This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Robustness under control sampling of reachability in fixed time for nonlinear control systems

Loïc Bourdin and Emmanuel Trélat L. Bourdin is with University of Limoges, XLIM Research Institute, F-87000 Limoges, France, [email protected]. Trélat is with Sorbonne Université, Laboratoire Jacques-Louis Lions, F-75005 Paris, France, [email protected].
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 nn, mm\in\mathbb{N}^{*} and T>0T>0 be fixed. In this work we consider the general nonlinear control system

x˙(t)=f(x(t),u(t),t),a.e. t[0,T],\dot{x}(t)=f(x(t),u(t),t),\quad\text{a.e.\ }t\in[0,T], (CS)

where the dynamics f:n×m×[0,T]nf:\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T]\to\mathbb{R}^{n} is a continuous mapping, of class C1\mathrm{C}^{1} 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 (x,u)(x,u) is a solution to (CS) if xAC([0,T],n)x\in\mathrm{AC}([0,T],\mathbb{R}^{n}) is an absolutely continuous function (called state or trajectory) and uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) is an essentially bounded measurable function (called control) such that x˙(t)=f(x(t),u(t),t)\dot{x}(t)=f(x(t),u(t),t) for a.e. t[0,T]t\in[0,T]. Throughout the paper we fix a starting point x0nx^{0}\in\mathbb{R}^{n} and a nonempty subset U\mathrm{U} of m\mathbb{R}^{m} standing for the set of control constraints. We say that a target point x1nx^{1}\in\mathbb{R}^{n} is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0} if there exists a solution (x,u)(x,u) to (CS), with uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), such that x(0)=x0x(0)=x^{0} and x(T)=x1x(T)=x^{1}.

We now sample the control uu over the time interval [0,T][0,T]: given a partition 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} of [0,T][0,T], consisting of real numbers satisfying 0=t0<t1<<tN1<tN=T0=t_{0}<t_{1}<\ldots<t_{N-1}<t_{N}=T, for some NN\in\mathbb{N}^{*}, we consider the set PC𝕋([0,T],m)\mathrm{PC}^{\mathbb{T}}([0,T],\mathbb{R}^{m}) of all possible piecewise constant functions u:[0,T]mu:[0,T]\rightarrow\mathbb{R}^{m} satisfying u(t)=uiu(t)=u_{i}, for some uimu_{i}\in\mathbb{R}^{m}, for every t[ti,ti+1)t\in[t_{i},t_{i+1}) and every i{0,,N1}i\in\{0,\ldots,N-1\}. We denote by 𝕋=maxi=0,,N1|ti+1ti|\|\mathbb{T}\|=\max_{i=0,\ldots,N-1}|t_{i+1}-t_{i}| the norm of the partition. We say that a target point x1nx^{1}\in\mathbb{R}^{n} is PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0} if there exists a solution (x,u)(x,u) to (CS), with uPC𝕋([0,T],U)u\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}), such that x(0)=x0x(0)=x^{0} and x(T)=x1x(T)=x^{1}.

In this paper we investigate the following question: assuming that a target point x1nx^{1}\in\mathbb{R}^{n} is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0} and given a partition 𝕋\mathbb{T} of [0,T][0,T], is the point x1x^{1} also PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0}? In other words, how robust is reachability in fixed time under control sampling? Without any specific assumption, even for small values of 𝕋\|\mathbb{T}\|, in general x1x^{1} fails to be PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0}, as shown in the following example.

Example 1.

Take T=n=m=1T=n=m=1, U=\mathrm{U}=\mathbb{R} and f(x,u,t)=1+(ut)2f(x,u,t)=1+(u-t)^{2} for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. The target point x1=1x^{1}=1 is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from the starting point x0=0x^{0}=0 with the control u(t)=tu(t)=t for a.e. t[0,T]t\in[0,T]. However there is no other control steering the control system from x0x^{0} to x1x^{1} in time TT. Therefore, given any partition 𝕋\mathbb{T} of [0,T][0,T], even with a small value of 𝕋\|\mathbb{T}\|, the target point x1x^{1} is not PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0}.

Our main result is the following.

Theorem 1.

Assume that U\mathrm{U} is convex and let x1nx^{1}\in\mathbb{R}^{n} be a target point that is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0} with a control uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}). If uu is weakly U\mathrm{U}-regular, then there exists a threshold δ>0\delta>0 such that x1x^{1} is PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta.

The key concept of weakly U\mathrm{U}-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 U\mathrm{U} and the C1\mathrm{C}^{1} smoothness assumption made on ff 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 U=m\mathrm{U}=\mathbb{R}^{m}), 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 U\mathrm{U} is a convex subset of m\mathbb{R}^{m}), thanks to the notion of strongly U\mathrm{U}-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 U\mathrm{U} is a general subset of m\mathbb{R}^{m}), thanks to the notion of weakly U\mathrm{U}-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 uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) is said to be admissible when there exists xAC([0,T],n)x\in\mathrm{AC}([0,T],\mathbb{R}^{n}), starting at x(0)=x0x(0)=x^{0}, such that (x,u)(x,u) is a solution to (CS). In that case the trajectory xx is unique and will be denoted by xux_{u}. The set 𝒰\mathcal{U} of all admissible controls is an open subset of L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) and the end-point mapping E:𝒰n\mathrm{E}:\mathcal{U}\to\mathbb{R}^{n} is the C1\mathrm{C}^{1} mapping defined by E(u)=xu(T)\mathrm{E}(u)=x_{u}(T) for every u𝒰u\in\mathcal{U}. Therefore a target point x1nx^{1}\in\mathbb{R}^{n} is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0} if and only if x1x^{1} belongs to the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set given by E(𝒰L([0,T],U))\mathrm{E}(\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U})). Reachability in time TT from x0x^{0} is thus related to a surjectivity property of E\mathrm{E}.

II-A Without control constraint

All results in this section are classical (see, e.g., [1, 5, 9, 25]). When U=m\mathrm{U}=\mathbb{R}^{m}, i.e., when there is no control constraint, some conditions ensuring surjectivity of E\mathrm{E} are well known. For instance, when the control system (CS) is linear and autonomous, i.e., f(x,u,t)=Ax+Bu+g(t)f(x,u,t)=Ax+Bu+g(t) for all (x,u,t)n×m×[0,T](x,u,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T], where An×nA\in\mathbb{R}^{n\times n} and Bn×mB\in\mathbb{R}^{n\times m} are constant matrices and gC([0,T],n)g\in\mathrm{C}([0,T],\mathbb{R}^{n}) is a continuous function, we have 𝒰=L([0,T],m)\mathcal{U}=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), and E\mathrm{E} is surjective if and only if the pair (A,B)(A,B) satisfies the classical Kalman condition. For a general nonlinear control system (CS), global surjectivity of E\mathrm{E} 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 u𝒰u\in\mathcal{U} is said to be strongly regular if the Fréchet differential DE(u):L([0,T],m)n\mathrm{D}\mathrm{E}(u):\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} is surjective, i.e., Ran(DE(u))=n\mathrm{Ran}(\mathrm{D}\mathrm{E}(u))=\mathbb{R}^{n}. A control u𝒰u\in\mathcal{U} is said to be weakly singular if it is not strongly regular, i.e., Ran(DE(u))\mathrm{Ran}(\mathrm{D}\mathrm{E}(u)) is a proper subspace of n\mathbb{R}^{n}.

Proposition 1.

If a control u𝒰u\in\mathcal{U} is strongly regular, then there exist an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a mapping V:𝒱𝒰V:\mathcal{V}\to\mathcal{U} of class C1\mathrm{C}^{1} satisfying V(xu(T))=uV(x_{u}(T))=u and E(V(z))=z\mathrm{E}(V(z))=z for every z𝒱z\in\mathcal{V}. In particular, any point of 𝒱\mathcal{V} is Lm\mathrm{L}^{\infty}_{\mathbb{R}^{m}}-reachable in time TT from x0x^{0}, and thus xu(T)x_{u}(T) belongs to the interior of the Lm\mathrm{L}^{\infty}_{\mathbb{R}^{m}}-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 E\mathrm{E} given by

DE(u)v=wvu(T)\mathrm{D}\mathrm{E}(u)\cdot v=w^{u}_{v}(T) (1)

for every u𝒰u\in\mathcal{U} and every vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), where wvuAC([0,T],n)w^{u}_{v}\in\mathrm{AC}([0,T],\mathbb{R}^{n}) is the unique solution to

{w˙(t)=xf(xu(t),u(t),t)w(t)+uf(xu(t),u(t),t)v(t),a.e. t[0,T],w(0)=0n.\left\{\begin{array}[]{l}\dot{w}(t)=\nabla_{x}f(x_{u}(t),u(t),t)w(t)\\ \qquad\qquad\qquad+\nabla_{u}f(x_{u}(t),u(t),t)v(t),\quad\text{a.e.\ }t\in[0,T],\\ w(0)=0_{\mathbb{R}^{n}}.\end{array}\right.

The Hamiltonian associated to (CS) is the function H:n×m×n×[0,T]H:\mathbb{R}^{n}\times\mathbb{R}^{m}\times\mathbb{R}^{n}\times[0,T]\to\mathbb{R} defined by

H(x,u,p,t)=p,f(x,u,t)nH(x,u,p,t)=\langle p,f(x,u,t)\rangle_{\mathbb{R}^{n}}

for all (x,u,p,t)n×m×n×[0,T](x,u,p,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times\mathbb{R}^{n}\times[0,T], where ,n\langle\cdot,\cdot\rangle_{\mathbb{R}^{n}} is the Euclidean scalar product in n\mathbb{R}^{n}.

Definition 2 (weak extremal lift).

A weak extremal lift of a pair (xu,u)(x_{u},u), where u𝒰u\in\mathcal{U}, is a triple (xu,u,p)(x_{u},u,p) where pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) (called adjoint vector) is a solution to the (linear) adjoint equation

p˙(t)=xH(xu(t),u(t),p(t),t)\dot{p}(t)=-\nabla_{x}H(x_{u}(t),u(t),p(t),t) (AE)

for a.e. t[0,T]t\in[0,T], satisfying the null Hamiltonian gradient condition

uH(xu(t),u(t),p(t),t)=0m\nabla_{u}H(x_{u}(t),u(t),p(t),t)=0_{\mathbb{R}^{m}} (NHG)

for a.e. t[0,T]t\in[0,T]. The weak extremal lift (xu,u,p)(x_{u},u,p) is said to be nontrivial if pp is nontrivial.

Proposition 2.

A control u𝒰u\in\mathcal{U} is weakly singular if and only if the pair (xu,u)(x_{u},u) 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 (xu,u)(x_{u},u), for some u𝒰u\in\mathcal{U}, has no nontrivial weak extremal lift, then any point of an open neighborhood of xu(T)x_{u}(T) is Lm\mathrm{L}^{\infty}_{\mathbb{R}^{m}}-reachable in time TT from x0x^{0}. Note that the contrapositive statement corresponds to a weak version of the geometric Pontryagin maximum principle: if xu(T)x_{u}(T), for some u𝒰u\in\mathcal{U}, belongs to the boundary of the Lm\mathrm{L}^{\infty}_{\mathbb{R}^{m}}-accessible set, then the pair (xu,u)(x_{u},u) admits a nontrivial weak extremal lift.

II-B With convex control constraints

When U\mathrm{U} is a proper subset of m\mathbb{R}^{m}, 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 22 and 33, 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 DE(u)\mathrm{D}\mathrm{E}(u) is surjective for some u𝒰u\in\mathcal{U}. When there are control constraints, a powerful approach is to use constrained versions of the implicit function theorem (as in [3, 4, 16]). When U\mathrm{U} is convex, the required hypothesis for a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is a conic surjectivity assumption made on DE(u)\mathrm{D}\mathrm{E}(u) as follows.

Definition 3 (strongly U\mathrm{U}-regular control).

Assume that U\mathrm{U} is convex. A control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is said to be strongly U\mathrm{U}-regular if DE(u)(𝒯LU[u])=n\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u])=\mathbb{R}^{n}, where 𝒯LU[u]\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] is the (convex) tangent cone to L([0,T],U)\mathrm{L}^{\infty}([0,T],\mathrm{U}) at uu defined by

𝒯LU[u]=+(L([0,T],U)u).\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]=\mathbb{R}_{+}(\mathrm{L}^{\infty}([0,T],\mathrm{U})-u).

The control uu is said to be weakly U\mathrm{U}-singular when it is not strongly U\mathrm{U}-regular, i.e., DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) is a proper convex subcone of n\mathbb{R}^{n}.

Proposition 3.

Assume that U\mathrm{U} is convex. If a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is strongly U\mathrm{U}-regular, then there exist an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a continuous mapping V:𝒱𝒰L([0,T],U)V:\mathcal{V}\to\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) satisfying V(xu(T))=uV(x_{u}(T))=u and E(V(z))=z\mathrm{E}(V(z))=z for every z𝒱z\in\mathcal{V}. In particular, any point in 𝒱\mathcal{V} is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0}, and thus xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-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 U\mathrm{U}-singular controls.

Definition 4 (weak U\mathrm{U}-extremal lift).

Assume that U\mathrm{U} is convex. A weak U\mathrm{U}-extremal lift of a pair (xu,u)(x_{u},u), where u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), is a triple (xu,u,p)(x_{u},u,p) where pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) (called adjoint vector) is a solution to the adjoint equation (AE) satisfying the Hamiltonian gradient condition

uH(xu(t),u(t),p(t),t)𝒩U[u(t)]\nabla_{u}H(x_{u}(t),u(t),p(t),t)\in\mathcal{N}_{\mathrm{U}}[u(t)] (HG)

for a.e. t[0,T]t\in[0,T], where

𝒩U[u(t)]={ϑmωU,ϑ,ωu(t)m0}\mathcal{N}_{\mathrm{U}}[u(t)]=\{\vartheta\in\mathbb{R}^{m}\mid\forall\omega\in\mathrm{U},\;\langle\vartheta,\omega-u(t)\rangle_{\mathbb{R}^{m}}\leq 0\}

is the normal cone to U\mathrm{U} at u(t)u(t). The weak U\mathrm{U}-extremal lift (xu,u,p)(x_{u},u,p) is said to be nontrivial if pp is nontrivial.

Proposition 4.

Assume that U\mathrm{U} is convex. A control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is weakly U\mathrm{U}-singular if and only if the pair (xu,u)(x_{u},u) admits a nontrivial weak U\mathrm{U}-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 U\mathrm{U} is convex, if a pair (xu,u)(x_{u},u), for some u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), has no nontrivial weak U\mathrm{U}-extremal lift, then any point of an open neighborhood of xu(T)x_{u}(T) is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0}. The contrapositive statement corresponds to a weak version of the geometric Pontryagin maximum principle in the presence of convex control constraints: when U\mathrm{U} is convex, if the point xu(T)x_{u}(T), for some u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), belongs to the boundary of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set, then the pair (xu,u)(x_{u},u) admits a nontrivial weak U\mathrm{U}-extremal lift.

Remark 1.

When U\mathrm{U} is convex, it is clear that, if a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is strongly U\mathrm{U}-regular, then it is strongly regular. The converse is not true in general, as shown in the next example, but is true when uu takes its values in the interior of U\mathrm{U} (see Proposition 7 further), in particular when U=m\mathrm{U}=\mathbb{R}^{m}.

Example 2.

Take T=n=m=1T=n=m=1, U=[1,1]\mathrm{U}=[-1,1] and f(x,u,t)=uf(x,u,t)=u for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. From the Hamiltonian characterizations, the constant control u1u\equiv 1 is strongly regular and weakly U\mathrm{U}-singular.

Remark 2.

Note that the conclusions of Propositions 1 and 3 are distinct: the local right-inverse mapping VV is of class C1\mathrm{C}^{1} 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 C1\mathrm{C}^{1} 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 C1\mathrm{C}^{1} smoothness does not seem to be an easy issue.

II-C With general control constraints

When U\mathrm{U} is convex and for a given control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), the set DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) consists of all elements wvu(T)w^{u}_{v}(T) (called the weak U\mathrm{U}-variation vectors associated with uu) generated by conic L\mathrm{L}^{\infty}-perturbations u+αvu+\alpha v of the control uu, where v𝒯LU[u]v\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] and α0\alpha\geq 0, in the sense that

wvu(T)=limα0+E(u+αv)E(u)α=DE(u)v.w^{u}_{v}(T)=\lim\limits_{\alpha\to 0^{+}}\dfrac{\mathrm{E}(u+\alpha v)-\mathrm{E}(u)}{\alpha}=\mathrm{D}\mathrm{E}(u)\cdot v.

The set DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) can be seen as a first-order conic convex approximation of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set at xu(T)x_{u}(T).

In the general control constraints case where U\mathrm{U} is not assumed to be convex, we can use sophisticated L1\mathrm{L}^{1}-perturbations, well known in the literature as needle-like variations. Precisely a needle-like variation u(τ,ω)αL([0,T],U)u^{\alpha}_{(\tau,\omega)}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) of a given control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is defined by

u(τ,ω)α(t)={ωalong [τ,τ+α),u(t)elsewhere,u^{\alpha}_{(\tau,\omega)}(t)=\left\{\begin{array}[]{ll}\omega&\text{along }[\tau,\tau+\alpha),\\[3.0pt] u(t)&\text{elsewhere,}\end{array}\right. (2)

for a.e. t[0,T]t\in[0,T] and for α0\alpha\geq 0, with (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}, where (fu)\mathcal{L}(f_{u}) stands for the full-measure set of all Lebesgue points in [0,T)[0,T) of the essentially bounded measurable function fu=f(xu,u,)f_{u}=f(x_{u},u,\cdot). In that framework, it is well known that u(τ,ω)αu^{\alpha}_{(\tau,\omega)} belongs to 𝒰\mathcal{U} for sufficiently small α0\alpha\geq 0 and that

limα0+E(u(τ,ω)α)E(u)α=w(τ,ω)u(T),\lim\limits_{\alpha\to 0^{+}}\dfrac{\mathrm{E}(u^{\alpha}_{(\tau,\omega)})-\mathrm{E}(u)}{\alpha}=w^{u}_{(\tau,\omega)}(T), (3)

where w(τ,ω)uAC([τ,T],n)w^{u}_{(\tau,\omega)}\in\mathrm{AC}([\tau,T],\mathbb{R}^{n}) is the unique solution to

{w˙(t)=xf(xu(t),u(t),t)w(t),a.e. t[τ,T],w(τ)=f(xu(τ),ω,τ)f(xu(τ),u(τ),τ).\left\{\begin{array}[]{l}\dot{w}(t)=\nabla_{x}f(x_{u}(t),u(t),t)w(t),\quad\text{a.e.\ }t\in[\tau,T],\\ w(\tau)=f(x_{u}(\tau),\omega,\tau)-f(x_{u}(\tau),u(\tau),\tau).\end{array}\right.

The elements w(τ,ω)u(T)w^{u}_{(\tau,\omega)}(T), with (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}, are called the strong U\mathrm{U}-variation vectors associated with uu.

Definition 5 (U\mathrm{U}-Pontryagin cone).

The U\mathrm{U}-Pontryagin cone of a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), denoted by PontU[u]\mathrm{Pont}_{\mathrm{U}}[u], is the smallest convex cone containing all strong U\mathrm{U}-variation vectors associated with uu.333In the literature, usually the U\mathrm{U}-Pontryagin cone of a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is defined as the smallest closed convex cone containing all strong U\mathrm{U}-variation vectors associated with uu (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 uu (see Section IV-E) generate (only) the U\mathrm{U}-Pontryagin cone of uu as defined in Definition 5 (i.e., without closure).

A strong U\mathrm{U}-variation vector associated with a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) 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 U\mathrm{U}-Pontryagin cone, which consists of all conic convex combinations of strong U\mathrm{U}-variation vectors associated with uu, can be generated by using multiple needle-like variations (see Section IV-E). Hence the set PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] can be seen as a first-order conic convex approximation of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set at xu(T)x_{u}(T). Note that, when U\mathrm{U} is convex, it is larger than DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) (in the sense of Remark 3) which leads to the following weakened notion of U\mathrm{U}-regularity.

Definition 6 (weakly U\mathrm{U}-regular control).

A control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is said to be weakly U\mathrm{U}-regular if PontU[u]=n\mathrm{Pont}_{\mathrm{U}}[u]=\mathbb{R}^{n}. The control uu is said to be strongly U\mathrm{U}-singular when it is not weakly U\mathrm{U}-regular, i.e., PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] is a proper convex subcone of n\mathbb{R}^{n}.

Although the U\mathrm{U}-Pontryagin cone of a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) cannot be written as the range of a differential DE(u)\mathrm{D}\mathrm{E}(u) 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 E\mathrm{E} to a multiple needle-like variation (see the proof in Section IV-E).

Proposition 5.

If a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is weakly U\mathrm{U}-regular, then there exist an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a mapping V:𝒱𝒰L([0,T],U)V:\mathcal{V}\to\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), that is continuous when endowing the codomain with the L1\mathrm{L}^{1}-metric, satisfying V(xu(T))=uV(x_{u}(T))=u and E(V(z))=z\mathrm{E}(V(z))=z for all z𝒱z\in\mathcal{V}. In particular any point in 𝒱\mathcal{V} is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0}, thus xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set.

Like in Sections II-A and II-B, we next provide a Hamiltonian characterization of strongly U\mathrm{U}-singular controls (see Proposition 6 below, proved in Section IV-F).

Definition 7 (strong U\mathrm{U}-extremal lift).

strong U\mathrm{U}-extremal lift of a pair (xu,u)(x_{u},u), where u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), is a triple (xu,u,p)(x_{u},u,p) where pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) (called adjoint vector) is a solution to the adjoint equation (AE) satisfying the Hamiltonian maximization condition

u(t)argmaxωUH(xu(t),ω,p(t),t)u(t)\in\operatorname*{arg\,max}_{\omega\in\mathrm{U}}H(x_{u}(t),\omega,p(t),t) (HM)

for a.e. t[0,T]t\in[0,T]. The strong U\mathrm{U}-extremal lift (xu,u,p)(x_{u},u,p) is said to be nontrivial if pp is nontrivial.

Proposition 6.

A control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is strongly U\mathrm{U}-singular if and only if the pair (xu,u)(x_{u},u) admits a nontrivial strong U\mathrm{U}-extremal lift.

From Propositions 5 and 6, if a pair (xu,u)(x_{u},u), where u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), has no nontrivial strong U\mathrm{U}-extremal lift, then any point of an open neighborhood of xu(T)x_{u}(T) is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from x0x^{0}. The contrapositive statement coincides exactly with the well known geometric Pontryagin maximum principle: if xu(T)x_{u}(T), for some u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), belongs to the boundary of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set, then the pair (xu,u)(x_{u},u) admits a nontrivial strong U\mathrm{U}-extremal lift.

Remark 3.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). Since PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] is convex, we have PontU[u]=n\mathrm{Pont}_{\mathrm{U}}[u]=\mathbb{R}^{n} if and only if its closure satisfies Clos(PontU[u])=n\mathrm{Clos}(\mathrm{Pont}_{\mathrm{U}}[u])=\mathbb{R}^{n}. When U\mathrm{U} is convex, we have DE(u)(𝒯LU[u])Clos(PontU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u])\subset\mathrm{Clos}(\mathrm{Pont}_{\mathrm{U}}[u]) and thus, if uu is strongly U\mathrm{U}-regular, then it is weakly U\mathrm{U}-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 T=n=m=1T=n=m=1, U=[1,1]\mathrm{U}=[-1,1] and f(x,u,t)=u3f(x,u,t)=u^{3} for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. From the Hamiltonian characterizations, the constant control u0u\equiv 0 is weakly U\mathrm{U}-regular and weakly U\mathrm{U}-singular.

Example 4.

Take T=n=1T=n=1, m=2m=2, U=[1,1]2\mathrm{U}=[-1,1]^{2} and f(x,(u1,u2),t)=u1u2f(x,(u_{1},u_{2}),t)=u_{1}u_{2} for all (x,(u1,u2),t)×2×[0,T](x,(u_{1},u_{2}),t)\in\mathbb{R}\times\mathbb{R}^{2}\times[0,T]. From the Hamiltonian characterizations, the constant control u02u\equiv 0_{\mathbb{R}^{2}} is weakly U\mathrm{U}-regular and weakly U\mathrm{U}-singular.

Remark 4.

No relationship can be established between strong regularity and weak U\mathrm{U}-regularity in general. One can check that Example 2 provides a control that is strongly regular and strongly U\mathrm{U}-singular, and that Example 3 provides a control that is weakly singular and weakly U\mathrm{U}-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 u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is strongly U\mathrm{U}-singular on the interval [0,T][0,T] with the starting point x0x^{0}, then it is also strongly U\mathrm{U}-singular on any subinterval [τ0,τ1][0,T][\tau^{0},\tau^{1}]\subset[0,T] of nonempty interior with the starting point xu(τ0)x_{u}(\tau^{0}). When U\mathrm{U} is convex, the same assertion is true when replacing “strongly U\mathrm{U}-singular” with “weakly U\mathrm{U}-singular”.

Remark 6.

Note that the conclusions of Propositions 3 and 5 are distinct. In Proposition 3, when U\mathrm{U} is convex and the control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is strongly U\mathrm{U}-regular, the controls allowing to reach an open neighborhood of xu(T)x_{u}(T) can be chosen close to uu in L\mathrm{L}^{\infty}-topology. In Proposition 5, when the control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is (only) weakly U\mathrm{U}-regular, closedness is obtained in the weaker L1\mathrm{L}^{1}-topology (this is because needle-like variations are L1\mathrm{L}^{1}-perturbations). There are similar subtleties in Section III due to the fact that piecewise constant functions are dense in L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) when endowed with the L1\mathrm{L}^{1}-norm (but not with the natural L\mathrm{L}^{\infty}-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 U\mathrm{U} is convex, the normal cone to U\mathrm{U} at any interior point of U\mathrm{U} is reduced to {0m}\{0_{\mathbb{R}^{m}}\}.

Proposition 7.

Let u𝒰L([0,T],Int(U))u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{Int}(\mathrm{U})), where Int(U)\mathrm{Int}(\mathrm{U}) is the interior of U\mathrm{U}.

  1. (i)

    If uu is strongly regular, then uu is weakly U\mathrm{U}-regular. The converse is not true in general (see Example 3 and Remark 4).

  2. (ii)

    When U\mathrm{U} is convex, uu is strongly regular if and only if uu is strongly U\mathrm{U}-regular.555This fact is obvious when uu belongs to Int(L([0,T],U))\mathrm{Int}(\mathrm{L}^{\infty}([0,T],\mathrm{U})) since then 𝒯LU[u]=L([0,T],m)\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). However note that the inclusion Int(L([0,T],U))L([0,T],Int(U))\mathrm{Int}(\mathrm{L}^{\infty}([0,T],\mathrm{U}))\subset\mathrm{L}^{\infty}([0,T],\mathrm{Int}(\mathrm{U})) may be strict (for a counterexample, take T=m=1T=m=1, U=[0,1]\mathrm{U}=[0,1] and u(t)=tu(t)=t for a.e. t[0,T]t\in[0,T]).

Remark 7.

By Remark 5 and Proposition 7, if a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) takes its values in Int(U)\mathrm{Int}(\mathrm{U}) along a subinterval [τ0,τ1][0,T][\tau^{0},\tau^{1}]\subset[0,T] of nonempty interior on which it is moreover strongly regular with the starting point xu(τ0)x_{u}(\tau^{0}), then uu is weakly U\mathrm{U}-regular (and even strongly U\mathrm{U}-regular if U\mathrm{U} is convex) on [0,T][0,T] with the starting point x0x^{0}.

The control system (CS) is said to be control-affine when f(x,u,t)=g(x,t)+B(x,t)uf(x,u,t)=g(x,t)+B(x,t)u for all (x,u,t)n×m×[0,T](x,u,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T], where g:n×[0,T]ng:\mathbb{R}^{n}\times[0,T]\to\mathbb{R}^{n} and B:n×[0,T]n×mB:\mathbb{R}^{n}\times[0,T]\to\mathbb{R}^{n\times m} are continuous mappings, of class C1\mathrm{C}^{1} with respect to their first variable. In that context we have

H(x,ω,p,t)H(x,u,p,t)=uH(x,u,p,t),ωumH(x,\omega,p,t)-H(x,u,p,t)=\langle\nabla_{u}H(x,u,p,t),\omega-u\rangle_{\mathbb{R}^{m}}

for all (x,u,ω,p,t)n×m×m×n×[0,T](x,u,\omega,p,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times\mathbb{R}^{m}\times\mathbb{R}^{n}\times[0,T] and the next proposition follows straightforwardly.

Proposition 8.

Assume that the control system (CS) is control-affine and let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}).

  1. (i)

    If uu is strongly conv(U)\mathrm{conv}(\mathrm{U})-regular, where conv(U)\mathrm{conv}(\mathrm{U}) is the convex hull of U\mathrm{U}, then uu is weakly U\mathrm{U}-regular.

  2. (ii)

    If uu is weakly U\mathrm{U}-regular, then uu is strongly regular. The converse is not true in general (see Example 2 and Remark 4).

  3. (iii)

    When U\mathrm{U} is convex, uu is weakly U\mathrm{U}-regular if and only if uu is strongly U\mathrm{U}-regular.

As a particular case of control-affine system, the control system (CS) is said to be linear when f(x,u,t)=A(t)x+B(t)u+g(t)f(x,u,t)=A(t)x+B(t)u+g(t) for all (x,u,t)n×m×[0,T](x,u,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T], where AC([0,T],n×n)A\in\mathrm{C}([0,T],\mathbb{R}^{n\times n}), BC([0,T],n×m)B\in\mathrm{C}([0,T],\mathbb{R}^{n\times m}) and gC([0,T],n)g\in\mathrm{C}([0,T],\mathbb{R}^{n}) are continuous functions. In that context 𝒰=L([0,T],m)\mathcal{U}=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) and E\mathrm{E} 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 U=m\mathrm{U}=\mathbb{R}^{m} 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 uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}). Then xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set if and only if uu is weakly U\mathrm{U}-regular.

Remark 8.

Assume that the control system (CS) is linear and autonomous (i.e., A()=AA(\cdot)=A and B()=BB(\cdot)=B are constant). Since 𝒰=L([0,T],m)\mathcal{U}=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) and E\mathrm{E} is affine, a control uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) is strongly regular if and only if DE(u)\mathrm{D}\mathrm{E}(u) is surjective, if and only if E\mathrm{E} is surjective, if and only if the pair (A,B)(A,B) satisfies the Kalman condition. This characterization does not depend on (T,x0,u)(T,x^{0},u). Hence, under the Kalman condition, any control uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) is strongly regular on any subinterval [τ0,τ1][0,T][\tau^{0},\tau^{1}]\subset[0,T] of nonempty interior and from any starting point. Thus, under the Kalman condition and using Remark 7, if a control uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) takes its values in Int(U)\mathrm{Int}(\mathrm{U}) along a subinterval [τ0,τ1][0,T][\tau^{0},\tau^{1}]\subset[0,T] of nonempty interior, then uu is weakly U\mathrm{U}-regular (and even strongly U\mathrm{U}-regular if U\mathrm{U} is convex) on [0,T][0,T] from any starting point.

We now introduce a last notion which will be instrumental in order to relax the convexity assumption made on U\mathrm{U} in our main result (see Remark 13 and 14 further for details).

Definition 8 (parameterization of U\mathrm{U}).

We say that U\mathrm{U} is parameterizable by a nonempty subset U\mathrm{U}^{\prime} of m\mathbb{R}^{m^{\prime}}, with mm^{\prime}\in\mathbb{N}^{*}, if there exists a C1\mathrm{C}^{1} mapping φ:mm\varphi:\mathbb{R}^{m^{\prime}}\to\mathbb{R}^{m} satisfying φ(U)=U\varphi(\mathrm{U}^{\prime})=\mathrm{U} and, for every uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), there exists uL([0,T],U)u^{\prime}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}^{\prime}) such that u=φuu=\varphi\circ u^{\prime}.

Example 5.

Using a standard measurable selection theorem, we see that the two-dimensional unit circle U={(u1,u2)2u12+u22=1}\mathrm{U}=\{(u_{1},u_{2})\in\mathbb{R}^{2}\mid u_{1}^{2}+u_{2}^{2}=1\} is parameterizable by the interval [0,2π][0,2\pi].

In the context of Definition 8, the control system (CS) has the same trajectories as the control system (CS’) given by

x˙(t)=f(x(t),u(t),t),a.e. t[0,T],\dot{x}^{\prime}(t)=f^{\prime}(x^{\prime}(t),u^{\prime}(t),t),\quad\text{a.e.\ }t\in[0,T], (CS’)

starting at the same initial point x0x^{0}, where the dynamics f:n×m×[0,T]nf^{\prime}:\mathbb{R}^{n}\times\mathbb{R}^{m^{\prime}}\times[0,T]\to\mathbb{R}^{n} is defined by f(x,u,t)=f(x,φ(u),t)f^{\prime}(x^{\prime},u^{\prime},t)=f(x^{\prime},\varphi(u^{\prime}),t) for all (x,u,t)n×m×[0,T](x^{\prime},u^{\prime},t)\in\mathbb{R}^{n}\times\mathbb{R}^{m^{\prime}}\times[0,T] and where U\mathrm{U}^{\prime} is the control constraint set. Precisely, for a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), any control uL([0,T],U)u^{\prime}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}^{\prime}) satisfying u=φuu=\varphi\circ u^{\prime} belongs to the set 𝒰\mathcal{U}^{\prime} of all admissible controls for (CS’), and xu=xux^{\prime}_{u^{\prime}}=x_{u}. Furthermore, by the Hamiltonian characterization, if uu is weakly U\mathrm{U}-regular for (CS), then uu^{\prime} is weakly U\mathrm{U}^{\prime}-regular for (CS’). We say that weak U\mathrm{U}-regularity is preserved by parameterization. However, when U\mathrm{U} and U\mathrm{U}^{\prime} are convex, strong U\mathrm{U}-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 u0u\equiv 0 is strongly U\mathrm{U}-regular. Considering the parameterization of U\mathrm{U} by itself, with the C1\mathrm{C}^{1} mapping φ:\varphi:\mathbb{R}\to\mathbb{R} defined by φ(u)=u3\varphi(u^{\prime})=u^{\prime 3} for every uu^{\prime}\in\mathbb{R}, we recover the control system considered in Example 3 in which the constant control u0u^{\prime}\equiv 0, which satisfies u=φuu=\varphi\circ u^{\prime}, is weakly U\mathrm{U}^{\prime}-singular.

III Robustness under control sampling of reachability in fixed time

When dealing with controls uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), 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 t[0,T]t\in[0,T]. Otherwise, when dealing with piecewise constant controls uPC𝕋([0,T],U)u\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}), for a given partition 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} of [0,T][0,T], 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 ti𝕋t_{i}\in\mathbb{T} and remains frozen on each sampling interval [ti,ti+1)[t_{i},t_{i+1}).

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 u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), we introduce the properties (4) and (Pu\mathrm{P}^{\prime}_{u}) defined by

δ>0,𝕋𝒫,𝕋δ,xu(T)E(𝒰PC𝕋([0,T],U))\exists\delta>0,\quad\forall\mathbb{T}\in\mathcal{P},\;\|\mathbb{T}\|\leq\delta,\\ x_{u}(T)\in\mathrm{E}(\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})) (4)

and

𝕋𝒫,xu(T)E(𝒰PC𝕋([0,T],U)),\exists\mathbb{T}\in\mathcal{P},\quad x_{u}(T)\in\mathrm{E}(\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})), (Pu\mathrm{P}^{\prime}_{u})

where 𝒫\mathcal{P} is the set of all partitions of [0,T][0,T] and where E(𝒰PC𝕋([0,T],U))\mathrm{E}(\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})) is the PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-accessible set. Example 1 shows that, for a given control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), Property (Pu\mathrm{P}^{\prime}_{u}) is not satisfied in general. Most of the literature focuses on establishing sufficient conditions for properties related to Property (Pu\mathrm{P}^{\prime}_{u}) (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 δ>0\delta>0 in Property (4) (which is not considered in Property (Pu\mathrm{P}^{\prime}_{u})) 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 δ>0\delta>0 implies that PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachability in time TT from x0x^{0} of the final point xu(T)x_{u}(T) is sensitive to small perturbations of the partition 𝕋\mathbb{T} of [0,T][0,T], in the sense of the next proposition (proved in Section V-A and illustrated in Remark 9 further).

Proposition 10.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). If Property (4) is not satisfied, then, for any partition 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} of [0,T][0,T] and any ε>0\varepsilon>0, there exists a partition 𝕋ε={tiε}i=0,,N\mathbb{T}^{\varepsilon}=\{t^{\varepsilon}_{i}\}_{i=0,\ldots,N} of [0,T][0,T] such that |tiεti|<ε|t^{\varepsilon}_{i}-t_{i}|<\varepsilon for all i{1,,N1}i\in\{1,\ldots,N-1\} and such that xu(T)x_{u}(T) is not PCU𝕋ε\mathrm{PC}^{\mathbb{T}^{\varepsilon}}_{\mathrm{U}}-reachable in time TT from x0x^{0}.

This section is organized as follows. In Section III-A we first investigate the condition that xu(T)x_{u}(T), for some u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set. Our main result (Theorem 1), which is valid under the stronger condition that uu is weakly U\mathrm{U}-regular, is discussed in Section III-B.

III-A Final point in the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). Here we focus on the condition that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set. The next example, based on a commensurability rigidity, shows that it is not a sufficient condition for Property (4).

Example 7.

Take T=4T=4, n=m=1n=m=1, U={0,1}\mathrm{U}=\{0,1\} and f(x,u,t)=uf(x,u,t)=u for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. The target point x1=πx^{1}=\pi is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from the starting point x0=0x^{0}=0 with the control u(t)=1u(t)=1 for a.e. t[0,π]t\in[0,\pi] and u(t)=0u(t)=0 for a.e. t[π,4]t\in[\pi,4]. By the Hamiltonian characterization, the control uu is weakly U\mathrm{U}-regular and thus x1=xu(T)x^{1}=x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set (see Proposition 5). However, for any given partition 𝕋\mathbb{T} of [0,T][0,T], x1x^{1} belongs to the PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-accessible set if and only if there exists a subfamily of sampling intervals associated with 𝕋\mathbb{T} whose sum of lengths is equal to π\pi. As a consequence, for any partition 𝕋\mathbb{T} of [0,T][0,T] containing only rational sampling times (with norm 𝕋\|\mathbb{T}\| arbitrarily small), x1x^{1} is not PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0}. We conclude that Property (4) is not satisfied (while Property (Pu\mathrm{P}^{\prime}_{u}) is).

In Example 7, the set U\mathrm{U} is not convex. However note that another counterexample, in which U\mathrm{U} is convex, is provided in Appendix -E.

Remark 9.

Example 7 illustrates Proposition 10 in the sense that, given any partition 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} of [0,T][0,T] (even such that the target point x1x^{1} is PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0}) and given any ε>0\varepsilon>0, there always exists a partition 𝕋ε={tiε}i=0,,N\mathbb{T}^{\varepsilon}=\{t^{\varepsilon}_{i}\}_{i=0,\ldots,N} of [0,T][0,T] containing only rational sampling times such that |tiεti|<ε|t^{\varepsilon}_{i}-t_{i}|<\varepsilon for all i{1,,N1}i\in\{1,\ldots,N-1\}, and thus such that x1x^{1} is not PCU𝕋ε\mathrm{PC}^{\mathbb{T}^{\varepsilon}}_{\mathrm{U}}-reachable in time TT from x0x^{0}. We provide in the following example a similar illustration of Proposition 10 with U\mathrm{U} convex.

Example 8.

Take T=4T=4, n=2n=2, m=1m=1, U=[0,1]\mathrm{U}=[0,1] and f((x1,x2),u,t)=(u,u2)f((x_{1},x_{2}),u,t)=(u,u^{2}) for all ((x1,x2),u,t)2××[0,T]((x_{1},x_{2}),u,t)\in\mathbb{R}^{2}\times\mathbb{R}\times[0,T]. Consider the starting point x0=02x^{0}=0_{\mathbb{R}^{2}}. The point xu(T)x_{u}(T) belongs to the segment {(x1,x2)20x1=x24}\{(x_{1},x_{2})\in\mathbb{R}^{2}\mid 0\leq x_{1}=x_{2}\leq 4\} if and only if the corresponding control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) takes its values in {0,1}\{0,1\}. As a consequence, by considering the target point x1=(π,π)x^{1}=(\pi,\pi), we find the same conclusions as in Example 7.

In the one-dimensional case n=1n=1, 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 n=1n=1, that U\mathrm{U} is convex777Actually assuming that U\mathrm{U} is connected is sufficient. and that 𝒰=L([0,T],m)\mathcal{U}=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). If xu(T)x_{u}(T), for some uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set, then Property (4) is satisfied.

III-B Comments on Theorem 1

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). This section focuses on the condition that uu is weakly U\mathrm{U}-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 U\mathrm{U} is convex, it is a sufficient condition for Property (4).

Example 9.

Take T=18T=18, n=2n=2, m=1m=1, U=[1,1]\mathrm{U}=[-1,1] and f((x1,x2),u,t)=(x2,u)f((x_{1},x_{2}),u,t)=(x_{2},u) for all ((x1,x2),u,t)2××[0,T]((x_{1},x_{2}),u,t)\in\mathbb{R}^{2}\times\mathbb{R}\times[0,T]. The target point x1=(0,0)x^{1}=(0,0) is LU\mathrm{L}^{\infty}_{\mathrm{U}}-reachable in time TT from the starting point x0=(78,0)x^{0}=(78,0) with the control u(t)=1u(t)=-1 for a.e. t[0,6]t\in[0,6], u(t)=t93u(t)=\frac{t-9}{3} for a.e. t[6,12]t\in[6,12] and u(t)=1u(t)=1 for a.e. t[12,18]t\in[12,18]. The control uu is weakly U\mathrm{U}-regular by Remark 8. Therefore, by Theorem 1, there exists δ>0\delta>0 such that x1x^{1} is PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta.

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 uu is strongly U\mathrm{U}-regular. This proof uses results of Section II-B (in particular, conic L\mathrm{L}^{\infty}-perturbations of uu) 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 uu is assumed to be (only) weakly U\mathrm{U}-regular. This proof uses results of Section II-C (in particular, needle-like variations of uu) 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 U\mathrm{U}-regularity assumption, relies on the conic implicit function theorem [3, Theorem 1]. However this theorem must be used in the Banach space Ls([0,T],m)\mathrm{L}^{s}([0,T],\mathbb{R}^{m}), for some 1<s<+1<s<+\infty, and not in L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). This is because it is not true that any function in L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) can be approximated in L\mathrm{L}^{\infty}-norm by piecewise constant functions, while it can be in Ls\mathrm{L}^{s}-norm with any 1s<+1\leq s<+\infty (see Appendix -F). This leads us to extend the end-point mapping to Ls([0,T],m)\mathrm{L}^{s}([0,T],\mathbb{R}^{m}) which makes no sense a priori because the control system (CS) is nonlinear.888For example, take n=m=1n=m=1, s=2s=2 and f(x,u,t)=u4f(x,u,t)=u^{4} for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. Then considering L2\mathrm{L}^{2}-controls makes no sense. To overcome this difficulty, we introduce in Appendix -H a truncated version of the dynamics ff, vanishing outside of a sufficiently large compact subset of n×m\mathbb{R}^{n}\times\mathbb{R}^{m}. Then the corresponding truncated end-point mapping is well defined on Ls([0,T],m)\mathrm{L}^{s}([0,T],\mathbb{R}^{m}), but is not Fréchet-differentiable when s=1s=1. However it is of class C1\mathrm{C}^{1} when 1<s<+1<s<+\infty and the surjectivity of the differential of the truncated end-point mapping in Ls\mathrm{L}^{s}-norm can be related to the surjectivity of the differential in L\mathrm{L}^{\infty}-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 U\mathrm{U}-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 U\mathrm{U} onto a piecewise constant control with values in U\mathrm{U}.

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.

  1. (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.

  2. (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.

  3. (iii)

    In [12, Remark 3.5] it is noted that if a point of the state space is normally reachable in time less than TT, then it belongs to the interior of the reachable set with piecewise-constant controls in time less than TT.

  4. (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 TT 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 TT^{\prime} (and indeed, inspection of the proof shows that, in general, TTT^{\prime}\neq T). Second, our main result (Theorem 1) states the existence of a threshold δ>0\delta>0 for which reachability (exactly at time TT) of a target point with a piecewise constant control is guaranteed for any partition 𝕋\mathbb{T} satisfying 𝕋δ\|\mathbb{T}\|\leq\delta. 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 \subset is not a total order over 𝒫\mathcal{P}, it may occur that 𝕋2𝕋1\|\mathbb{T}_{2}\|\leq\|\mathbb{T}_{1}\| while 𝕋1𝕋2\mathbb{T}_{1}\not\subset\mathbb{T}_{2}. In the above references, it is not guaranteed that reachability of a target point with a 𝕋1\mathbb{T}_{1}-piecewise constant control implies reachability with a 𝕋2\mathbb{T}_{2}-piecewise constant control. With the conclusion of Theorem 1, when 𝕋1δ\|\mathbb{T}_{1}\|\leq\delta, it is guaranteed.

Remark 13.

The convexity assumption made on U\mathrm{U} in Theorem 1 can be relaxed. Indeed let us prove that Theorem 1 is still true when U\mathrm{U} is assumed to be (only) convex by parameterization, i.e., when U\mathrm{U} is parameterizable (see Definition 8) by a nonempty convex subset U\mathrm{U}^{\prime} of m\mathbb{R}^{m^{\prime}} for some mm^{\prime}\in\mathbb{N}^{*} (see examples in Remark 14). In that context, for a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) that is weakly U\mathrm{U}-regular, there exists u𝒰L([0,T],U)u^{\prime}\in\mathcal{U}^{\prime}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}^{\prime}) such that u=φuu=\varphi\circ u^{\prime} and uu^{\prime} is weakly U\mathrm{U}^{\prime}-regular for the control system (CS’). Since U\mathrm{U}^{\prime} is convex, there exists by Theorem 1 a threshold δ>0\delta>0 such that xu(T)=xv𝕋(T)x^{\prime}_{u^{\prime}}(T)=x^{\prime}_{v^{\prime}_{\mathbb{T}}}(T), for some v𝕋𝒰PC𝕋([0,T],U)v^{\prime}_{\mathbb{T}}\in\mathcal{U}^{\prime}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}^{\prime}), for all partitions 𝕋\mathbb{T} satisfying 𝕋δ\|\mathbb{T}\|\leq\delta. Introducing v𝕋=φv𝕋𝒰PC𝕋([0,T],U)v_{\mathbb{T}}=\varphi\circ v^{\prime}_{\mathbb{T}}\in\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}), we obtain that xu(T)=xu(T)=xv𝕋(T)=xv𝕋(T)x_{u}(T)=x^{\prime}_{u^{\prime}}(T)=x^{\prime}_{v^{\prime}_{\mathbb{T}}}(T)=x_{v_{\mathbb{T}}}(T) for all partitions 𝕋\mathbb{T} satisfying 𝕋δ\|\mathbb{T}\|\leq\delta.

Remark 14.

If U\mathrm{U} is convex by parameterization (see Remark 13), then U\mathrm{U} must be connected. Actually a quite large class of connected sets are convex by parameterization. For example, in the two-dimensioncal case m=2m=2, the unit circle U={(u1,u2)2u12+u22=1}\mathrm{U}=\{(u_{1},u_{2})\in\mathbb{R}^{2}\mid u_{1}^{2}+u_{2}^{2}=1\}, the donut-shaped set U={(u1,u2)21u12+u224}\mathrm{U}=\{(u_{1},u_{2})\in\mathbb{R}^{2}\mid 1\leq u_{1}^{2}+u_{2}^{2}\leq 4\} or the cross-shaped set U=([1,1]×{0})({0}×[1,1])\mathrm{U}=([-1,1]\times\{0\})\cup(\{0\}\times[-1,1]) 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 U\mathrm{U} is strongly nonconnected, i.e., when it can be written as U=U1U2\mathrm{U}=\mathrm{U}_{1}\cup\mathrm{U}_{2}, where U1\mathrm{U}_{1} and U2\mathrm{U}_{2} are nonempty, and there exists a C1\mathrm{C}^{1} mapping Θ:m\Theta:\mathbb{R}^{m}\to\mathbb{R} taking the value 0 on U1\mathrm{U}_{1} and the value 11 on U2\mathrm{U}_{2}.999For example, when m=1m=1, the set (,0][1,+)(-\infty,0]\cup[1,+\infty) is strongly nonconnected, while the set (,0)(0,+)(-\infty,0)\cup(0,+\infty) is nonconnected (but not strongly). An open question is to extend Theorem 1 to sets U\mathrm{U} that are neither convex by parameterization, nor strongly nonconnected. We emphasize that our proof of Theorem 1, when U\mathrm{U} is convex, uses the averaging operators introduced in Appendix -G, which project any control with values in U\mathrm{U} onto a piecewise constant control with values in U\mathrm{U} (see Proposition 13). When U\mathrm{U} 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 ff is of class C1\mathrm{C}^{1} with respect to uu. Actually this assumption is required (only) when uf\nabla_{u}f has to be considered (such as in Sections II-A and II-B where we use conic L\mathrm{L}^{\infty}-perturbations). When using needle-like variations (which are L1\mathrm{L}^{1}-perturbations) such as in Section II-C, it is only required that ff is of class C1\mathrm{C}^{1} with respect to xx and is Lipschitz continuous with respect to (x,u)(x,u) on any compact subset of n×m×[0,T]\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T]. 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 φ\varphi (resp., Θ\Theta) that is (only) Lipschitz continuous on any compact subset of m\mathbb{R}^{m^{\prime}} (resp., of m\mathbb{R}^{m}).

Remark 16.

As far as we know, the U\mathrm{U}-Pontryagin cone of a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) cannot be written as the range of a differential DE(u)\mathrm{D}\mathrm{E}(u) taken in an appropriate sense. Indeed, we explain in Section IV-E how PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] can be generated using multiple needle-like variations which are Ls\mathrm{L}^{s}-perturbations for any 1s<+1\leq s<+\infty. Nevertheless, even using truncated dynamics in order to work in Ls([0,T],m)\mathrm{L}^{s}([0,T],\mathbb{R}^{m}) for some 1s<+1\leq s<+\infty, we explain in Appendix -H that the truncated end-point mapping is not Fréchet-differentiable when s=1s=1 and, when 1<s<+1<s<+\infty, the Fréchet differential of the truncated end-point mapping generates (only) weak U\mathrm{U}-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 PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] 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 ff 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 (𝒵,d𝒵)(\mathcal{Z},\mathrm{d}_{\mathcal{Z}}) is a metric set, we denote by B𝒵(z,ρ)\mathrm{B}_{\mathcal{Z}}(z,\rho) (resp. B¯𝒵(z,ρ)\overline{\mathrm{B}}_{\mathcal{Z}}(z,\rho)) the open ball (resp. closed ball) centered at some z𝒵z\in\mathcal{Z} of some radius ρ0\rho\geq 0.

IV-A Proof of Proposition 1

Let u𝒰u\in\mathcal{U} be strongly regular. By Definition 1 there exists a nn-tuple v¯={vj}j=1,,n\overline{v}=\{v_{j}\}_{j=1,\ldots,n} of elements of L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) such that DE(u)vj=ej\mathrm{D}\mathrm{E}(u)\cdot v_{j}=e_{j} for all j{1,,n}j\in\{1,\ldots,n\}, where {ej}j=1,,n\{e_{j}\}_{j=1,\ldots,n} is the canonical basis of n\mathbb{R}^{n}. We define the mapping Φ:n×[β,β]nn\Phi:\mathbb{R}^{n}\times[-\beta,\beta]^{n}\longrightarrow\mathbb{R}^{n} by

Φ(z,α¯)=E(u+j=1nαjvj)z\Phi(z,\overline{\alpha})=\mathrm{E}\bigg{(}u+\displaystyle\sum_{j=1}^{n}\alpha_{j}v_{j}\bigg{)}-z

for all (z,α¯)n×[β,β]n(z,\overline{\alpha})\in\mathbb{R}^{n}\times[-\beta,\beta]^{n}, where β>0\beta>0 is small enough to guarantee that u+j=1nαjvj𝒰u+\sum_{j=1}^{n}\alpha_{j}v_{j}\in\mathcal{U} for all α¯[β,β]n\overline{\alpha}\in[-\beta,\beta]^{n}, which is possible because 𝒰\mathcal{U} is an open subset of L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). The mapping Φ\Phi is of class C1\mathrm{C}^{1} and satisfies Φ(xu(T),0n)=0n\Phi(x_{u}(T),0_{\mathbb{R}^{n}})=0_{\mathbb{R}^{n}} and Φα¯(xu(T),0n)=Idn\frac{\partial\Phi}{\partial\overline{\alpha}}(x_{u}(T),0_{\mathbb{R}^{n}})=\mathrm{Id}_{\mathbb{R}^{n}} which is invertible. By the implicit function theorem, there exists an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a C1\mathrm{C}^{1} mapping α¯:𝒱[β,β]n\overline{\alpha}:\mathcal{V}\to[-\beta,\beta]^{n} satisfying α¯(xu(T))=0n\overline{\alpha}(x_{u}(T))=0_{\mathbb{R}^{n}} and Φ(z,α¯(z))=0n\Phi(z,\overline{\alpha}(z))=0_{\mathbb{R}^{n}} for all z𝒱z\in\mathcal{V}. Then it suffices to introduce the C1\mathrm{C}^{1} mapping V:𝒱𝒰V:\mathcal{V}\to\mathcal{U} defined by V(z)=u+j=1nαj(z)vjV(z)=u+\sum_{j=1}^{n}\alpha_{j}(z)v_{j} for all z𝒱z\in\mathcal{V}.

IV-B Proof of Proposition 2

Lemma 1.

Let u𝒰u\in\mathcal{U} and pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) be a solution to (AE). The following statements are equivalent:

  1. (i)

    (xu,u,p)(x_{u},u,p) is a weak extremal lift of the pair (xu,u)(x_{u},u);

  2. (ii)

    p(T),DE(u)vn=0\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}=0 for all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}).

Proof.

We set hv(t)=p(t),wvu(t)nh_{v}(t)=\langle p(t),w^{u}_{v}(t)\rangle_{\mathbb{R}^{n}} for all t[0,T]t\in[0,T] and all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), where wvuw^{u}_{v} is defined after (1). Therefore (ii) is equivalent to hv(T)=0h_{v}(T)=0 for all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). For all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), note that hv(0)=0h_{v}(0)=0 and, using the adjoint equation (AE), that

h˙v(t)=uH(xu(t),u(t),p(t),t),v(t)m\dot{h}_{v}(t)=\langle\nabla_{u}H(x_{u}(t),u(t),p(t),t),v(t)\rangle_{\mathbb{R}^{m}}

for a.e. t[0,T]t\in[0,T]. 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 h˙v(t)=0\dot{h}_{v}(t)=0 for a.e. t[0,T]t\in[0,T] and thus hv(T)=hv(0)=0h_{v}(T)=h_{v}(0)=0 for all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), which gives (ii). Now, assuming (ii), we have  0TuH(xu(t),u(t),p(t),t),v(t)m𝑑t=hv(T)=0\int_{0}^{T}\langle\nabla_{u}H(x_{u}(t),u(t),p(t),t),v(t)\rangle_{\mathbb{R}^{m}}dt=h_{v}(T)=0 for every vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). We deduce the null Hamiltonian gradient condition (NHG), which gives (i). ∎

Let us prove Proposition 2. Let u𝒰u\in\mathcal{U}. First, assume that uu is weakly singular, i.e., Ran(DE(u))\mathrm{Ran}(\mathrm{D}\mathrm{E}(u)) is a proper subspace of n\mathbb{R}^{n}. Hence there exists ψn\{0n}\psi\in\mathbb{R}^{n}\backslash\{0_{\mathbb{R}^{n}}\} such that ψ,DE(u)vn=0\langle\psi,\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}=0 for all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). Considering pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) the unique solution to (AE) ending at p(T)=ψp(T)=\psi (in particular pp is not trivial), we obtain that p(T),DE(u)vn=0\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}=0 for all vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). By Lemma 1, (xu,u,p)(x_{u},u,p) is a nontrivial weak extremal lift of (xu,u)(x_{u},u). Conversely, assume that uu is strongly regular, i.e., Ran(DE(u))=n\mathrm{Ran}(\mathrm{D}\mathrm{E}(u))=\mathbb{R}^{n}. By contradiction let us assume that (xu,u)(x_{u},u) admits a nontrivial weak extremal lift (xu,u,p)(x_{u},u,p). Then there exists vL([0,T],m)v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) such that DE(u)v=p(T)\mathrm{D}\mathrm{E}(u)\cdot v=p(T). It follows from Lemma 1 that p(T)n2=0\|p(T)\|_{\mathbb{R}^{n}}^{2}=0 and thus p(T)=0np(T)=0_{\mathbb{R}^{n}}. Since the adjoint equation (AE) is linear, it follows that pp is trivial, which raises a contradiction.

IV-C Proof of Proposition 3

Lemma 2.

Assume that U\mathrm{U} is convex and let uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}). We have

𝒯LU[u]={vL([0,T],m)β>0,u+βvL([0,T],U)}.\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]=\big{\{}v\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m})\mid\\ \exists\beta>0,\;u+\beta v\in\mathrm{L}^{\infty}([0,T],\mathrm{U})\big{\}}.

Furthermore, for every JJ\in\mathbb{N}^{*}, we have

u+j=1JαjvjL([0,T],U)u+\sum_{j=1}^{J}\alpha_{j}v_{j}\in\mathrm{L}^{\infty}([0,T],\mathrm{U})

for every αj[0,βjJ]\alpha_{j}\in[0,\frac{\beta_{j}}{J}], where vj𝒯LU[u]v_{j}\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] and βj>0\beta_{j}>0 is such that u+βjvjL([0,T],U)u+\beta_{j}v_{j}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) for every j{1,,J}j\in\{1,\ldots,J\}.

Lemma 2 is obvious. Assume that U\mathrm{U} is convex and let us prove Proposition 3. Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) be strongly U\mathrm{U}-regular. By Definition 3, there exists a 2n2n-tuple v¯={vj}j=1,,2n\overline{v}=\{v_{j}\}_{j=1,\ldots,2n} of elements of 𝒯LU[u]\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] such that

DE(u)vj=ejandDE(u)vn+j=ej\mathrm{D}\mathrm{E}(u)\cdot v_{j}=e_{j}\quad\text{and}\quad\mathrm{D}\mathrm{E}(u)\cdot v_{n+j}=-e_{j} (5)

for every j{1,,n}j\in\{1,\ldots,n\}, where {ej}j=1,,n\{e_{j}\}_{j=1,\ldots,n} is the canonical basis of n\mathbb{R}^{n}. We define the map Φ:n×[0,β]2nn\Phi:\mathbb{R}^{n}\times[0,\beta]^{2n}\longrightarrow\mathbb{R}^{n} by

Φ(z,α¯)=E(u+j=12nαjvj)z\Phi(z,\overline{\alpha})=\mathrm{E}\bigg{(}u+\displaystyle\sum_{j=1}^{2n}\alpha_{j}v_{j}\bigg{)}-z

for all (z,α¯)n×[0,β]2n(z,\overline{\alpha})\in\mathbb{R}^{n}\times[0,\beta]^{2n}, where β>0\beta>0 is small enough to guarantee that u+j=12nαjvj𝒰L([0,T],U)u+\sum_{j=1}^{2n}\alpha_{j}v_{j}\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) for every α¯[0,β]2n\overline{\alpha}\in[0,\beta]^{2n}, which is possible by Lemma 2 and because 𝒰\mathcal{U} is an open subset of L([0,T],m)\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). The mapping Φ\Phi is of class C1\mathrm{C}^{1} and satisfies Φ(xu(T),02n)=0n\Phi(x_{u}(T),0_{\mathbb{R}^{2n}})=0_{\mathbb{R}^{n}} and Φα¯(xu(T),02n)+2n=n\frac{\partial\Phi}{\partial\overline{\alpha}}(x_{u}(T),0_{\mathbb{R}^{2n}})\cdot\mathbb{R}^{2n}_{+}=\mathbb{R}^{n} thanks to (5). From the conic implicit function theorem [3, Theorem 1], there exists an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a continuous mapping α¯:𝒱[0,β]2n\overline{\alpha}:\mathcal{V}\to[0,\beta]^{2n} satisfying α¯(xu(T))=02n\overline{\alpha}(x_{u}(T))=0_{\mathbb{R}^{2n}} and Φ(z,α¯(z))=0n\Phi(z,\overline{\alpha}(z))=0_{\mathbb{R}^{n}} for all z𝒱z\in\mathcal{V}. Then it suffices to introduce the continuous mapping V:𝒱𝒰L([0,T],U)V:\mathcal{V}\to\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) defined by V(z)=u+j=12nαj(z)vjV(z)=u+\sum_{j=1}^{2n}\alpha_{j}(z)v_{j} for all z𝒱z\in\mathcal{V}.

IV-D Proof of Proposition 4

Lemma 3.

Assume that U\mathrm{U} is convex. Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) and pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) be a solution to (AE). The following statements are equivalent:

  1. (i)

    (xu,u,p)(x_{u},u,p) is a weak U\mathrm{U}-extremal lift of the pair (xu,u)(x_{u},u);

  2. (ii)

    p(T),DE(u)(vu)n0\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot(v-u)\rangle_{\mathbb{R}^{n}}\leq 0 for all vL([0,T],U)v\in\mathrm{L}^{\infty}([0,T],\mathrm{U});

  3. (iii)

    p(T),DE(u)vn0\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}\leq 0 for all v𝒯LU[u]v\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u].

Proof.

The equivalence between (ii) and (iii) follows from the definition of 𝒯LU[u]\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] (see Definition 3). Note that (ii) is equivalent to hvu(T)0h_{v-u}(T)\leq 0 for all vL([0,T],U)v\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) (see the definition of hvuh_{v-u} 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 h˙vu(t)0\dot{h}_{v-u}(t)\leq 0 for a.e. t[0,T]t\in[0,T] and thus hvu(T)hvu(0)=0h_{v-u}(T)\leq h_{v-u}(0)=0 for all vL([0,T],U)v\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), which gives (ii). Now, assuming (ii), we have 0TuH(xu(t),u(t),p(t),t),v(t)u(t)m𝑑t=hvu(T)0\int_{0}^{T}\langle\nabla_{u}H(x_{u}(t),u(t),p(t),t),v(t)-u(t)\rangle_{\mathbb{R}^{m}}dt=h_{v-u}(T)\leq 0 for every vL([0,T],U)v\in\mathrm{L}^{\infty}([0,T],\mathrm{U}). Then, for any Lebesgue point τ[0,T)\tau\in[0,T) of uH(xu,u,p,)L([0,T],m)\nabla_{u}H(x_{u},u,p,\cdot)\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) and of uH(xu,u,p,),umL([0,T],)\langle\nabla_{u}H(x_{u},u,p,\cdot),u\rangle_{\mathbb{R}^{m}}\in\mathrm{L}^{\infty}([0,T],\mathbb{R}) and for any ωU\omega\in\mathrm{U}, taking the needle-like variation v=u(τ,ω)αL([0,T],U)v=u^{\alpha}_{(\tau,\omega)}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) as defined in (2), we get that 1αττ+αuH(xu(t),u(t),p(t),t),ωu(t)m𝑑t0\frac{1}{\alpha}\int_{\tau}^{\tau+\alpha}\langle\nabla_{u}H(x_{u}(t),u(t),p(t),t),\omega-u(t)\rangle_{\mathbb{R}^{m}}dt\leq 0 for every α>0\alpha>0 small enough. Taking the limit α0+\alpha\to 0^{+}, since τ\tau is an appropriate Lebesgue point, we obtain that uH(xu(τ),u(τ),p(τ),τ),ωu(τ)m0\langle\nabla_{u}H(x_{u}(\tau),u(\tau),p(\tau),\tau),\omega-u(\tau)\rangle_{\mathbb{R}^{m}}\leq 0. Since τ\tau and ω\omega have been chosen arbitrarily, the Hamiltonian gradient condition (HG) is satisfied, which gives (i). ∎

Assume that U\mathrm{U} is convex and let us prove Proposition 4. Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). Firstly, assume that uu is weakly U\mathrm{U}-singular, i.e., DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) is a proper subcone of n\mathbb{R}^{n}. Hence 0n0_{\mathbb{R}^{n}} belongs to its boundary and, since DE(u)(𝒯LU[u])\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]) is also convex, by a standard separation argument, there exists ψn\{0n}\psi\in\mathbb{R}^{n}\backslash\{0_{\mathbb{R}^{n}}\} such that ψ,DE(u)vn0\langle\psi,\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}\leq 0 for all v𝒯LU[u]v\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]. Considering pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) the unique solution to (AE) ending at p(T)=ψp(T)=\psi (in particular pp is not trivial), we obtain that p(T),DE(u)vn0\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot v\rangle_{\mathbb{R}^{n}}\leq 0 for all v𝒯LU[u]v\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u]. By Lemma 3, (xu,u,p)(x_{u},u,p) is a nontrivial weak U\mathrm{U}-extremal lift of (xu,u)(x_{u},u). Conversely, assume that uu is strongly U\mathrm{U}-regular, i.e., DE(u)(𝒯LU[u])=n\mathrm{D}\mathrm{E}(u)(\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u])=\mathbb{R}^{n}. By contradiction let us assume that (xu,u)(x_{u},u) admits a nontrivial weak U\mathrm{U}-extremal lift (xu,u,p)(x_{u},u,p). There exists v𝒯LU[u]v\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] such that DE(u)v=p(T)\mathrm{D}\mathrm{E}(u)\cdot v=p(T). By Lemma 3 we get that p(T)n20\|p(T)\|_{\mathbb{R}^{n}}^{2}\leq 0 and thus p(T)=0np(T)=0_{\mathbb{R}^{n}}. Since the adjoint equation (AE) is linear, it follows that pp is trivial, which raises a contradiction.

IV-E Proof of Proposition 5

Given uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) and 1s<+1\leq s<+\infty, we define

NLs(u,ρ,M)=B¯Ls(u,ρ)B¯L(0L,M)\mathrm{N}_{\mathrm{L}^{s}}(u,\rho,M)=\overline{\mathrm{B}}_{\mathrm{L}^{s}}(u,\rho)\cap\overline{\mathrm{B}}_{\mathrm{L}^{\infty}}(0_{\mathrm{L}^{\infty}},M)

for every MuLM\geq\|u\|_{\mathrm{L}^{\infty}} and every ρ>0\rho>0, which corresponds to a usual Ls\mathrm{L}^{s}-neighborhood of uu, truncated with a uniform L\mathrm{L}^{\infty}-bound. The following lemmas follow from standard techniques in ordinary differential equations theory.

Lemma 4.

Let 1s<+1\leq s<+\infty and u𝒰u\in\mathcal{U}. For any MuLM\geq\|u\|_{\mathrm{L}^{\infty}}, there exists ρM>0\rho^{M}>0 such that NLs(u,ρM,M)𝒰\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M)\subset\mathcal{U} and xvxuC1\|x_{v}-x_{u}\|_{\mathrm{C}}\leq 1 for all vNLs(u,ρM,M)v\in\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M) . Moreover the restriction of E\mathrm{E} to NLs(u,ρM,M)\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M) is Lipschitz continuous when endowing NLs(u,ρM,M)\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M) with the Ls\mathrm{L}^{s}-metric.

Definition 9 (Multiple needle-like variation).

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). A package χ=(τ¯,ω¯)(fu)Q×UR\chi=(\overline{\tau},\overline{\omega})\in\mathcal{L}(f_{u})^{Q}\times\mathrm{U}^{R}, with QQ, RR\in\mathbb{N}^{*}, QRQ\leq R, consists of:

  • a QQ-tuple τ¯={τq}q=1,,Q(fu)Q\overline{\tau}=\{\tau_{q}\}_{q=1,\ldots,Q}\in\mathcal{L}(f_{u})^{Q} such that 0τ1<τ2<<τQ<T0\leq\tau_{1}<\tau_{2}<\ldots<\tau_{Q}<T;

  • a RR-tuple ω¯={ωqr}q=1,,Qr=1,,RqUR\overline{\omega}=\{\omega^{r}_{q}\}^{r=1,\ldots,R_{q}}_{q=1,\ldots,Q}\in\mathrm{U}^{R} with RqR_{q}\in\mathbb{N}^{*} for all q{1,,Q}q\in\{1,\ldots,Q\}, and R=q=1QRqR=\sum_{q=1}^{Q}R_{q}.

The multiple needle-like variation uχα¯L([0,T],U)u^{\overline{\alpha}}_{\chi}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) of the control uu is defined by

uχα¯(t)={ωqralong [τq+=1r1αq,τq+=1rαq),r{1,,Rq},q{1,,Q},u(t)elsewhere,u^{\overline{\alpha}}_{\chi}(t)=\left\{\begin{array}[]{ll}\omega^{r}_{q}&\text{along }[\tau_{q}+\sum_{\ell=1}^{r-1}\alpha^{\ell}_{q},\tau_{q}+\sum_{\ell=1}^{r}\alpha^{\ell}_{q}),\\[5.0pt] &\quad\quad\;\forall r\in\{1,\ldots,R_{q}\},\;\forall q\in\{1,\ldots,Q\},\\[3.0pt] u(t)&\text{elsewhere,}\end{array}\right.

for a.e. t[0,T]t\in[0,T] and for all α¯+R\overline{\alpha}\in\mathbb{R}^{R}_{+} sufficiently small so that the intervals do not overlap.

Remark 17.

Let 1s<+1\leq s<+\infty and consider the framework of Definition 9. The mapping α¯uχα¯\overline{\alpha}\mapsto u^{\overline{\alpha}}_{\chi} is continuous when endowing L([0,T],U)\mathrm{L}^{\infty}([0,T],\mathrm{U}) with the Ls\mathrm{L}^{s}-metric. Taking M=uL+ω¯(m)RM=\|u\|_{\mathrm{L}^{\infty}}+\|\overline{\omega}\|_{(\mathbb{R}^{m})^{R}} and considering ρM>0\rho^{M}>0 given in Lemma 4, there exists β>0\beta>0 sufficiently small so that uχα¯NLs(u,ρM,M)𝒰u^{\overline{\alpha}}_{\chi}\in\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M)\subset\mathcal{U} for all α¯[0,β]R\overline{\alpha}\in[0,\beta]^{R}.

Lemma 5.

In the frameworks of Definition 9 and of Remark 17, the mapping Ψ:[0,β]Rn\Psi:[0,\beta]^{R}\to\mathbb{R}^{n}, defined by Ψ(α¯)=E(uχα¯)\Psi(\overline{\alpha})=\mathrm{E}(u^{\overline{\alpha}}_{\chi}) for all α¯[0,β]R\overline{\alpha}\in[0,\beta]^{R}, satisfies Ψ(0R)=xu(T)\Psi(0_{\mathbb{R}^{R}})=x_{u}(T) and is of class C1\mathrm{C}^{1} with

Ψαqr(0R)=wτq,ωqru(T)\dfrac{\partial\Psi}{\partial\alpha^{r}_{q}}(0_{\mathbb{R}^{R}})=w^{u}_{\tau_{q},\omega^{r}_{q}}(T)

for every r{1,,Rq}r\in\{1,\ldots,R_{q}\} and every q{1,,Q}q\in\{1,\ldots,Q\}.

Remark 18.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). Note that, for any Lebesgue point τq(fu)\tau_{q}\in\mathcal{L}(f_{u}) considered in a multiple needle-like variation (see Definition 9), it is possible to consider several values ωqrU\omega^{r}_{q}\in\mathrm{U} for r=1,,Rqr=1,\ldots,R_{q} with RqR_{q}\in\mathbb{N}^{*}. This additional degree of freedom is essential in order to generate the U\mathrm{U}-Pontryagin cone of uu with multiple needle-like variations, as developed in the next remark.

Remark 19.

The U\mathrm{U}-Pontryagin cone of a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) is generated by multiple needle-like variations as follows. Consider some zPontU[u]z\in\mathrm{Pont}_{\mathrm{U}}[u]. Definition 5 gives

z=q=1Q~λqw(τq,ωq)u(T)z=\sum_{q=1}^{\widetilde{Q}}\lambda_{q}w^{u}_{(\tau_{q},\omega_{q})}(T)

for some Q~\widetilde{Q}\in\mathbb{N}^{*}, where λq0\lambda_{q}\geq 0 and (τq,ωq)(fu)×U(\tau_{q},\omega_{q})\in\mathcal{L}(f_{u})\times\mathrm{U} for all q{1,,Q~}q\in\{1,\ldots,\widetilde{Q}\}. By gathering the Lebesgue points τq\tau_{q} that are equal (and thus gathering the corresponding values ωq\omega_{q}, see Remark 18), we construct a package χ=(τ¯,ω¯)(fu)Q×UR\chi=(\overline{\tau},\overline{\omega})\in\mathcal{L}(f_{u})^{Q}\times\mathrm{U}^{R} as in Definition 9 (with QR=Q~Q\leq R=\widetilde{Q}) and

z=q=1Qr=1Rqλqrw(τq,ωqr)u(T).z=\sum_{q=1}^{Q}\sum_{r=1}^{R_{q}}\lambda^{r}_{q}w^{u}_{(\tau_{q},\omega^{r}_{q})}(T).

Denoting by λ¯={λqr}q=1,,Qr=1,,Rq+R\overline{\lambda}=\{\lambda^{r}_{q}\}^{r=1,\ldots,R_{q}}_{q=1,\ldots,Q}\in\mathbb{R}^{R}_{+}, we introduce the C1\mathrm{C}^{1} mapping Ψ:[0,β]n\Psi^{\prime}:[0,\beta^{\prime}]\to\mathbb{R}^{n}, defined by Ψ(α)=Ψ(αλ¯)\Psi^{\prime}(\alpha)=\Psi(\alpha\overline{\lambda}) for all α[0,β]\alpha\in[0,\beta^{\prime}], where Ψ\Psi is the mapping defined in Lemma 5 and where β>0\beta^{\prime}>0 is sufficiently small to guarantee that αλ¯[0,β]R\alpha\overline{\lambda}\in[0,\beta]^{R} for all α[0,β]\alpha\in[0,\beta^{\prime}]. We finally get that

limα0+E(uχαλ¯)E(u)α=z\lim\limits_{\alpha\to 0^{+}}\dfrac{\mathrm{E}(u^{\alpha\overline{\lambda}}_{\chi})-\mathrm{E}(u)}{\alpha}=z

because Ψα(0)=DΨ(0R)λ¯=z\frac{\partial\Psi^{\prime}}{\partial\alpha}(0)=\mathrm{D}\Psi(0_{\mathbb{R}^{R}})\cdot\overline{\lambda}=z.

Now let us prove Proposition 5. Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) be weakly U\mathrm{U}-regular. Thus PontU[u]=n\mathrm{Pont}_{\mathrm{U}}[u]=\mathbb{R}^{n} contains eje_{j} and ej-e_{j} for all j{1,,n}j\in\{1,\ldots,n\}, where {ej}j=1,,n\{e_{j}\}_{j=1,\ldots,n} is the canonical basis of n\mathbb{R}^{n}. For all j{1,,n}j\in\{1,\ldots,n\}, Definition 5 gives

ej=q=1Q~j+λqj+w(τqj+,ωqj+)u(T)e_{j}=\sum_{q=1}^{\widetilde{Q}^{j+}}\lambda^{j+}_{q}w^{u}_{(\tau^{j+}_{q},\omega^{j+}_{q})}(T)

for some Q~j+\widetilde{Q}^{j+}\in\mathbb{N}^{*}, where λqj+0\lambda^{j+}_{q}\geq 0 and (τqj+,ωqj+)(fu)×U(\tau^{j+}_{q},\omega^{j+}_{q})\in\mathcal{L}(f_{u})\times\mathrm{U} for all q{1,,Q~j+}q\in\{1,\ldots,\widetilde{Q}^{j+}\}, and

ej=q=1Q~jλqjw(τqj,ωqj)u(T)-e_{j}=\sum_{q=1}^{\widetilde{Q}^{j-}}\lambda^{j-}_{q}w^{u}_{(\tau^{j-}_{q},\omega^{j-}_{q})}(T)

for some Q~j\widetilde{Q}^{j-}\in\mathbb{N}^{*}, where λqj0\lambda^{j-}_{q}\geq 0 and (τqj,ωqj)(fu)×U(\tau^{j-}_{q},\omega^{j-}_{q})\in\mathcal{L}(f_{u})\times\mathrm{U} for all q{1,,Q~j}q\in\{1,\ldots,\widetilde{Q}^{j-}\}. By gathering the Lebesgue points τqj±\tau^{j\pm}_{q} which are equal (and thus gathering the corresponding values ωqj±\omega^{j\pm}_{q}, see Remark 18), we construct a package χ=(τ¯,ω¯)(fu)Q×UR\chi=(\overline{\tau},\overline{\omega})\in\mathcal{L}(f_{u})^{Q}\times\mathrm{U}^{R} as in Definition 9 (with QR=j=1n(Q~j++Q~j)Q\leq R=\sum_{j=1}^{n}(\widetilde{Q}^{j+}+\widetilde{Q}^{j-})). Considering the C1\mathrm{C}^{1} mapping Ψ\Psi defined in Lemma 5, it is clear, in the same spirit as in Remark 19, that each vector eje_{j} and ej-e_{j} belong to DΨ(0R)+R\mathrm{D}\Psi(0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}, and thus DΨ(0R)+R=n\mathrm{D}\Psi(0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}=\mathbb{R}^{n}. Now we define the mapping Φ:n×[0,β]Rn\Phi:\mathbb{R}^{n}\times[0,\beta]^{R}\longrightarrow\mathbb{R}^{n} by Φ(z,α¯)=Ψ(α¯)z\Phi(z,\overline{\alpha})=\Psi(\overline{\alpha})-z for all (z,α¯)n×[0,β]R(z,\overline{\alpha})\in\mathbb{R}^{n}\times[0,\beta]^{R}. The mapping Φ\Phi is of class C1\mathrm{C}^{1} and satisfies Φ(xu(T),0R)=0n\Phi(x_{u}(T),0_{\mathbb{R}^{R}})=0_{\mathbb{R}^{n}} and Φα¯(xu(T),0R)+R=DΨ(0R)+R=n\frac{\partial\Phi}{\partial\overline{\alpha}}(x_{u}(T),0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}=\mathrm{D}\Psi(0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}=\mathbb{R}^{n}. From the conic implicit function theorem [3, Theorem 1], there exists an open neighborhood 𝒱\mathcal{V} of xu(T)x_{u}(T) and a continuous mapping α¯:𝒱[0,β]R\overline{\alpha}:\mathcal{V}\to[0,\beta]^{R} satisfying α¯(xu(T))=0R\overline{\alpha}(x_{u}(T))=0_{\mathbb{R}^{R}} and Φ(z,α¯(z))=0n\Phi(z,\overline{\alpha}(z))=0_{\mathbb{R}^{n}} for all z𝒱z\in\mathcal{V}. Then it suffices to introduce the mapping V:𝒱𝒰L([0,T],U)V:\mathcal{V}\to\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) defined by V(z)=uχα¯(z)V(z)=u^{\overline{\alpha}(z)}_{\chi} for all z𝒱z\in\mathcal{V}. By Remark 17, the mapping VV is continuous when endowing its codomain with the L1\mathrm{L}^{1}-metric.

IV-F Proof of Proposition 6

Lemma 6.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) and pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) be a solution to (AE). The following statements are equivalent:

  1. (i)

    (xu,u,p)(x_{u},u,p) is a strong U\mathrm{U}-extremal lift of the pair (xu,u)(x_{u},u);

  2. (ii)

    p(T),zn0\langle p(T),z\rangle_{\mathbb{R}^{n}}\leq 0 for all zPontU[u]z\in\mathrm{Pont}_{\mathrm{U}}[u];

  3. (iii)

    p(T),w(τ,ω)u(T)n0\langle p(T),w^{u}_{(\tau,\omega)}(T)\rangle_{\mathbb{R}^{n}}\leq 0 for all (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}.

Proof.

The equivalence between (ii) and (iii) follows from Definition 5. For all (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}, we set h(τ,ω)(t)=p(t),w(τ,ω)u(t)nh_{(\tau,\omega)}(t)=\langle p(t),w^{u}_{(\tau,\omega)}(t)\rangle_{\mathbb{R}^{n}} for all t[τ,T]t\in[\tau,T] which is constant thanks to (AE). Note that (iii), which can be written as h(τ,ω)(T)0h_{(\tau,\omega)}(T)\leq 0 for all (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}, is equivalent to h(τ,ω)(τ)0h_{(\tau,\omega)}(\tau)\leq 0 for all (τ,ω)(fu)×U(\tau,\omega)\in\mathcal{L}(f_{u})\times\mathrm{U}, which exactly corresponds to the Hamiltonian maximization condition (HM), giving (i). ∎

Let us prove Proposition 6. Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}). First, assume that uu is strongly U\mathrm{U}-singular, i.e., PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] is a proper subcone of n\mathbb{R}^{n}. Hence 0n0_{\mathbb{R}^{n}} belongs to its boundary and, since PontU[u]\mathrm{Pont}_{\mathrm{U}}[u] is also convex, by a standard separation argument, there exists ψn\{0n}\psi\in\mathbb{R}^{n}\backslash\{0_{\mathbb{R}^{n}}\} such that ψ,zn0\langle\psi,z\rangle_{\mathbb{R}^{n}}\leq 0 for all zPontU[u]z\in\mathrm{Pont}_{\mathrm{U}}[u]. Considering pAC([0,T],n)p\in\mathrm{AC}([0,T],\mathbb{R}^{n}) the unique solution to (AE) ending at p(T)=ψp(T)=\psi (in particular pp is not trivial), we obtain that p(T),zn0\langle p(T),z\rangle_{\mathbb{R}^{n}}\leq 0 for all zPontU[u]z\in\mathrm{Pont}_{\mathrm{U}}[u]. By Lemma 6, (xu,u,p)(x_{u},u,p) is a nontrivial strong U\mathrm{U}-extremal lift of (xu,u)(x_{u},u). Conversely, assume that uu is weakly U\mathrm{U}-regular, i.e., PontU[u]=n\mathrm{Pont}_{\mathrm{U}}[u]=\mathbb{R}^{n}. By contradiction, let us assume that (xu,u)(x_{u},u) admits a nontrivial strong U\mathrm{U}-extremal lift (xu,u,p)(x_{u},u,p). Since p(T)PontU[u]=np(T)\in\mathrm{Pont}_{\mathrm{U}}[u]=\mathbb{R}^{n}, it follows from Lemma 6 that p(T)n20\|p(T)\|_{\mathbb{R}^{n}}^{2}\leq 0 and thus p(T)=0np(T)=0_{\mathbb{R}^{n}}. Since the adjoint equation (AE) is linear, it follows that pp is trivial, which raises a contradiction.

IV-G Proof of Proposition 9 (only the sufficient condition)

First step: assume that U\mathrm{U} is convex and that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set. Let us prove that uu is strongly U\mathrm{U}-regular (and thus is weakly U\mathrm{U}-regular by Proposition 8). By contradiction assume that uu is weakly U\mathrm{U}-singular. By Proposition 4, let (xu,u,p)(x_{u},u,p) be a nontrivial weak U\mathrm{U}-extremal lift of the pair (xu,u)(x_{u},u). Since the adjoint equation (AE) is linear, we know that p(T)0np(T)\neq 0_{\mathbb{R}^{n}}. Since xu(T)x_{u}(T) belongs to the interior of E(L([0,T],U))\mathrm{E}(\mathrm{L}^{\infty}([0,T],\mathrm{U})), there exist γ>0\gamma>0 sufficiently small and vL([0,T],U)v\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) such that xu(T)+γp(T)=E(v)x_{u}(T)+\gamma p(T)=\mathrm{E}(v). Since the control system (CS) is linear, E\mathrm{E} is affine and thus γp(T)=E(v)E(u)=DE(u)(vu)\gamma p(T)=\mathrm{E}(v)-\mathrm{E}(u)=\mathrm{D}\mathrm{E}(u)\cdot(v-u). Then γp(T)n2=p(T),DE(u)(vu)n0\gamma\|p(T)\|^{2}_{\mathbb{R}^{n}}=\langle p(T),\mathrm{D}\mathrm{E}(u)\cdot(v-u)\rangle_{\mathbb{R}^{n}}\leq 0 by Lemma 3, and thus p(T)=0np(T)=0_{\mathbb{R}^{n}}, which raises a contradiction.

Second step: in the general control constraints case, assume that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set. Then xu(T)x_{u}(T) belongs to the interior of the Lconv(U)\mathrm{L}^{\infty}_{\mathrm{conv}(\mathrm{U})}-accessible set. Since uL([0,T],U)L([0,T],conv(U))u\in\mathrm{L}^{\infty}([0,T],\mathrm{U})\subset\mathrm{L}^{\infty}([0,T],\mathrm{conv}(\mathrm{U})), we infer from the first step that uu is strongly conv(U)\mathrm{conv}(\mathrm{U})-regular. We deduce that uu is weakly U\mathrm{U}-regular from Proposition 8.

V Proofs of results of Section III

V-A Proof of Proposition 10

Remark 20.

Given a partition 𝕋\mathbb{T} of [0,T][0,T], it is clear that a target point x1nx^{1}\in\mathbb{R}^{n} is PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0} if and only if x1x^{1} is PCU𝕋\mathrm{PC}^{\mathbb{T}^{\prime}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for at least one partition 𝕋\mathbb{T}^{\prime} of [0,T][0,T] such that 𝕋𝕋\mathbb{T}^{\prime}\subset\mathbb{T}, if and only if x1x^{1} is PCU𝕋\mathrm{PC}^{\mathbb{T}^{\prime}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for all partitions 𝕋\mathbb{T}^{\prime} of [0,T][0,T] such that 𝕋𝕋\mathbb{T}\subset\mathbb{T}^{\prime}.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) and assume that Property (4) is not satisfied. Let 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} be a partition of [0,T][0,T] and ε>0\varepsilon>0. Since Property (4) is not satisfied, there exists a partition 𝕋={ti}i=0,,N\mathbb{T}^{\prime}=\{t^{\prime}_{i}\}_{i=0,\ldots,N^{\prime}} of [0,T][0,T] such that 𝕋<2ε\|\mathbb{T}^{\prime}\|<2\varepsilon and such that xu(T)x_{u}(T) is not PCU𝕋\mathrm{PC}^{\mathbb{T}^{\prime}}_{\mathrm{U}}-reachable in time TT from x0x^{0}. For any i{1,,N1}i\in\{1,\ldots,N-1\}, the intersection 𝕋(tiε,ti+ε)\mathbb{T}^{\prime}\cap(t_{i}-\varepsilon,t_{i}+\varepsilon) is not empty and we select tiεt^{\varepsilon}_{i} one of its elements. For i=0i=0 (resp. i=Ni=N), we choose t0ε=0t^{\varepsilon}_{0}=0 (resp. tNε=Tt^{\varepsilon}_{N}=T). Consider the partition 𝕋ε={tiε}i=0,,N\mathbb{T}^{\varepsilon}=\{t^{\varepsilon}_{i}\}_{i=0,\ldots,N} of [0,T][0,T]. Since 𝕋ε𝕋\mathbb{T}^{\varepsilon}\subset\mathbb{T}^{\prime}, we know from Remark 20 that xu(T)x_{u}(T) is not PCU𝕋ε\mathrm{PC}^{\mathbb{T}^{\varepsilon}}_{\mathrm{U}}-reachable in time TT from x0x^{0}.

V-B Proof of Proposition 11

Lemma 7 (Approximated reachability).

Given any u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) and any ε>0\varepsilon>0, there exists a threshold δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exists v𝒰PC𝕋([0,T],U)v\in\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that xv(T)xu(T)nε\|x_{v}(T)-x_{u}(T)\|_{\mathbb{R}^{n}}\leq\varepsilon.

Proof.

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) and ε>0\varepsilon>0. Take s=1s=1 and M=uLM=\|u\|_{\mathrm{L}^{\infty}} in Lemma 4 and let LM>0L^{M}>0 being a positive Lipschitz constant of E\mathrm{E} restricted to NL1(u,ρM,M)\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M) endowed with the L1\mathrm{L}^{1}-metric. By Proposition 12, there exists δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exists vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that vuL1min(ρM,εLM)\|v-u\|_{\mathrm{L}^{1}}\leq\min(\rho^{M},\frac{\varepsilon}{L^{M}}) and vLuL=M\|v\|_{\mathrm{L}^{\infty}}\leq\|u\|_{\mathrm{L}^{\infty}}=M. Since vNL1(u,ρM,M)𝒰v\in\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M)\subset\mathcal{U}, from Lemma 4, we have xv(T)xu(T)n=E(v)E(u)nLMvuL1ε\|x_{v}(T)-x_{u}(T)\|_{\mathbb{R}^{n}}=\|\mathrm{E}(v)-\mathrm{E}(u)\|_{\mathbb{R}^{n}}\leq L^{M}\|v-u\|_{\mathrm{L}^{1}}\leq\varepsilon. ∎

Let us prove Proposition 11. Let uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) be such that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set. There exist uu^{\prime}, u′′L([0,T],U)u^{\prime\prime}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) such that xu(T)<xu(T)<xu′′(T)x_{u^{\prime}}(T)<x_{u}(T)<x_{u^{\prime\prime}}(T). We infer from Lemma 7 that there exists δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exist vv^{\prime}, v′′PC𝕋([0,T],U)v^{\prime\prime}\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that xv(T)xu(T)xv′′(T)x_{v^{\prime}}(T)\leq x_{u}(T)\leq x_{v^{\prime\prime}}(T). Now let us fix such a partition 𝕋\mathbb{T} of [0,T][0,T] which satisfies 𝕋δ\|\mathbb{T}\|\leq\delta. In view of the above, we know that xu(T)x_{u}(T) belongs to the convex hull of E(PC𝕋([0,T],U))\mathrm{E}(\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})). On the other hand, since U\mathrm{U} is convex, PC𝕋([0,T],U)\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) is convex and thus is a connected set. Since E\mathrm{E} is continuous on 𝒰=L([0,T],m)\mathcal{U}=\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}), we deduce that E(PC𝕋([0,T],U))\mathrm{E}(\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})) is a connected set of \mathbb{R}, and thus is convex. We have proved that xu(T)E(PC𝕋([0,T],U))x_{u}(T)\in\mathrm{E}(\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U})).

V-C Proof of Theorem 1 under strong U\mathrm{U}-regularity

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) be a control such that x1=xu(T)=E(u)x^{1}=x_{u}(T)=\mathrm{E}(u). Let M=xuC+uL+1M=\|x_{u}\|_{\mathrm{C}}+\|u\|_{\mathrm{L}^{\infty}}+1 and let us fix some 1<s<+1<s<+\infty. Using the truncated dynamics fMf^{M} introduced in Appendix -H, we have xuM=xux^{M}_{u}=x_{u} and DEM(u)=DE(u)\mathrm{D}\mathrm{E}^{M}(u)=\mathrm{D}\mathrm{E}(u) (see Remark 21). Assume that uu is strongly U\mathrm{U}-regular. By Definition 3, there exists a 2n2n-tuple v¯={vj}j=1,,2n\overline{v}=\{v_{j}\}_{j=1,\ldots,2n} of elements of 𝒯LU[u]\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] such that

DEM(u)vj=DE(u)vj=ejDEM(u)vn+j=DE(u)vn+j=ej\begin{split}&\mathrm{D}\mathrm{E}^{M}(u)\cdot v_{j}=\mathrm{D}\mathrm{E}(u)\cdot v_{j}=e_{j}\\ &\mathrm{D}\mathrm{E}^{M}(u)\cdot v_{n+j}=\mathrm{D}\mathrm{E}(u)\cdot v_{n+j}=-e_{j}\end{split} (6)

for every j{1,,n}j\in\{1,\ldots,n\}, where {ej}j=1,,n\{e_{j}\}_{j=1,\ldots,n} is the canonical basis of n\mathbb{R}^{n}. We define the mapping Ψ:Ls([0,T],m)×Ls([0,T],m)2n×+2nn\Psi:\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\times\mathrm{L}^{s}([0,T],\mathbb{R}^{m})^{2n}\times\mathbb{R}^{2n}_{+}\longrightarrow\mathbb{R}^{n} by

Ψ(y,z¯,α¯)=EM(y+j=12nαjzj)\Psi(y,\overline{z},\overline{\alpha})=\mathrm{E}^{M}\bigg{(}y+\displaystyle\sum_{j=1}^{2n}\alpha_{j}z_{j}\bigg{)}

for all (y,z¯,α¯)Ls([0,T],m)×Ls([0,T],m)2n×+2n(y,\overline{z},\overline{\alpha})\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\times\mathrm{L}^{s}([0,T],\mathbb{R}^{m})^{2n}\times\mathbb{R}^{2n}_{+}. This mapping satisfies Ψ(u,v¯,02n)=EM(u)=xuM(T)=xu(T)=x1\Psi(u,\overline{v},0_{\mathbb{R}^{2n}})=\mathrm{E}^{M}(u)=x^{M}_{u}(T)=x_{u}(T)=x^{1}. Furthermore, since EM:Ls([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\to\mathbb{R}^{n} is of class C1\mathrm{C}^{1} (see Proposition 14), the mapping Ψ\Psi is also of class C1\mathrm{C}^{1} and we infer from (6) that Ψα¯(u,v¯,02n)+2n=n\frac{\partial\Psi}{\partial\overline{\alpha}}(u,\overline{v},0_{\mathbb{R}^{2n}})\cdot\mathbb{R}^{2n}_{+}=\mathbb{R}^{n}. By the conic implicit function theorem [3, Theorem 1], there exists a continuous mapping α¯:B¯Ls(u,η)×B¯(Ls)2n(v¯,η)+2n\overline{\alpha}:\overline{\mathrm{B}}_{\mathrm{L}^{s}}(u,\eta)\times\overline{\mathrm{B}}_{(\mathrm{L}^{s})^{2n}}(\overline{v},\eta)\to\mathbb{R}^{2n}_{+}, with η>0\eta>0, satisfying α¯(u,v¯)=02n\overline{\alpha}(u,\overline{v})=0_{\mathbb{R}^{2n}} and Ψ(y,z¯,α¯(y,z¯))=x1\Psi(y,\overline{z},\overline{\alpha}(y,\overline{z}))=x^{1} for all (y,z¯)B¯Ls(u,η)×B¯(Ls)2n(v¯,η)(y,\overline{z})\in\overline{\mathrm{B}}_{\mathrm{L}^{s}}(u,\eta)\times\overline{\mathrm{B}}_{(\mathrm{L}^{s})^{2n}}(\overline{v},\eta).

By Lemma 9, there exists a threshold δ>0\delta>0 such that 𝕋(u)B¯Ls(u,η)\mathcal{I}^{\mathbb{T}}(u)\in\overline{\mathrm{B}}_{\mathrm{L}^{s}}(u,\eta) and 𝕋(v¯)B¯(Ls)2n(v¯,η)\mathcal{I}^{\mathbb{T}}(\overline{v})\in\overline{\mathrm{B}}_{(\mathrm{L}^{s})^{2n}}(\overline{v},\eta), and thus

Ψ(𝕋(u),𝕋(v¯),α¯(𝕋(u),𝕋(v¯)))=x1\Psi\Big{(}\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v}),\overline{\alpha}\big{(}\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v})\big{)}\Big{)}=x^{1}

for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, where 𝕋\mathcal{I}^{\mathbb{T}} is the averaging operator introduced in Appendix -G. For any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, we define the control

V𝕋=u+j=12nαj(𝕋(u),𝕋(v¯))vjL([0,T],m).V^{\mathbb{T}}=u+\displaystyle\sum_{j=1}^{2n}\alpha_{j}(\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v}))v_{j}\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}).

Using the linearity of the averaging operators, we obtain the piecewise constant control

𝕋(V𝕋)=𝕋(u)+j=12nαj(𝕋(u),𝕋(v¯))𝕋(vj)PC𝕋([0,T],m)\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}})=\mathcal{I}^{\mathbb{T}}(u)\\ +\displaystyle\sum_{j=1}^{2n}\alpha_{j}(\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v}))\mathcal{I}^{\mathbb{T}}(v_{j})\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathbb{R}^{m})

which satisfies

EM(𝕋(V𝕋))=Ψ(𝕋(u),𝕋(v¯),α¯(𝕋(u),𝕋(v¯)))=x1\mathrm{E}^{M}(\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}}))=\Psi\Big{(}\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v}),\overline{\alpha}\big{(}\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v})\big{)}\Big{)}=x^{1}

for all partitions 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta. If necessary we take a smaller value of δ>0\delta>0 to have 𝕋(u)uLs\|\mathcal{I}^{\mathbb{T}}(u)-u\|_{\mathrm{L}^{s}} and 𝕋(v¯))v¯(Ls)2n\|\mathcal{I}^{\mathbb{T}}(\overline{v}))-\overline{v}\|_{(\mathrm{L}^{s})^{2n}} small enough (by Lemma 9), and thus α¯(𝕋(u),𝕋(v¯))2n\|\overline{\alpha}(\mathcal{I}^{\mathbb{T}}(u),\mathcal{I}^{\mathbb{T}}(\overline{v}))\|_{\mathbb{R}^{2n}} small enough as well, to get that:

  1. (i)

    𝕋(V𝕋)LV𝕋LuL+1M\|\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}})\|_{\mathrm{L}^{\infty}}\leq\|V^{\mathbb{T}}\|_{\mathrm{L}^{\infty}}\leq\|u\|_{\mathrm{L}^{\infty}}+1\leq M (here we used in particular Lemma 8);

  2. (ii)

    𝕋(V𝕋)uLs𝕋(V𝕋)𝕋(u)Ls+𝕋(u)uLsV𝕋uLs+𝕋(u)uLsρM\|\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}})-u\|_{\mathrm{L}^{s}}\leq\|\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}})-\mathcal{I}^{\mathbb{T}}(u)\|_{\mathrm{L}^{s}}+\|\mathcal{I}^{\mathbb{T}}(u)-u\|_{\mathrm{L}^{s}}\leq\|V^{\mathbb{T}}-u\|_{\mathrm{L}^{s}}+\|\mathcal{I}^{\mathbb{T}}(u)-u\|_{\mathrm{L}^{s}}\leq\rho^{M} where ρM>0\rho^{M}>0 is given in Lemma 4 (here also we used Lemma 8);

  3. (iii)

    V𝕋V^{\mathbb{T}} is with values in U\mathrm{U} (which is possible by Lemma 2 with J=2nJ=2n and using that vj𝒯LU[u]v_{j}\in\mathcal{T}_{\mathrm{L}^{\infty}_{\mathrm{U}}}[u] for all j{1,,2n}j\in\{1,\ldots,2n\}), and thus so is 𝕋(V𝕋)\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}}) by Proposition 13;

for all partitions 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta.

We are now in a position to conclude the proof. Let us fix a partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta and, for the ease of notations, let us denote simply by V=𝕋(V𝕋)PC𝕋([0,T],m)V=\mathcal{I}^{\mathbb{T}}(V^{\mathbb{T}})\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathbb{R}^{m}) and recall that EM(V)=x1\mathrm{E}^{M}(V)=x^{1}. Since VV is with values in U\mathrm{U} from the above item (iii), we have VPC𝕋([0,T],U)V\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}). By the above items (i) and (ii) and by Lemma 4, we have VNLs(u,ρM,M)𝒰V\in\mathrm{N}_{\mathrm{L}^{s}}(u,\rho^{M},M)\subset\mathcal{U} and xVxuC1\|x_{V}-x_{u}\|_{\mathrm{C}}\leq 1. We infer that xVCxuC+1M\|x_{V}\|_{\mathrm{C}}\leq\|x_{u}\|_{\mathrm{C}}+1\leq M and, since VLM\|V\|_{\mathrm{L}^{\infty}}\leq M from the above item (i), we obtain from Remark 21 that xVM=xVx^{M}_{V}=x_{V} and thus E(V)=xV(T)=xVM(T)=EM(V)=x1\mathrm{E}(V)=x_{V}(T)=x^{M}_{V}(T)=\mathrm{E}^{M}(V)=x^{1}. The proof is complete.

V-D Proof of Theorem 1 under weak U\mathrm{U}-regularity

Let u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) be a control such that x1=xu(T)=E(u)x^{1}=x_{u}(T)=\mathrm{E}(u). Assume that uu is weakly U\mathrm{U}-regular and, by contradiction, that Property (4) is not satisfied. Then there exists a sequence (𝕋k)k(\mathbb{T}_{k})_{k\in\mathbb{N}} of partitions of [0,T][0,T] such that 𝕋k0\|\mathbb{T}_{k}\|\to 0 as k+k\to+\infty and such that x1x^{1} is not PCU𝕋k\mathrm{PC}^{\mathbb{T}_{k}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for all kk\in\mathbb{N}.

We first introduce several notations. Since uu is weakly U\mathrm{U}-regular, considering {ej}j=1,,n\{e_{j}\}_{j=1,\ldots,n} the canonical basis of n\mathbb{R}^{n}, we construct a package χ=(τ¯,ω¯)(fu)Q×UR\chi=(\overline{\tau},\overline{\omega})\in\mathcal{L}(f_{u})^{Q}\times\mathrm{U}^{R} as in the proof of Proposition 5. Now take s=1s=1 and M=uL+ω¯(m)RM=\|u\|_{\mathrm{L}^{\infty}}+\|\overline{\omega}\|_{(\mathbb{R}^{m})^{R}} and consider ρM>0\rho^{M}>0 given in Lemma 4. As in Remark 17, there exists β>0\beta>0 sufficiently small so that uχα¯NL1(u,ρM2,M)u^{\overline{\alpha}}_{\chi}\in\mathrm{N}_{\mathrm{L}^{1}}(u,\frac{\rho^{M}}{2},M) for all α¯[0,β]R\overline{\alpha}\in[0,\beta]^{R}. In particular we have uχα¯𝒰L([0,T],U)u^{\overline{\alpha}}_{\chi}\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) for all α¯[0,β]R\overline{\alpha}\in[0,\beta]^{R}. Consider the C1\mathrm{C}^{1} mapping Ψ:[0,β]Rn\Psi:[0,\beta]^{R}\to\mathbb{R}^{n}, defined by Ψ(α¯)=E(uχα¯)\Psi(\overline{\alpha})=\mathrm{E}(u^{\overline{\alpha}}_{\chi}) for all α¯[0,β]R\overline{\alpha}\in[0,\beta]^{R}, which satisfies Ψ(0R)=xu(T)\Psi(0_{\mathbb{R}^{R}})=x_{u}(T) and DΨ(0R)+R=n\mathrm{D}\Psi(0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}=\mathbb{R}^{n} as in the proof of Proposition 5.

We define the C1\mathrm{C}^{1} mapping Φ:n×[0,β]Rn\Phi:\mathbb{R}^{n}\times[0,\beta]^{R}\to\mathbb{R}^{n} by Φ(z,α¯)=Ψ(α¯)z\Phi(z,\overline{\alpha})=\Psi(\overline{\alpha})-z for all (z,α¯)n×[0,β]R(z,\overline{\alpha})\in\mathbb{R}^{n}\times[0,\beta]^{R}. It follows from the above arguments that Φα¯(xu(T),0R)+R=n\frac{\partial\Phi}{\partial\overline{\alpha}}(x_{u}(T),0_{\mathbb{R}^{R}})\cdot\mathbb{R}^{R}_{+}=\mathbb{R}^{n} and, since Φ(xu(T),0R)=0n\Phi(x_{u}(T),0_{\mathbb{R}^{R}})=0_{\mathbb{R}^{n}}, the conic implicit function theorem [3, Theorem 1] provides the existence of a continuous mapping α¯:B¯n(xu(T),η)[0,β]R\overline{\alpha}:\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta)\to[0,\beta]^{R}, with η>0\eta>0, such that α¯(xu(T))=0R\overline{\alpha}(x_{u}(T))=0_{\mathbb{R}^{R}} and Φ(z,α¯(z))=0n\Phi(z,\overline{\alpha}(z))=0_{\mathbb{R}^{n}} for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta).

The mapping V:B¯n(xu(T),η)NL1(u,ρM2,M)V:\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta)\to\mathrm{N}_{\mathrm{L}^{1}}(u,\frac{\rho^{M}}{2},M), defined by V(z)=uχα¯(z)V(z)=u^{\overline{\alpha}(z)}_{\chi} for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta), is such that V(z)𝒰L([0,T],U)V(z)\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}) for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). When endowing the codomain with the L1\mathrm{L}^{1}-metric, the continuity of VV follows from the continuity of α¯\overline{\alpha} and from Remark 17. Finally note that xV(z)(T)=E(V(z))=E(uχα¯(z))=Ψ(α¯(z))=Φ(z,α¯(z))+z=zx_{V(z)}(T)=\mathrm{E}(V(z))=\mathrm{E}(u^{\overline{\alpha}(z)}_{\chi})=\Psi(\overline{\alpha}(z))=\Phi(z,\overline{\alpha}(z))+z=z for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta).

In what follows we denote by LM>0L^{M}>0 a positive Lipschitz constant of E\mathrm{E} restricted to NL1(u,ρM,M)\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M) endowed with the L1\mathrm{L}^{1}-metric (see Lemma 4). By contradiction, assume that, for all kk\in\mathbb{N}, there exists some zkB¯n(xu(T),η)z_{k}\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta) such that

min(ρM2,ηLM)<V(zk)𝕋k(V(zk))L1,\min\bigg{(}\frac{\rho^{M}}{2},\frac{\eta}{L^{M}}\bigg{)}<\|V(z_{k})-\mathcal{I}^{\mathbb{T}_{k}}(V(z_{k}))\|_{\mathrm{L}^{1}},

where 𝕋k\mathcal{I}^{\mathbb{T}_{k}} is the averaging operator introduced in Appendix -G. By compactness of B¯n(xu(T),η)\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta), up to a subsequence (that we do not relabel), the sequence (zk)k(z_{k})_{k\in\mathbb{N}} converges to some zB¯n(xu(T),η)z^{\prime}\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). We infer from Lemma 8 that

min(ρM2,ηLM)<2V(zk)V(z)L1+V(z)𝕋k(V(z))L1\min\bigg{(}\frac{\rho^{M}}{2},\frac{\eta}{L^{M}}\bigg{)}<2\|V(z_{k})-V(z^{\prime})\|_{\mathrm{L}^{1}}\\ +\|V(z^{\prime})-\mathcal{I}^{\mathbb{T}_{k}}(V(z^{\prime}))\|_{\mathrm{L}^{1}}

for every kk\in\mathbb{N}, raising a contradiction when k+k\to+\infty by continuity of VV and by Lemma 9. We conclude that there exists KK\in\mathbb{N} such that

V(z)𝕋K(V(z))L1min(ρM2,ηLM)\|V(z)-\mathcal{I}^{\mathbb{T}_{K}}(V(z))\|_{\mathrm{L}^{1}}\leq\min\left(\frac{\rho^{M}}{2},\frac{\eta}{L^{M}}\right) (7)

for every zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). Since V(z)NL1(u,ρM2,M)V(z)\in\mathrm{N}_{\mathrm{L}^{1}}(u,\frac{\rho^{M}}{2},M), we deduce from (7) and from Lemma 8 that 𝕋K(V(z))NL1(u,ρM,M)\mathcal{I}^{\mathbb{T}_{K}}(V(z))\in\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M) for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). Since V(z)L([0,T],U)V(z)\in\mathrm{L}^{\infty}([0,T],\mathrm{U}), we infer from Proposition 13 that 𝕋K(V(z))PC𝕋K([0,T],U)\mathcal{I}^{\mathbb{T}_{K}}(V(z))\in\mathrm{PC}^{\mathbb{T}_{K}}([0,T],\mathrm{U}) for all zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta).

To conclude the proof of Theorem 1, we define :B¯n(xu(T),η)n\mathcal{B}:\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta)\to\mathbb{R}^{n} by

(z)=xu(T)+zx𝕋K(V(z))(T)=E(u)+E(V(z))E(𝕋K(V(z)))\mathcal{B}(z)=x_{u}(T)+z-x_{\mathcal{I}^{\mathbb{T}_{K}}(V(z))}(T)\\ =\mathrm{E}(u)+\mathrm{E}\Big{(}V(z)\Big{)}-\mathrm{E}\Big{(}\mathcal{I}^{\mathbb{T}_{K}}(V(z))\Big{)}

for every zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). By Lemma 8 and thanks to the continuities of the mapping VV and of the restriction of E\mathrm{E} on NL1(u,ρM,M)\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M) endowed with the L1\mathrm{L}^{1}-metric, \mathcal{B} is continuous. Furthermore, since V(z)V(z) and 𝕋K(V(z))\mathcal{I}^{\mathbb{T}_{K}}(V(z)) both belong to NL1(u,ρM,M)\mathrm{N}_{\mathrm{L}^{1}}(u,\rho^{M},M), we have

(z)xu(T)n=E(V(z))E(𝕋K(V(z)))nLMV(z)𝕋K(V(z))L1η\|\mathcal{B}(z)-x_{u}(T)\|_{\mathbb{R}^{n}}=\left\|\mathrm{E}\Big{(}V(z)\Big{)}-\mathrm{E}\Big{(}\mathcal{I}^{\mathbb{T}_{K}}(V(z))\Big{)}\right\|_{\mathbb{R}^{n}}\\ \leq L^{M}\|V(z)-\mathcal{I}^{\mathbb{T}_{K}}(V(z))\|_{\mathrm{L}^{1}}\leq\eta

for every zB¯n(xu(T),η)z\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta), where we have used (7). Therefore \mathcal{B} is a continuous mapping from B¯n(xu(T),η)\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta) with values in B¯n(xu(T),η)\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta). By the Brouwer fixed-point theorem, \mathcal{B} has a fixed-point zB¯n(xu(T),η)z^{*}\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(x_{u}(T),\eta), and thus

x𝕋K(V(z))(T)=xu(T)=x1.x_{\mathcal{I}^{\mathbb{T}_{K}}(V(z^{*}))}(T)=x_{u}(T)=x^{1}.

Since 𝕋K(V(z))𝒰PC𝕋K([0,T],U)\mathcal{I}^{\mathbb{T}_{K}}(V(z^{*}))\in\mathcal{U}\cap\mathrm{PC}^{\mathbb{T}_{K}}([0,T],\mathrm{U}), x1x^{1} is PCU𝕋K\mathrm{PC}^{\mathbb{T}_{K}}_{\mathrm{U}}-reachable in time TT from x0x^{0}, 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 u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), the condition that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set is not a sufficient condition for Property (4), even if U\mathrm{U} is convex.

Take T=n=m=2T=n=m=2 and U=2\mathrm{U}=\mathbb{R}^{2}. Take g1C([0,2],)g_{1}\in\mathrm{C}([0,2],\mathbb{R}) be a continuous function that is positive on the interval [0,1)[0,1) and vanishing on the interval [1,2][1,2]. Take g2L([0,2],)g_{2}\in\mathrm{L}^{\infty}([0,2],\mathbb{R}) be arbitrarily fixed and g3AC([0,2],)g_{3}\in\mathrm{AC}([0,2],\mathbb{R}) be defined by g3(t)=0tg1(ξ)g2(ξ)𝑑ξg_{3}(t)=\int_{0}^{t}g_{1}(\xi)g_{2}(\xi)\,d\xi for all t[0,2]t\in[0,2]. Note that g3g_{3} is constant on the interval [1,2][1,2]. We denote by GG the corresponding constant values. We set x0=02x^{0}=0_{\mathbb{R}^{2}} and the expression of f((x1,x2),(u1,u2),t)f((x_{1},x_{2}),(u_{1},u_{2}),t) by

(g1(t)u1+g1(2t)((x1G)2+x22)u1(x1g3(t))2+g1(2t)((x1G)2+x22)u2)\begin{pmatrix}g_{1}(t)u_{1}+g_{1}(2-t)\Big{(}(x_{1}-G)^{2}+x_{2}^{2}\Big{)}u_{1}\\ \Big{(}x_{1}-g_{3}(t)\Big{)}^{2}+g_{1}(2-t)\Big{(}(x_{1}-G)^{2}+x_{2}^{2}\Big{)}u_{2}\end{pmatrix}

for all ((x1,x2),(u1,u2),t)2×2×[0,2]((x_{1},x_{2}),(u_{1},u_{2}),t)\in\mathbb{R}^{2}\times\mathbb{R}^{2}\times[0,2].

Claim 1.

The point (G,0)(G,0) is an equilibrium of the control system on the interval [1,2][1,2], independently of the control.

Proof.

Since g1(t)=0g_{1}(t)=0 and g3(t)=Gg_{3}(t)=G for all t[1,2]t\in[1,2], we have f((G,0),u,t)=02f((G,0),u,t)=0_{\mathbb{R}^{2}} for all (u,t)2×[1,2](u,t)\in\mathbb{R}^{2}\times[1,2]. ∎

Claim 2.

Let uL([0,2],2)u\in\mathrm{L}^{\infty}([0,2],\mathbb{R}^{2}) satisfying u1(t)=g2(t)u_{1}(t)=g_{2}(t) for a.e. t[0,1]t\in[0,1]. Then u𝒰u\in\mathcal{U} and xu=(g3,0)x_{u}=(g_{3},0). In particular xu(2)=(G,0)x_{u}(2)=(G,0).

Proof.

Since g1(2ξ)=0g_{1}(2-\xi)=0 for all ξ[0,1]\xi\in[0,1], it holds that xu,1(t)=0tg1(ξ)u1(ξ)𝑑ξ=0tg1(ξ)g2(ξ)𝑑ξ=g3(t)x_{u,1}(t)=\int_{0}^{t}g_{1}(\xi)u_{1}(\xi)\,d\xi=\int_{0}^{t}g_{1}(\xi)g_{2}(\xi)\,d\xi=g_{3}(t) for all t[0,1]t\in[0,1]. From the second coordinate, we obtain that xu,2(t)=0x_{u,2}(t)=0 for all t[0,1]t\in[0,1]. Since xu(1)=(g3(1),0)=(G,0)x_{u}(1)=(g_{3}(1),0)=(G,0), we get from Claim 1 that xu(t)=(G,0)=(g3(t),0)x_{u}(t)=(G,0)=(g_{3}(t),0) for all t[1,2]t\in[1,2]. ∎

Claim 3.

Let u𝒰u\in\mathcal{U} such that xu(T)=(G,0)x_{u}(T^{\prime})=(G,0) for some T[1,2]T^{\prime}\in[1,2]. Then u1(t)=g2(t)u_{1}(t)=g_{2}(t) for a.e. t[0,1]t\in[0,1].

Proof.

By Claim 1, xu(1)=(G,0)x_{u}(1)=(G,0). Since g1(2ξ)=0g_{1}(2-\xi)=0 for all ξ[0,1]\xi\in[0,1], we get that 0=xu,2(1)=01(xu,1(ξ)g3(ξ))2𝑑ξ0=x_{u,2}(1)=\int_{0}^{1}(x_{u,1}(\xi)-g_{3}(\xi))^{2}\,d\xi and thus xu,1(t)=g3(t)x_{u,1}(t)=g_{3}(t) for all t[0,1]t\in[0,1]. Derivating this equality leads to g1(t)u1(t)=g1(t)g2(t)g_{1}(t)u_{1}(t)=g_{1}(t)g_{2}(t) for a.e. t[0,1]t\in[0,1]. Since g1g_{1} is positive on the interval [0,1)[0,1), we get that u1(t)=g2(t)u_{1}(t)=g_{2}(t) for a.e. t[0,1]t\in[0,1]. ∎

Claim 4.

The end-point mapping is surjective.

Proof.

Let x12x^{1}\in\mathbb{R}^{2}. Let us prove that there exists u𝒰u\in\mathcal{U} such that E(u)=xu(2)=x1\mathrm{E}(u)=x_{u}(2)=x^{1}. If x1=(G,0)x^{1}=(G,0), from Claim 2, it is sufficient to take any control uL([0,2],2)u\in\mathrm{L}^{\infty}([0,2],\mathbb{R}^{2}) which satisfies u1(t)=g2(t)u_{1}(t)=g_{2}(t) for a.e. t[0,1]t\in[0,1]. In the rest of this proof, we focus on the case x1(G,0)x^{1}\neq(G,0).

Consider a function g4L([0,32],)g_{4}\in\mathrm{L}^{\infty}([0,\frac{3}{2}],\mathbb{R}) such that the measure of {t[0,1]g4(t)g2(t)}\{t\in[0,1]\mid g_{4}(t)\neq g_{2}(t)\} is positive and such that the L\mathrm{L}^{\infty}-norm of g4g2g_{4}-g_{2} on [0,1][0,1] is small enough to guarantee that any control uL([0,2],2)u\in\mathrm{L}^{\infty}([0,2],\mathbb{R}^{2}) which satisfies u1(t)=g4(t)u_{1}(t)=g_{4}(t) for a.e. t[0,32]t\in[0,\frac{3}{2}] is admissible, i.e., u𝒰u\in\mathcal{U}. This is possible by Claim 2, since 𝒰\mathcal{U} is an open subset of L([0,2],2)\mathrm{L}^{\infty}([0,2],\mathbb{R}^{2}). Take such a control uu (which is only determined on the interval [0,32][0,\frac{3}{2}] at this step). By Claim 3, xu(32)(G,0)x_{u}(\frac{3}{2})\neq(G,0). Consider now a C1\mathrm{C}^{1} function ϱ:[32,2]2\varrho:[\frac{3}{2},2]\to\mathbb{R}^{2} which satisfies ϱ(32)=xu(32)\varrho(\frac{3}{2})=x_{u}(\frac{3}{2}), ϱ(2)=x1\varrho(2)=x^{1} and ϱ(t)(G,0)\varrho(t)\neq(G,0) for all t[32,2]t\in[\frac{3}{2},2]. We determine the control uu on [0,32][0,\frac{3}{2}] as

u1(t)=ϱ1˙(t)ϱ¯(t),u2(t)=ϱ2˙(t)(ϱ1(t)g3(t))2ϱ¯(t),u_{1}(t)=\dfrac{\dot{\varrho_{1}}(t)}{\overline{\varrho}(t)},\quad u_{2}(t)=\dfrac{\dot{\varrho_{2}}(t)-(\varrho_{1}(t)-g_{3}(t))^{2}}{\overline{\varrho}(t)},

where ϱ¯(t)=g1(2t)((ϱ1(t)G)2+ϱ2(t)2)\overline{\varrho}(t)=g_{1}(2-t)((\varrho_{1}(t)-G)^{2}+\varrho_{2}(t)^{2}) for a.e. t[32,2]t\in[\frac{3}{2},2]. The control uu belongs to L([0,2],2)\mathrm{L}^{\infty}([0,2],\mathbb{R}^{2}) and xu=ϱx_{u}=\varrho along [32,2][\frac{3}{2},2]. Thus E(u)=xu(2)=ϱ(2)=x1\mathrm{E}(u)=x_{u}(2)=\varrho(2)=x^{1}. ∎

Let us prove that the converse of the geometric Pontryagin maximum principle is not true in general. Take a control uL([0,T],2)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{2}) which satisfies u1(t)=g2(t)u_{1}(t)=g_{2}(t) for a.e. t[0,1]t\in[0,1]. By Claims 2 and 4, we have u𝒰u\in\mathcal{U} and xu(2)x_{u}(2) belongs to the interior of the L2\mathrm{L}^{\infty}_{\mathbb{R}^{2}}-accessible set. Consider the constant function p:[0,2]2p:[0,2]\to\mathbb{R}^{2} defined by p(t)=(0,1)02p(t)=(0,1)\neq 0_{\mathbb{R}^{2}} for all t[0,2]t\in[0,2]. One can easily check that (xu,u,p)(x_{u},u,p) is a nontrivial strong 2\mathbb{R}^{2}-extremal lift of (xu,u)(x_{u},u) and thus uu is strongly 2\mathbb{R}^{2}-singular by Proposition 6.

We now prove that, given a control u𝒰L([0,T],U)u\in\mathcal{U}\cap\mathrm{L}^{\infty}([0,T],\mathrm{U}), the condition that xu(T)x_{u}(T) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set is not a sufficient condition for Property (4), even if U\mathrm{U} is convex. Take g2(t)=tg_{2}(t)=t for a.e. t[0,1]t\in[0,1] (which is not piecewise constant). Even if (G,0)(G,0) belongs to the interior of the LU\mathrm{L}^{\infty}_{\mathrm{U}}-accessible set (Claim 4), we easily infer from Claim 3 that (G,0)(G,0) is not PCU𝕋\mathrm{PC}^{\mathbb{T}}_{\mathrm{U}}-reachable in time TT from x0x^{0} for any partition 𝕋\mathbb{T} of [0,T][0,T]. Hence Property (Pu\mathrm{P}^{\prime}_{u}) is not satisfied, and neither is the stronger Property (4).

-F A general result on Ls\mathrm{L}^{s}-approximation by piecewise constant functions

Proposition 12.

Let 1s<+1\leq s<+\infty. Given any uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) and any ε>0\varepsilon>0, there exists a threshold δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exists vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that vuLsε\|v-u\|_{\mathrm{L}^{s}}\leq\varepsilon and vLuL\|v\|_{\mathrm{L}^{\infty}}\leq\|u\|_{\mathrm{L}^{\infty}}.

Proof.

Let uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) and ε>0\varepsilon>0. By the Lusin theorem [19], there exists a compact subset Kε[0,T]\mathrm{K}_{\varepsilon}\subset[0,T] such that (2uL)sμ([0,T]\Kε)εs/2(2\|u\|_{\mathrm{L}^{\infty}})^{s}\mu([0,T]\backslash\mathrm{K}_{\varepsilon})\leq\varepsilon^{s}/2, where μ\mu is the Lebesgue measure, and such that uu is continuous on Kε\mathrm{K}_{\varepsilon}. By uniform continuity of uu on Kε\mathrm{K}_{\varepsilon}, there exists δ>0\delta>0 such that u(ξ2)u(ξ1)mε(2T)1/s\|u(\xi_{2})-u(\xi_{1})\|_{\mathbb{R}^{m}}\leq\frac{\varepsilon}{(2T)^{1/s}} for all ξ1\xi_{1}, ξ2Kε\xi_{2}\in\mathrm{K}_{\varepsilon} satisfying |ξ2ξ1|δ|\xi_{2}-\xi_{1}|\leq\delta. Now, let 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} be a partition of [0,T][0,T] such that 𝕋δ\|\mathbb{T}\|\leq\delta. We set

I={i{0,,N1}μ(Kε[ti,ti+1)>0}.I=\{i\in\{0,\ldots,N-1\}\mid\mu(\mathrm{K}_{\varepsilon}\cap[t_{i},t_{i+1})>0\}.

For every iIi\in I, we consider some ξiKε[ti,ti+1)\xi_{i}\in\mathrm{K}_{\varepsilon}\cap[t_{i},t_{i+1}) such that u(ξi)Uu(\xi_{i})\in\mathrm{U} and u(ξi)muL\|u(\xi_{i})\|_{\mathbb{R}^{m}}\leq\|u\|_{\mathrm{L}^{\infty}}. We also consider some ωU\omega\in\mathrm{U} such that ωmuL\|\omega\|_{\mathbb{R}^{m}}\leq\|u\|_{\mathrm{L}^{\infty}}. We now define

v(t)={u(ξi)ift[ti,ti+1) with iI,ωift[ti,ti+1) with iI,v(t)=\left\{\begin{array}[]{lcl}u(\xi_{i})&\text{if}&t\in[t_{i},t_{i+1})\text{ with }i\in I,\\ \omega&\text{if}&t\in[t_{i},t_{i+1})\text{ with }i\notin I,\end{array}\right.

for every t[0,T]t\in[0,T]. In particular we have vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) and vLuL\|v\|_{\mathrm{L}^{\infty}}\leq\|u\|_{\mathrm{L}^{\infty}}. Finally we get that

vuLss\displaystyle\|v-u\|^{s}_{\mathrm{L}^{s}} =\displaystyle= [0,T]\Kεv(t)u(t)ms𝑑t\displaystyle\int_{[0,T]\backslash\mathrm{K}_{\varepsilon}}\|v(t)-u(t)\|_{\mathbb{R}^{m}}^{s}\;dt
+i=0N1Kε[ti,ti+1)v(t)u(t)ms𝑑t\displaystyle+\displaystyle\sum_{i=0}^{N-1}\int_{\mathrm{K}_{\varepsilon}\cap[t_{i},t_{i+1})}\|v(t)-u(t)\|_{\mathbb{R}^{m}}^{s}\;dt
\displaystyle\leq (2uL)sμ([0,T]\Kε)\displaystyle(2\|u\|_{\mathrm{L}^{\infty}})^{s}\mu([0,T]\backslash\mathrm{K}_{\varepsilon})
+iIKε[ti,ti+1)u(ξi)u(t)ms𝑑t\displaystyle+\displaystyle\sum_{i\in I}\int_{\mathrm{K}_{\varepsilon}\cap[t_{i},t_{i+1})}\|u(\xi_{i})-u(t)\|_{\mathbb{R}^{m}}^{s}\;dt
\displaystyle\leq εs2+εs2TiIμ(Kε[ti,ti+1))εs,\displaystyle\dfrac{\varepsilon^{s}}{2}+\dfrac{\varepsilon^{s}}{2T}\displaystyle\sum_{i\in I}\mu(\mathrm{K}_{\varepsilon}\cap[t_{i},t_{i+1}))\leq\varepsilon^{s},

which concludes the proof. ∎

Note that Proposition 12 is not true with s=+s=+\infty, as shown in the following Fuller-type example [10].

Example 10.

Take T=1T=1, m=1m=1 and U=\mathrm{U}=\mathbb{R}. Consider the oscillating function uL([0,T],U)u\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) defined by u(t)=1u(t)=1 for a.e. t(1k+1,1k]t\in(\frac{1}{k+1},\frac{1}{k}] for all even kk\in\mathbb{N}^{*} and u(t)=0u(t)=0 for a.e. t(1k+1,1k]t\in(\frac{1}{k+1},\frac{1}{k}] for all odd kk\in\mathbb{N}^{*}. We have vuL12\|v-u\|_{\mathrm{L}^{\infty}}\geq\frac{1}{2} for all vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) and all partitions 𝕋\mathbb{T} of [0,T][0,T].

Corollary 1.

Let 1s<+1\leq s<+\infty. Given any uLs([0,T],U)u\in\mathrm{L}^{s}([0,T],\mathrm{U}) and any ε>0\varepsilon>0, there exists a threshold δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exists vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that vuLsε\|v-u\|_{\mathrm{L}^{s}}\leq\varepsilon.

Proof.

Let uLs([0,T],U)u\in\mathrm{L}^{s}([0,T],\mathrm{U}) and ε>0\varepsilon>0. We fix some ωU\omega\in\mathrm{U} and we define Ck={t[0,T]u(t)mk}C_{k}=\{t\in[0,T]\mid\|u(t)\|_{\mathbb{R}^{m}}\geq k\} and

uk(t)={u(t)iftCk,ωiftCk,u_{k}(t)=\left\{\begin{array}[]{lcl}u(t)&\text{if}&t\notin C_{k},\\ \omega&\text{if}&t\in C_{k},\end{array}\right.

for a.e. t[0,T]t\in[0,T] and for every kk\in\mathbb{N}. In particular ukL([0,T],U)u_{k}\in\mathrm{L}^{\infty}([0,T],\mathrm{U}) for every kk\in\mathbb{N}. It is clear that (uk(t)u(t))k(u_{k}(t)-u(t))_{k\in\mathbb{N}} converges to 0m0_{\mathbb{R}^{m}} as k+k\to+\infty and that uk(t)u(t)mωm+u(t)m\|u_{k}(t)-u(t)\|_{\mathbb{R}^{m}}\leq\|\omega\|_{\mathbb{R}^{m}}+\|u(t)\|_{\mathbb{R}^{m}} for a.e. t[0,T]t\in[0,T]. By the Lebesgue dominated convergence theorem, we get that ukuLs0\|u_{k}-u\|_{\mathrm{L}^{s}}\rightarrow 0 as k+k\to+\infty. Hence, there exists kk\in\mathbb{N} such that ukuLsε2\|u_{k}-u\|_{\mathrm{L}^{s}}\leq\frac{\varepsilon}{2}. By Proposition 12, there exists δ>0\delta>0 such that, for any partition 𝕋\mathbb{T} of [0,T][0,T] satisfying 𝕋δ\|\mathbb{T}\|\leq\delta, there exists vPC𝕋([0,T],U)v\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) such that vukLsε2\|v-u_{k}\|_{\mathrm{L}^{s}}\leq\frac{\varepsilon}{2} and thus vuLsvukLs+ukuLsε\|v-u\|_{\mathrm{L}^{s}}\leq\|v-u_{k}\|_{\mathrm{L}^{s}}+\|u_{k}-u\|_{\mathrm{L}^{s}}\leq\varepsilon. ∎

-G Averaging operators

For any partition 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} of [0,T][0,T], we define the averaging operator 𝕋:L1([0,T],m)PC𝕋([0,T],m)\mathcal{I}^{\mathbb{T}}:\mathrm{L}^{1}([0,T],\mathbb{R}^{m})\to\mathrm{PC}^{\mathbb{T}}([0,T],\mathbb{R}^{m}) by

𝕋(u)(t)=1ti+1tititi+1u(ξ)𝑑ξ\mathcal{I}^{\mathbb{T}}(u)(t)=\dfrac{1}{t_{i+1}-t_{i}}\int_{t_{i}}^{t_{i+1}}u(\xi)\,d\xi (8)

for every t[ti,ti+1)t\in[t_{i},t_{i+1}), every i{0,,N1}i\in\{0,\ldots,N-1\} and every uL1([0,T],m)u\in\mathrm{L}^{1}([0,T],\mathbb{R}^{m}). The aim of this section is to establish several useful properties of the averaging operators.

Let 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} be a partition of [0,T][0,T]. The averaging operator 𝕋\mathcal{I}^{\mathbb{T}} is linear and projects any integrable function onto a piecewise constant function respecting the partition 𝕋\mathbb{T} (by averaging its value on each sampling interval [ti,ti+1)[t_{i},t_{i+1})). Furthermore we have

𝕋(u)(t)mu(t)m\|\mathcal{I}^{\mathbb{T}}(u)(t)\|_{\mathbb{R}^{m}}\leq\|u(t)\|_{\mathbb{R}^{m}} (9)

for a.e. t[0,T]t\in[0,T] and all uL1([0,T],m)u\in\mathrm{L}^{1}([0,T],\mathbb{R}^{m}).

Lemma 8.

Let 1s+1\leq s\leq+\infty. For any partition 𝕋\mathbb{T} of [0,T][0,T], we have 𝕋(u)LsuLs\|\mathcal{I}^{\mathbb{T}}(u)\|_{\mathrm{L}^{s}}\leq\|u\|_{\mathrm{L}^{s}} for all uLs([0,T],m)u\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m}).

Proof.

Let 𝕋={ti}i=0,,N\mathbb{T}=\{t_{i}\}_{i=0,\ldots,N} be a partition of [0,T][0,T] and let uLs([0,T],m)u\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m}). When s=+s=+\infty, the inequality 𝕋(u)LuL\|\mathcal{I}^{\mathbb{T}}(u)\|_{\mathrm{L}^{\infty}}\leq\|u\|_{\mathrm{L}^{\infty}} follows from (9). When 1s<+1\leq s<+\infty, we get from the Hölder inequality that

1ti+1tititi+1u(ξ)𝑑ξms1ti+1tititi+1u(ξ)ms𝑑ξ\left\|\dfrac{1}{t_{i+1}-t_{i}}\int_{t_{i}}^{t_{i+1}}u(\xi)d\xi\right\|_{\mathbb{R}^{m}}^{s}\!\!\!\leq\dfrac{1}{t_{i+1}-t_{i}}\int_{t_{i}}^{t_{i+1}}\|u(\xi)\|^{s}_{\mathbb{R}^{m}}d\xi

for all i{0,,N1}i\in\{0,\ldots,N-1\}, and thus 𝕋(u)LssuLss\|\mathcal{I}^{\mathbb{T}}(u)\|^{s}_{\mathrm{L}^{s}}\leq\|u\|^{s}_{\mathrm{L}^{s}}. ∎

The next lemma is instrumental in order to approximate with a Ls\mathrm{L}^{s}-norm (with any 1s<+1\leq s<+\infty) any control uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}) with piecewise constant controls.

Lemma 9.

Let 1s<+1\leq s<+\infty. Given any uLs([0,T],m)u\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m}), we have 𝕋(u)uLs0\|\mathcal{I}^{\mathbb{T}}(u)-u\|_{\mathrm{L}^{s}}\rightarrow 0 as 𝕋0\|\mathbb{T}\|\to 0.

Proof.

Let uLs([0,T],U)u\in\mathrm{L}^{s}([0,T],\mathrm{U}). Seeing (9) as a domination assumption and thanks to the Lebesgue dominated convergence theorem, we only need to prove that 𝕋(u)(τ)u(τ)\mathcal{I}^{\mathbb{T}}(u)(\tau)\rightarrow u(\tau) as 𝕋0\|\mathbb{T}\|\rightarrow 0 for a.e. τ[0,T]\tau\in[0,T]. For this purpose we set r(t)=0tu(ξ)𝑑ξr(t)=\int_{0}^{t}u(\xi)\,d\xi for every t[0,T]t\in[0,T] and let τ[0,T)\tau\in[0,T) being a Lebesgue point such that rr is derivable at τ\tau with r˙(τ)=u(τ)\dot{r}(\tau)=u(\tau). Given any ε>0\varepsilon>0, there exists δ>0\delta>0 such that

r(t)r(τ)tτu(τ)mε2\left\|\dfrac{r(t)-r(\tau)}{t-\tau}-u(\tau)\right\|_{\mathbb{R}^{m}}\leq\dfrac{\varepsilon}{2}

for every t[τδ,τ+δ][0,T]\{τ}t\in[\tau-\delta,\tau+\delta]\cap[0,T]\backslash\{\tau\}. Take 𝕋\mathbb{T} a partition of [0,T][0,T] such that 𝕋δ\|\mathbb{T}\|\leq\delta. There exists i{0,,N1}i\in\{0,\ldots,N-1\} such that τ[ti,ti+1)\tau\in[t_{i},t_{i+1}). Then

𝕋(u)(τ)u(τ)m=r(ti+1)r(ti)ti+1tiu(τ)mr(ti+1)r(τ)ti+1τu(τ)m|ti+1τti+1ti|+r(τ)r(ti)τtiu(τ)m|τtiti+1ti|ε,\|\mathcal{I}^{\mathbb{T}}(u)(\tau)-u(\tau)\|_{\mathbb{R}^{m}}=\left\|\dfrac{r(t_{i+1})-r(t_{i})}{t_{i+1}-t_{i}}-u(\tau)\right\|_{\mathbb{R}^{m}}\\ \leq\left\|\dfrac{r(t_{i+1})-r(\tau)}{t_{i+1}-\tau}-u(\tau)\right\|_{\mathbb{R}^{m}}\left|\dfrac{t_{i+1}-\tau}{t_{i+1}-t_{i}}\right|\\[3.0pt] +\left\|\dfrac{r(\tau)-r(t_{i})}{\tau-t_{i}}-u(\tau)\right\|_{\mathbb{R}^{m}}\left|\dfrac{\tau-t_{i}}{t_{i+1}-t_{i}}\right|\leq\varepsilon,

which concludes the proof. ∎

Our objective now is to prove that, when U\mathrm{U} is convex, the averaging operators project any integrable function with values in U\mathrm{U} onto a piecewise constant function with values in U\mathrm{U}.

Lemma 10.

Assume that U\mathrm{U} is convex. If uL1([0,1],U)u\in\mathrm{L}^{1}([0,1],\mathrm{U}), then 01u(ξ)𝑑ξU\int_{0}^{1}u(\xi)d\xi\in\mathrm{U}.

Proof.

Let uL1([0,1],U)u\in\mathrm{L}^{1}([0,1],\mathrm{U}) and let us prove that u~U\widetilde{u}\in\mathrm{U} where u~\widetilde{u} is defined by u~=01u(ξ)𝑑ξ\widetilde{u}=\int_{0}^{1}u(\xi)d\xi. We first give a simpler argument when U\mathrm{U} is furthermore assumed to be closed. In that context, by the Hilbert projection theorem, we have

u~projU(u~),u(ξ)projU(u~)m0\langle\widetilde{u}-\mathrm{proj}_{\mathrm{U}}(\widetilde{u}),u(\xi)-\mathrm{proj}_{\mathrm{U}}(\widetilde{u})\rangle_{\mathbb{R}^{m}}\leq 0

for a.e. ξ[0,1]\xi\in[0,1], where projU(u~)U\mathrm{proj}_{\mathrm{U}}(\widetilde{u})\in\mathrm{U} is the projection of u~\widetilde{u} onto U\mathrm{U}. Integrating the above inequality over [0,1][0,1] yields u~projU(u~)m20\|\widetilde{u}-\mathrm{proj}_{\mathrm{U}}(\widetilde{u})\|_{\mathbb{R}^{m}}^{2}\leq 0 and thus u~=projU(u~)U\widetilde{u}=\mathrm{proj}_{\mathrm{U}}(\widetilde{u})\in\mathrm{U}.

Now we remove the closedness assumption made on U\mathrm{U}. Let us prove that u~U\widetilde{u}\in\mathrm{U} by strong induction on the dimension dd\in\mathbb{N} of the nonempty convex set U\mathrm{U}. If d=0d=0, the set U\mathrm{U} is reduced to a singleton and the result is trivial. Now consider that d1d\geq 1 and assume that the result is true at all steps from 0 to d1d-1. By contradiction assume that u~U\widetilde{u}\notin\mathrm{U}. By separation, there exists ψm\{0m}\psi\in\mathbb{R}^{m}\backslash\{0_{\mathbb{R}^{m}}\} such that ψ,ωu~m0\langle\psi,\omega-\widetilde{u}\rangle_{\mathbb{R}^{m}}\leq 0 for all ωU\omega\in\mathrm{U}. We infer that the null integral 01ψ,u(ξ)u~m𝑑ξ\int_{0}^{1}\langle\psi,u(\xi)-\widetilde{u}\rangle_{\mathbb{R}^{m}}d\xi has a nonpositive integrand. Thus this integrand is zero almost everywhere on [0,1][0,1]. Therefore uu is with values in the convex set U(u~+ψ)\mathrm{U}\cap(\widetilde{u}+\psi^{\bot}), where ψ\psi^{\bot} stands for the standard hyperplane defined by orthogonality with the nonzero vector ψ\psi. Since U(u~+ψ)\mathrm{U}\cap(\widetilde{u}+\psi^{\bot}) is a nonempty convex set of dimension strictly inferior than dd, thanks to our induction hypothesis we get that u~U(u~+ψ)\widetilde{u}\in\mathrm{U}\cap(\widetilde{u}+\psi^{\bot}), 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 U\mathrm{U} is convex. If uL1([0,T],U)u\in\mathrm{L}^{1}([0,T],\mathrm{U}), then 𝕋(u)PC𝕋([0,T],U)\mathcal{I}^{\mathbb{T}}(u)\in\mathrm{PC}^{\mathbb{T}}([0,T],\mathrm{U}) for any partition 𝕋\mathbb{T} of [0,T][0,T].

-H Truncated end-point mapping and Ls\mathrm{L}^{s}-differential

For every M>0M>0, we fix a mapping ΛM:n×m\Lambda^{M}:\mathbb{R}^{n}\times\mathbb{R}^{m}\to\mathbb{R} of class C1\mathrm{C}^{1} satisfying

ΛM(x,u)={1 if (x,u)B¯n(0,2M)×B¯m(0,2M),0 if (x,u)Bn(0,3M)×Bm(0,3M).\Lambda^{M}(x,u)=\left\{\begin{array}[]{l}1\text{ if }(x,u)\in\overline{\mathrm{B}}_{\mathbb{R}^{n}}(0,2M)\times\overline{\mathrm{B}}_{\mathbb{R}^{m}}(0,2M),\\[3.0pt] 0\text{ if }(x,u)\notin\mathrm{B}_{\mathbb{R}^{n}}(0,3M)\times\mathrm{B}_{\mathbb{R}^{m}}(0,3M).\end{array}\right.

Let M>0M>0. When replacing the dynamics ff in the control system (CS) by the truncated dynamics fMf^{M}, defined by fM(x,u,t)=ΛM(x,u)f(x,u,t)f^{M}(x,u,t)=\Lambda^{M}(x,u)f(x,u,t) for all (x,u,t)n×m×[0,T](x,u,t)\in\mathbb{R}^{n}\times\mathbb{R}^{m}\times[0,T] we obtain a new control system that we denote by (CSM\mathrm{CS}^{M}). The main difference is that, for any control uL1([0,T],m)u\in\mathrm{L}^{1}([0,T],\mathbb{R}^{m}) (even unbounded), there exists a trajectory xAC([0,T],n)x\in\mathrm{AC}([0,T],\mathbb{R}^{n}), starting at x(0)=x0x(0)=x^{0}, such that x˙(t)=fM(x(t),u(t),t)\dot{x}(t)=f^{M}(x(t),u(t),t) for a.e. t[0,T]t\in[0,T]. In that case the trajectory xx is unique and will be denoted by xuMx^{M}_{u}. We now introduce, for any 1s+1\leq s\leq+\infty, the truncated end-point mapping EM:Ls([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} defined by EM(u)=xuM(T)\mathrm{E}^{M}(u)=x^{M}_{u}(T) for all uLs([0,T],m)u\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m}). Note that the next proposition, derived from standard techniques in ordinary differential equations theory, is true for any 1<s+1<s\leq+\infty. The case s=1s=1 is discussed in Remark 22.

Proposition 14.

Let 1<s+1<s\leq+\infty and M>0M>0. The truncated end-point mapping EM:Ls([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} is of class C1\mathrm{C}^{1} and its Fréchet differential is given by

DEM(u)v=wvu,M(T)\mathrm{D}\mathrm{E}^{M}(u)\cdot v=w^{u,M}_{v}(T) (10)

for all uu, vLs([0,T],m)v\in\mathrm{L}^{s}([0,T],\mathbb{R}^{m}), where wvu,MAC([0,T],n)w^{u,M}_{v}\in\mathrm{AC}([0,T],\mathbb{R}^{n}) is the unique solution to

{w˙(t)=xfM(xuM(t),u(t),t)w(t)+ufM(xuM(t),u(t),t)v(t),a.e. t[0,T],w(0)=0n.\left\{\begin{array}[]{l}\dot{w}(t)=\nabla_{x}f^{M}(x^{M}_{u}(t),u(t),t)w(t)\\ \qquad\qquad+\nabla_{u}f^{M}(x^{M}_{u}(t),u(t),t)v(t),\quad\text{a.e.\ }t\in[0,T],\\ w(0)=0_{\mathbb{R}^{n}}.\end{array}\right.
Remark 21.

Let 1<s+1<s\leq+\infty. For a given control u𝒰u\in\mathcal{U}, note that DE(u)\mathrm{D}\mathrm{E}(u) given in (1) admits a natural extension (still denoted by) DE(u):Ls([0,T],m)n\mathrm{D}\mathrm{E}(u):\mathrm{L}^{s}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n}. The nontruncated setting is related to the truncated one as follows:

  1. (i)

    Let u𝒰u\in\mathcal{U} and M>0M>0 be such that xuCM\|x_{u}\|_{\mathrm{C}}\leq M and uLM\|u\|_{\mathrm{L}^{\infty}}\leq M. Then xuM=xux^{M}_{u}=x_{u} and DEM(u)=DE(u)\mathrm{D}\mathrm{E}^{M}(u)=\mathrm{D}\mathrm{E}(u) when considering the above extension of DE(u)\mathrm{D}\mathrm{E}(u).

  2. (ii)

    Let uL([0,T],m)u\in\mathrm{L}^{\infty}([0,T],\mathbb{R}^{m}). If there exists M>0M>0 such that xuMCM\|x^{M}_{u}\|_{\mathrm{C}}\leq M and uLM\|u\|_{\mathrm{L}^{\infty}}\leq M, then u𝒰u\in\mathcal{U} and xu=xuMx_{u}=x^{M}_{u}.

Remark 22.

Let M>0M>0. In the case s=1s=1, it can be proved that the truncated end-point mapping EM:L1([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{1}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} is Gateaux-differentiable and its Gateaux differential is given by (10). However it is not Fréchet-differentiable (and thus not of class C1\mathrm{C}^{1}) in general, as shown in the next example.

Example 11.

Take T=n=m=1T=n=m=1, U=\mathrm{U}=\mathbb{R} and f(x,u,t)=u2f(x,u,t)=u^{2} for all (x,u,t)××[0,T](x,u,t)\in\mathbb{R}\times\mathbb{R}\times[0,T]. Consider the starting point x0=0x^{0}=0 and the constant control u0u\equiv 0. In that context, with M=s=1M=s=1, it is clear that xuM0x^{M}_{u}\equiv 0 and that the Gateaux differential DGEM(u):L1([0,T],m)n\mathrm{D}^{\mathrm{G}}\mathrm{E}^{M}(u):\mathrm{L}^{1}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} of the truncated end-point mapping EM:L1([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{1}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} at uu, given by the expression (10), is null. Now, taking the needle-like variation u(0,1)αu^{\alpha}_{(0,1)}, as defined in (2), associated with the pair (0,1)(fu)×U(0,1)\in\mathcal{L}(f_{u})\times\mathrm{U}, we obtain that

limα0+EM(u+u(0,1)α)EM(u)DGEM(u)u(0,1)αu(0,1)αL1=1.\lim\limits_{\alpha\to 0^{+}}\dfrac{\mathrm{E}^{M}(u+u^{\alpha}_{(0,1)})-\mathrm{E}^{M}(u)-\mathrm{D}^{\mathrm{G}}\mathrm{E}^{M}(u)\cdot u^{\alpha}_{(0,1)}}{\|u^{\alpha}_{(0,1)}\|_{\mathrm{L}^{1}}}=1.

Therefore EM:L1([0,T],m)n\mathrm{E}^{M}:\mathrm{L}^{1}([0,T],\mathbb{R}^{m})\rightarrow\mathbb{R}^{n} is not Fréchet-differentiable at uu.

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.