Gianna Stefani and Pierluigi Zezza

17A Hamiltonian approach to sufficiency in optimal control with minimal regularity conditions: Part I

Gianna Stefani, Dipartimento di Matematica e Informatica, Via di Santa Marta 3 50139 FIRENZE, Italy, [email protected]

Pierluigi Zezza, Dipartimento di Scienze per lEconomia e lImpresa, Via delle Pandette 9 50137FIRENZE, Italy, [email protected]

Abstract: In this paper, we develop a Hamiltonian approach to sufficient conditions in optimal control problems. We extend the known conditions for C2 maximized Hamiltonians into two directions: on the one hand we explain the role of a super Hamiltonian (i.e., a Hamiltonian which is greater or equal to the maximized one), on the other we develop the theory under some minimal regularity assumptions. The results we present enclose many known results and they can be used to tackle new problems.

Keywords: Sufficient condition, strong local minima, Hamiltonian methods

AMS classification: 49K15, 49J30. See www.ams.org/msc

17.1Introduction

In this paper, we develop the first part of a larger project whose goal is to describe how a Hamiltonian approach can be a key instrument to prove sufficient conditions in optimal control problems (OCPs). Hamiltonian methods have been widely used in OCPs to state sufficient conditions ensuring the strong local optimality of a reference trajectory.

The main feature of this approach is that it allows us to compare the costs of different trajectories by lifting them to the cotangent bundle and hence independently of the control values. The seminal idea goes back to 1879 when K. Weierstrass discovered a method that enables one to establish a strong minimum property for solutions of EulerLagranges equations, that is, for stationary curves. We can summarize Weierstrasss method as a combination of a local convexity assumption on the integrand and of an embedding of the given curve in a suitable field of nonintersecting stationary curves. The extension to optimal control requires some efforts since one has to deal with the maximized Hamiltonian coming from the Pontryagin Maximum Principle (PMP), introduced in 1956 by Lev Pontryagin and which is an extension to OCPs of the EulerLagranges equation for the Calculus of Variations (CV); one of the main difficulties being that the maximized Hamiltonian in OCPs does not possess the same regularity properties as the one in the CV.

The goal of this project is to propose a unified Hamiltonian approach to sufficient optimality conditions for OCPs whose state evolves on a manifold M. These results include some known ones, emphasizing their common features, and will allow us to tackle new problems.

The leading ideas of the project are the following:

To compare the costs of neighboring admissible trajectories by lifting them to the cotangent bundle TM by using the symplectic properties of TM.

TodefineinTM a suitable Hamiltonian flow t emanating from a horizontal Lagrangian submanifold Λ. This flow is the one of the maximized Hamiltonian when this last is at least C2.

To estimate the variation of the cost of admissible trajectories by the variation of a function of their final points and, if it is the case, of their final times.

To obtain a suitable second-order approximation (2nd variation) in the form of a coordinate-free linear quadratic (LQ) problem and to require its coercivity.

To show that t (the derivative of t along the reference extremal) is, up to an isomorphism, the linear Hamiltonian flow associated to the LQ problem.

To use the coercivity of the 2nd variation to substitute the manifold described by the transversality conditions, Λ0, by a horizontal one Λ. This can be obtained by adding a penalty term, which reduces the problem to a problem with free initial point and whose 2nd variation is still coercive, see M. Hestenes [16]. This allows us to overcome a difficult point: in OCPs, for lack of controllability, itmay happen that the projection of the flow starting from Λ0 at t = 0 is not locally onto for a nontrivial time interval.

To deduce for a problem with free initial point and fixed final point that the projection on M oft emanating from Λ is locally invertible so that we can go back to the first issue and we can compare the costs of neighboring admissible trajectories by lifting them to the cotangent bundle.

To use again the coercivity of the 2nd variation to complete the proof for the general case.

The project goes back to some initial papers, [4] and [3], where we began the use of this kind of approach to OCPs where the maximized Hamiltonian is at least C2. In [4] the authors studied a Bolza problem in n on a fixed time interval and with general endpoint constraints. They assumed that the data are uniformly quasi- C2, see Definition 1 therein, where the quasi refers to the fact that the data are time-dependent; moreover, under the strengthened Legendre condition, assuming that in a neighborhood of the reference Pontryagin extremal the local maximum of the Hamiltonian coincides with the global maximum they proved that the maximized Hamiltonian is quasi-C2. This kind of regularity allowed them to look for second-order conditions, which guarantee that the abstract theory can be applied to obtain strong local optimality.

When the state evolves on manifolds, the first issue is to prove an invariant version of second-order conditions; this problem was addressed in [3] for a Mayer problem on a fixed time interval. This result can be obtained by pulling back the control problem to a neighborhood of the initial point ̂x0; in this way the second-order approximation leads naturally to derive a Hamiltonian formulation of the second variation as a LQ OCP on T̂x0M with control functions in L2. This formulation is coordinate-free and hence invariant.

In the book [2], dedicated to the geometric approach to control problems, the smooth case is investigated in a Hamiltonian setting for a Bolza problem with fixed endpoints; in [9] the authors present a numerical algorithm to compute the first point where the trajectory ceases to be locally optimal.

The assumption that the data are uniformly quasi-C2 can be weakened as shown in [5] by imposing conditions directly on the Hamiltonian flow (Theorem 2.3 therein); these conditions make sense even if the second variation does not exist and, when the second variation exists, they are implied by its coercivity.

Here we develop the first part of the project and we relax the regularity assumptions on the Hamiltonian in such a way that they guarantee the existence and the regularity of the flow so that this technique can be applied to a larger class of different OCPs. The Hamiltonian we consider can be either the maximized Hamiltonian or a suitable super-Hamiltonian, which is not necessarily equal to the maximized one but which will be needed in the study of problems where the control contains a singular arc.

The regularity assumptions we propose allow us to use the symplectic properties of Hamiltonian flows (Lemma 3.3) to state abstract sufficient optimality conditions for a Bolza problem (Theorem3.9) and we show their possible applications to a minimum time problem (Theorem 3.16).

In the final Section 17.4 we briefly summarize the OCPs where our assumptions are satisfied and this approach has been used.

In a forthcoming paper we will give a more detailed description of this approach and larger set of references; furthermore we will give some suggestions about possible extensions and further investigations. For different approaches here we quote only [8, 14, 19] and reference therein.

17.1.1The problem

We consider the following optimization problem:

minimize

subject to

The final time T can be fixed or variable, M is a n-dimensional connected paracompact smooth manifold M and the state endpoint constraints N0 and Nf are smooth connected embedded submanifolds of M. We assume that c0, cf , f 0, f are defined on open sets and that they are C.

We take smooth time-independent data because we are interested in the irregularities arising from the maximization of the Hamiltonian.

The couple (ξ, υ) is called admissible if it is a solution of (1.2a), which satisfies the constraints (1.2b)(1.2c)(1.2d); in this case we refer to ξ as an admissible trajectory and to υ as the associated control.

We consider as a candidate optimal solution a given reference admissible trajectory ̂ξ, which is identified by the triple (̂T, , ̂υ) and we study its strong local optimality according to the following Definition 1.1.

Definition 1.1 (Strong local minimizer).

The reference admissible trajectory is a strong local minimizer of the above considered problem (1.1), if there are neighborhoods U × M of the graph of ̂ξ, denoted by Γ, and ? ×M of (̂T, (̂T)) such that ̂ξ is a minimizer among all the admissible trajectories ξ satisfying

Γξ U, (T, ξ(T)) ?

independently of the associated control.

When the final time is fixed, this definition reduces to the usual definition of strong local minimum which uses the C0 topology to describe a neighborhood of the admissible trajectories; whereas in the case of variable end time our definition is local with respect to the graph of ̂ξ and also with respect to both the final time and the final point. We need the neighborhood ? × M of (̂T, (̂T)), since (T, ξ(T)) may belong to U being T ̂T.

This notion has been called time-state-local optimality in [23, 24], where a stronger version of optimality, called state-local optimality, is also used.

17.2Notations and preliminary results

We recall that for every connected paracompact smooth manifold M (see [18]), there is a Riemannian structure which induces a distance d M on M and the metric topology is the same as the original topology. This distance allows us to talk about Lipschitz property for maps on manifolds.

The next definition will be used to describe the main regularity assumption, which is a strengthening of the usual Carathéodory-type assumption.

Definition 2.1. Assume that M, N are finite-dimensional Riemannian smooth manifolds and that J is an open interval in . We will say that the map G: J × M N is a LipschitzCarathéodory map if it satisfies the following:

i. For almost every t J the map x G(t, x) is locally Lipschitz.

ii. For each x M the map t G(t, x) is bounded measurable.

iii. For any compact set K M there is an essentially bounded measurable function m such that

dN(G(t, x), G(t, y)) m(t) dM(x, y), x, y K .

When G is a time-dependent vector field and hence N = TM is the tangent space to M, this hypothesis assures existence, uniqueness, and Lipschitz continuity in (t0, z0) of the solutions of the differential equation

ζ̇(t) = G(t, ζ), ζ(t0) = z0, t0 J .

Moreover we will use it in an analogous but simpler way when N = to obtain thatthe Lipschitz continuity of

Let f be a vector field on the manifold M and φ: M be a smooth function.The action of f on φ (directional derivative or Lie derivative) evaluated at a point x is denoted with one of the two expressions

Lf φ(x) = ⟨dφ(x), f(x)⟩ .

For any C2-function φ such that dφ(x) = 0 the second derivative D2φ(x) is well defined as a bilinear symmetric form on TxM.

Finally we identify any bilinear form Q on a vector space V with a linear form Q: V V and we write

Q(υ, w) := ⟨Qυ, w⟩, Q[υ]2 := Q(υ, υ) .

17.2.1Symplectic notations

For a general introduction to symplectic geometry and its application to variational problems we refer to [7] while for more specific applications to optimal control we refer to [1].

Denote by π : TM M the cotangent bundle, it is well known that TM possesses a canonically defined symplectic structure, given by the symplectic form σ = ds(), where denotes an element of TM and s is the Liouville canonical one-form s() = π.

If qi are local coordinates on the base manifold M and pi the fiber coordinates, then in local coordinates we can write

where d denotes the exterior derivative and denotes the exterior product.

We recall that thanks to the symplectic structurewe can associate, with each timedependent locally defined Hamiltonian Ht : TM , a unique vector field on TM defined by the action

This vector field defines a corresponding Hamiltonian system

and we will denote its flow by (t, ) (t, ), that is, (t, ) is the solution at time t of system (2.1). Note that for any time-dependent object we will use the notation

M(t, ) = Mt() = M(t)

whenever one of the variables is to be considered as fixed.

17.2.2The Pontryagin maximum principle

We assume that the candidate optimal solution is a state extremal, ξ̂that is, it satisfiesthe PMP, see Assumption 2.2.

For simpler notations we set

To state the PMP we introduce three Hamiltonians; if U × U is the common domain ofboth f and f 0 in M × U, let Ω := π1(U) TM.

We define

(1) the (control-dependent) pre-Hamiltonian F : Ω × U as

F : (, u) , f(π(), u)⟩ p0 f 0(π(), u) ,

(2) the time-dependent reference Hamiltonian ̂F : [0, ̂T] × Ω as

̂F : (t, ) F(, ̂υ(t)) ,

(3) the maximized Hamiltonian as

Let us remark that the maximized Hamiltonian could take the value +.

All the above Hamiltonians depend on the parameter p0, which can take the values {0, 1} characterizing, respectively, the abnormal and normal case. We have assumed p0 0 to state the PMP in its standard form, some authors assume p0 0 in which case you have to substitute p0 by p0 in all the statements to obtain a Pontryagin minimum principle.

Assumption 2.2 (Pontryagin Maximum Principle). There is a nontrivial couple (p0,), where p0 {0, 1} and: [0, ̂T] TMis a lifting of ̂ξ to TM, (i.e., satisfying the Hamiltonian system

the transversality conditions

and the maximization property

Moreover ̂Ft((t)) is constant and it is zero when the end time T is variable.

is the Pontryagin extremal associated to the state extremal ξ̂For simpler notations we set

It is not difficult to see that the two functions p0 c0 and p0 cf act on N 0 and N f , respectively, but they can be extended to an open set in M in such a way that the transversality conditions (2.3) and (2.4) hold on the whole tangent space. Namely we denote by α, β : M two locally defined functions such that

The transversality conditions can also be expressed as

where means orthogonal with respect to the dual coupling.

In the normal case α, β are cost functions equivalent to the original ones while in the abnormal case they are extensions of the zero function on the constraints.

When p0 = 0 all the costs disappear from our conditions and indeed we will studya problem with a zero cost. Proving that is a strict minimizer will imply that it isisolated among the admissible trajectories.

17.3A Hamiltonian approach to optimality

In this section, we extend the known results for a C2 maximized Hamiltonian in two different directions. We first prove that the Cartan formmaintains its symplectic properties under weaker regularity assumptions on the Hamiltonian, Section 17.3.1. Afterward,motivated by the singular case where the maximized Hamiltonian does not possess the required regularity properties, we introduce a super-Hamiltonian which can be used to compare the costs of admissible trajectories by lifting them to the cotangent bundle, Section 17.3.2. Using these results we prove abstract sufficient conditions, Section 17.3.3.

17.3.1The Cartan form

Let α : M be a smooth function and let Λ be the graph of dα on a contractible open set. Λ is a Lagrangian submanifold of TM and it is known that

s|Λ = d(α π) .

Let ? be an open set in TM and J be an open interval, with J [0, ̂T], and ? Λ, we consider a time-dependent Hamiltonian

H: J × ? × TM .

Associated to this Hamiltonian we consider the Cartan form on J × ?

ω = s Hdt .

The following assumption describes a set of minimal regularity conditions which the Hamiltonian H has to verify to develop our approach. In specific examples, the properties of H will be different but they will imply this kind of regularity.

Assumption 3.1. Assume that

1.The flow

(t, ) J × Λ (t, ) TM

is well defined and Lipschitz continuous.

2.The function

is LipschitzCarathéodory (Definition 2.1).

Thanks to Assumption 3.1 the Cartan form defines a Lipschitz function θ : J × Λ , which will play a crucial role. Let

and hence

Let us now prove that, also in this case, the classical property of the Cartan form holds, that is, the form ω is exact on J × Λ.

We will consider differential one-forms with possibly L coefficients; these forms are called (Whitney) flat forms and their properties are presented in [15]. The general theory is fully presented in Whitneys monograph [28], see also [13]. We do not use the results in their full extent, what we really need is the chain rule for Lipschitz functions as it can be found in Lemma 4.6.3 in [17] and a suitable version of the Stokes theorem; an explicit proof is in [4] but the essential elements needed for the proof are contained in the cited books, [13] for the general case or [15] for the Lipschitz one.

The following lemma is the first step to prove that the form ω is exact on J × Λ.

Lemma 3.2. For every given t J the form s is exact on Λ and

Proof. We can equivalently prove that for every Lipschitz curve γ : [0, 1] Λ

Let us consider the set

Δ := {(τ, s): 0 τ t, 0 s 1}

and the map

ϕ: (τ, s) (τ, γ(s)) .

By the Stokes theorem, we have

moreover

by equating the two right-hand sides we obtain

We are now able to prove the main result of this section.

Lemma 3.3. The differential form ω is exact on J × Λ and

ω = d θ a.e. (t, ) J × Λ .

Proof. We can equivalently prove that for every Lipschitz curve

μ : t [0, 1] (t, γ(t)) J × Λ

we have

By definition, we have that

and by (3.2) and (3.3)

When the state projection of the Hamiltonian flow is invertible we obtain some important properties of the function θ.We recall that a homeomorphism between metric spaces is said to be bi-Lipschitz if it is Lipschitz with Lipschitz inverse.

Lemma 3.4. Suppose that, for a given t J, the function πt is bi-Lipschitz then on πt(Λ)

(i)d (θt (πt)1) = t (πt)1.

(ii)The function αt := θt (πt)1 is a C1 function.

(iii)Λt := t(Λ) is the graph of dαt.

Proof.

Since θt (πt)1 is Lipschitz then by the chain rule we can prove (i) by proving thattheir integrals coincide over any Lipschitz curve γ : [0, 1] πt(Λ).By (3.3) we have

which proves statement (i). (ii) follows immediately from (i). To prove (iii) it is sufficient to notice that

17.3.2The super-Hamiltonian and its properties

Throughout this section we assume that α satisfies (2.5) and that Λ is the graph of dα on a contractible neighborhood of ̂x0. Moreover we assume that H satisfies Assumption 3.1 and the following Assumption 3.5, this last motivating the name super-Hamiltonian. We underline that these assumptions concern jointly the super-Hamiltonian and the horizontal Lagrangian manifold Λ.

Assumption 3.5. The super-Hamiltonian H satisfies the following:

As a consequence the Pontryagin extremalis also a solution of the system

Remark 3.6. If the maximized Hamiltonian Fmax satisfies Assumption 3.1, then it satisfies also Assumption 3.5; for example, in the classical case when F max is C2 we can take H := Fmax for any Λ.

From Assumption 3.1 it follows that there exists an interval I := [0, T] with T > ̂T such that t0) is defined for t I and we have t(̂0) =(t) on [0, ̂T]. Moreover from the compactness of the time interval I, it follows that there is an open neighborhood ?0of ̂0 such that, without loss of generality, we can redefine Λ := Λ ̂?0, to obtain that

is defined on I × Λ and is defined on [0, ̂T] × Λ, where is ̂̂the flow of

To prove the main theorem we need a final crucial assumption concerning the invertibility of the flow projection of the Hamiltonian onto the state space.

Assumption 3.7. The function

idI × π: (t, ) I × Λ (t, πt ()) I × M

is invertible and bi-Lipschitz between I × Λ and an open set U of I × M containing thegraph of.

To verify this assumption one can use one of the available inverse function theorems for locally Lipschitz functions based on the invertibility of πt in [0, ̂T] × {̂0}. We refer, for example, to the one proposed by F. Clarke in [12].

Remark 3.8. For OCPs where the initial point is free the problem is always normal, moreover the initial Lagrangian manifold Λ0 is itself horizontal and we take it as Λ. In general, however, this is not the case and one has to define an appropriate α; in our approach we expect that the suitable α can be obtained from the second-order conditions as it is the case in the applications we describe in the final Section 17.4.

We can now state the main theorem, which compares the reference cost with the cost of an admissible neighboring trajectory.

Theorem 3.9. Under Assumption 3.13.53.7, let ξ : [0, T] M be an admissible trajectory whose graph is contained in U and let ρ(t) := (πt)1 ξ(t) then

Proof. Let μ := t ρ be the lift of ξ to TM, by Lemma 3.3 we obtain

If moreover we isolate, on the left-hand side, the terms that describe the costs of the two trajectories we obtain

where we have used the definitions of F and ̂F. Now, by the (3.4a) property of the super-Hamiltonian, by adding to both sides β(ξ(T)) βxf ), and recalling that α = p0 c0 on the initial manifold and β = p0 cf on the final manifold we obtain

17.3.3Abstract sufficient optimality conditions

We are now able to state an abstract theorem, which reduces the strong local optimality of an admissible reference trajectory ̂ξ to the local optimality of a suitable function of the right endpoint (̂T, ̂xf ) of the graph of.Let

Φ: (t, x) U β(x) + θ(t, (πt)1 (x)) = (β + αt)(x) ,

where αt = t(πt)1 was defined in part (ii) of Lemma 3.4. By means of this function we can rewrite the inequality (3.6) as

This relation will allow us to characterize the minima of the function p0J by studying the function Φ at the reference point (̂T, ̂xf ). Indeed we can now state a sufficient optimality condition.

Theorem 3.10. In the normal case, p0 = 1, under Assumptions 3.13.53.7, if there exists a neighborhood ? ofT, ̂xf ) in Nf such thatT, ̂xf ) is a minimizer of Φ restricted to ? thenT, ̂ξ , ̂υ) is a strong local minimizer.

Proof. It follows immediately from (3.7).

Remark 3.11. When the initial point is free we have that α = c0, if, moreover, the final point and time are fixed then the right-hand side of (3.7) is zero and we can prove that ̂ξ is a strong local minimizer by only verifying Assumptions 3.13.53.7 for the given Λ. In the other cases we have to find a suitable Λ and to prove the minimality property of Φ at (̂T, ̂xf ).

Remark 3.12. Concerning the abnormal case, p0 = 0, if one can prove that the minimum is strict then, as a by-product, it follows that ̂ξ is isolated among the admissible trajectories. To prove that ̂ξ is a strict strong local minimizer the first step is to require that (̂T, ̂xf ) is a strict local minimizer for Φ, in this case, if we have another minimizer (T, ξ, υ), we can conclude that T = ̂T and ξ(T) = ̂xf . From the proof of Theorem 3.9 wehave

By Assumption (3.4a) we obtain that

F(μ(t), υ(t)) Ht μ(t) = 0, a.e. t [0, ̂T] .

On the other hand, since μ = t (πt)1 ξ we have that

If

then one can conclude that μ = ̂λ since they both satisfy the same Hamiltonian equation with the same boundary conditions. Unfortunately Assumption (3.4a) is too mild to prove (3.8). We can strengthen it by assuming that

where ? is a neighborhood of the range of ; we note that this is true if Ht = Fmax. From this new assumption we deduce that

hence by (3.10)

where υ is the vertical derivative along the fiber. This is not the only way to obtain local uniqueness of the strong minimizer; for the case of singular control see for example [10, 24].

To state necessary and/ or sufficient condition for (̂T, ̂xf ) to be a local minimizer for Φ we compute its first and second derivatives. We note that, by Lemma 3.4, the function Φ is a C1 function of x at t fixed and Lemma 3.13 states that the point (̂T, ̂xf ) is a critical point for the function Φ without further assumptions on the data. On the other hand to obtain the existence of second derivatives we will require stronger regularity assumptions on the data which here we do not specify. We underline that since the first derivatives are zero, then the second ones are well defined as a quadratic function on × T̂xfM.

Lemma 3.13. Under Assumptions 3.13.53.7, we have that

(i)xΦ(t, x) = dβ(x) +t (πt)1 (x) = d(β + αt)(x), t I,

(ii)tΦ(t, x) = Ht t (πt)1 (x) = Ht (dαt)(x), a.e. t I.

Moreover

dΦT, ̂xf) = 0 .

Proof. (i) follows immediately from Lemma (3.4). To prove (ii) from (3.2) and Lemma (3.4) we have

Computing these derivatives at (̂T, ̂xf ), by the transversality condition (2.4), we obtain

By Assumption (3.4b) and by the PMP with variable final time we have

tΦT, ̂xf) = ĤTf) = ̂F̂Tf) = 0 .

When the final time is fixed we are not interested in this last derivative.

Lemma 3.14. Let Assumptions 3.13.53.7 hold true and assume moreover that Φ is C2 in a neighborhood ? ofT, ̂xf ), then

  1. xxΦT, ̂xf ) [δx]2 = D2(β + α̂T)(̂xf )[δx]2 = σ ((dα̂T)δx, d(β)δx),
  2. txΦT, ̂xf ) δx = LδxL̂Tα̂Txf) = σ f ), (dα̂T)δx),

= (tHT, ̂xf) + σ ((dα̂T)̂T ,f ))).

Notice that since α̂T = ̂T (πĤT)1 then (dα̂T) = ̂T (πĤT) 1 .

Proof.

1. It follows directly from Lemma 3.13, part (i).

2./3. From Lemma 3.13, part (ii), taking into account that and some known symplectic equalities, that can be easily obtained in coordinates, we get

and

17.3.4The minimum time problem

In this section, we apply Theorem 3.9 to the minimum time problem which can be obtained from the general one by setting

c0 = cf = 0, f 0 = 1 .

For the minimum time problem with fixed final endpoint, directly from Theorem3.9,we obtain a sufficient optimality condition which requires only mild regularity of the super-Hamiltonian near the final time.

Lemma 3.15. Let Assumptions 3.13.53.7 hold true and let the function

t Ht t (πt)1xf )

be Lipschitz in a neighborhood of ̂T in . If p0 = 1 (normal case) thenT, ) is a strong local minimum for the minimum time problem with fixed final endpoint.

This lemma is a consequence of the following theorem for the minimum time problem with variable endpoint where we require a similar regularity assumption as in Lemma 3.15 and a second geometric assumption which is verified when the final point is fixed. This theorem is motivated by the results in [24] and [10].

Theorem 3.16. Let Assumptions 3.13.53.7 hold true and assume moreover that

(i)The function (t, x) Ht t (πt)1 (x) is Lipschitz in a neighborhood ofT, ̂xf ) in × Nf .

(ii) (t, x) ?T, ̂xf ) , x Nf

if p0 = 1 (normal case) thenT, ̂ξ) is a strong local minimum for the minimum timeproblem.

Proof. Let k be the dimension of the submanifod Nf , we can take local coordinates centered at ̂xf such that Nf is diffeomorphic to the plane given by the last n k coordinates equal to zero so that we can consider the problem on ×n with ̂xf = 0 and ξ(T) k.

For the minimum time problem, in the normal case, equation (3.7) reads

By contradiction assume there exist admissible trajectories ξn : [0, tn] M, tn < ̂T, whose graph is contained in U such that

Consider the curve

which is such that γ(tn) = ξn(tn) and γT) = 0. By Lemma 3.3 and Assumption (ii) weget from equation (3.12)

Let L > 0 be the Lipschitz constant of assumption (i). Since (πĤT)1xf) = ̂0 by (3.4b) and Assumption (i) we can write

Dividing by ̂T tn > 0 we obtain

which yield a contradiction by (3.13).

17.4Final comments

This general unified Hamiltonian approach can be used in different situations, which were already addressed in some published papers, where second-order conditions were also developed and appropriate H and Λ defined to obtain sufficient conditions for strong local optimality, here we briefly summarize them.

Bangbang control

In this case, H := F max is continuous and piecewise C and Lipschitz. The case when there are a finite number of simple switches was studied in [6] for a Mayerproblem with variable endpoints on a fixed time interval and in [20] for a Bolza problem, while the corresponding minimum time problem was studied in [23]. The double switch case was addressed in [21] for a Mayer problem and in [22] for the minimum time problem. The numerical analysis with Maple of a case study is in [27].

Totally singular control

In [26], for a Mayer problem and a single-input system, the maximized Hamiltonian does not define a regular flow and hence the author first introduced the idea of a super-Hamiltonian needed to prove sufficiency. In [24] and [25] the result for the minimum time problem were obtained as a by-product of problems with controls containing bang and singular arcs. Furthermore in [11] the minimum time problem was studied in the multi-input case for a system where the controlled vector fields generate an involutive Lie algebra, while in [10] this last assumption is removed. We notice that the super-Hamiltonian, and hence its flow, is C in aneighborhood of the range of but it satisfies Assumption 3.5 only starting from a submanifold of TM containing Λ, as we require.

Bangsingular control

The last application considered is the case when the reference trajectory contains both singular and bang arcs. The obtained results concern the minimum time problem for a single-input control system. In [24] a bang-singular-bang trajectory in a problem with fixed endpoints was addressed, while in [25] a bang-singular trajectory with the initial point fixed and the final one constrained to the integral line of the controlled vector field was considered. In this case the flow of the super-Hamiltonian is sufficiently regular and satisfies Assumption 3.5 only starting from Λ.

Bibliography

[1]A. A. Agrachev and R. V. Gamkrelidze, The exponential representation of flows and the chronological calculus. Mat. Sb. (N.S.), 107(149)(4):467532, 639, 1978.

[2]A. 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.

[3]A. A. Agrachev, G. Stefani, and P. Zezza. An invariant second variation in optimal control. Internat. J. Control, 71(5):689715, 1998.

[4]A. A. Agrachev, G. Stefani, and P. Zezza. Strong minima in optimal control. Tr. Mat. Inst. Steklova, 220 (Optim. Upr., Differ. Uravn. i Gladk. Optim.):826, 1998.

[5]A. A. Agrachev, G. Stefani, and P. Zezza. A Hamiltonian approach to strong minima in optimalcontrol. In Differential geometry and control (Boulder, CO, 1997), volume 64 of Proc. Sympos.Pure Math., pages 1122. Amer. Math. Soc., Providence, RI, 1999.

[6]A. A. Agrachev, G. Stefani, and P. Zezza. Strong optimality for a bang-bang trajectory. SIAM J. Control Optim., 41(4):9911014 (electronic), 2002.

[7]V. I. Arnold, V. V. Kozlov, and A. I. Neishtadt. Mathematical aspects of classical and celestial mechanics, volume 3 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, third edition, 2006. [Dynamical systems. III], Translated from the Russian original by E. Khukhro.

[8]M. S. Aronna, J. F. Bonnans, A. V. Dmitruk, and P. A. Lotito. Quadratic order conditions for bangsingular extremals. Numer. Algebra Control Optim., 2(3):511546, 2012.

[9]B. Bonnard, J.-B. Caillau, and E. Trélat. Second order optimality conditions in the smooth case and applications in optimal control. ESAIM Control Optim. Calc. Var., 13(2):207236 (electronic), 2007.

[10]F. Chittaro and G. Stefani. Minimumtime strong optimality of a singular arc: the multiinput non involutive case. ESAIM: COCV, DOI: http://dx.doi.org/10.1051/cocv/2015026.

[11]F. C. Chittaro and G. Stefani. Singular extremals in multi-input time-optimal problems: a sufficient condition. Control Cybernet., 39(4):10291068, 2010.

[12]F. H. Clarke. On the inverse function theorem. Pacific J. Math., 64(1):97102, 1976.

[13]H. Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.

[14]U. Felgenhauer. On stability of bang-bang type controls. SIAM J. Control Optim., 41(6):18431867 (electronic), 2003.

[15]J. Heinonen. Lectures on Lipschitz analysis, volume 100 of Report. University of Jyväskylä Department of Mathematics and Statistics. University of Jyväskylä, Jyväskylä, 2005.

[16]M. R. Hestenes. Applications of the theory of quadratic forms in Hilbert space to the calculus of variations. Pacific J. Math., 1:525581, 1951.

[17]T. Iwaniec and G. Martin. Geometric function theory and non-linear analysis. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 2001.

[18]J. M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.

[19]H. Maurer and N. P. Osmolovskii. Second order sufficient conditions for time-optimal bangbang control. SIAM J. Control Optim., 42(6):22392263 (electronic), 2004.

[20]L. Poggiolini. On local state optimality of bang-bang extremal. Rend. Semin. Mat. Univ. Politec. Torino, 64(1):123, 2006.

[21]L. Poggiolini and M. Spadini. Strong local optimality for a bang-bang trajectory in a Mayer problem. SIAM J. Control Optim., 49(1):140161, 2011.

[22]L. Poggiolini and M. Spadini. Bangbang trajectories with a double switching time in the minimum time problem. ESAIM: COCV, DOI: http://dx.doi.org/10.1051/cocv/2015021.

[23]L. Poggiolini and G. Stefani. State-local optimality of a bang-bang trajectory: a Hamiltonian approach. Systems Control Lett., 53(34):269279, 2004.

[24]L. Poggiolini and G. Stefani. Bang-singular-bang extremals: sufficient optimality conditions. J. Dyn. Control Syst., 17(4):469514, 2011.

[25]L. Poggiolini and G. Stefani. A case study in strong optimality and structural stability of bangsingular extremals. In Geometric control theory and sub-Riemannian geometry, volume 5 ofSpringer INdAM Ser., pages 333350. Springer, Cham, 2014.

[26]G. Stefani. Strong optimality of singular trajectories. In Geometric control and nonsmooth analysis, volume 76 of Ser. Adv. Math. Appl. Sci., pages 300326. World Sci. Publ., Hackensack, NJ, 2008.

[27]G. Stefani and P. Zezza. Time optimality of a bang-bang trajectory with Maple. In Lagrangianand Hamiltonian methods for nonlinear control 2003, pages 135140. IFAC, Laxenburg, 2003.

[28]H. Whitney. Geometric integration theory. Princeton University Press, Princeton, N. J., 1957.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset