Skip to main content
Log in

State-Constrained Control-Affine Parabolic Problems I: First and Second Order Necessary Optimality Conditions

  • Published:
Set-Valued and Variational Analysis Aims and scope Submit manuscript

Abstract

In this paper we consider an optimal control problem governed by a semilinear heat equation with bilinear control-state terms and subject to control and state constraints. The state constraints are of integral type, the integral being with respect to the space variable. The control is multidimensional. The cost functional is of a tracking type and contains a linear term in the control variables. We derive second order necessary conditions relying on the concept of alternative costates and quasi-radial critical directions. The appendix provides an example illustrating the applicability of our results.

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. Aronna, M.S., Bonnans, J.F., Kröner, A: State-constrained control-affine parabolic problems II: Second-order sufficient optimality conditions. 1909.05056 (2019)

  2. Goh, B.S.: Necessary conditions for singular extremals involving multiple control variables. SIAM J. Control 4, 716–731 (1966)

    Article  MathSciNet  Google Scholar 

  3. Kelley, H.J.: A second variation test for singular extremals. AIAA J. 2, 1380–1382 (1964)

    Article  MathSciNet  Google Scholar 

  4. Dmitruk, A.V.: Quadratic conditions for a weak minimum for singular regimes in optimal control problems. Soviet Math. Doklady 18(2), 418–422 (1977)

    MATH  Google Scholar 

  5. Aronna, M.S., Bonnans, J.F., Dmitruk, A.V., Lotito, P.A.: Quadratic order conditions for bang-singular extremals. Numerical Algebra, Control and Optimization, AIMS Journal 2(3), 511–546 (2012). https://doi.org/10.3934/naco.2012.2.511

    Article  MathSciNet  MATH  Google Scholar 

  6. Maurer, H.: On optimal control problems with bounded state variables and control appearing linearly. SIAM J. Control Optimization 15(3), 345–362 (1977)

    Article  MathSciNet  Google Scholar 

  7. McDanell, J.P., Powers, W.F.: Necessary conditions for joining optimal singular and nonsingular subarcs. SIAM J. Control 9, 161–173 (1971)

    Article  MathSciNet  Google Scholar 

  8. Maurer, H., Kim, J.-H.R., Vossen, G.: On a state-constrained control problem in optimal production and maintenance, pp 289–308. Springer US, Boston, MA (2005). https://doi.org/10.1007/0-387-25805-1_17

    MATH  Google Scholar 

  9. Schättler, H: Local fields of extremals for optimal control problems with state constraints of relative degree 1. J. Dyn. Control Syst. 12(4), 563–599 (2006). https://doi.org/10.1007/s10883-006-0005-y

    Article  MathSciNet  MATH  Google Scholar 

  10. Aronna, M.S., Bonnans, J.F., Goh, B.S.: Second order analysis of control-affine problems with scalar state constraint. Math. Program. 160(1-2, Ser. A), 115–147 (2016). https://doi.org/10.1007/s10107-015-0976-0

    Article  MathSciNet  MATH  Google Scholar 

  11. Casas, E., Tröeltzsch, F., Unger, A.: Second order sufficient optimality conditions for a nonlinear elliptic control problem. J. for Analysis and its Applications (ZAA) 15, 687–707 (1996)

    MathSciNet  MATH  Google Scholar 

  12. Bonnans, J.F.: Second-order analysis for control constrained optimal control problems of semilinear elliptic systems. Appl. Math. Optim. 38(3), 303–325 (1998). https://doi.org/10.1007/s002459900093

    Article  MathSciNet  MATH  Google Scholar 

  13. Casas, E., Mateos, M., Tröltzsch, F.: Necessary and sufficient optimality conditions for optimization problems in function spaces and applications to control theory. In: Proceedings of 2003 MODE-SMAI Conference, ESAIM Proceedings, vol 13, pp 18–30. EDP Sciences (2003)

  14. Casas, E., Tröltzsch, F.: Recent advances in the analysis of pointwise state-constrained elliptic optimal control problems. ESAIM Control Optim. Calc. Var. 16(3), 581–600 (2010). https://doi.org/10.1051/cocv/2009010

    Article  MathSciNet  MATH  Google Scholar 

  15. Raymond, J.-P., Tröltzsch, F.: Second order sufficient optimality conditions for nonlinear parabolic control problems with state constraints. Discrete Contin. Dynam. Systems 6(2), 431–450 (2000)

    Article  MathSciNet  Google Scholar 

  16. Krumbiegel, K., Rehberg, J.: Second order sufficient optimality conditions for parabolic optimal control problems with pointwise state constraints. SIAM J. Control Optim. 51(1), 304–331 (2013). https://doi.org/10.1137/120871687

    Article  MathSciNet  MATH  Google Scholar 

  17. Casas, E., de Los Reyes, J.C., Tröltzsch, F.: Sufficient second-order optimality conditions for semilinear control problems with pointwise state constraints. SIAM J. Optim. 19(2), 616–643 (2008)

    Article  MathSciNet  Google Scholar 

  18. de Los Reyes, J.C., Merino, P., Rehberg, J., Tröltzsch, F.: Optimality conditions for state-constrained PDE control problems with time-dependent controls. Control. Cybern. 37(1), 5–38 (2008)

    MathSciNet  MATH  Google Scholar 

  19. Bonnans, J.F., Jaisson, P.: Optimal control of a parabolic equation with time-dependent state constraints. SIAM J. Control Optim. 48 (7), 4550–4571 (2010). https://doi.org/10.1137/080744608

    Article  MathSciNet  MATH  Google Scholar 

  20. Bonnans, J.F., Shapiro, A.: Perturbation analysis of optimization problems. Springer Series in Operations Research. Springer-Verlag, New York (2000)

    Book  Google Scholar 

  21. Lions, J.-L., Magenes, E.: Problèmes aux limites non homogènes et applications. Vol. 1. Dunod, Paris (1968)

    MATH  Google Scholar 

  22. Aubin, J.-P.: Un théorème de compacité. C. R. Acad. Sci. Paris 256, 5042–5044 (1963)

    MathSciNet  MATH  Google Scholar 

  23. Lions, J.-L.: Quelques méthodes de résolution des problèmes aux limites non linéaires. Dunod, Paris (1969)

    MATH  Google Scholar 

  24. Lions, J.-L.: Contrôle des systèmes distribués singuliers. Méthodes Mathématiques de l’Informatique, vol. 13. Gauthier-Villars, Montrouge (1983)

  25. Bonnans, J.F., Hermant, A.: Second-order analysis for optimal control problems with pure state constraints and mixed control-state constraints. Ann. Inst. H. Poincaré Anal. Non Linéaire 26(2), 561–598 (2009). https://doi.org/10.1016/j.anihpc.2007.12.002

    Article  MathSciNet  MATH  Google Scholar 

  26. Maurer, H.: On the minimum principle for optimal control problems with state constraints. Schriftenreihe des Rechenzentrum 41, Universität Münster (1979)

  27. Evans, L.C.: Partial differential equations. Amer. Math Soc., Providence, RI (1998). Graduate Studies in Mathematics 19

    MATH  Google Scholar 

  28. Dmitruk, A.V.: Jacobi type conditions for singular extremals. Control & Cybernetics 37(2), 285–306 (2008)

    MathSciNet  MATH  Google Scholar 

  29. Lieberman, G.M.: Second order parabolic differential equations. World Scientific Publishing Co., Inc., River Edge, NJ (1996). https://doi.org/10.1142/3302

    Book  Google Scholar 

  30. Bonnans, J., Giorgi, D., Grélard, V, Heymann, B., Maindrault, S., Martinon, P., Tissot, O., Liu, J.: Bocop – A collection of examples. INRIA. http://www.bocop.org (2017)

Download references

Acknowledgements

The authors thank the two referees for their careful reading and useful remarks.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. Soledad Aronna.

Additional information

Publisher’s Note

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

The first author was supported by FAPERJ, CNPq and CAPES (Brazil) and by the Alexander von Humboldt Foundation (Germany). The second author thanks the ‘Laboratoire de Finance pour les Marchès de l’Energie’ for its support. The second and third authors were supported by a public grant as part of the Investissement d’avenir project, reference ANR-11-LABX-0056-LMH, LabEx LMH, in a joint call with Gaspard Monge Program for optimization, operations research and their interactions with data sciences. This is the first part of a work on optimality conditions for a control problem of a semilinear heat equation. More precisely, the full version, available at arXiv:1906.00237v1, has been divided in two, resulting in the current manuscript (that corresponds to Part I) and arXiv:1909.05056 (which is Part II).

Appendices

Appendix A: Strong Solutions of the Heat Equation

We consider the heat equation with Dirichlet boundary condition:

$$ \dot y - {\Delta} y = f \text{in \textit{Q}}, y(x,0)= y_0(x); y = h \text{on \ensuremath{{\Sigma}}}. $$
(A.1)

We have the following result, see Lieberman [29, Thm 7.32, p. 182]:

Theorem A.1

Let r ≥ 2, wW2,1,r(Q) and fLr(Q). Setting y0 := w(⋅, 0) and h := τΣw (trace of w over Σ), equation Eq. A.1has a unique solution yW2,1,r(Q). In addition there exists C > 0 such that

$$ \| y \|_{W^{2,1,r}(Q)} \leq C \left( \| f\|_{L^r(Q)} + \| w \|_{W^{2,1,r}(Q)} \right). $$
(A.2)

Corollary A.2

Given r ≥ 2, \(y_{0}\in {W^{1,r}_{0}({\Omega })\cap W^{2,r}({\Omega })}\) and fLr(Q), equation Eq. A.1has, for h = 0, a unique solution yW2,1,r(Q) that satisfies

$$ \| y \|_{W^{2,1,r}(Q)} \leq C \left( \| f\|_{L^r(Q)} + \| y_0 \|_{W^{2,r}({\Omega})} \right). $$
(A.3)

Proof

Apply Theorem A.1 with w(x,t) := y0(x). It is clear that wW2,1,r(Q) and that w has trace y0 at time 0 and zero trace over Σ. The conclusion follows. □

By the standard Sobolev embeddings, we have the continuous inclusion

$$ W^{2,1,r}(Q) \subset W^{1,r}(Q) \subset L^{\infty}(Q), \quad \text{if \ensuremath{r>n+1}.} $$
(A.4)

This allows to prove the following.

Theorem A.3

Assume that \(u\in L^{\infty }(0,T)\), \(y_{0}\in {W^{1,r}_{0}({\Omega })\cap W^{2,r}({\Omega }) }\) and fLr(Q), with r > n + 1. Then the state Eq. 2.1 has a unique solution y[u, y0, f] in W2,1,r(Q), and the mapping y[u, y0, f] is of class \(C^{\infty }\) from \(L^{\infty }(0,T) \times {W^{1,r}_{0}({\Omega })\cap W^{2,r}({\Omega }) }\times L^{r}({\Omega })\) into W2,1,r(Q).

Proof

We have that g := −Δy0 belongs to Lr(Ω). Let \(y^{\pm }_{0}\) be the unique solution of \(-{\Delta } y^{\pm }_{0} =g^{\pm }\) in Ω, where \(g^{+}:=\max \limits (g,0)\) and \(g^{-}:=-\min \limits (g,0)\), with homogeneous Dirichlet condition on the boundary. Set \(f^{+} := \max \limits (f,0)\) and \(f^{-}:=-\min \limits (f,0)\). Denote by y+ (resp., y) the solution of the state Eq. 2.1 when (y0,f) is \((y_{0}^{+},f^{+})\) (resp. \((y_{0}^{-},f^{-})\)). By the monotonicity results in Lemma 2.3, we have that − yyy+. Now let y++, y−− denote the solutions of the state Eq. 2.1 when (y0,f) is \((y_{0}^{+},f^{+})\), \((y_{0}^{-},f^{-}),\) respectively and, in addition, γ = 0. We claim that − y−−≤−yyy+y++. Indeed, for zY, set \(H_{u} z := \dot z - {\Delta } z - z {\sum }_{i} u_{i} b_{i}\). Then

$$ H_u y^{+} = f^+-\gamma (y^+)^3 \leq f^+ = H_u y^{++}. $$
(A.5)

Since y+ and y++ have the same initial conditions, it follows that y+y++. In an analogous way, it can be proved that − y−−≤−y.

Since \(y_{0}^{\pm } \in {W^{1,r}_{0}({\Omega })\cap W^{2,r}({\Omega }) }\) and f±Lr(Q), by Corollary A.2, y++ and y−− belong to W2,1,r(Q) and, therefore, since r > n + 1, they are also elements of \(L^{\infty }(Q)\). So, \(y \in L^{\infty }(Q)\). Consequently, Huy = fγy3Lr(Ω) and, by Theorem A.1 again, yW2,1,r(Q).

We recall that, for r > n + 1, Yr denotes the set of elements of W2,1,r(Q) with zero trace on Σ, and \({Y^{0}_{r}}\) denotes the trace of Yr at time zero. Endowed with the “trace norm”, \({Y^{0}_{r}}\) is a Banach space that contains \( W^{1,r}_{0}({\Omega }) \cap W^{2,r}({\Omega })\) in view of the proof of the above Corollary A.2 (by Lions [24, p. 20], \({Y^{0}_{r}}\) is a subset of W2 − 2/r,r(Ω)). That (u,y0,f)↦y[u,y0,f] is of class \(C^{\infty }\) is a consequence of the Implicit Function Theorem applied to the mapping F from \(Y_{r}\times L^{\infty }(0,T) \times {Y_{r}^{0}} \times L^{r}(Q)\) into \(L^{r}(Q)\times {Y^{0}_{r}}\), defined by $$ F(y,u,y_0,f) := (H_u y + \gamma y^3, y(0)- y_0). $$

$$ F(y,u,y_0,f) := (H_u y + \gamma y^3, y(0)- y_0). $$
(A.6)

The key step is to prove that the partial derivative DyF is bijective; this can be done easily, taking advantage of the fact that \(W^{2,1,r}(Q) \subset L^{\infty }(Q)\) when r > n + 1. □

Appendix B: An Example

Since we made a number of hypotheses about the optimal trajectory, especially at junction points, it is useful to give an example where these hypotheses are satisfied. For that purpose we discuss a particular case in which the original optimal control problem can be reduced to the optimal control of a scalar ODE.

Let Ω = (0, 1), and denote by \(c_{1}(x):=\sqrt 2 {\sin \limits } \pi x\) the first (normalized) eigenvector of the Laplace operator.

We assume that γ = 0, the control is scalar (m = 1), b0 ≡ 0 and b1 ≡ 1 in Ω, and that f ≡ 0 in Q. Then the state equation with initial condition c1 reads

$$ \dot y(x,t) - {\Delta} y(x,t) = u(t) y(x,t); \quad (x,t) \in (0,1)\times (0,T), \quad y(x,0) = c_1(x),\quad x\in {\Omega}. $$
(B.1)

It is easily seen that the state satisfies y(x,t) = y1(t)c1(x), where y1 is solution of

$$ \dot y_1(t) + \pi^2 y_1(t) = u(t) y_1(t); \quad t \in (0,T), \quad y_1(0) = y_{10}=1. $$
(B.2)

We set T = 3 and consider the state constraint Eq. 3.17 with q = 1 and d1 := − 2, and the cost function Eq. 2.5 with α1 = 0.

The state constraint reduces to

$$ y_1(t) \leq 2,\quad t\in [0,3]. $$
(B.3)

As target functions take ydT := c1 and \(y_{d}(x,t) := \hat {y}_{d}(t) c_{1}(x)\) with

$$ \hat{y}_d(t) := \left\{ \begin{array}{lll} 1.5e^{t} &\quad \text{for }t\in (0, \log 2),\\ 3 &\quad \text{for }t\in (\log 2, 1),\\ 4- t &\quad \text{for }t\in (1, 3). \end{array}\right. $$
(B.4)

We assume that the lower and upper bounds for the control are \(\check {u} := -1\) and \(\hat {u} :=\pi ^{2}+1\). We will check that the optimal control is

$$ \bar{u}(t) := \left\{ \begin{array}{ll} \hat{u} & \quad \text{for } t\in (0, \log 2), \\ \pi^2 & \quad \text{for } t\in (\log 2, 2),\\ \pi^2 -1/\hat{y}_d & \quad \text{for } t\in (2, 3). \end{array}\right. $$
(B.5)

Thus, for the optimal state we have

$$ \bar{y}_1(t) := \left\{ \begin{array}{lll} e^t & \quad\text{for } t\in (0, \log 2), &\\ 2 & \quad\text{for } t\in (\log 2,2), &\\ 4-t & \quad \text{for } t\in (2, 3). & \end{array}\right. $$
(B.6)

The above control is feasible. The trajectory \((\bar {u},\bar {y})\) is optimal since for any t ∈ (0,T), the state \(\bar {y}_{1}(t)\) has the best possible value (in order to approach \(\hat y_{d}\) and minimize the cost function) that respects the state constraint.

Let us check Hypothesis 4.1 for this example. Conditions 1 and 2 are obviously satisfied. For the constraint qualification in Condition 3 consider the linearized state equation with unique z1[v]:

$$ \dot z_1=(\bar{u} - \pi^2) z_1 + v\bar{y}_1; \quad z_1(0)=0, $$
(B.7)

with \(v(t):= \check {u}-\bar {u}(t) < 0\). One easily checks that z1[v](t) < 0 for all t > 0. Hence, we can find ε > 0 such that

$$ g_1(\bar{y}(\cdot,t)) + g_1^{\prime}(\bar{y}(\cdot,t))z_1[v](\cdot,t) = \bar{y}_1(t) -2 + z_1(t) < -\varepsilon,\quad \text{for all } t\in (0,T). $$
(B.8)

Conditions 4 holds, since

$$ M(t)=\bar M_1(t)={\int\limits}_{\Omega} c_1(x) \bar{y}(x,t)\text{d} x=\bar{y}_1(t)>0\quad \text{for } t \in (0,T). $$
(B.9)

For Condition 5 we have

$$ \operatorname{dist}(t,I^C_1)= \begin{cases} \log 2 -t & \text{for } t \in (0, \log 2), \\ 0 & \text{for } t\in (\log 2, 2), \\ t- 2 & \text{for } t \in (2,3), \end{cases} $$
(B.10)

and hence,

$$ g_1(\bar{y}(\cdot,t))=\bar{y}_1(t)-2 \le - \operatorname{dist}(t,I^{C}_1). $$
(B.11)

Conditions 6 and 8 hold by the choice of the control in Eq. B.5. Condition 7 holds by definition.

We solve this problem numerically using BOCOP [30] and get the optimal control and state given in Fig. 1.

Fig. 1
figure 1

Optimal control and state for the example

We now discuss the second order optimality condition for this example. The costate equation is

$$ -\dot p +Ap = c_1 (\bar{y}_1-\hat y_d) + c_1\dot{\mu}_1,\quad p(\cdot,T)= \bar{y}(T)-y_{dT} = 0 $$
(B.12)

with A as defined in Eq. 2.20. Since \(\bar {y}\) and yd are colinear to c1, it follows that p(x,t) = p1(t)c1(x), and

$$ -\dot p_1 + \pi^2 p_1 = \bar{u} p_1 + \bar{y}_1 - \hat{y}_{d} + \dot{\mu}_1; \quad p_1(3)=0. $$
(B.13)

Over (2,3), \(\dot {\mu }_{1}=0\) (sate constraint not active) and \(\bar {y}_{1} = \hat {y}_{d}\), therefore p1 and p identically vanish. Over \((\log 2,2)\), \(\bar {u}\) is out of bounds and therefore

$$ 0 = {\int\limits}_{\Omega} p(x,t) \bar{y}(x,t) = p_1(t) \bar{y}_1(t) {\int\limits}_{\Omega} c_1(x)^2 = 2p_1(t). $$
(B.14)

It follows that p1 and p also vanish on \((\log 2,2)\) and that

$$ \dot \mu_1 = -(\bar{y}_1 - \hat{y}_{d} ) >0, \quad \text{a.a. }t\in (\log 2,2). $$
(B.15)

Over \((0,\log 2),\) the control attains its upper bound, then

$$ -\dot p_{1} = p_{1} - \frac{1}{2} e^{t} $$
(B.16)

with final condition \(p_{1}(\log 2)=0\), so that

$$ p_1(t) = \frac{e^{t}}{4} - e^{-t}. $$
(B.17)

As expected, p1 is negative.

Next, the linearized state equation at \((\bar {u},\bar {y})\) reads

$$ \dot z - {\Delta} z = \bar{u} z + v \bar{y}; \quad z(\cdot,0) = 0. $$
(B.18)

Since \(\bar {y}=\bar {y}_{1}(t) c_{1}(x)\), we deduce that z = z1(t)c1(x), with z1 solution of

$$ \dot z_1 + \pi^{2} z = \bar{u} z_{1} + v \bar{y}_{1}; \quad z_{1}(0) = 0. $$
(B.19)

Therefore if (v,z) satisfy the linearized state equation

$$ \begin{aligned} {\mathcal Q}[p](z,v) &= {\int\limits}_{Q} (z^2 + p v z)\text{d} x \text{d} t + {\int\limits}_{\Omega} z(x,T)^2 \text{d} x = {\int\limits}_0^3 (z_1(t)^2 + p_1(t) v(t) z_1(t) ) \text{d} t + z_1(3)^2. \end{aligned} $$
(B.20)

If in addition v is a critical direction, since v = 0 and z1 = 0 a.e. on (0, 2), and p1(t) = 0 on (2, 3), we get

$$ \begin{aligned} {\mathcal Q}[p](z,v) &= {\int\limits}_2^3 z_1(t)^2 \text{d} t + z_1(3)^2. \end{aligned} $$
(B.21)

Thus, \({\mathcal Q}\) is non-negative for any critical directions (z[v],v), in accordance with the second-order necessary condition of Theorem 4.7.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aronna, M.S., Bonnans, J.F. & Kröner, A. State-Constrained Control-Affine Parabolic Problems I: First and Second Order Necessary Optimality Conditions. Set-Valued Var. Anal 29, 383–408 (2021). https://doi.org/10.1007/s11228-020-00560-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11228-020-00560-2

Keywords

Navigation