Skip to main content
Log in

On the computation of Nash and Pareto equilibria for some bi-objective control problems for the wave equation

  • Published:
Advances in Computational Mathematics Aims and scope Submit manuscript

Abstract

This paper deals with the numerical implementation of a systematic method for solving bi-objective optimal control problems for wave equations. More precisely, we look for Nash and Pareto equilibria which respectively correspond to appropriate noncooperative and cooperative strategies in multi-objective optimal control. The numerical methods described here consist of a combination of the following: finite element techniques for space approximation; finite difference schemes for time discretization; gradient algorithms for the solution of the discrete control problems. The efficiency of the computational methods is illustrated by the results of some numerical experiments.

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.

Similar content being viewed by others

References

  1. Araruna, F.D., Fernández-Cara, E., Silva, L.C.: Hierarchic control for the wave equation. J. Optimiz. Theory App. 178(4), 264—288 (2018)

    MathSciNet  MATH  Google Scholar 

  2. Ciarlet, P.H.G.: Introduction to Numerical Linear Algebra and Optimisation. Cambridge University Press, Cambridge (1989)

    Google Scholar 

  3. Bristeau, M.O., Glowinski, R., Mantel, B., Periaux, J., Sefrioui, M.: Genetic Algorithms for Electromagnetic Backscattering Multiobjective Optimization. In: Rahmat-samii, Y., Michielssen, E. (eds.) Electromagnetic Optimization by Genetic Algorithms, pp 399—434. John Wiley, New York (1999)

  4. Carvalho, P.P., Fernández-Cara, E.: On the Computation of Nash and Pareto Equilibria for Some Bi-objective Control Problems, J. Scientific Computing. Springer, US (2019)

    MATH  Google Scholar 

  5. Díaz, J.I.: On the von Neumann problem and the approximate controllability of Stackelberg-Nash strategies for some environmental problems. Rev. R. Acad. Cien., Serie A. Math. 96(3), 343–356 (2002)

    MathSciNet  MATH  Google Scholar 

  6. Díaz, J.I., Lions, J.-L.: On the approximate controllability of Stackelberg-Nash strategies. In: Ocean Circulation and Pollution Control: a Mathematical and Numerical Investigation, pp 17–27. Springer, Berlin (2004)

  7. Glowinski, R.: Numerical Methods for Nonlinear Variational Problems. Springer, New York (1984)

    MATH  Google Scholar 

  8. Hecht, F.: New development in freefem++. J. Numer.Math. 20, 251–265 (2012)

    MathSciNet  MATH  Google Scholar 

  9. Lions, J.-L.: Contrôle de Pareto de systèmes distribués. Le cas d’évolution. C. R. Acad. Sc. Paris, Sé,rie I 302(11), 413–417 (1986)

    MathSciNet  MATH  Google Scholar 

  10. Lions, J.-L.: Some remarks on Stackelberg’s optimization. Math. Mod. Meth. Appl. Sci. 4, 477–487 (1994)

    MathSciNet  MATH  Google Scholar 

  11. Nash, J.F.: Noncooperative games. Ann. Math. 54, 286–295 (1951)

    MathSciNet  MATH  Google Scholar 

  12. Pareto, V.: Cours D’économie Politique, Rouge. Switzerland, Laussane (1896)

    Google Scholar 

  13. Ramos, A.M., Glowinski, R., Periaux, J.: Nash equilibria for the multi-objective control of linear partial partial differential equations. J. Optim. Theory Appl 112(3), 457–498 (2002)

    MathSciNet  MATH  Google Scholar 

  14. Ramos, A.M., Glowinski, R., Periaux, J.: Pointwise control of the burgers equation and related Nash equilibria problems: a computational approach. J. Optim. Theory Appl. 112, 499–516 (2002)

    MathSciNet  MATH  Google Scholar 

  15. Ramos, A.M., Roubíc̆ek, T.: Nash equilibria in noncooperative predator-prey games. J. Appl Math Optim 56, 211–241 (2007)

    MathSciNet  MATH  Google Scholar 

  16. Von Stackelberg, H.: Marktform und Gleichgewicht. Springer, Berlin Germany (1934)

    MATH  Google Scholar 

Download references

Acknowledgments

This paper was partially written during a stay of the first author at the Institute of Mathematics of the University of Sevilla (IMUS). He is indebted to this Institute for its assistance.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pitágoras Pinheiro de Carvalho.

Additional information

Communicated by: Enrique Zuazua

Publisher’s note

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

Appendix: Existence and uniqueness of a solution to (6)

Appendix: Existence and uniqueness of a solution to (6)

Let us consider the mapping:

$$ (v_{1}, v_{2}) \in \mathcal{U} \mapsto \left( \frac{\partial J_{1}}{\partial v_{1}}(v_{1}, v_{2}), \frac{\partial J_{2}}{\partial v_{2}}(v_{1}, v_{2}) \right) \in \mathcal{U} . $$
(34)

From the linearity of (2) and the fact that J1 and J2 are quadratic, it is obvious that there exist a unique bounded linear operator \({\mathscr{A}}\in {\mathscr{L}}(\mathcal {U} ; \mathcal {U})\) and a unique \( b \in \mathcal {U} \) such that:

$$ \left( \frac{\partial J_{1}}{\partial v_{1}}(v_{1}, v_{2}), \frac{\partial J_{2}}{\partial v_{2}}(v_{1}, v_{2}) \right) = \mathscr{A}(v_{1}, v_{2}) - b \ \ \ \ \forall (v_{1},v_{2}) \in \mathcal{U}. $$

Let us identify the mapping \({\mathscr{A}}\). For every \((v_{1}, v_{2}) \in \mathcal {U}\), one has:

where the ϕi are given by

(35)

and y is the solution to

(36)

Proposition 1

The mapping \({\mathscr{A}}\) is linear, continuous, and self-adjoint. Furthermore, if μ1 and μ2 are sufficiently large (depending on Ω and T), \({\mathscr{A}}\) is strongly positive in \(\mathcal {U}\), that is, there exists α0 > 0 such that:

$$ \left( \mathscr{A}(v_{1},v_{2}),(v_{1},v_{2})\right)_{\mathcal{U}} \geq \alpha_{0} \|(v_{1},v_{2})\|_{\mathcal{U}}^{2} \quad \ \ \forall (v_{1},v_{2}) \in \mathcal{U}. $$
(37)

Proof

The argument is adapted from [13], Proposition 4.1.

Let us see that \(\mathcal {A}\) is self-adjoint. Indeed, for any \((v_{1}, v_{2}), (w_{1}, w_{2}) \in \mathcal {U}\), one has:

$$ \begin{array}{@{}rcl@{}} \left( \mathscr{A}(v_{1},v_{2}),(w_{1},w_{2})\right)_{\mathcal{U}} = \iint_{\mathcal{O}_{1} \times (0,T)} \left( \mu_{1} v_{1} + \phi_{1} \right)w_{1} dx dt \\ \noalign{} \phantom{\left( \mathscr{A}(v_{1},v_{2}),(w_{1},w_{2})\right)_{\mathcal{U}}} + \iint_{\mathcal{O}_{2} \times (0,T)} \left( \mu_{2} v_{2} + \phi_{2} \right)w_{2} dx dt . \end{array} $$
(38)

If y and z denote the solutions to (36) corresponding to (v1,v2) and (w1,w2), we have:

(39)

Thus,

$$ \begin{array}{@{}rcl@{}} \left( \mathscr{A}(v_{1},v_{2}),(w_{1},w_{2})\right)_{\mathcal{U}} &=& \sum\limits_{i=1}^{2} \mu_{i} \iint_{\mathcal{O}_{i} \times (0,T)} v_{i} w_{i} dx dt \\ \noalign{} \phantom{\left( \mathscr{A}(v_{1},v_{2}),(w_{1},w_{2})\right)_{\mathcal{U}}} &+& \iint_{(\mathcal{O}_{1,d} \cup \mathcal{O}_{2,d}) \times (0,T)} y z dx dt \\ \noalign{} \phantom{\left( \mathscr{A}(v_{1},v_{2}),(w_{1},w_{2})\right)_{\mathcal{U}}} &=& \Bigl((v_{1},v_{2}),\mathscr{A}(w_{1},w_{2})\Bigr)_{\mathcal{U}} . \end{array} $$
(40)

Therefore, \(\mathcal {A}\) is self-adjoint.

For completeness, let us recall the proof of strong positiveness. Thus, note that, for any \((v_{1},v_{2}) \in \mathcal {U}\):

$$ \left( \mathscr{A}(v_{1},v_{2}),(v_{1},v_{2})\right)_{\mathcal{U}} = \sum\limits_{i=1}^{2} \iint_{\mathcal{O}_{i} \times (0,T)} \left( \mu_{i} v_{i} + \phi_{i} \right)v_{i} dx dt . $$
(41)

Let z1 (resp. z2) be the solution to (36) with v2 = 0 (resp. v1 = 0). Then, from standard integration by parts, we see that:

$$ \iint_{\mathcal{O}_{i} \times (0,T)} \phi_{i} v_{i} dx dt = \iint_{Q} \phi_{i} (z_{i,tt} - {\varDelta} z_{i}) dx dt = \iint_{\mathcal{O}_{i,d} \times (0,T)} y z_{i} dx dt. $$
(42)

Consequently, there exists C0 depending on Ω and T such that:

$$ | \iint_{\mathcal{O}_{i} \times (0,T)} \phi_{i} v_{i} dx dt | \leq C_{0} \|(v_{1},v_{2})\|_{\mathcal{U}} \|v_{i}\|_{\mathcal{U}_{i}}, \ \ i=1,2. $$
(43)

From (41) and (43), we see that, if \(\min \limits (\mu _{1},\mu _{2}) > \sqrt {2} C_{0}\), then

$$ \left( \mathscr{A}(v_{1},v_{2}),(v_{1},v_{2})\right)_{\mathcal{U}} \geq \left[\min(\mu_{1},\mu_{2}) - \sqrt{2} C_{0} \right] \|(v_{1},v_{2})\|^{2}_{\mathcal{U}} \quad \forall (v_{1},v_{2}) \in \mathcal{U}, $$

whence (37) holds. □

Remark 3

From (41) and (42) we also observe that, if \(\mathcal {O}_{1,d}\) and \(\mathcal {O}_{2,d}\) coincide and we set \(\mathcal {O}_{d} = \mathcal {O}_{i,d}\), then

$$ \left( \mathscr{A}(v_{1},v_{2}),(v_{1},v_{2})\right)_{\mathcal{U}} \geq \min(\mu_{1},\mu_{2}) \|(v_{1},v_{2})\|^{2}_{\mathcal{U}} + \iint_{\mathcal{O}_{d} \times (0,T)} |y|^{2} dx dt \\ $$

for all \((v_{1},v_{2}) \in \mathcal {U}\). Therefore, in this case, \({\mathscr{A}}\) is always strongly positive, regardless of the sizes of the μi.

Let us identify b, that is, the nonhomogeneous term in the affine mapping (34). One has:

where \(\bar {\phi }_{i}\) is the solution to

(44)

and \(\bar {y}\) is the solution of

$$ \left\{ \begin{array} [c]{ll} \bar{y}_{tt} - {\varDelta} \bar{y} = f \ \ & \ \ \text{in} \ Q, \\ \bar{y} = 0 \ \ \ \ & \ \text{on} \ {\varSigma}_{1}, \\ \frac{\partial \bar{y}}{\partial n} = 0 \ & \ \text{on} \ {\varSigma}_{2}, \\ \bar{y}(x,0) = y_{0}(x), \ \ \ \ \ \bar{y}_{t}(x,0) = y_{1}(x) \ \ & \ \ \text{in} \ {\varOmega} . \end{array} \right. $$
(45)

Now, if we define by

$$ a(v, w) := \left( \mathscr{A}v , w \right)_{\mathcal{U}} \quad \ \ \forall v, w \in \mathcal{U} $$

and by

$$ L(w) := \left( b, w\right)_{\mathcal{U}} \ \quad \ \forall w \in \mathcal{U}, $$

we deduce that (6) is equivalent to

$$ a(v, w) = L(w) \quad \forall v \in \mathcal{U}. $$
(46)

From Proposition 1, we see that a(⋅,⋅) is bilinear, continuous, and symmetric. Moreover, if μ1 and μ2 are large enough, it is also \(\mathcal {U}\)-elliptic. On the other hand, L is linear and continuous. Hence, from Lax-Milgram Theorem, the existence and uniqueness of a solution to (46) is ensured if the μi are sufficiently large.

Remark 4

Let us fix λ ∈ (0, 1) and let us consider the system (18)–(20). We can adapt the previous argument and prove that there exist other \({\mathscr{A}}\) and b such that (21) holds. Now, we get:

$$ \begin{array}{@{}rcl@{}} \left( \mathscr{A}(v_{1}, v_{2}),(v_{1},v_{2}) \right)_{\mathcal{U}} &=& \iint_{\mathcal{O}_{1}\times(0,T)}\left( \lambda \mu_{1}v_{1} + \lambda \phi_{1} + (1-\lambda)\phi_{2})\right)v_{1} dx dt \\ \noalign{} \phantom{\left( \mathscr{A}(v_{1}, v_{2}),(v_{1},v_{2}) \right)_{\mathcal{U}}} &+& \iint_{\mathcal{O}_{2}\times(0,T)}\left( (1 - \lambda) \mu_{2}v_{2} + \lambda \phi_{1} + (1-\lambda)\phi_{2}\right)v_{2} dx dt \\ \noalign{} \phantom{\left( \mathscr{A}(v_{1}, v_{2}),(v_{1},v_{2}) \right)_{\mathcal{U}}} &\geq& \left( {\min\left( \lambda \mu_{1}, (1-\lambda)\mu_{2} \right)} - \sqrt{2}C_{1}\right)||(v_{1}, v_{2})||^{2}_{\mathcal{U}} \ , \end{array} $$

where C1 only depends on Ω and T. Thus, if \(\min \limits (\mu _{1},\mu _{2}) > \displaystyle \frac {\sqrt {2}C_{1}}{\min \limits (\lambda , 1 - \lambda )}\), the system (18)–(20) possesses exactly one solution.

1.1 Existence and uniqueness of a solution to the semilinear system (23)–(25)

In this section, we consider the semilinear state (i). As before, we assume that is globally Lipschitz-continuous. Thus, there exists CF > 0 such that:

Note that the couple (v1,v2) solves (23)–(25) if and only if it is a fixed point of the nonlinear mapping \({\varLambda }: \mathcal {U} \mapsto \mathcal {U}\), where

$$ \left\{ \begin{array}{l} \displaystyle {\varLambda}(v) = \left( {\varLambda}_{1}(v),{\varLambda}_{2}(v)\right), \ \ {\varLambda}_{i}(v) = -\frac{1}{\mu_{i}} \phi_{i} |_{\mathcal{O}_{i} \times (0,T)} , \\ \phi_{i} \text{is the solution to~(24) for} i=1,2, \\ y \ \text{is} \ \text{the} \ \text{solution} \ \text{to}~(23). \end{array} \right. $$

It is not difficult to check that there exists \(C({\varOmega },T,C_{F},\|f\|_{L^{2}(Q)})\) such that, if

$$ \min(\mu_{1},\mu_{2}) > C({\varOmega},T,C_{F},\|f\|_{L^{2}(Q)}), $$

the mapping Λ is a contraction. Indeed, from the usual energy estimates, setting \(\mu _{0} := \min \limits (\mu _{1},\mu _{2})\), it is clear that:

$$ \begin{array}{@{}rcl@{}} \| {\varLambda}(v) - {\varLambda}(\tilde v) \|_{\mathcal{U}}^{2}& \leq& \sum\limits_{i=1}^{2} \frac{1}{{\mu_{i}}^{2}} \iint_{\mathcal{O}_{i} \times (0,T)} | (\phi_{1},\phi_{2}) - (\tilde \phi_{1}, \tilde \phi_{2}) |^{2} dx dt \\ \noalign{} \phantom{\| {\varLambda}(v) - {\varLambda}(\tilde v) \|_{\mathcal{U}}^{2}} &\leq& {\mu_{0}}^{-2} C({\varOmega},T) \| (\phi_{1},\phi_{2}) - (\tilde \phi_{1}, \tilde \phi_{2}) \|_{L^{2}(Q) \times L^{2}(Q)}^{2} \\ \noalign{} \phantom{\| {\varLambda}(v) - {\varLambda}(\tilde v) \|_{\mathcal{U}}^{2}} &\leq& {\mu_{0}}^{-2} C({\varOmega},T,C_{F}) \| y - \tilde y \|_{L^{2}(Q)}^{2} \\ \noalign{} \phantom{\| {\varLambda}(v) - {\varLambda}(\tilde v) \|_{\mathcal{U}}^{2}} &\leq& {\mu_{0}}^{-2} C({\varOmega},T,C_{F},\|f\|_{L^{2}(Q)}) \| v - \tilde v \|_{\mathcal{U}}^{2} \end{array} $$

for all \(v, \tilde v \in \mathcal {U}\), where the notation is self-explanatory.

As a consequence, we find that, if μ1 and μ2 are large enough, (23)–(25) possesses a unique solution. In other words, (6) is uniquely solvable.

Remark 5

For any fixed λ ∈ (0, 1), we can also consider the system (28)–(30). Arguing in a similar way, we can deduce that there exists \(C({\varOmega },T,C_{F},\|f\|_{L^{2}(Q)},\lambda )\) such that, for greater values of \(\min \limits (\mu _{1},\mu _{2})\), this system possesses exactly one solution.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

de Carvalho, P., Fernández-Cara, E. & Ferrel, J.B.L. On the computation of Nash and Pareto equilibria for some bi-objective control problems for the wave equation. Adv Comput Math 46, 73 (2020). https://doi.org/10.1007/s10444-020-09812-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10444-020-09812-z

Keywords

Mathematics Subject Classification (2010)

Navigation