Robustness under control sampling of reachability in fixed time for nonlinear control systems
Abstract
Under a regularity assumption we prove that reachability in fixed time for nonlinear control systems is robust under control sampling.
Index Terms:
Nonlinear control systems, reachability, sampled-data controls, piecewise constant controls, regular controls.I Introduction and main result
Let , and be fixed. In this work we consider the general nonlinear control system
(CS) |
where the dynamics is a continuous mapping, of class with respect to its first two variables.111This regularity assumption can be relaxed at several occasions in the paper (see Remark 15 for details). We say that a pair is a solution to (CS) if is an absolutely continuous function (called state or trajectory) and is an essentially bounded measurable function (called control) such that for a.e. . Throughout the paper we fix a starting point and a nonempty subset of standing for the set of control constraints. We say that a target point is -reachable in time from if there exists a solution to (CS), with , such that and .
We now sample the control over the time interval : given a partition of , consisting of real numbers satisfying , for some , we consider the set of all possible piecewise constant functions satisfying , for some , for every and every . We denote by the norm of the partition. We say that a target point is -reachable in time from if there exists a solution to (CS), with , such that and .
In this paper we investigate the following question: assuming that a target point is -reachable in time from and given a partition of , is the point also -reachable in time from ? In other words, how robust is reachability in fixed time under control sampling? Without any specific assumption, even for small values of , in general fails to be -reachable in time from , as shown in the following example.
Example 1.
Take , and for all . The target point is -reachable in time from the starting point with the control for a.e. . However there is no other control steering the control system from to in time . Therefore, given any partition of , even with a small value of , the target point is not -reachable in time from .
Our main result is the following.
Theorem 1.
Assume that is convex and let be a target point that is -reachable in time from with a control . If is weakly -regular, then there exists a threshold such that is -reachable in time from for any partition of satisfying .
The key concept of weakly -regular control is defined, commented and characterized in Section II, in relation with local reachability results. Theorem 1 is discussed in detail in Section III. In particular we emphasize here that the convexity assumption made on and the smoothness assumption made on can both be relaxed (see Remarks 13, 14 and 15 for details). All proofs are done in Sections IV and V.
II Recap on local reachability results
This section gathers in a concise way a number of local reachability results, helpful for various purposes all along this paper. Most of these results are well known in the literature (see, e.g., [1, 5, 9, 18, 21, 22, 25] and references therein), while others are less known or even new.
In Section II-A we deal with the unconstrained control case (i.e., when ), recalling how the implicit function theorem can provide local reachability results thanks to the notion of strongly regular control. In Section II-B we show how to extend this approach under convex control constraints (i.e., when is a convex subset of ), thanks to the notion of strongly -regular control and to a conic version of the implicit function theorem. In Section II-C we treat the general control constraints case (i.e., when is a general subset of ), thanks to the notion of weakly -regular control and using needle-like variations. These different notions lead to distinct results, that we comment further in Section II-D.
We first recall some basic facts and terminology. A control is said to be admissible when there exists , starting at , such that is a solution to (CS). In that case the trajectory is unique and will be denoted by . The set of all admissible controls is an open subset of and the end-point mapping is the mapping defined by for every . Therefore a target point is -reachable in time from if and only if belongs to the -accessible set given by . Reachability in time from is thus related to a surjectivity property of .
II-A Without control constraint
All results in this section are classical (see, e.g., [1, 5, 9, 25]). When , i.e., when there is no control constraint, some conditions ensuring surjectivity of are well known. For instance, when the control system (CS) is linear and autonomous, i.e., for all , where and are constant matrices and is a continuous function, we have , and is surjective if and only if the pair satisfies the classical Kalman condition. For a general nonlinear control system (CS), global surjectivity of cannot be ensured in general. But, thanks to the implicit function theorem, local surjectivity can be established (see Proposition 1 below, proved in Section IV-A).
Definition 1 (strongly222With respect to the existing literature, we add the word “strongly”, in contrast to the notion of “weakly” regular control defined in Section II-C. regular control).
A control is said to be strongly regular if the Fréchet differential is surjective, i.e., . A control is said to be weakly singular if it is not strongly regular, i.e., is a proper subspace of .
Proposition 1.
If a control is strongly regular, then there exist an open neighborhood of and a mapping of class satisfying and for every . In particular, any point of is -reachable in time from , and thus belongs to the interior of the -accessible set.
A Hamiltonian characterization of weakly singular controls (recalled in Proposition 2 further) can be derived from the expression of the Fréchet differential of given by
(1) |
for every and every , where is the unique solution to
The Hamiltonian associated to (CS) is the function defined by
for all , where is the Euclidean scalar product in .
Definition 2 (weak extremal lift).
A weak extremal lift of a pair , where , is a triple where (called adjoint vector) is a solution to the (linear) adjoint equation
(AE) |
for a.e. , satisfying the null Hamiltonian gradient condition
(NHG) |
for a.e. . The weak extremal lift is said to be nontrivial if is nontrivial.
Proposition 2.
A control is weakly singular if and only if the pair admits a nontrivial weak extremal lift.
While the definition of weakly singular control is quite abstract, the above classical Hamiltonian characterization (proved in Section IV-B) is practical. For example one can easily prove that the control in Example 1 is weakly singular.
As a consequence of Propositions 1 and 2, if a pair , for some , has no nontrivial weak extremal lift, then any point of an open neighborhood of is -reachable in time from . Note that the contrapositive statement corresponds to a weak version of the geometric Pontryagin maximum principle: if , for some , belongs to the boundary of the -accessible set, then the pair admits a nontrivial weak extremal lift.
II-B With convex control constraints
When is a proper subset of , i.e., when there are control constraints, reachability properties are more difficult to establish in general. We refer the reader to [8] for conic-type conditions for autonomous linear control systems, to [6, 17] for single-input control-affine systems in dimensions and , and to [4, 24] for more general systems.
In Section II-A we have recalled that, in the absence of control constraint, local reachability can be ensured thanks to the classical implicit function theorem, by assuming that is surjective for some . When there are control constraints, a powerful approach is to use constrained versions of the implicit function theorem (as in [3, 4, 16]). When is convex, the required hypothesis for a control is a conic surjectivity assumption made on as follows.
Definition 3 (strongly -regular control).
Assume that is convex. A control is said to be strongly -regular if , where is the (convex) tangent cone to at defined by
The control is said to be weakly -singular when it is not strongly -regular, i.e., is a proper convex subcone of .
Proposition 3.
Assume that is convex. If a control is strongly -regular, then there exist an open neighborhood of and a continuous mapping satisfying and for every . In particular, any point in is -reachable in time from , and thus belongs to the interior of the -accessible set.
The proof of Proposition 3, based on the conic implicit function theorem [3, Theorem 1], is provided in Section IV-C.
Similar results to Proposition 3 are known in the literature. For example it echoes results obtained in [4, 16] in which the sufficient condition is settled as a constrained controllability property of the linearized control system. Such a condition is however not easy to check in practice.
As in the unconstrained control case (Section II-A), we next provide a practical Hamiltonian characterization of weakly -singular controls.
Definition 4 (weak -extremal lift).
Assume that is convex. A weak -extremal lift of a pair , where , is a triple where (called adjoint vector) is a solution to the adjoint equation (AE) satisfying the Hamiltonian gradient condition
(HG) |
for a.e. , where
is the normal cone to at . The weak -extremal lift is said to be nontrivial if is nontrivial.
Proposition 4.
Assume that is convex. A control is weakly -singular if and only if the pair admits a nontrivial weak -extremal lift.
The proof of Proposition 4, using in particular needle-like variations (recalled in Section II-C), is done in Section IV-D.
As a consequence of Propositions 3 and 4, when is convex, if a pair , for some , has no nontrivial weak -extremal lift, then any point of an open neighborhood of is -reachable in time from . The contrapositive statement corresponds to a weak version of the geometric Pontryagin maximum principle in the presence of convex control constraints: when is convex, if the point , for some , belongs to the boundary of the -accessible set, then the pair admits a nontrivial weak -extremal lift.
Remark 1.
When is convex, it is clear that, if a control is strongly -regular, then it is strongly regular. The converse is not true in general, as shown in the next example, but is true when takes its values in the interior of (see Proposition 7 further), in particular when .
Example 2.
Take , and for all . From the Hamiltonian characterizations, the constant control is strongly regular and weakly -singular.
Remark 2.
Note that the conclusions of Propositions 1 and 3 are distinct: the local right-inverse mapping is of class in Proposition 1 (in the unconstrained control case), while it is (only) continuous in Proposition 3 (in the convex control constraints case). In the latter, obtaining smoothness is an open question. Indeed, in all references on constrained implicit function theorems we found (such as [3]), the continuity of the local right-inverse mapping is established, but obtaining smoothness does not seem to be an easy issue.
II-C With general control constraints
When is convex and for a given control , the set consists of all elements (called the weak -variation vectors associated with ) generated by conic -perturbations of the control , where and , in the sense that
The set can be seen as a first-order conic convex approximation of the -accessible set at .
In the general control constraints case where is not assumed to be convex, we can use sophisticated -perturbations, well known in the literature as needle-like variations. Precisely a needle-like variation of a given control is defined by
(2) |
for a.e. and for , with , where stands for the full-measure set of all Lebesgue points in of the essentially bounded measurable function . In that framework, it is well known that belongs to for sufficiently small and that
(3) |
where is the unique solution to
The elements , with , are called the strong -variation vectors associated with .
Definition 5 (-Pontryagin cone).
The -Pontryagin cone of a control , denoted by , is the smallest convex cone containing all strong -variation vectors associated with .333In the literature, usually the -Pontryagin cone of a control is defined as the smallest closed convex cone containing all strong -variation vectors associated with (see, e.g., [18]). As explained in Remark 3, considering the closure (or not) has no impact on the notions and results presented in this paper. Nevertheless we emphasize that the multiple needle-like variations of the control (see Section IV-E) generate (only) the -Pontryagin cone of as defined in Definition 5 (i.e., without closure).
A strong -variation vector associated with a control is generated in (3) by using a single needle-like variation (2): this is standard in the literature. What is less standard is that, actually, the -Pontryagin cone, which consists of all conic convex combinations of strong -variation vectors associated with , can be generated by using multiple needle-like variations (see Section IV-E). Hence the set can be seen as a first-order conic convex approximation of the -accessible set at . Note that, when is convex, it is larger than (in the sense of Remark 3) which leads to the following weakened notion of -regularity.
Definition 6 (weakly -regular control).
A control is said to be weakly -regular if . The control is said to be strongly -singular when it is not weakly -regular, i.e., is a proper convex subcone of .
Although the -Pontryagin cone of a control cannot be written as the range of a differential taken in an appropriate sense (see Remark 16), the next proposition can be obtained by applying the conic implicit function theorem [3, Theorem 1] to a restriction of to a multiple needle-like variation (see the proof in Section IV-E).
Proposition 5.
If a control is weakly -regular, then there exist an open neighborhood of and a mapping , that is continuous when endowing the codomain with the -metric, satisfying and for all . In particular any point in is -reachable in time from , thus belongs to the interior of the -accessible set.
Like in Sections II-A and II-B, we next provide a Hamiltonian characterization of strongly -singular controls (see Proposition 6 below, proved in Section IV-F).
Definition 7 (strong -extremal lift).
A strong -extremal lift of a pair , where , is a triple where (called adjoint vector) is a solution to the adjoint equation (AE) satisfying the Hamiltonian maximization condition
(HM) |
for a.e. . The strong -extremal lift is said to be nontrivial if is nontrivial.
Proposition 6.
A control is strongly -singular if and only if the pair admits a nontrivial strong -extremal lift.
From Propositions 5 and 6, if a pair , where , has no nontrivial strong -extremal lift, then any point of an open neighborhood of is -reachable in time from . The contrapositive statement coincides exactly with the well known geometric Pontryagin maximum principle: if , for some , belongs to the boundary of the -accessible set, then the pair admits a nontrivial strong -extremal lift.
Remark 3.
Let . Since is convex, we have if and only if its closure satisfies . When is convex, we have and thus, if is strongly -regular, then it is weakly -regular.444This fact can also be derived from the Hamiltonian characterizations. The converse is not true in general, as shown in the following two examples.
Example 3.
Take , and for all . From the Hamiltonian characterizations, the constant control is weakly -regular and weakly -singular.
Example 4.
Take , , and for all . From the Hamiltonian characterizations, the constant control is weakly -regular and weakly -singular.
Remark 4.
No relationship can be established between strong regularity and weak -regularity in general. One can check that Example 2 provides a control that is strongly regular and strongly -singular, and that Example 3 provides a control that is weakly singular and weakly -regular. We refer to Propositions 7 and 8 further for relationships in special cases.
Remark 5.
It follows from the Hamiltonian characterization that, if a control is strongly -singular on the interval with the starting point , then it is also strongly -singular on any subinterval of nonempty interior with the starting point . When is convex, the same assertion is true when replacing “strongly -singular” with “weakly -singular”.
Remark 6.
Note that the conclusions of Propositions 3 and 5 are distinct. In Proposition 3, when is convex and the control is strongly -regular, the controls allowing to reach an open neighborhood of can be chosen close to in -topology. In Proposition 5, when the control is (only) weakly -regular, closedness is obtained in the weaker -topology (this is because needle-like variations are -perturbations). There are similar subtleties in Section III due to the fact that piecewise constant functions are dense in when endowed with the -norm (but not with the natural -norm).
II-D Additional comments and results
The next proposition, which seems to be new, follows straightforwardly from the Hamiltonian characterizations and from the fact that, when is convex, the normal cone to at any interior point of is reduced to .
Proposition 7.
Let , where is the interior of .
- (i)
-
(ii)
When is convex, is strongly regular if and only if is strongly -regular.555This fact is obvious when belongs to since then . However note that the inclusion may be strict (for a counterexample, take , and for a.e. ).
Remark 7.
The control system (CS) is said to be control-affine when for all , where and are continuous mappings, of class with respect to their first variable. In that context we have
for all and the next proposition follows straightforwardly.
Proposition 8.
Assume that the control system (CS) is control-affine and let .
-
(i)
If is strongly -regular, where is the convex hull of , then is weakly -regular.
- (ii)
-
(iii)
When is convex, is weakly -regular if and only if is strongly -regular.
As a particular case of control-affine system, the control system (CS) is said to be linear when for all , where , and are continuous functions. In that context and is affine. An example given in Appendix -E shows that the converse of the geometric Pontryagin maximum principle stated at the end of Section II-C is not true in general.666Since in that example, it also shows that the converses of the weak versions of the geometric Pontryagin maximum principle stated at the end of Sections II-A and II-B are also not true in general. However, for linear control systems, the converse is true, as stated in the next proposition (proved in Section IV-G).
Proposition 9.
Assume that the control system (CS) is linear and let . Then belongs to the interior of the -accessible set if and only if is weakly -regular.
Remark 8.
Assume that the control system (CS) is linear and autonomous (i.e., and are constant). Since and is affine, a control is strongly regular if and only if is surjective, if and only if is surjective, if and only if the pair satisfies the Kalman condition. This characterization does not depend on . Hence, under the Kalman condition, any control is strongly regular on any subinterval of nonempty interior and from any starting point. Thus, under the Kalman condition and using Remark 7, if a control takes its values in along a subinterval of nonempty interior, then is weakly -regular (and even strongly -regular if is convex) on from any starting point.
We now introduce a last notion which will be instrumental in order to relax the convexity assumption made on in our main result (see Remark 13 and 14 further for details).
Definition 8 (parameterization of ).
We say that is parameterizable by a nonempty subset of , with , if there exists a mapping satisfying and, for every , there exists such that .
Example 5.
Using a standard measurable selection theorem, we see that the two-dimensional unit circle is parameterizable by the interval .
In the context of Definition 8, the control system (CS) has the same trajectories as the control system (CS’) given by
(CS’) |
starting at the same initial point , where the dynamics is defined by for all and where is the control constraint set. Precisely, for a control , any control satisfying belongs to the set of all admissible controls for (CS’), and . Furthermore, by the Hamiltonian characterization, if is weakly -regular for (CS), then is weakly -regular for (CS’). We say that weak -regularity is preserved by parameterization. However, when and are convex, strong -regularity may not be preserved by parameterization, as shown in the following example.
Example 6.
Consider the framework of Example 2. By the Hamiltonian characterization, the constant control is strongly -regular. Considering the parameterization of by itself, with the mapping defined by for every , we recover the control system considered in Example 3 in which the constant control , which satisfies , is weakly -singular.
III Robustness under control sampling of reachability in fixed time
When dealing with controls , the control system (CS) is said to be with permanent controls, in the sense that the control value can be modified at any real time . Otherwise, when dealing with piecewise constant controls , for a given partition of , the control system (CS) is said to be with sampled-data controls (see [20]) which are a particular case of nonpermanent controls, in the sense that the control value can be modified only at the sampling times and remains frozen on each sampling interval .
In [7] we proved that the optimal sampled-data control of a general unconstrained linear-quadratic problem converges pointwisely to the optimal permanent control when the norm of the corresponding partition converges to zero. In an ongoing work we extend this result to a general nonlinear setting, moreover under convex control constraints and with fixed endpoint. For this purpose, robustness under control sampling of reachability in fixed time of the fixed endpoint has to be investigated. This issue has motivated the present work.
For any control , we introduce the properties (4) and () defined by
(4) |
and
() |
where is the set of all partitions of and where is the -accessible set. Example 1 shows that, for a given control , Property () is not satisfied in general. Most of the literature focuses on establishing sufficient conditions for properties related to Property () (see Remark 12 further for details and references). One of the novelties of the present work is to provide sufficient conditions for the stronger Property (4). The interest of the threshold in Property (4) (which is not considered in Property ()) is twofold. On one hand, its existence is instrumental to extend the convergence result obtained in [7] to a general nonlinear setting under convex control constraints and with fixed endpoint, precisely in order to guarantee that the corresponding optimal sampled-data control problem is feasible for partitions of sufficiently small norm. On the other hand, the nonexistence of such a threshold implies that -reachability in time from of the final point is sensitive to small perturbations of the partition of , in the sense of the next proposition (proved in Section V-A and illustrated in Remark 9 further).
Proposition 10.
Let . If Property (4) is not satisfied, then, for any partition of and any , there exists a partition of such that for all and such that is not -reachable in time from .
This section is organized as follows. In Section III-A we first investigate the condition that , for some , belongs to the interior of the -accessible set. Our main result (Theorem 1), which is valid under the stronger condition that is weakly -regular, is discussed in Section III-B.
III-A Final point in the interior of the -accessible set
Let . Here we focus on the condition that belongs to the interior of the -accessible set. The next example, based on a commensurability rigidity, shows that it is not a sufficient condition for Property (4).
Example 7.
Take , , and for all . The target point is -reachable in time from the starting point with the control for a.e. and for a.e. . By the Hamiltonian characterization, the control is weakly -regular and thus belongs to the interior of the -accessible set (see Proposition 5). However, for any given partition of , belongs to the -accessible set if and only if there exists a subfamily of sampling intervals associated with whose sum of lengths is equal to . As a consequence, for any partition of containing only rational sampling times (with norm arbitrarily small), is not -reachable in time from . We conclude that Property (4) is not satisfied (while Property () is).
In Example 7, the set is not convex. However note that another counterexample, in which is convex, is provided in Appendix -E.
Remark 9.
Example 7 illustrates Proposition 10 in the sense that, given any partition of (even such that the target point is -reachable in time from ) and given any , there always exists a partition of containing only rational sampling times such that for all , and thus such that is not -reachable in time from . We provide in the following example a similar illustration of Proposition 10 with convex.
Example 8.
Take , , , and for all . Consider the starting point . The point belongs to the segment if and only if the corresponding control takes its values in . As a consequence, by considering the target point , we find the same conclusions as in Example 7.
In the one-dimensional case , the next proposition is obtained (see the proof in Section V-B based on the fact that one-dimensional connected sets are convex).
Proposition 11.
Assume that , that is convex777Actually assuming that is connected is sufficient. and that . If , for some , belongs to the interior of the -accessible set, then Property (4) is satisfied.
III-B Comments on Theorem 1
Let . This section focuses on the condition that is weakly -regular. Example 7 shows that it is not a sufficient for Property (4) in general. However note that our main result (Theorem 1) states that, when is convex, it is a sufficient condition for Property (4).
Example 9.
In this paper we provide two different proofs of Theorem 1. A first proof is done in Section V-C, under the stronger condition that is strongly -regular. This proof uses results of Section II-B (in particular, conic -perturbations of ) and, as explained in Remark 10 further, we resort to truncated dynamics. In the second proof, given in Section V-D, we treat the case where is assumed to be (only) weakly -regular. This proof uses results of Section II-C (in particular, needle-like variations of ) and, as explained in Remark 11, we resort to the Brouwer fixed-point theorem. We think the two proofs are interesting, not only for pedagogical reasons but also because the different techniques that we introduce may be useful for other issues. Note that both proofs use, at some step, the conic implicit function theorem [3, Theorem 1] and averaging operators which project any integrable function onto a piecewise constant function.
Remark 10.
The first proof of Theorem 1, given in Section V-C under the strong -regularity assumption, relies on the conic implicit function theorem [3, Theorem 1]. However this theorem must be used in the Banach space , for some , and not in . This is because it is not true that any function in can be approximated in -norm by piecewise constant functions, while it can be in -norm with any (see Appendix -F). This leads us to extend the end-point mapping to which makes no sense a priori because the control system (CS) is nonlinear.888For example, take , and for all . Then considering -controls makes no sense. To overcome this difficulty, we introduce in Appendix -H a truncated version of the dynamics , vanishing outside of a sufficiently large compact subset of . Then the corresponding truncated end-point mapping is well defined on , but is not Fréchet-differentiable when . However it is of class when and the surjectivity of the differential of the truncated end-point mapping in -norm can be related to the surjectivity of the differential in -norm of the nontruncated end-point mapping. This is a key technical point in the first proof of Theorem 1.
Remark 11.
The second proof of Theorem 1, given in Section V-D under the weak -regularity assumption, relies on the conic implicit function theorem [3, Theorem 1] applied to the end-point mapping restricted to a multiple needle-like variation (as in the proof of Proposition 5). This second proof of Theorem 1 also uses the Brouwer fixed-point theorem. Like in [15, Lemma 3.1] or in [2, Lemma 7], the main idea is that, under appropriate assumptions, local surjectivity of a continuous mapping is preserved under small perturbations. In our context, local surjectivity of the above restriction of the end-point mapping is preserved under the perturbation due to the composition with an averaging operator (see Appendix -G) which project any control with values in onto a piecewise constant control with values in .
Remark 12.
Theorem 1 establishes robustness under control sampling of reachability in fixed time. If one does not fix the final time, robustness under control sampling of reachability is already known, and this remark is dedicated to the remarkable series of papers [12, 13, 14, 23] by Grasse and Sussmann (see also references therein) on reachability and controllability with piecewise constant controls.
-
(i)
It is established in [23, Theorem 4.2] that normal reachability of a target point in the state space implies normal reachability with a piecewise constant control. Roughly speaking, normal reachability is reachability under a surjectivity assumption which is similar to the notion of regularity considered in the present work.
-
(ii)
With another point of view (not based on a surjectivity property), it is established in [14, Theorem 3.17] that, under global controllability, the controllability can be achieved with piecewise constant controls.
-
(iii)
In [12, Remark 3.5] it is noted that if a point of the state space is normally reachable in time less than , then it belongs to the interior of the reachable set with piecewise-constant controls in time less than .
-
(iv)
It is proved in [13, Corollary 4.4] that, if the initial condition belongs to the interior of the reachable set, then this reachable set coincides with the reachable set with piecewise constant controls.
Our main result (Theorem 1) differs from the above results for two reasons. First, as underlined above, the final time is fixed in our work, while it is not in the abovementioned references. For instance, in [23, Theorem 4.2], normal reachability with a piecewise constant control is established, but a priori for a different final time (and indeed, inspection of the proof shows that, in general, ). Second, our main result (Theorem 1) states the existence of a threshold for which reachability (exactly at time ) of a target point with a piecewise constant control is guaranteed for any partition satisfying . The existence of this threshold (which is not considered in the abovementioned works) is of particular interest when considering refinements of partitions (for convergence results for instance) and for robustness of reachability under small perturbations of the partition (see Proposition 10). Furthermore, since the inclusion is not a total order over , it may occur that while . In the above references, it is not guaranteed that reachability of a target point with a -piecewise constant control implies reachability with a -piecewise constant control. With the conclusion of Theorem 1, when , it is guaranteed.
Remark 13.
The convexity assumption made on in Theorem 1 can be relaxed. Indeed let us prove that Theorem 1 is still true when is assumed to be (only) convex by parameterization, i.e., when is parameterizable (see Definition 8) by a nonempty convex subset of for some (see examples in Remark 14). In that context, for a control that is weakly -regular, there exists such that and is weakly -regular for the control system (CS’). Since is convex, there exists by Theorem 1 a threshold such that , for some , for all partitions satisfying . Introducing , we obtain that for all partitions satisfying .
Remark 14.
If is convex by parameterization (see Remark 13), then must be connected. Actually a quite large class of connected sets are convex by parameterization. For example, in the two-dimensioncal case , the unit circle , the donut-shaped set or the cross-shaped set are nonconvex connected sets that are convex by parameterization. For these sets, the conclusion of Theorem 1 holds true. However, adapting Example 7, note that the conclusion of Theorem 1 fails in general if is strongly nonconnected, i.e., when it can be written as , where and are nonempty, and there exists a mapping taking the value on and the value on .999For example, when , the set is strongly nonconnected, while the set is nonconnected (but not strongly). An open question is to extend Theorem 1 to sets that are neither convex by parameterization, nor strongly nonconnected. We emphasize that our proof of Theorem 1, when is convex, uses the averaging operators introduced in Appendix -G, which project any control with values in onto a piecewise constant control with values in (see Proposition 13). When is not convex, one has to consider other operators: one way may be to follow the approach based on the Lusin theorem [19] as developed in Appendix -F.
Remark 15.
Several statements in the present paper do not require that the dynamics is of class with respect to . Actually this assumption is required (only) when has to be considered (such as in Sections II-A and II-B where we use conic -perturbations). When using needle-like variations (which are -perturbations) such as in Section II-C, it is only required that is of class with respect to and is Lipschitz continuous with respect to on any compact subset of . In particular the conclusion of Theorem 1 remains true in that context.101010By Remark 15, Definition 8 (resp., the notion of strongly nonconnected set introduced in Remark 14) can be relaxed by considering a mapping (resp., ) that is (only) Lipschitz continuous on any compact subset of (resp., of ).
Remark 16.
As far as we know, the -Pontryagin cone of a control cannot be written as the range of a differential taken in an appropriate sense. Indeed, we explain in Section IV-E how can be generated using multiple needle-like variations which are -perturbations for any . Nevertheless, even using truncated dynamics in order to work in for some , we explain in Appendix -H that the truncated end-point mapping is not Fréchet-differentiable when and, when , the Fréchet differential of the truncated end-point mapping generates (only) weak -variation vectors. We conclude this comment by referring to the work of Gamkrelidze in [11] in which classical controls are embedded in the set of Radon measures. With this nonstandard approach, it is proved that is contained in the range of the differential of the end-point mapping considered on the set of Radon measures. Unfortunately the above embedding has a convexification effect on the dynamics and, as a result, the inclusion is (only) strict in general.
IV Proofs of results of Section II
This section is dedicated to proving the results of Section II. Most of the following proofs are known in the literature. They are recalled here because the techniques and results developed hereafter will be helpful at several occasions in Section V (devoted to proving the new results presented in Section III).
In what follows, when is a metric set, we denote by (resp. ) the open ball (resp. closed ball) centered at some of some radius .
IV-A Proof of Proposition 1
Let be strongly regular. By Definition 1 there exists a -tuple of elements of such that for all , where is the canonical basis of . We define the mapping by
for all , where is small enough to guarantee that for all , which is possible because is an open subset of . The mapping is of class and satisfies and which is invertible. By the implicit function theorem, there exists an open neighborhood of and a mapping satisfying and for all . Then it suffices to introduce the mapping defined by for all .
IV-B Proof of Proposition 2
Lemma 1.
Let and be a solution to (AE). The following statements are equivalent:
-
(i)
is a weak extremal lift of the pair ;
-
(ii)
for all .
Proof.
We set for all and all , where is defined after (1). Therefore (ii) is equivalent to for all . For all , note that and, using the adjoint equation (AE), that
for a.e. . Now let us to prove that (i) is equivalent to (ii). First let us assume (i). From the null Hamiltonian gradient condition (NHG), we have for a.e. and thus for all , which gives (ii). Now, assuming (ii), we have for every . We deduce the null Hamiltonian gradient condition (NHG), which gives (i). ∎
Let us prove Proposition 2. Let . First, assume that is weakly singular, i.e., is a proper subspace of . Hence there exists such that for all . Considering the unique solution to (AE) ending at (in particular is not trivial), we obtain that for all . By Lemma 1, is a nontrivial weak extremal lift of . Conversely, assume that is strongly regular, i.e., . By contradiction let us assume that admits a nontrivial weak extremal lift . Then there exists such that . It follows from Lemma 1 that and thus . Since the adjoint equation (AE) is linear, it follows that is trivial, which raises a contradiction.
IV-C Proof of Proposition 3
Lemma 2.
Assume that is convex and let . We have
Furthermore, for every , we have
for every , where and is such that for every .
Lemma 2 is obvious. Assume that is convex and let us prove Proposition 3. Let be strongly -regular. By Definition 3, there exists a -tuple of elements of such that
(5) |
for every , where is the canonical basis of . We define the map by
for all , where is small enough to guarantee that for every , which is possible by Lemma 2 and because is an open subset of . The mapping is of class and satisfies and thanks to (5). From the conic implicit function theorem [3, Theorem 1], there exists an open neighborhood of and a continuous mapping satisfying and for all . Then it suffices to introduce the continuous mapping defined by for all .
IV-D Proof of Proposition 4
Lemma 3.
Assume that is convex. Let and be a solution to (AE). The following statements are equivalent:
-
(i)
is a weak -extremal lift of the pair ;
-
(ii)
for all ;
-
(iii)
for all .
Proof.
The equivalence between (ii) and (iii) follows from the definition of (see Definition 3). Note that (ii) is equivalent to for all (see the definition of in the proof of Lemma 1). Now let us prove that (i) is equivalent to (ii). First let us assume (i). We infer from the Hamiltonian gradient condition (HG) that for a.e. and thus for all , which gives (ii). Now, assuming (ii), we have for every . Then, for any Lebesgue point of and of and for any , taking the needle-like variation as defined in (2), we get that for every small enough. Taking the limit , since is an appropriate Lebesgue point, we obtain that . Since and have been chosen arbitrarily, the Hamiltonian gradient condition (HG) is satisfied, which gives (i). ∎
Assume that is convex and let us prove Proposition 4. Let . Firstly, assume that is weakly -singular, i.e., is a proper subcone of . Hence belongs to its boundary and, since is also convex, by a standard separation argument, there exists such that for all . Considering the unique solution to (AE) ending at (in particular is not trivial), we obtain that for all . By Lemma 3, is a nontrivial weak -extremal lift of . Conversely, assume that is strongly -regular, i.e., . By contradiction let us assume that admits a nontrivial weak -extremal lift . There exists such that . By Lemma 3 we get that and thus . Since the adjoint equation (AE) is linear, it follows that is trivial, which raises a contradiction.
IV-E Proof of Proposition 5
Given and , we define
for every and every , which corresponds to a usual -neighborhood of , truncated with a uniform -bound. The following lemmas follow from standard techniques in ordinary differential equations theory.
Lemma 4.
Let and . For any , there exists such that and for all . Moreover the restriction of to is Lipschitz continuous when endowing with the -metric.
Definition 9 (Multiple needle-like variation).
Let . A package , with , , , consists of:
-
•
a -tuple such that ;
-
•
a -tuple with for all , and .
The multiple needle-like variation of the control is defined by
for a.e. and for all sufficiently small so that the intervals do not overlap.
Remark 17.
Lemma 5.
Remark 18.
Let . Note that, for any Lebesgue point considered in a multiple needle-like variation (see Definition 9), it is possible to consider several values for with . This additional degree of freedom is essential in order to generate the -Pontryagin cone of with multiple needle-like variations, as developed in the next remark.
Remark 19.
The -Pontryagin cone of a control is generated by multiple needle-like variations as follows. Consider some . Definition 5 gives
for some , where and for all . By gathering the Lebesgue points that are equal (and thus gathering the corresponding values , see Remark 18), we construct a package as in Definition 9 (with ) and
Denoting by , we introduce the mapping , defined by for all , where is the mapping defined in Lemma 5 and where is sufficiently small to guarantee that for all . We finally get that
because .
Now let us prove Proposition 5. Let be weakly -regular. Thus contains and for all , where is the canonical basis of . For all , Definition 5 gives
for some , where and for all , and
for some , where and for all . By gathering the Lebesgue points which are equal (and thus gathering the corresponding values , see Remark 18), we construct a package as in Definition 9 (with ). Considering the mapping defined in Lemma 5, it is clear, in the same spirit as in Remark 19, that each vector and belong to , and thus . Now we define the mapping by for all . The mapping is of class and satisfies and . From the conic implicit function theorem [3, Theorem 1], there exists an open neighborhood of and a continuous mapping satisfying and for all . Then it suffices to introduce the mapping defined by for all . By Remark 17, the mapping is continuous when endowing its codomain with the -metric.
IV-F Proof of Proposition 6
Lemma 6.
Let and be a solution to (AE). The following statements are equivalent:
-
(i)
is a strong -extremal lift of the pair ;
-
(ii)
for all ;
-
(iii)
for all .
Proof.
Let us prove Proposition 6. Let . First, assume that is strongly -singular, i.e., is a proper subcone of . Hence belongs to its boundary and, since is also convex, by a standard separation argument, there exists such that for all . Considering the unique solution to (AE) ending at (in particular is not trivial), we obtain that for all . By Lemma 6, is a nontrivial strong -extremal lift of . Conversely, assume that is weakly -regular, i.e., . By contradiction, let us assume that admits a nontrivial strong -extremal lift . Since , it follows from Lemma 6 that and thus . Since the adjoint equation (AE) is linear, it follows that is trivial, which raises a contradiction.
IV-G Proof of Proposition 9 (only the sufficient condition)
First step: assume that is convex and that belongs to the interior of the -accessible set. Let us prove that is strongly -regular (and thus is weakly -regular by Proposition 8). By contradiction assume that is weakly -singular. By Proposition 4, let be a nontrivial weak -extremal lift of the pair . Since the adjoint equation (AE) is linear, we know that . Since belongs to the interior of , there exist sufficiently small and such that . Since the control system (CS) is linear, is affine and thus . Then by Lemma 3, and thus , which raises a contradiction.
Second step: in the general control constraints case, assume that belongs to the interior of the -accessible set. Then belongs to the interior of the -accessible set. Since , we infer from the first step that is strongly -regular. We deduce that is weakly -regular from Proposition 8.
V Proofs of results of Section III
V-A Proof of Proposition 10
Remark 20.
Given a partition of , it is clear that a target point is -reachable in time from if and only if is -reachable in time from for at least one partition of such that , if and only if is -reachable in time from for all partitions of such that .
Let and assume that Property (4) is not satisfied. Let be a partition of and . Since Property (4) is not satisfied, there exists a partition of such that and such that is not -reachable in time from . For any , the intersection is not empty and we select one of its elements. For (resp. ), we choose (resp. ). Consider the partition of . Since , we know from Remark 20 that is not -reachable in time from .
V-B Proof of Proposition 11
Lemma 7 (Approximated reachability).
Given any and any , there exists a threshold such that, for any partition of satisfying , there exists such that .
Proof.
Let us prove Proposition 11. Let be such that belongs to the interior of the -accessible set. There exist , such that . We infer from Lemma 7 that there exists such that, for any partition of satisfying , there exist , such that . Now let us fix such a partition of which satisfies . In view of the above, we know that belongs to the convex hull of . On the other hand, since is convex, is convex and thus is a connected set. Since is continuous on , we deduce that is a connected set of , and thus is convex. We have proved that .
V-C Proof of Theorem 1 under strong -regularity
Let be a control such that . Let and let us fix some . Using the truncated dynamics introduced in Appendix -H, we have and (see Remark 21). Assume that is strongly -regular. By Definition 3, there exists a -tuple of elements of such that
(6) |
for every , where is the canonical basis of . We define the mapping by
for all . This mapping satisfies . Furthermore, since is of class (see Proposition 14), the mapping is also of class and we infer from (6) that . By the conic implicit function theorem [3, Theorem 1], there exists a continuous mapping , with , satisfying and for all .
By Lemma 9, there exists a threshold such that and , and thus
for any partition of satisfying , where is the averaging operator introduced in Appendix -G. For any partition of satisfying , we define the control
Using the linearity of the averaging operators, we obtain the piecewise constant control
which satisfies
for all partitions of satisfying . If necessary we take a smaller value of to have and small enough (by Lemma 9), and thus small enough as well, to get that:
-
(i)
(here we used in particular Lemma 8);
- (ii)
- (iii)
for all partitions of satisfying .
We are now in a position to conclude the proof. Let us fix a partition of satisfying and, for the ease of notations, let us denote simply by and recall that . Since is with values in from the above item (iii), we have . By the above items (i) and (ii) and by Lemma 4, we have and . We infer that and, since from the above item (i), we obtain from Remark 21 that and thus . The proof is complete.
V-D Proof of Theorem 1 under weak -regularity
Let be a control such that . Assume that is weakly -regular and, by contradiction, that Property (4) is not satisfied. Then there exists a sequence of partitions of such that as and such that is not -reachable in time from for all .
We first introduce several notations. Since is weakly -regular, considering the canonical basis of , we construct a package as in the proof of Proposition 5. Now take and and consider given in Lemma 4. As in Remark 17, there exists sufficiently small so that for all . In particular we have for all . Consider the mapping , defined by for all , which satisfies and as in the proof of Proposition 5.
We define the mapping by for all . It follows from the above arguments that and, since , the conic implicit function theorem [3, Theorem 1] provides the existence of a continuous mapping , with , such that and for all .
The mapping , defined by for all , is such that for all . When endowing the codomain with the -metric, the continuity of follows from the continuity of and from Remark 17. Finally note that for all .
In what follows we denote by a positive Lipschitz constant of restricted to endowed with the -metric (see Lemma 4). By contradiction, assume that, for all , there exists some such that
where is the averaging operator introduced in Appendix -G. By compactness of , up to a subsequence (that we do not relabel), the sequence converges to some . We infer from Lemma 8 that
for every , raising a contradiction when by continuity of and by Lemma 9. We conclude that there exists such that
(7) |
for every . Since , we deduce from (7) and from Lemma 8 that for all . Since , we infer from Proposition 13 that for all .
To conclude the proof of Theorem 1, we define by
for every . By Lemma 8 and thanks to the continuities of the mapping and of the restriction of on endowed with the -metric, is continuous. Furthermore, since and both belong to , we have
for every , where we have used (7). Therefore is a continuous mapping from with values in . By the Brouwer fixed-point theorem, has a fixed-point , and thus
Since , is -reachable in time from , raising a contradiction.
-E An example
We develop here an example inspired from [13, Section II], showing that the converse of the geometric Pontryagin maximum principle is not true in general and that, given a control , the condition that belongs to the interior of the -accessible set is not a sufficient condition for Property (4), even if is convex.
Take and . Take be a continuous function that is positive on the interval and vanishing on the interval . Take be arbitrarily fixed and be defined by for all . Note that is constant on the interval . We denote by the corresponding constant values. We set and the expression of by
for all .
Claim 1.
The point is an equilibrium of the control system on the interval , independently of the control.
Proof.
Since and for all , we have for all . ∎
Claim 2.
Let satisfying for a.e. . Then and . In particular .
Proof.
Since for all , it holds that for all . From the second coordinate, we obtain that for all . Since , we get from Claim 1 that for all . ∎
Claim 3.
Let such that for some . Then for a.e. .
Proof.
By Claim 1, . Since for all , we get that and thus for all . Derivating this equality leads to for a.e. . Since is positive on the interval , we get that for a.e. . ∎
Claim 4.
The end-point mapping is surjective.
Proof.
Let . Let us prove that there exists such that . If , from Claim 2, it is sufficient to take any control which satisfies for a.e. . In the rest of this proof, we focus on the case .
Consider a function such that the measure of is positive and such that the -norm of on is small enough to guarantee that any control which satisfies for a.e. is admissible, i.e., . This is possible by Claim 2, since is an open subset of . Take such a control (which is only determined on the interval at this step). By Claim 3, . Consider now a function which satisfies , and for all . We determine the control on as
where for a.e. . The control belongs to and along . Thus . ∎
Let us prove that the converse of the geometric Pontryagin maximum principle is not true in general. Take a control which satisfies for a.e. . By Claims 2 and 4, we have and belongs to the interior of the -accessible set. Consider the constant function defined by for all . One can easily check that is a nontrivial strong -extremal lift of and thus is strongly -singular by Proposition 6.
We now prove that, given a control , the condition that belongs to the interior of the -accessible set is not a sufficient condition for Property (4), even if is convex. Take for a.e. (which is not piecewise constant). Even if belongs to the interior of the -accessible set (Claim 4), we easily infer from Claim 3 that is not -reachable in time from for any partition of . Hence Property () is not satisfied, and neither is the stronger Property (4).
-F A general result on -approximation by piecewise constant functions
Proposition 12.
Let . Given any and any , there exists a threshold such that, for any partition of satisfying , there exists such that and .
Proof.
Let and . By the Lusin theorem [19], there exists a compact subset such that , where is the Lebesgue measure, and such that is continuous on . By uniform continuity of on , there exists such that for all , satisfying . Now, let be a partition of such that . We set
For every , we consider some such that and . We also consider some such that . We now define
for every . In particular we have and . Finally we get that
which concludes the proof. ∎
Example 10.
Take , and . Consider the oscillating function defined by for a.e. for all even and for a.e. for all odd . We have for all and all partitions of .
Corollary 1.
Let . Given any and any , there exists a threshold such that, for any partition of satisfying , there exists such that .
Proof.
Let and . We fix some and we define and
for a.e. and for every . In particular for every . It is clear that converges to as and that for a.e. . By the Lebesgue dominated convergence theorem, we get that as . Hence, there exists such that . By Proposition 12, there exists such that, for any partition of satisfying , there exists such that and thus . ∎
-G Averaging operators
For any partition of , we define the averaging operator by
(8) |
for every , every and every . The aim of this section is to establish several useful properties of the averaging operators.
Let be a partition of . The averaging operator is linear and projects any integrable function onto a piecewise constant function respecting the partition (by averaging its value on each sampling interval ). Furthermore we have
(9) |
for a.e. and all .
Lemma 8.
Let . For any partition of , we have for all .
Proof.
Let be a partition of and let . When , the inequality follows from (9). When , we get from the Hölder inequality that
for all , and thus . ∎
The next lemma is instrumental in order to approximate with a -norm (with any ) any control with piecewise constant controls.
Lemma 9.
Let . Given any , we have as .
Proof.
Let . Seeing (9) as a domination assumption and thanks to the Lebesgue dominated convergence theorem, we only need to prove that as for a.e. . For this purpose we set for every and let being a Lebesgue point such that is derivable at with . Given any , there exists such that
for every . Take a partition of such that . There exists such that . Then
which concludes the proof. ∎
Our objective now is to prove that, when is convex, the averaging operators project any integrable function with values in onto a piecewise constant function with values in .
Lemma 10.
Assume that is convex. If , then .
Proof.
Let and let us prove that where is defined by . We first give a simpler argument when is furthermore assumed to be closed. In that context, by the Hilbert projection theorem, we have
for a.e. , where is the projection of onto . Integrating the above inequality over yields and thus .
Now we remove the closedness assumption made on . Let us prove that by strong induction on the dimension of the nonempty convex set . If , the set is reduced to a singleton and the result is trivial. Now consider that and assume that the result is true at all steps from to . By contradiction assume that . By separation, there exists such that for all . We infer that the null integral has a nonpositive integrand. Thus this integrand is zero almost everywhere on . Therefore is with values in the convex set , where stands for the standard hyperplane defined by orthogonality with the nonzero vector . Since is a nonempty convex set of dimension strictly inferior than , thanks to our induction hypothesis we get that , which raises a contradiction. ∎
From Lemma 10 and applying a simple affine change of variable in (8), we obtain the next proposition.
Proposition 13.
Assume that is convex. If , then for any partition of .
-H Truncated end-point mapping and -differential
For every , we fix a mapping of class satisfying
Let . When replacing the dynamics in the control system (CS) by the truncated dynamics , defined by for all we obtain a new control system that we denote by (). The main difference is that, for any control (even unbounded), there exists a trajectory , starting at , such that for a.e. . In that case the trajectory is unique and will be denoted by . We now introduce, for any , the truncated end-point mapping defined by for all . Note that the next proposition, derived from standard techniques in ordinary differential equations theory, is true for any . The case is discussed in Remark 22.
Proposition 14.
Let and . The truncated end-point mapping is of class and its Fréchet differential is given by
(10) |
for all , , where is the unique solution to
Remark 21.
Let . For a given control , note that given in (1) admits a natural extension (still denoted by) . The nontruncated setting is related to the truncated one as follows:
-
(i)
Let and be such that and . Then and when considering the above extension of .
-
(ii)
Let . If there exists such that and , then and .
Remark 22.
Let . In the case , it can be proved that the truncated end-point mapping is Gateaux-differentiable and its Gateaux differential is given by (10). However it is not Fréchet-differentiable (and thus not of class ) in general, as shown in the next example.
Example 11.
Take , and for all . Consider the starting point and the constant control . In that context, with , it is clear that and that the Gateaux differential of the truncated end-point mapping at , given by the expression (10), is null. Now, taking the needle-like variation , as defined in (2), associated with the pair , we obtain that
Therefore is not Fréchet-differentiable at .
References
- [1] A. Agrachev and Y.L. Sachkov. Control theory from the geometric viewpoint, volume 87 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2004. Control Theory and Optimization, II.
- [2] A. A. Agrachev and M. Caponigro. Dynamics control by a time-varying feedback. J. Dyn. Control Syst., 16(2):149–162, 2010.
- [3] P. Antoine and H. Zouaki. Étude locale de l’ensemble des points critiques d’un problème d’optimisation paramétré. C. R. Acad. Sci. Paris Sér. I Math., 310(7):587–590, 1990.
- [4] W. Bian and J.R.L. Webb. Constrained open mapping theorems and applications. J. London Math. Soc. (2), 60(3):897–911, 1999.
- [5] B. Bonnard and M. Chyba. Singular trajectories and their role in control theory, volume 40 of Mathématiques & Applications (Berlin) [Mathematics & Applications]. Springer-Verlag, Berlin, 2003.
- [6] U. Boscain and B. Piccoli. Optimal syntheses for control systems on 2-D manifolds, volume 43 of Mathématiques & Applications (Berlin) [Mathematics & Applications]. Springer-Verlag, Berlin, 2004.
- [7] L. Bourdin and E. Trélat. Linear-quadratic optimal sampled-data control problems: convergence result and Riccati theory. Automatica J. IFAC, 79:273–281, 2017.
- [8] R.F. Brammer. Controllability in linear autonomous systems with positive controllers. SIAM J. Control, 10:339–353, 1972.
- [9] J.-M. Coron. Control and nonlinearity, volume 136 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2007.
- [10] A.T. Fuller. Relay control systems optimized for various performance criteria. IFAC Proceedings Volumes, 1(1):520 – 529, 1960. 1st International IFAC Congress on Automatic and Remote Control, Moscow, USSR, 1960.
- [11] R. V. Gamkrelidze. Principles of optimal control theory. Plenum Press, New York-London, revised edition, 1978. Translated from the Russian by Karol Malowski, Translation edited by and with a foreword by Leonard D. Berkovitz, Mathematical Concepts and Methods in Science and Engineering, Vol. 7.
- [12] K.A. Grasse. On the relation between small-time local controllability and normal self-reachability. Math. Control Signals Systems, 5(1):41–66, 1992.
- [13] K.A. Grasse. Reachability of interior states by piecewise constant controls. Forum Math., 7(5):607–628, 1995.
- [14] K.A. Grasse and H.J. Sussmann. Global controllability by nice controls. In Nonlinear controllability and optimal control, volume 133 of Monogr. Textbooks Pure Appl. Math., pages 33–79. Dekker, New York, 1990.
- [15] Kevin A. Grasse. Perturbations of nonlinear controllable systems. SIAM J. Control Optim., 19(2):203–220, 1981.
- [16] J. Klamka. Constrained controllability of nonlinear systems. J. Math. Anal. Appl., 201(2):365–374, 1996.
- [17] A.J. Krener and H. Schättler. The structure of small-time reachable sets in low dimensions. SIAM J. Control Optim., 27(1):120–147, 1989.
- [18] E.B. Lee and L. Markus. Foundations of optimal control theory. John Wiley & Sons, Inc., New York-London-Sydney, 1967.
- [19] N. Lusin. Sur les propriétés des fonctions mesurables. C. R. Acad. Sci. Paris, 154:1688–1690, 1912.
- [20] Dragan Nesić and Andrew R. Teel. Sampled-data control of nonlinear systems: an overview of recent results. In Perspectives in robust control (Newcastle, 2000), volume 268 of Lect. Notes Control Inf. Sci., pages 221–239. Springer, London, 2001.
- [21] L.S. Pontryagin, V.G. Boltyanskii, R.V. Gamkrelidze, and E.F. Mishchenko. 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, 1962.
- [22] E.D. Sontag. Mathematical control theory, volume 6 of Texts in Applied Mathematics. Springer-Verlag, New York, second edition, 1998. Deterministic finite-dimensional systems.
- [23] H.J. Sussmann. Reachability by means of nice controls. In Proceedings of the 26th IEEE Conference on Decision and Control, pages 1368–1373, 1987.
- [24] H.J. Sussmann and V. Jurdjevic. Controllability of nonlinear systems. J. Differential Equations, 12:95–116, 1972.
- [25] E. Trélat. Contrôle optimal. Mathématiques Concrètes. [Concrete Mathematics]. Vuibert, Paris, 2005. Théorie & applications. [Theory and applications].
Loïc Bourdin was born in 1986. He received his PhD degree in Applied Mathematics from the University of Pau (France) in 2013. Since 2014, he is associate professor at the University of Limoges (France) and since 2016, he is member of SMAI-MODE group. His mathematical interests are control theory, optimal control, shape optimization problems and nonsmooth analysis. |
Emmanuel Trélat was born in 1974. He is full professor at Sorbonne Université and director of Laboratoire Jacques-Louis Lions. He is the Editor in Chief of the journal ESAIM: Control Optim. Calc. Var., and is Associate Editor of several other journals. He has been awarded the SIAM Outstanding Paper Prize (2006), Maurice Audin Prize (2010), Felix Klein Prize (European Math. Society, 2012), Blaise Pascal Prize (french Academy of Science, 2014), Big Prize Madame Victor Noury (french Academy of Science, 2016). His research interests range over control theory in finite and infinite dimension, optimal control, stabilization, geometry, numerical analysis. |