Skip to main content

Hamilton–Jacobi–Bellman Equations on Multi-domains

  • Chapter
  • First Online:
Book cover Control and Optimization with PDE Constraints

Part of the book series: International Series of Numerical Mathematics ((ISNM,volume 164))

Abstract

A system of Hamilton–Jacobi (HJ) equations on a partition of \(\mathbb {R}^{d}\) is considered, and a uniqueness and existence result of viscosity solution is analyzed. While the notion of viscosity solution is by now well known, the question of uniqueness of solution, when the Hamiltonian is discontinuous, remains an important issue. A uniqueness result has been derived for a class of problems, where the behavior of the solution, in the region of discontinuity of the Hamiltonian, is assumed to be irrelevant and can be ignored (see (Camilli, Siconolfi in Adv. Differ. Equ. 8(6):733–768, 2003)). Here, we provide a new uniqueness result for a more general class of Hamilton–Jacobi equations.

The work has been partially supported by the European Union under the 7th Framework Programme “FP7-PEOPLE-2010-ITN”, GA number 264735-SADCO.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
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. Y. Achdou, F. Camilli, A. Cutri, N. Tchou, Hamilton–Jacobi equations on networks. Nonlinear Differ. Equ. Appl. (2012). doi:10.1007/s00030-012-0158-1

  2. J.-P. Aubin, A. Cellina, Differential Inclusions. Comprehensive Studies in Mathematics, vol. 264 (Springer, Berlin, 1984)

    Book  Google Scholar 

  3. M. Bardi, I. Capuzzo-Dolcetta, Optimal Control and Viscosity Solutions of Hamilton–Jacobi–Bellman Equations. Systems and Control: Foundations and Applications (Birkhäuser, Boston, 1997)

    Book  Google Scholar 

  4. G. Barles, A. Briani, E. Chasseigne, A Bellman approach for two-domains optimal control problems in \(\mathbb {R}^{N}\). To appear in ESAIM Control Optim. Calc. Var. (2012)

    Google Scholar 

  5. R.C. Barnard, P.R. Wolenski, Flow invariance on stratified domains. Set-Valued Var. Anal. (2013). doi:10.1007/s11228-013-0230-y

  6. A. Bressan, Y. Hong, Optimal control problems on stratified domains. Netw. Heterog. Media 2(2), 313–331 (2007)

    Article  MathSciNet  Google Scholar 

  7. F. Camilli, A. Siconolfi, Hamilton–Jacobi equations with measurable dependence on the state variable. Adv. Differ. Equ. 8(6), 733–768 (2003)

    MathSciNet  MATH  Google Scholar 

  8. F.H. Clarke, Optimization and Nonsmooth Analysis (Society for Industrial and Applied Mathematics, Philadelphia, 1990)

    Book  Google Scholar 

  9. F.H. Clarke, Yu.S. Ledyaev, R.J. Stern, P.R. Wolenski, Nonsmooth Analysis and Control Theory (Springer, Berlin, 1998)

    MATH  Google Scholar 

  10. G. Dal Maso, H. Frankowska, Value function for Bolza problem with discontinuous Lagrangian and Hamilton–Jacobi inequalities. ESAIM Control Optim. Calc. Var. 5, 369–394 (2000)

    Article  MathSciNet  Google Scholar 

  11. A.F. Filippov, Differential Equations with Discontinuous Right-Hand Sides (Kluwer Academic, Norwell, 1988)

    Book  Google Scholar 

  12. H. Frankowska, Lower semicontinuous solutions of Hamilton–Jacobi–Bellman equations. SIAM J. Control Optim. 31(1), 257–272 (1993)

    Article  MathSciNet  Google Scholar 

  13. C. Imbert, R. Monneau, H. Zidani, A Hamilton–Jacobi approach to junction problems and application to traffic flows. ESAIM Control Optim. Calc. Var. e-first (2011)

    Google Scholar 

  14. H. Ishii, A boundary value problem of the Dirichlet type for Hamilton–Jacobi equations. Ann. Sc. Norm. Super. Pisa, Cl. Sci. 16(1), 105–135 (1989). (eng)

    MathSciNet  MATH  Google Scholar 

  15. Z. Rao, A. Siconolfi, H. Zidani, Transmission conditions on interfaces for Hamilton–Jacobi–Bellman equations (2013 submitted)

    Google Scholar 

  16. P. Soravia, Boundary value problems for Hamilton–Jacobi equations with discontinuous Lagrangian. Indiana Univ. Math. J. 51(2), 451–477 (2002)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors are grateful to Peter Wolenski and Antonio Siconolfi for many helpful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hasnaa Zidani .

Editor information

Editors and Affiliations

Appendices

Appendix A

Proof of Lemma 2.2

Note that although G is only usc on \(\mathbb {R}^{d}\), G is Lipschitz continuous on Γ j since G is the convexification of a finite group of Lipschitz continuous multifunctions on Γ j. For any xΓ j, there exists α>0 and a diffeomorphism \(g\in C^{1,1}(\mathbb {R}^{d})\) such that

$$\overline{B(x,\alpha)\cap\varGamma_j}=\bigl\{x:g(x)=0\bigr\}\quad \mbox{and}\quad \nabla g(y)\neq0,\quad \forall y\in\overline{B(x,\alpha)}. $$

We can take g as the signed distance function to Γ j for instance. Then there exists β>0 such that

$$\bigl\|\nabla g(y)\bigr\|\geq\beta,\quad \forall y\in B(x,\alpha)\cap\varGamma_j. $$

For any \(w\in G(x)\cap \mathcal {T}_{\varGamma_{j}}(x)\), by the Lipschitz continuity of G there exists vG(y) such that

$$\|w-v\|\leq L_G\|x-y\|, $$

where L G is the Lipschitz constant of G(⋅). Since \(w\in \mathcal {T}_{\varGamma_{j}}(x)\), we have

$$w\cdot\nabla g(x)=0. $$

Then

$$\bigl\|v\cdot\nabla g(x)\bigr\|=\bigl\|(v-w)\cdot\nabla g(x)\bigr\|\leq L_G\|\nabla g\| \|x-y\|. $$

Thus,

$$ \begin{aligned} \bigl\|v\cdot\nabla g(y)\bigr\|&\leq\bigl\|v\cdot\nabla g(x)\bigr\|+\bigl\| v\cdot\bigl( \nabla g(y)-\nabla g(x)\bigr)\bigr\| \\ &\leq\bigl(L_G\|\nabla g\|+\|G\|L_g' \bigr)\|x-y\|, \end{aligned} $$

where \(L_{g}'\) is the Lipschitz constant of ∇g(⋅). We consider the following three cases:

  • If v⋅∇g(y)=0, then \(v\in \mathcal {T}_{\varGamma_{j}}(y)\) and we deduce that

    $$w\in G(y)\cap \mathcal {T}_{\varGamma_j}(y)+L_G\|x-y\|B(0,1). $$
  • If v⋅∇g(y):=−γ<0, let p:=δg(y)/∥∇g(y)∥, then by (H4)(iv),

    $$p\in G(y)\quad \mbox{and}\quad p\cdot\nabla g(y):=\tilde{\beta}\geq\delta\beta>0. $$

We set

$$q:=\frac{\tilde{\beta}}{\tilde{\beta}+\gamma}v +\frac{\gamma}{\tilde {\beta}+\gamma}p, $$

then q⋅∇g(y)=0, i.e. \(q\in \mathcal {T}_{\varGamma_{j}}(y)\). And since G(y) is convex, we have qG(y). Then we obtain

$$ \begin{aligned} \|w-q\|&\leq\|w-v\|+\|v-q\| \\ &\leq L_G\|x-y\|+\frac{\gamma}{\tilde{\beta}+\gamma}\|v-p\| \\ &\leq\biggl(L_G+\frac{L_G\|\nabla g\|+\|G\|L_g'}{\delta\beta}2\|G\| \biggr)\|x-y\|, \end{aligned} $$

where we deduce that

$$ w\in G(y)\cap \mathcal {T}_{\varGamma_j}(y)+ L\|x-y\|B(0,1), $$
(A.1)

with \(L:=L_{G}+2\|G\|(L_{G}\|\nabla g\|+\|G\|L_{g}')/\delta\beta\).

If v⋅∇g(y)>0, then by the same argument taking p=−δg(y)/∥∇g(y)∥, (A.1) holds true as well.

Finally, (A.1) implies the local Lipschitz continuity of \(G(\cdot)\cap \mathcal {T}_{\varGamma_{j}}(\cdot)\) on Γ j with the local constant L. □

Proof of Lemma 3.6

For k=1,…,m+, we set

For each \(k\in\mathbb{K}(x)\), we have \(x\in \mathcal {M}_{k}\). Up to a subsequence, there exists 0≤λ k≤1 and \(p_{k}\in \mathbb {R}^{d}\) so that

$$\frac{\mu^n_k}{t_n}\rightarrow \lambda_k,\qquad \sum _{k\in\mathbb{K}(x)}\lambda_k=1,\qquad \frac{1}{\mu^n_k}\int _{J^n_k}\dot{y}(s)ds\rightarrow p_k $$

as n→+∞. By Proposition 3.4 and the Lipschitz continuity of \(F^{\mathit {new}}_{k}\), we have

$$ \begin{aligned} p_k&=\lim_{n\rightarrow \infty}\frac{1}{\mu^n_k} \int_{J^n_k}\dot{y}(s)ds \\ &\in\lim_{n\rightarrow \infty}\frac{1}{\mu^n_k}\int_{J^n_k} F^{\mathit {new}}_k\bigl(y(s)\bigr)ds \\ &\subseteq\lim_{n\rightarrow \infty} \biggl[\frac{1}{\mu^n_k}\int _{J^n_k}F^{\mathit {new}}_k(x)ds+ \frac{1}{\mu^n_k}\int _{J^n_k}L_k\bigl\|y(s)-x\bigr\|B(0,1)ds \biggr] \\ &\subseteq\lim_{n\rightarrow \infty} \biggl[F^{\mathit {new}}_k(x)+L_k \|F\| \biggl[\frac{1}{\mu^n_k}\int_{J^n_k}sds \biggr]B(0,1) \biggr]=F^{\mathit {new}}_k(x). \end{aligned} $$

We then have

$$ \begin{aligned} p&=\lim_{n\rightarrow \infty}\frac{y(t_n)-x}{t_n} =\lim _{n\rightarrow \infty}\frac{1}{t_n}\int^{t_n}_0 \dot{y}(s)ds \\ &=\sum_{k\in\mathbb{K}(x)}\lim_{n\rightarrow \infty} \frac{\mu^n_k}{t_n} \biggl[\frac{1}{\mu^n_k}\int_{J^n_k} \dot{y}(s)ds \biggr] \\ &=\sum_{k\in\mathbb{K}(x)}\lambda_k p_k \in\sum_{k\in\mathbb{K}(x)}\lambda_k F^{\mathit {new}}_k(x) \subseteq \text {co }\bigcup_{k\in\mathbb{K}(x)} F^{\mathit {new}}_k(x). \end{aligned} $$

Now set \(\mathcal {M}:=\cup_{k\in\mathbb{K}(x)}\mathcal {M}_{k}\), and since \(y(t_{n})\in \mathcal {M}\) for all large n, we have \(p\in \mathcal {T}_{\overline{\mathcal {M}}}(x)\). Then we obtain

$$p\in\biggl(\text {co }\bigcup_{k\in\mathbb{K}(x)}F^{\mathit {new}}_k(x) \biggr) \cap \mathcal {T}_{\overline{\mathcal {M}}}(x). $$

The fact that \(F^{\mathit {new}}_{k}(z)\subseteq \mathcal {T}_{\mathcal {M}_{k}}(z)\) whenever \(z\in \mathcal {M}_{k}\) implies

$$F^{\mathit {new}}_k(x)\cap \mathcal {T}_{\overline{\mathcal {M}}}(x)=F^{\mathit {new}}_k(x) \cap \mathcal {T}_{\overline{\mathcal {M}}_k}(x) $$

whenever \(x\in\overline{\mathcal {M}}_{k}\). Hence

 □

Appendix B

We review the background in nonsmooth analysis required in our analysis. A closed set \(\mathcal {C}\subseteq \mathbb {R}^{d}\) is called proximally smooth of radius δ>0 provided the distance function \(d_{\mathcal {C}}(x):=\inf_{c\in \mathcal {C}}\|c-x\|\) is differentiable on the open neighborhood \(\mathcal {C}+\delta B(0,1)\) of \(\mathcal {C}\). For any \(c\in \mathcal {C}\), we denote the Clarke normal cone by \(\mathcal{N}_{\mathcal {C}}(c)\). Recall the tangent cone \(\mathcal {T}_{\mathcal {C}}(c)\) at \(c\in \mathcal {C}\) is defined as

$$ \mathcal {T}_{\mathcal {C}}(c)= \biggl\{v:\liminf_{t\rightarrow 0^-} \frac{d_{\mathcal {C}}(c+tv)}{t}=0 \biggr\}, $$

and in the case of \(\mathcal {C}\) proximally smooth, equals the Clarke tangent cone as the negative polar of \(\mathcal{N}_{\mathcal {C}}(c)\):

$$v\in \mathcal {T}_{\mathcal {C}}(c)\quad\Longleftrightarrow\quad\langle\zeta,v\rangle \leq0\quad\forall\zeta\in\mathcal{N}_{\mathcal {C}}(c). $$

If \(\mathcal {M}\) is an embedded C 2 manifold, \(\mathcal {C}:=\overline{\mathcal {M}}\), and \(c\in \mathcal {M}\), then \(\mathcal {T}_{\mathcal {C}}(c)\) agrees with the usual tangent space \(\mathcal {T}_{\mathcal {M}}(c)\) to \(\mathcal {M}\) at c from differential geometry (see [9, Proposition 1.9]). If in addition \(\overline{\mathcal {M}}\) is proximally smooth, then for each \(x\in\overline{\mathcal {M}}\), the tangent cone \(\mathcal {T}_{\overline{\mathcal {M}}}(x)\) is closed and convex, and thus has a relative interior denoted by \(\text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}}(x)\). Its relative boundary is defined as \(\text {\emph {r-bdry} }\mathcal {T}_{\overline{\mathcal {M}}}(x):=\mathcal {T}_{\overline{\mathcal {M}}}(x)\backslash \text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}}(x)\).

Another key assumption on the multi-domains is each domain being relatively wedged. A set \(\mathcal {C}\subseteq \mathbb {R}^{N}\) is wedged (see [9, p.166]) if at every \(x\in \text {bdry }\mathcal {C}\), \(\mathit{int}\mathcal {T}_{\overline{\mathcal {C}}}\not=\emptyset\). If \(\mathcal {C}=\overline{\mathcal {M}}\) is the closure of an embedded manifold \(\mathcal {M}\), then \(\mathcal {C}\) relatively wedged means the dimension of \(\text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\) is equal to d k.

The following result is [5, Lemma 3.1] and is the key geometrical ingredient that permits the construction of boundary trajectories of (DI).

Lemma B.1

If \(x\in\overline{\mathcal {M}_{k}}\backslash \mathcal {M}_{k}\)and \(v\in \text {\emph {r-bdry} }\mathcal {T}_{\overline{\mathcal {M}_{k}}}(x)\), then there exists an indexjfor which \(\mathcal {M}_{j}\subseteq \overline{\mathcal {M}_{j}}\), \(x\in\overline{\mathcal {M}}_{j}\), and \(v\in \mathcal {T}_{\overline{\mathcal {M}_{k}}}(x)\). Of course in this case, one hasd j<d k.

Proof

See [5, Lemma 3.1]. □

We finally give a sketch of the proof for Lemma 3.8.

Proof

A key fact is that for any \(p\in G(x)\cap \text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\), there exist τ>0 and a C 1 trajectory \(y(\cdot):[t,\tau]\to{ \mathcal {M}}_{k}\cup\{x\}\) so that

$$ y(t)=x \quad\mbox{and}\quad\dot{y}(t)=p. $$
(B.1)

Let k be such that \(x\in\overline{\mathcal {M}}_{k}\) and \(p\in G(x)\cap \mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\). If \(p\in \text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\), then the result follows by the key fact (B.1). If \(p\notin \text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\), then \(p\in \text {\emph {r-bdry} }\mathcal {T}_{\overline{\mathcal {M}}_{k}}(x)\) and hence by Lemma B.1, there exists another subdomain \(\mathcal {M}_{j}\subseteq \overline{\mathcal {M}}_{k}\) with \(x\in\overline{\mathcal {M}}_{j}\) and \(p\in \mathcal {T}_{\overline{\mathcal {M}}_{j}}(x)\). If \(p\in \text {\emph {r-int} }\mathcal {T}_{\overline{\mathcal {M}}_{j}}(x)\), then the result follows from (B.1), otherwise the argument just given can be repeated with k replaced by j. The process must eventually terminate since the dimension is decreasing at each step. □

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Basel

About this chapter

Cite this chapter

Rao, Z., Zidani, H. (2013). Hamilton–Jacobi–Bellman Equations on Multi-domains. In: Bredies, K., Clason, C., Kunisch, K., von Winckel, G. (eds) Control and Optimization with PDE Constraints. International Series of Numerical Mathematics, vol 164. Birkhäuser, Basel. https://doi.org/10.1007/978-3-0348-0631-2_6

Download citation

Publish with us

Policies and ethics