+m}\times \textrm{IR}^{n+p}\) be such that \(dF.\begin{pmatrix}Y_1\\ Y_2\end{pmatrix}=\begin{pmatrix}0\\ 0\end{pmatrix}\), i.e., \(E_1Y_1+E_2^\top Y_2=0\) and \(E_2Y_1=0\). Then \(0 = Y_1^\top E_1 Y_1+Y_1^\top E_2^\top Y_2 = Y_1^\top E_1 Y_1\) because \(Y_1^\top E_2^\top =0\), and thus \(Y_1=0\) because \(E_1\) is negative definite. Then, \(E_2^\top Y_2=0\) implies \(Y_2=0\) because \(E_2^\top \) is injective.

  3. Actually, this matrix is exactly the matrix of the linearized system (43) studied further.

  4. It suffices to prove that 0 is the only possible closure point of the (bounded) function \(h_1-h_2\) at \(+\infty \). Let \(a\in \textrm{IR}\) be such a closure point: there exists a sequence \(x_k\rightarrow +\infty \) such that \(h_1(x_k)-h_2(x_k)\rightarrow a\). Since the sequence \((h_2(x_k))_{k\in \textrm{IN}}\) is bounded, there exists a subsequence such that \(h_2(x_{k_p})\rightarrow 0\). Hence, by assumption, \(h_1(x_{k_p})-h_2(x_{k_p})\rightarrow 0\), and thus \(a=0\).

References

  1. Abou-Kandil H, Freiling G, Ionescu V, Jank G (2003) Matrix Riccati equations. Systems & control: foundations & applications. Birkhäuser Verlag, Basel, p xx+572. ISBN: 3-7643-0085-X

  2. Aftalion A (2017) How to run 100 meters. SIAM J Appl Math 77(4):1320–1334

    Article  MathSciNet  MATH  Google Scholar 

  3. Aftalion A, Bonnans J-F (2014) Optimization of running strategies based on anaerobic energy and variations of velocity. SIAM J Appl Math 74(5):1615–1636

    Article  MathSciNet  MATH  Google Scholar 

  4. Aftalion A, Trélat E (2020) How to build a new athletic track to break records. R Soc Open Sci 7:200007, p 10

  5. Aftalion A, Trélat E (2021) Pace and motor control optimization for a runner. J Math Biol 83(1):9

    Article  MathSciNet  MATH  Google Scholar 

  6. Agrachev A, Sachkov Y (2004) Control theory from the geometric viewpoint. Encyclopaedia of mathematical sciences. Control theory and optimization, II., vol 87. Springer-Verlag, Berlin, p xiv+412

    MATH  Google Scholar 

  7. Anderson BDO, Kokotovic PV (1987) Optimal control problems over large time intervals. Autom J IFAC 23(3):355–363

    Article  MathSciNet  MATH  Google Scholar 

  8. Bonnard B, Caillau J-B, Trélat E (2007) Second order optimality conditions in the smooth case and applications in optimal control. ESAIM Control Optim Calc Var 13(2):207–236

    Article  MathSciNet  MATH  Google Scholar 

  9. Bonnard B, Chyba M (2003) Singular trajectories and their role in control theory. Mathématiques & Applications (Berlin) [Mathematics & Applications], vol 40. Springer-Verlag, Berlin, p xvi+357

    MATH  Google Scholar 

  10. Boyd S, Vandenberghe L (2004) Convex optimization. Cambridge University Press, Cambridge, p xiv+716

    Book  MATH  Google Scholar 

  11. Brogliato B, Lozano R, Maschke B, Egeland O (2007) Dissipative systems analysis and control, 2nd edn. Springer-Verlag, London Ltd, London

    Book  MATH  Google Scholar 

  12. Cannarsa P, Sinestrari C (2004) Semiconcave functions, Hamilton-Jacobi equations, and optimal control, vol 58. Progress in nonlinear differential equations and their applications. Birkhäuser Boston Inc, Boston MA, p xiv+304

    Book  MATH  Google Scholar 

  13. Carlson DA, Haurie AB, Leizarowitz A (1991) Infinite horizon optimal control. Springer-Verlag, Berlin. Deterministic and stochastic systems, Second revised and enlarged edition of the 1987 original

  14. Chitour Y, Jean F, Trélat E (2006) Genericity results for singular curves. J Differ Geom 73(1):45–73

    Article  MathSciNet  MATH  Google Scholar 

  15. Chitour Y, Jean F, Trélat E (2008) Singular trajectories of control-affine systems. SIAM J Control Optim 47(2):1078–1095

    Article  MathSciNet  MATH  Google Scholar 

  16. Clarke FH (2013) Functional analysis, calculus of variations and optimal control. Graduate texts in mathematics, vol 264. Springer, London, p xiv+591

    Book  MATH  Google Scholar 

  17. Coron J-M, Trélat E (2004) Global steady-state controllability of one-dimensional semilinear heat equations. SIAM J Control Optim 43(2):549–569

    Article  MathSciNet  MATH  Google Scholar 

  18. Damm T, Grüne L, Stieler M, Worthmann K (2014) An exponential turnpike theorem for dissipative discrete time optimal control problems. SIAM J. Control Optim. 52:1935–1957

    Article  MathSciNet  MATH  Google Scholar 

  19. Fathi A (2016) Weak KAM theorem in Lagrangian dynamics. Cambridge studies in advanced mathematics, Hardcover

  20. Faulwasser T, Flasskamp K, Ober-Blöbaum S, Worthmann K (2019) Towards velocity turnpikes in optimal control of mechanical systems. IFAC-PapersOnLine, 52(16):490–495. In Proc. 11th IFAC Symposium on Nonlinear Control Systems, NOLCOS

  21. Faulwasser T, Flasskamp K, Ober-Blöbaum S, Worthmann K(2021) A dissipativity characterization of velocity turnpikes in optimal control problems for mechanical systems. IFAC-PapersOnLine 54(9):624–629. In 24th International Symposium on Mathematical Theory of Networks and Systems MTNS 2020: Cambridge, United Kingdom

  22. Faulwasser T, Grüne L (2022) Turnpike properties in optimal control. In: Handbook of numerical Analysis, vol 23. pp 367–400

  23. Faulwasser T, Grüne L, Humaloja JP, Schaller M (2020) The interval turnpike property for adjoints. Pure and Applied Functional Analysis, in press, arXiv:2005.12120

  24. Faulwasser T, Korda M, Jones CN, Bonvin D (2017) On turnpike and dissipativity properties of continuous-time optimal control problems. Autom J IFAC 81:297–304

    Article  MathSciNet  MATH  Google Scholar 

  25. Fourer R, Gay DM, Kernighan BW (2002) AMPL: a modeling language for mathematical programming, 2nd edn. Duxbury Press, Scituate, p 540

    Google Scholar 

  26. Grüne L (2013) Economic receding horizon control without terminal constraints. Automatica 49:725–734

    Article  MathSciNet  MATH  Google Scholar 

  27. Grüne L, Müller MA (2016) On the relation between strict dissipativity and turnpike properties. Syst Control Lett 90:45–53

    Article  MathSciNet  MATH  Google Scholar 

  28. Grüne L, Guglielmi R (2018) Turnpike properties and strict dissipativity for discrete time linear quadratic optimal control problems. SIAM J Control Optim 56(2):1282–1302

    Article  MathSciNet  MATH  Google Scholar 

  29. Grüne L, Schaller M, Schiela A (2020) Exponential sensitivity and turnpike analysis for linear quadratic optimal control of general evolution equations. J Differ Equ 268(12):7311–7341

    Article  MathSciNet  MATH  Google Scholar 

  30. Jurdjevic V (1997) Geometric control theory, vol 52. Cambridge studies in advanced mathematics. Cambridge University Press, Cambridge, p xviii+492

    MATH  Google Scholar 

  31. Keller JB (1974) Optimal velocity in a race. Am Math Mon 81:474–480

    Article  MathSciNet  MATH  Google Scholar 

  32. Kučera V (1972) A contribution to matrix quadratic equations. IEEE Trans Autom Control 17(3):344–347

    Article  MathSciNet  MATH  Google Scholar 

  33. Perko L (2001) Differential equations and dynamical systems, vol 7, 3rd edn. Texts in applied mathematics. Springer-Verlag, New York, p xiv+553

    Book  MATH  Google Scholar 

  34. Pontryagin LS, Boltyanskii VG, Gamkrelidze RV, Mishchenko EF (1962) 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-London

  35. Porretta A, Zuazua E (2013) Long time versus steady state optimal control. SIAM J Control Optim 51(6):4242–4273

    Article  MathSciNet  MATH  Google Scholar 

  36. Rapaport A, Cartigny P (2004) Turnpike theorems by a value function approach. ESAIM Control Optim Calc Var 10(1):123–141

    Article  MathSciNet  MATH  Google Scholar 

  37. Rapaport A, Cartigny P (2005) Competition between most rapid approach paths: necessary and sufficient conditions. J Optim Theory Appl 124(1):1–27

    Article  MathSciNet  MATH  Google Scholar 

  38. Sontag ED (1998) Mathematical control theory, deterministic finite-dimensional systems, vol 6, 2nd edn. Texts in applied mathematics. Springer-Verlag, New York

    MATH  Google Scholar 

  39. Tikhonov AN, Arsenin VY (1977) Solutions of ill-posed problems. V. H. Winston & Sons, Washington, D.C.: John Wiley & Sons, New York-Toronto, Ont.-London. Translated from the Russian, Preface by translation editor Fritz John, Scripta Series in Mathematics

  40. Trélat E (2000) Some properties of the value function and its level sets for affine control systems with quadratic cost. J Dyn Control Syst 6(4):511–541

    Article  MathSciNet  MATH  Google Scholar 

  41. Trélat E (2005) Contrôle optimal, théorie & applications. Mathématiques Concrètes. Vuibert, Paris

    MATH  Google Scholar 

  42. Trélat E (2012) Optimal control and applications to aerospace: some results and challenges. J Optim Theory Appl 154(3):713–758

    Article  MathSciNet  MATH  Google Scholar 

  43. Trélat E, Zhang C (2018) Integral and measure-turnpike properties for infinite-dimensional optimal control systems. Math Control Signals Syst 30:34

    Article  MathSciNet  MATH  Google Scholar 

  44. Trélat E, Zhang C, Zuazua E (2018) Steady-state and periodic exponential turnpike property for optimal control problems in Hilbert spaces. SIAM J Control Optim 56(2):1222–1252

    Article  MathSciNet  MATH  Google Scholar 

  45. Trélat E, Zuazua E (2015) The turnpike property in finite-dimensional nonlinear optimal control. J Differ Equ 258(1):81–114

    Article  MathSciNet  MATH  Google Scholar 

  46. Wächter A, Biegler LT (2006) On the implementation of an interior-point filter line-search algorithm for large-scale nonlinear programming. Math Program 106:25–57

    Article  MathSciNet  MATH  Google Scholar 

  47. Wilde RR, Kokotovic PV (1972) A dichotomy in linear control theory. IEEE Trans Autom Control 17(3):382–383

    Article  MATH  Google Scholar 

  48. Willems JC (1972) Dissipative dynamical systems, Part I: General theory. Arch Ration Mech Anal 45:321–351

    Article  MATH  Google Scholar 

  49. Zaslavski AJ (2006) Turnpike properties in the calculus of variations and optimal control, vol 80. Nonconvex optimization and its applications. Springer, New York

    MATH  Google Scholar 

  50. Zaslavski AJ (2015) Turnpike theory of continuous-time linear optimal control problems, vol 104. Springer optimization and its applications. Springer, Cham

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emmanuel Trélat.

Additional information

Publisher's Note

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

A Appendix

A Appendix

This appendix is devoted to illustrating the comments done in Remark 2. We take a very simple example and we give numerical simulations showing the competition between two global turnpikes, or, at the level of the initialization of numerical methods, between local and global turnpikes.

We consider the one-dimensional optimal control problem

$$\begin{aligned} \begin{aligned}&\dot{x} = -3x + 3x^3 + u, \qquad x(0)=x_0,\quad x(T)=x_f\text { or free} \\&\min \int _0^T \left( (x(t)-x_d)^2 + (u(t)-u_d)^2 \right) dt \end{aligned} \end{aligned}$$
(89)

The corresponding static problem is

$$\begin{aligned} \min _{(x,u)\ \mid \ -3x + 3x^3 + u=0} \left( (x-x_d)^2 + (u-u_d)^2\right) . \end{aligned}$$
(90)

1.1 A.1 Competition between two (global) turnpikes

We take \(x_d=1\) and \(u_d=3.47197\). The choice of \(u_d\) is done so that the static problem (90) has two global minima, at \({\bar{x}}_1=-1.347372066\) and \({\bar{x}}_2=0.5939615956\), see Fig. 4.

Fig. 4
figure 4

Plot of the function \(x\mapsto (x-1)^2 + (3x - 3x^3-3.47197)^2\)

In Fig. 5, in dashed blue, we have computed the optimal trajectory with \(x_0=-5\), \(x_f=-1\), \(T=10\): we observe a turnpike phenomenon around \({\bar{x}}_1\). In solid red, we have computed the optimal trajectory with \(x_0=2\), \(x_f=1\), \(T=10\): we observe a turnpike phenomenon around \({\bar{x}}_2\).

Fig. 5
figure 5

Global turnpikes around \({\bar{x}}_1\) and \({\bar{x}}_2\)

The fact that the turnpike phenomenon is either around \({\bar{x}}_1\) or around \({\bar{x}}_2\) depends on the terminal conditions. For instance, if \(x_0\) and \(x_f\) are close to \({\bar{x}}_1\) (resp., \({\bar{x}}_2\)) then the optimal trajectory will make a turnpike around \({\bar{x}}_1\) (resp., \({\bar{x}}_2\)). But when the terminal conditions are farther, it is not clear to predict the behavior of the optimal trajectory.

Since the minima are global, we expect to observe a competition between both turnpikes, depending on the terminal conditions (see [37]). Let us provide some numerical simulations illustrating this competition. To facilitate the understanding, we consider the problem with \(x_f\) free. It is interesting to note that the numerical result strongly depends on the initialization of the numerical method. Here, to compute numerically the optimal solutions of (89), we use AMPL (see [25]) combined with IpOpt (see [46]): the trajectory and the controls are discretized (the control is piecewise constant and the trajectory is piecewise linear, on a given subdivision that is chosen fine enough), and we initialize the trajectory with the same constant value over all the subdivision. Then according to the value of this initialization, we can make emerge such or such turnpike, and all solutions are anyway optimal.

Fig. 6
figure 6

\(x_0=-2\), \(x_f\) free, \(T=20\)

In Fig. 6, we have taken \(T=20\), \(x_0=-2\) (and \(x_f\) is let free). In dashed blue, we have initialized the trajectory to the constant trajectory \({\bar{x}}_1\), and we then obtain an optimal trajectory which stays essentially near \({\bar{x}}_1=-1.347372066\), with a kind of “hesitation” towards \({\bar{x}}_2=0.5939615956\) near the end. In solid red, we have initialized the trajectory to the constant trajectory \({\bar{x}}_2\), and we obtain a trajectory staying essentially near \({\bar{x}}_2\), with a kind of “hesitation” towards \({\bar{x}}_1\) near the beginning. We stress that the two solutions are optimal: both have a cost \(C\simeq 6.2822\). We could make emerge other similar trajectories, which “hesitate” between the two turnpikes. All of them are optimal, or, at least, “quasi-optimal” (there is a small error due to switches from one turnpike to the other).

1.2 A.3 Local versus global turnpike

We take \(x_d=1\) and \(u_d=1\). The choice of \(u_d\) is now such that the static problem (90) has a unique global solution \({\bar{x}} = 0.781538640850898\), see Fig. 7. But it has also a local solution \({\bar{x}}_{loc}=-1.10551208794920\).

Fig. 7
figure 7

Plot of the function \(x\mapsto (x-1)^2 + (3x - 3x^3-1)^2\)

The global minimum \({\bar{x}}\) is a global turnpike, while the local minimum \({\bar{x}}\) is a local turnpike. In the numerical simulations, when performing either an optimization or a Newton method (shooting), we compute local solutions. Hence, we must expect that the numerical results depend on the initialization. To check global optimality, we have to compare the costs.

Fig. 8
figure 8

\(x_0=-2\), \(x_f=-1\), \(T=10\)

In Fig. 8, in dashed blue, we have initialized the code with the constant trajectory \({\bar{x}}_{loc}\). We observe a turnpike around \({\bar{x}}_{loc}=-1.10551208794920\). The cost is \(C\simeq 7.008\). But this is a local turnpike only. This trajectory that we obtain is only locally optimal, and is not globally optimal. In solid red, we have initialized the code with the constant trajectory \({\bar{x}}\). We observe a turnpike around \({\bar{x}} = 0.781538640850898\). The cost is \(C\simeq 3.825\) and is lower than the one of the previous one. Here, we have actually computed the globally optimal trajectory. This is the global turnpike.

It is also interesting to see what happens if we take T much smaller. Let us take \(T=2\). In Fig. 9, in dashed blue, we have initialized the code with the constant trajectory \({\bar{x}}\). We observe a trend to the turnpike around \({\bar{x}} = 0.781538640850898\). The cost is \(C\simeq 17.792\). But this trajectory, now, is not globally optimal. In solid red, we have initialized the code with the constant trajectory \({\bar{x}}_{loc}\). We observe a turnpike around \({\bar{x}}_{loc}=-1.347372066\). The cost is \(C\simeq 16.322\). It is the globally optimal trajectory.

Fig. 9
figure 9

\(x_0=-2\), \(x_f=-1\), \(T=2\)

We can search a time \(2<T_c<10\) for which both previous initializations give equivalent turnpikes around \({\bar{x}}_{loc}\) and \({\bar{x}}\) (with same cost). This is done hereafter. What is important is that, in large time T, we have indeed the global turnpike around \({\bar{x}}\).

In Fig. 10, at the left, the code is initialized with \(x(t)\equiv -1.1\), in order to promote the turnpike around \(\bar{x}_{loc}=-1.347372066\). At the right, the code is initialized with \(x(t)\equiv 0.78\), in order to promote the turnpike around \({\bar{x}} = 0.781538640850898\). We have computed trajectories for the following successives values of T: 2.5, 2.7, 2.9, 3.1, 3.3. We observe that the blue and cyan trajectories at the top are optimal (see the value of their cost); and that the red, green and black trajectories at the bottom are optimal. The bifurcation occurs around \(T=2.9\).

Finally, in Fig. 11, we represent the global optimal trajectory. For \(T\lesssim 2.9\), the global optimal trajectory makes a turnpike around \({\bar{x}}_{loc}=-1.347372066\), which is a local minimizer of the optimal static problem. For \(T\geqslant 2.9\) we have a bifurcation and the global optimal trajectory makes a turnpike around \({\bar{x}} = 0.781538640850898\), which is the global minimizer of the optimal static problem.

Fig. 10
figure 10

\(x_0=-2\), \(x_f=-1\)

Fig. 11
figure 11

\(x_0=-2\), \(x_f=-1\)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Trélat, E. Linear turnpike theorem. Math. Control Signals Syst. 35, 685–739 (2023). https://doi.org/10.1007/s00498-023-00354-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00498-023-00354-5

Keywords

Navigation