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

Differential Flatness by Pure Prolongation: Necessary and Sufficient Conditions

Jean Lévine Centre Automatique et Systèmes (CAS), MINES Paris - PSL, 75006 Paris, France, e-mail: [email protected]
(July 2023)
Abstract

In this article, we introduce the notion of differential flatness by pure prolongation: loosely speaking, a system admits this property if, and only if, there exists a pure prolongation of finite order such that the prolonged system is feedback linearizable. We obtain Lie-algebraic necessary and sufficient conditions for a general nonlinear multi-input system to satisfy this property. These conditions are comprised of the involutivity and relative invariance of a pair of filtrations of distributions of vector fields. An algorithm computing the minimal prolongation lengths of the input channels that achieve the system linearization, yielding the associated flat outputs, is deduced. Examples that show the efficiency and computational tractability of the approach are then presented.

Keywords— differential flatness; Lie-Bäcklund isomorphism; Lie brackets; distributions of vector fields; prolongation of vector fields; dynamic feedback linearization.

1 Introduction

The notions of static feedback linearization [17, 14] and dynamic feedback linearization of a nonlinear system, whose preliminary results were reported in [6, 7, 32, 1] (see also [33, 34, 11, 2, 12]), were at the origin of a so far uninterrupted thread of studies in nonlinear system theory. In particular, during the last three decades, they gave birth to the concept of differential flatness, that plays a prominent part in motion planning and trajectory tracking problems and their applications (see [24, 9, 10] and [21] for a thorough presentation).

Sufficient or necessary conditions in special cases, as well as a general differential flatness characterization, although without an upper bound on the number of iterations of the corresponding algorithm, have been obtained (see e.g. [21, 22] for a historical review). Nevertheless, the question of obtaining computationally tractable general necessary and sufficient conditions for dynamic feedback linearization as well as for differential flatness, remains open. In this paper, we restrict our study to the class of differentially flat systems by pure prolongation, i.e. , roughly speaking, the class of nn-dimensional systems with mm inputs

x˙=f(x,u1,,um)\dot{x}=f(x,u_{1},\ldots,u_{m})

for which there exists a multi-index 𝐣(j1,,jm)\mathbf{j}\triangleq(j_{1},\ldots,j_{m}) such that the prolonged system

x˙=f(x,u1,,um),\displaystyle\dot{x}=f(x,u_{1},\ldots,u_{m}), (1.1)
ui(ji+1)=vi,i=1,,m,\displaystyle u_{i}^{(j_{i}+1)}=v_{i},\quad i=1,\ldots,m,

of dimension n+i=1mjin+\sum_{i=1}^{m}j_{i}, where we have denoted by ui(ji)u_{i}^{(j_{i})} the jij_{i}-th order time derivative of uiu_{i}, is locally static feedback linearizable.

The corresponding transformation is called pure dynamic extension in [33, 34] in the context of finding upper bounds on the indices j1,,jmj_{1},\ldots,j_{m}, if they exist. We prefer using here the word prolongation initially introduced in [7], in reference to É. Cartan [5] (see also [32]).

This special class of differentially flat systems, initially introduced in the framework of dynamic feedback linearization [6, 7, 32, 1], has also been studied in [11, 12], under the name of linearization by prolongation, again in terms of upper bounds of prolongation orders, and in [2] where an algorithm, yielding a sufficient condition, is obtained.

However, the necessary and sufficient conditions presented in this paper, tackled via the here introduced concept of equivalence by pure prolongation, are original to the author’s knowledge. Moreover, they are easily checkable by an algorithm that only involves Lie bracket and linear algebra operations, yielding the minimal pure prolongations if they exist.

Our main contributions are:

  • 1.

    The introduction of the equivalence relation of systems by pure prolongation, strictly coarser than the well-known equivalence by diffeomorphism and feedback (see e.g. [17, 15, 16, 28, 20]) and strictly finer than the Lie-Bäcklund equivalence (see e.g. [10, 21]). For this equivalence relation, the set of flat systems by pure prolongation is identified with the equivalence class of 0 (modulo the trivial vector field)(see definition 4.2);

  • 2.

    Proposition 3.2, extending and specifying results implicitly present in various forms in [7, 11, 2, 12], where we prove that, whatever the prolongation 𝐣(j1,,jm)\mathbf{j}\triangleq(j_{1},\ldots,j_{m}), the filtration of prolonged distributions, built on the successive Lie brackets of the prolonged drift with the prolonged control vector fields, is decomposable into the direct sum of two filtrations of distributions. The first one, denoted by {Δk(𝐣)}k0\{\Delta_{k}^{(\mathbf{j})}\}_{k\geq 0}, is, at each point of the prolonged manifold, generated by vector fields belonging to the original tangent bundle, of dimension n+mn+m (in a suitably defined vertical bundle, see (3.13)); The second one, denoted by {Γk(𝐣)}k0\{\Gamma_{k}^{(\mathbf{j})}\}_{k\geq 0}, is generated by the sequence of prolonged control vector fields of decreasing orders up to 1, i.e. Γk(𝐣)=i=1,,m{ui(jir)r=0,,min(ji1,k)}\Gamma_{k}^{(\mathbf{j})}=\bigoplus_{i=1,\ldots,m}\{\frac{\partial}{\partial u_{i}^{(j_{i}-r)}}\mid r=0,\ldots,\min(j_{i}-1,k)\} (see (3.12));

  • 3.

    Theorem 4.1 giving the necessary and sufficient conditions: (i) Δk(𝐣)\Delta_{k}^{(\mathbf{j})} must be involutive with locally constant dimension and (ii) invariant by Γk(𝐣)\Gamma_{k}^{(\mathbf{j})} for all kk, and (iii) dimΔk(𝐣)\dim\Delta_{k}^{(\mathbf{j})} must be equal to n+mn+m for all kk large enough (strong controllability);

  • 4.

    Formula (4.11) and theorem 4.2 giving the prolongation lengths, if ithey exist, or otherwise a criterion of non flatness by pure prolongation;

  • 5.

    Algorithm 4.2 whose input comprises the system vector fields and whose output is either the list of minimal prolongation lengths or the claim that the system is not flat by pure prolongation, deduced in a finite number of steps.

The paper is organized as follows: In section 2, we present the necessary recalls of basic results on differential flatness and feedback linearization. Then we introduce and study the purely prolonged distributions and the associated vertical ones in section 3. The equivalence by pure prolongation and the definition of flatness by pure prolongation, followed by its necessary and sufficient conditions and by the pure prolongation algorithm are then presented in section 4. A series of five examples, four with two inputs (sections 5.1, 5.2, 5.3 and 5.4) and one with three inputs (section 5.5) then illustrate our results in section 5. Moreover, the pendulum example (section 5.4), is proven to be non flat by pure prolongation, though known to be differentially flat [10, 21], thus proving that the set of flat systems by pure prolongation is strictly contained in the set of differentially flat ones. The paper ends with concluding remarks and an appendix establishing a comparison formula between prolonged and non prolonged Lie brackets.

2 Recalls on Differential Flatness and Feedback Linearization

Consider a non-linear system over a smooth nn-dimensional manifold XX given by

x˙=f(x,u)\dot{x}=f(x,u) (2.1)

where xx is the nn-dimensional state vector, umu\in\mathbb{R}^{m} the input or control vector, with mnm\leq n, and ff a CC^{\infty} vector field in the tangent bundle TX\mathrm{T}X of XX for each umu\in\mathbb{R}^{m} and whose dependence on uu is of class CC^{\infty}.

Before going further, we need the following notations, valid all along this paper:

  • Boldface letters 𝐣,𝐤,,\mathbf{j},\mathbf{k},\ldots, are systematically used to denote multi-integers, i.e.

    𝐣(j1,j2,,jm),𝐤(k1,k2,,km),\mathbf{j}\triangleq(j_{1},j_{2},\ldots,j_{m}),\quad\mathbf{k}\triangleq(k_{1},k_{2},\ldots,k_{m}),\quad\ldots
  • We denote the minimum of two arbitrary integers kk and ll by klmin{k,l}k\lor l\triangleq\min\{k,l\}, and their maximum by klmax(k,l)k\wedge l\triangleq\max(k,l).

    Also, for every 𝐣(j1,,jm)m\mathbf{j}\triangleq\left(j_{1},\ldots,j_{m}\right)\in\mathbb{N}^{m} and 𝐤(k1,,km)m\mathbf{k}\triangleq\left(k_{1},\ldots,k_{m}\right)\in\mathbb{N}^{m}, we use the componentwise minimum notation

    𝐣𝐤(j1k1,,jmkm)=(min(j1,k1),,min(jm,km)).\mathbf{j}\bigvee\mathbf{k}\triangleq\left(j_{1}\vee k_{1},\ldots,j_{m}\vee k_{m}\right)=\left(\min(j_{1},k_{1}),\ldots,\min(j_{m},k_{m})\right). (2.2)

    Accordingly, the componentwise maximum is denoted by

    𝐣𝐤(j1k1,,jmkm)=(max(j1,k1),,max(jm,km)).\mathbf{j}\bigwedge\mathbf{k}\triangleq\left(j_{1}\wedge k_{1},\ldots,j_{m}\wedge k_{m}\right)=\left(\max(j_{1},k_{1}),\ldots,\max(j_{m},k_{m})\right). (2.3)

    If k1==km=k k_{1}=\cdots=k_{m}=k \in\mathbb{N}, we also use the notations 𝐣k\mathbf{j}\bigwedge k and 𝐣k\mathbf{j}\bigvee k, with kk in regular math typeface.

  • Overlined symbols will denote a collection of successive time derivatives of a time-dependent function as follows. Given a multi-integer 𝐤=(k1,,km)m\mathbf{k}=(k_{1},\ldots,k_{m})\in\mathbb{N}^{m} and a locally defined CC^{\infty} function tξ(t)mt\mapsto\xi(t)\in\mathbb{R}^{m},

    • ξ¯(𝐤)\overline{\xi}^{(\mathbf{k})} denotes the vector (ξ1,ξ˙1,,ξ1(k1),,ξm,ξ˙m,,ξm(km))\left(\xi_{1},\dot{\xi}_{1},\ldots,\xi_{1}^{(k_{1})},\ldots,\xi_{m},\dot{\xi}_{m},\ldots,\xi_{m}^{(k_{m})}\right) of dimension m+|𝐤|m+|\mathbf{k}|, with |𝐤|i=1mki|\mathbf{k}|\triangleq\sum_{i=1}^{m}k_{i} and ξi(j)djξidtj\xi_{i}^{(j)}\triangleq\frac{d^{j}\xi_{i}}{dt^{j}}, j=1,,kij=1,\ldots,k_{i}, i=1,,mi=1,\ldots,m;

    • ξ¯\overline{\xi} denotes the infinite sequence

      ξ¯(ξ,ξ˙,ξ¨,)(ξi(k);i=1,,m;k0)m,\overline{\xi}\triangleq(\xi,\dot{\xi},\ddot{\xi},\ldots)\triangleq\left(\xi_{i}^{(k)}\;;\;i=1,\dots,m\;;k\geq 0\right)\in\mathbb{R}^{m}_{\infty},

      where mm×m×\mathbb{R}^{m}_{\infty}\triangleq\mathbb{R}^{m}\times\mathbb{R}^{m}\times\cdots is the product of an infinite number of copies of m\mathbb{R}^{m}.

2.1 Recalls on Lie-Bäcklund Equivalence

For a detailed presentation of the topics of this section, the reader may refer e.g. to [10, 29, 21, 22, 30].

Definition 2.1.

The infinite order jet space prolongation of system (2.1) is given by the pair (X×m,Cf)(X\times\mathbb{R}^{m}_{\infty},C_{f}), where X×mX\times\mathbb{R}^{m}_{\infty} is the product of XX with an infinite number of copies of m\mathbb{R}^{m}, with coordinates (x,u¯)(x,\overline{u}), endowed with the Cartan field

Cf=f(x,u)x+j0i=1mui(j+1)ui(j),C_{f}=f(x,u)\frac{\partial}{\partial x}+\sum_{j\geq 0}\sum_{i=1}^{m}u_{i}^{(j+1)}\frac{\partial}{\partial u_{i}^{(j)}}, (2.4)

defined on TX×Tm\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m}_{\infty}, the tangent bundle of X×mX\times\mathbb{R}^{m}_{\infty}.

Definition 2.2 (Lie-Bäcklund equivalence).

Consider two systems:

x˙=g(x,u)andy˙=γ(y,v)\dot{x}=g(x,u)\quad\mathrm{and}\quad\dot{y}=\gamma(y,v) (2.5)

and their respective prolongations (X×m,Cg)(X\times\mathbb{R}^{m}_{\infty},C_{g}), with coordinates (x,u¯)(x,\overline{u}) and Cartan field

Cg=g(x,u)x+j0i=1mui(j+1)ui(j),C_{g}=g(x,u)\frac{\partial}{\partial x}+\sum_{j\geq 0}\sum_{i=1}^{m}u_{i}^{(j+1)}\frac{\partial}{\partial u_{i}^{(j)}}, (2.6)

and (Y×μ,Cγ)(Y\times\mathbb{R}^{\mu}_{\infty},C_{\gamma}), with coordinates (y,v¯)(y,\overline{v}), and Cartan field

Cγ=γ(y,v)y+j0i=1μvi(j+1)vi(j).C_{\gamma}=\gamma(y,v)\frac{\partial}{\partial y}+\sum_{j\geq 0}\sum_{i=1}^{\mu}v_{i}^{(j+1)}\frac{\partial}{\partial v_{i}^{(j)}}. (2.7)

We say that they are Lie-Bäcklund equivalent at a pair of points (x0,u¯0)(x_{0},\overline{u}_{0}) and (y0,v¯0)(y_{0},\overline{v}_{0}) if, and only if, there exists neighborhoods 𝒩x0,u¯0X×m{\mathcal{N}}_{x_{0},\overline{u}_{0}}\subset X\times\mathbb{R}^{m}_{\infty} and 𝒩y0,v¯0Y×μ{\mathcal{N}}_{y_{0},\overline{v}_{0}}\subset Y\times\mathbb{R}^{\mu}_{\infty} and a CC^{\infty} isomorphism111Recall that a continuous function and, a fortiori, differentiable, resp. CC^{\infty}, depends, by definition of the source and target product topologies, on a finite number of components of its variables, namely Φ\Phi (resp. Ψ\Psi) depends an a finite number of components of (y,v¯)(y,\overline{v}) (resp. (x,u¯)(x,\overline{u})) (see e.g. [19, 36, 21]). Φ:𝒩y0,v¯0𝒩x0,u¯0\Phi:{\mathcal{N}}_{y_{0},\overline{v}_{0}}\rightarrow{\mathcal{N}}_{x_{0},\overline{u}_{0}} satisfying Φ(y0,v¯0)=(x0,u¯0)\Phi(y_{0},\overline{v}_{0})=(x_{0},\overline{u}_{0}), with CC^{\infty} inverse Ψ\Psi, such that the respective Cartan fields are Φ\Phi and Ψ\Psi related, i.e. ΦCγ=Cg\Phi_{\ast}C_{\gamma}=C_{g} in 𝒩x0,u¯0{\mathcal{N}}_{x_{0},\overline{u}_{0}} and ΨCg=Cγ\Psi_{\ast}C_{g}=C_{\gamma} in 𝒩y0,v¯0{\mathcal{N}}_{y_{0},\overline{v}_{0}}.

In other words, the two systems are Lie-Bäcklund equivalent at the points (x0,u¯0)(x_{0},\overline{u}_{0}) and (y0,v¯0)(y_{0},\overline{v}_{0}) if there exist neighborhoods of these points where every integral curve of the first system is mapped to an integral curve of the second one and conversely, with the same time parameterization. Clearly, this relation is an equivalence relation.

We recall, without proof, a most important result from [24] (see also [9, 10, 21]) giving an interpretation of the Lie-Bäcklund equivalence in terms of diffeomorphism in finite dimension and endogenous dynamic feedback, that will be useful later on.

Theorem 2.1 (Martin [24]).

If the two systems (2.5) are Lie-Bäcklund equivalent at a given pair of points, then (i) and (ii) must be satisfied:

  • (i)

    m=μm=\mu, i.e. they must have the same number of independent inputs;

  • (ii)

    there exist

    • an endogenous dynamic feedback222A dynamic feedback is said endogenous if, and only if, the closed-loop system and the original one are Lie-Bäcklund equivalent, i.e. if, and only if, the extended state zz can be locally expressed as a smooth function of xx, uu and a finite number of time derivatives of uu (see [24, 9, 10, 21]).

      u=α(x,z,w),z˙=β(x,z,w),u=\alpha(x,z,w),\quad\dot{z}=\beta(x,z,w), (2.8)

      where zz belongs to ZZ, a finite dimensional smooth manifold,

    • a multi-integer333Recall that we denote by v(𝐫)(v1(r1),,vm(rm))(dr1v1dtr1,,drmvmdtrm)v^{(\mathbf{r})}\triangleq\left(v_{1}^{(r_{1})},\ldots,v_{m}^{(r_{m})}\right)\triangleq\left(\frac{\mathrm{d}^{r_{1}}v_{1}}{\mathrm{d}t^{r_{1}}},\ldots,\frac{\mathrm{d}^{r_{m}}v_{m}}{\mathrm{d}t^{r_{m}}}\right). 𝐫(r1,,rm)\mathbf{r}\triangleq\left(r_{1},\ldots,r_{m}\right),

    • and a local diffeomorphism χ:Y×|𝐫|X×Z\chi:Y\times\mathbb{R}^{|\mathbf{r}|}\rightarrow X\times Z,

    all defined in a neighborhood of the considered points, such that the extended system

    y˙=γ(y,v),v(𝐫)=w\dot{y}=\gamma(y,v),\quad v^{(\mathbf{r})}=w (2.9)

    and the closed-loop one

    x˙=g(x,α(x,z,w)),z˙=β(x,z,w)\dot{x}=g(x,\alpha(x,z,w)),\quad\dot{z}=\beta(x,z,w) (2.10)

    are χ\chi-related for all wmw\in\mathbb{R}^{m}, i.e.

    (x,z)=χ(y,v,v˙,,v(𝐫𝟏)),(y,v,v˙,,v(𝐫𝟏))=χ1(x,z)(x,z)=\chi(y,v,\dot{v},\ldots,v^{(\mathbf{r-1})}),\qquad(y,v,\dot{v},\ldots,v^{(\mathbf{r-1})})=\chi^{-1}(x,z) (2.11)

    and

    g^=χγ^,γ^=χ1g^\hat{g}=\chi_{\ast}\hat{\gamma},\qquad\hat{\gamma}=\chi^{-1}_{\ast}\hat{g} (2.12)

    where we have denoted

    g^(x,z,w)g(x,α(x,z,w))x+β(x,z,w)z\displaystyle\hat{g}(x,z,w)\triangleq g(x,\alpha(x,z,w))\frac{\partial}{\partial x}+\beta(x,z,w)\frac{\partial}{\partial z}
    γ^(y,v,v˙,,v(𝐫𝟏),w)γ(y,v)y+i=1mj=0ri1vi(j+1)vi(j)+wivi(ri).\displaystyle\hat{\gamma}(y,v,\dot{v},\ldots,v^{(\mathbf{r-1})},w)\triangleq\gamma(y,v)\frac{\partial}{\partial y}+\sum_{i=1}^{m}\sum_{j=0}^{r_{i}-1}v_{i}^{(j+1)}\frac{\partial}{\partial v_{i}^{(j)}}+w_{i}\frac{\partial}{\partial v_{i}^{(r_{i})}}.

2.2 Recalls on Differential Flatness

Definition 2.3.

We say that system (2.1) is differentially flat (or, more shortly, flat) at the pair of points (x0,u¯0)(x_{0},\overline{u}_{0}) and y¯0m\overline{y}_{0}\in\mathbb{R}^{m}_{\infty} if and only if, it is Lie-Bäcklund equivalent to the trivial system (m,τ)\mathbb{R}^{m}_{\infty},\tau) where τ\tau is the trivial Cartan field

τj0i=1myi(j+1)yi(j)\tau\triangleq\sum_{j\geq 0}\sum_{i=1}^{m}y_{i}^{(j+1)}\frac{\partial}{\partial y_{i}^{(j)}} (2.13)

at the considered points.

Otherwise stated, the locally defined flat output y=Ψ(x,u¯)y=\Psi(x,\overline{u}) is such that (x,u¯)=Φ(y¯)(Φ1(y¯),Φ0(y¯),Φ1(y¯),)(x,\overline{u})=\Phi(\overline{y})\triangleq(\Phi_{-1}(\overline{y}),\Phi_{0}(\overline{y}),\Phi_{1}(\overline{y}),\ldots), with444The first component of Φ\Phi, corresponding to the xx component, is denoted by Φ1\Phi_{-1} so that the component Φi\Phi_{i} effectively corresponds to the iith time derivative of uu. ddtΦ1(y¯)f(Φ1(y¯),Φ0(y¯))\frac{\mathrm{d}}{\mathrm{d}t}{\Phi}_{-1}(\overline{y})\equiv f(\Phi_{-1}(\overline{y}),\Phi_{0}(\overline{y})) for all sufficiently differentiable function y:ty(t)my:t\in\mathbb{R}\mapsto y(t)\in\mathbb{R}^{m}.

This definition immediately implies that a system is flat if, and only if, there exists a generalized output y=Ψ(x,u¯)y=\Psi(x,\overline{u}) of dimension mm, depending at most on a finite number of derivatives of uu, with independent derivatives of all orders, such that xx and u¯\overline{u} can be expressed in terms of yy and a finite number of its successive derivatives, i.e. (x,u¯)=Φ(y¯)(x,\overline{u})=\Phi(\overline{y}), and such that the system equation ddtΦ1(y¯)=fΦ(y¯)\frac{\mathrm{d}}{\mathrm{d}t}{\Phi_{-1}}(\overline{y})=f\circ\Phi(\overline{y}) is identically satisfied for all sufficiently differentiable y:my:\mathbb{R}\rightarrow\mathbb{R}^{m}.

For a flat system, with the notations of theorem 2.1, the vector field γ\gamma, or γ^\widehat{\gamma} indifferently, corresponds to the linear system in Brunovský canonical form

yi(ri+1)=wi,i=1,,m,y_{i}^{(r_{i}+1)}=w_{i},\qquad i=1,\ldots,m, (2.14)

and CγC_{\gamma}, defined by (2.7), is the trivial Cartan field Cγ=τC_{\gamma}=\tau, with τ\tau given by (2.13).

Remark 2.1.

In view of Definition 2.2, since the trivial Cartan field (2.13) is the infinite jet prolongation of y˙i=ui\dot{y}_{i}=u_{i}, i=1,,mi=1,\ldots,m, system (2.1) is flat if, and only if, it is Lie-Bäcklund equivalent to mm simple integrators in parallel. We may also remark that, since the Cartan field is equal to 0 (modulo τ)\tau), the class of flat systems may be identified with the Lie-Bäcklund equivalence class of 0 (modulo τ\tau).

Theorem 2.1 reads:

Corollary 2.1.

If system (2.1) is flat at a given point, there exists an endogenous dynamic feedback of the form (2.8), a multi-integer 𝐫(r1,,rm)\mathbf{r}\triangleq\left(r_{1},\ldots,r_{m}\right) and a finite dimensional local diffeomorphism χ\chi such that systems (2.10), with ff in place of gg, and (2.14) are χ\chi-related for all wmw\in\mathbb{R}^{m}.

Consequently, a flat system can be transformed, by diffeomorphism and feedback of a suitably extended space, in a linear controllable system, which motivates the recalls of the next section.

2.3 Recalls on Feedback Linearization

Feedback linearizable systems [17, 15] (see also [16, 28, 23, 20]) constitute a subclass of differentially flat systems. They correspond to the equivalence class of linear controllable systems with respect to the following finer equivalence relation, called equivalence by diffeomorphism and feedback.

Definition 2.4.

The two systems given by (2.5) are said equivalent by diffeomorphism and feedback if, and only if, there exists

  • a local diffeomorphism φ\varphi from a neighborhood of an equilibrium point of XX (which may be chosen, without loss of generality, as the origin 0X0\in X) to a suitable neighborhood of 0Y0\in Y,

  • and a static feedback v=α(x,u)v=\alpha(x,u), α\alpha being invertible with respect to uu for all xx in the above mentioned neighborhood of the origin, i.e. 𝗋𝖺𝗇𝗄(αu)(x,u)=m\mathsf{rank}\left(\frac{\partial\alpha}{\partial u}\right)(x,u)=m for all xx and uu as above,

such that φ1(f(x,u))=γ(φ×α)(x,u)=γ(y,v)\varphi^{-1}_{\ast}(f(x,u))=\gamma\circ(\varphi\times\alpha)(x,u)=\gamma(y,v),

Indeed, this equivalence implies that n=dimX=dimYn=\dim X=\dim Y and that both uu and vv are mm-dimensional.

Definition 2.5 (Feedback Linearizability).

System (2.1) is said static feedback linearizable or, shortly, feedback linearizable if the context allows, if, and only if, it is equivalent by diffeomorphism and feedback to a linear system in Brunovský controllability canonical form

yi(ri)=vi,i=1,,m,y_{i}^{(r_{i})}=v_{i},\quad i=1,\ldots,m, (2.15)

where the multi-integer 𝐫(r1,,rm)\mathbf{r}\triangleq(r_{1},\ldots,r_{m}), whose components rir_{i} are called the controllability indices of system (2.15), satisfies |𝐫|i=1mri=n=dimX|\mathbf{r}|\triangleq\sum_{i=1}^{m}r_{i}=n=\dim X.

Indeed, since the nn-dimensional vector

y¯(𝐫𝟏)(y1,,y1(r11),,ym,,ym(rm1))\overline{y}^{(\mathbf{r-1})}\triangleq(y_{1},\ldots,y_{1}^{(r_{1}-1)},\ldots,y_{m},\ldots,y_{m}^{(r_{m}-1)})

and xx are diffeomorphic (again we have noted 𝐫𝟏(r11,,rm1)\mathbf{r-1}\triangleq(r_{1}-1,\ldots,r_{m}-1)), it is immediate to verify that (y1,,ym)(y_{1},\ldots,y_{m}) is a flat output and that every feedback linearizable system is differentially flat.

These systems have been first characterized by [17, 14, 15] for control-affine systems, i.e. systems given by f(x,u)=f0(x)+i=1muifi(x)f(x,u)=f_{0}(x)+\sum_{i=1}^{m}u_{i}f_{i}(x). More generally, it can be easily proven that systems of the form (2.1) are feedback linearizable if, and only if, the first-order control-affine prolongation

x˙=f(x,u),\displaystyle\dot{x}=f(x,u), (2.16)
u˙i=ui(1),i=1,,m\displaystyle\dot{u}_{i}=u_{i}^{(1)},\quad i=1,\ldots,m

with state x¯(𝟎)(x,u)=(x1,,xn,u1,,um)X(𝟎)X×m\overline{x}^{(\mathbf{0})}\triangleq(x,u)=(x_{1},\ldots,x_{n},u_{1},\ldots,u_{m})\in X^{(\mathbf{0})}\triangleq X\times\mathbb{R}^{m} and control vector u(𝟏)(u1(1),,um(1))mu^{(\mathbf{1})}\triangleq(u_{1}^{(1)},\ldots,u_{m}^{(1)})\in\mathbb{R}^{m}, is feedback linearizable (see e.g. [6, 31, 33, 20]).

Indeed, in the local coordinates555We introduce the superscript (0) from now on to get ready to work with higher order prolongations (see section 3). x¯(𝟎)\overline{x}^{(\mathbf{0})} of X(𝟎)X^{(\mathbf{0})}, denoting the associated vector fields by

g0(𝟎)(x¯(𝟎))i=1nfi(x,u)xi,gi(0)(x¯(𝟎))ui,i=1,,m,g_{0}^{(\mathbf{0})}(\overline{x}^{(\mathbf{0})})\triangleq\sum_{i=1}^{n}f_{i}(x,u)\frac{\partial}{\partial x_{i}},\qquad g_{i}^{(0)}(\overline{x}^{(\mathbf{0})})\triangleq\frac{\partial}{\partial u_{i}},\quad i=1,\ldots,m, (2.17)

defined on the tangent bundle TX(𝟎)=TX×Tm\mathrm{T}X^{(\mathbf{0})}=\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m}, system (2.16) reads

x¯˙(𝟎)=g0(𝟎)(x¯(𝟎))+i=1mui(1)gi(0)(x¯(𝟎))\dot{\overline{x}}^{(\mathbf{0})}=g_{0}^{(\mathbf{0})}(\overline{x}^{(\mathbf{0})})+\sum_{i=1}^{m}u_{i}^{(1)}g_{i}^{(0)}(\overline{x}^{(\mathbf{0})}) (2.18)

with the usual abuse of notations identifying a vector field expressed in local coordinates with its associated (Lie derivative) first order partial differential operator.

Until now, for simplicity’s sake, a system (2.1) will always be considered in the form (2.18), even if the vector-field ff is already given in control-affine form. For the sake of coherence, we set u=(u1,,um)u(𝟎)=(u1(0),,um(0))u=(u_{1},\ldots,u_{m})\triangleq u^{(\mathbf{0})}=(u_{1}^{(0)},\ldots,u_{m}^{(0)}), so that u˙(𝟎)=u(𝟏)\dot{u}^{(\mathbf{0})}=u^{(\mathbf{1})}.

2.4 Recalls on Lie brackets and Distributions

The Lie bracket [η,γ][\eta,\gamma] of two arbitrary vector fields η\eta and γ\gamma of TX(𝟎)\mathrm{T}X^{(\mathbf{0})} is given, in the x¯(𝟎)\overline{x}^{(\mathbf{0})}-coordinates, by [η,γ]i=1n+mj=1n+m(ηjγixj(0)γjηixj(0))xi(0)[\eta,\gamma]\triangleq\sum_{i=1}^{n+m}\sum_{j=1}^{n+m}\left(\eta_{j}\frac{\partial\gamma_{i}}{\partial x_{j}^{(0)}}-\gamma_{j}\frac{\partial\eta_{i}}{\partial x_{j}^{(0)}}\right)\frac{\partial}{\partial x_{i}^{(0)}}, with x¯(𝟎)=(x,u(𝟎))(x1(0),,xn+m(0))\overline{x}^{(\mathbf{0})}=(x,u^{(\mathbf{0})})\triangleq(x_{1}^{(0)},\ldots,x_{n+m}^{(0)}).

For iterated Lie brackets, we use the notation adηγ[η,γ]\mathrm{ad}_{\eta}\gamma\triangleq[\eta,\gamma] and adηkγ[η,adηk1γ]\mathrm{ad}_{\eta}^{k}\gamma\triangleq[\eta,\mathrm{ad}_{\eta}^{k-1}\gamma] for k1k\geq 1, with the convention that adη0γ=γ\mathrm{ad}_{\eta}^{0}\gamma=\gamma. In addition, if Γ\Gamma is an arbitrary distribution of vector fields on TX(𝟎)\mathrm{T}X^{(\mathbf{0})}, we note adηkΓ{adηkγ:γΓ}\mathrm{ad}_{\eta}^{k}\Gamma\triangleq\{\mathrm{ad}_{\eta}^{k}\gamma:\gamma\in\Gamma\}.

The distribution Γ\Gamma is said involutive if, and only if, [η,γ]Γ[\eta,\gamma]\in\Gamma for every pair of vector fields η,γΓ\eta,\gamma\in\Gamma, to which case we note [Γ,Γ]Γ[\Gamma,\Gamma]\subset\Gamma, or Γ¯=Γ\overline{\Gamma}=\Gamma, where Γ¯\overline{\Gamma} denotes the involutive closure of Γ\Gamma, i.e. the smallest involutive distribution containing Γ\Gamma.

If the distribution Γ\Gamma is locally generated by pp vector fields γ1,,γp\gamma_{1},\ldots,\gamma_{p}, with pp arbitrary, we write Γ{γ1,,γp}\Gamma\triangleq\{\gamma_{1},\ldots,\gamma_{p}\}. We also denote by Γ(ξ){γ1(ξ),,γp(ξ)}\Gamma(\xi)\triangleq\{\gamma_{1}(\xi),\ldots,\gamma_{p}(\xi)\} the vector space generated by the vectors γ1(ξ),,γp(ξ)\gamma_{1}(\xi),\ldots,\gamma_{p}(\xi) at a point ξX(𝟎)\xi\in X^{(\mathbf{0})}.

Consider the (𝟎\mathbf{0}th-order or non prolonged) filtration of distributions built on the vector fields (2.17)666As before, the superscript (0) is used to indicate that the distributions Gk(𝟎)G_{k}^{(\mathbf{0})} and the related indices ρk(𝟎)\rho_{k}^{(\mathbf{0})} and κk(𝟎)\kappa_{k}^{(\mathbf{0})} are built on the non prolonged vector fields (2.17) and to distinguish them from the prolonged distributions Gk(𝐣)G_{k}^{(\mathbf{j})} of arbitrary 𝐣\mathbf{j}th order, 𝐣m\mathbf{j}\in\mathbb{N}^{m}, and related indices, ρk(𝐣)\rho_{k}^{(\mathbf{j})} and κk(𝐣)\kappa_{k}^{(\mathbf{j})}, later introduced in sections 3 and 4.:

G0(𝟎){g1(𝟎),,gm(𝟎)},Gk+1(𝟎)Gk(𝟎)+adg0(𝟎)Gk(𝟎),k0,G_{0}^{(\mathbf{0})}\triangleq\{g_{1}^{(\mathbf{0})},\ldots,g_{m}^{(\mathbf{0})}\},\qquad G_{k+1}^{(\mathbf{0})}\triangleq G_{k}^{(\mathbf{0})}+\mathrm{ad}_{g_{0}^{(\mathbf{0})}}G_{k}^{(\mathbf{0})},\quad\forall k\geq 0, (2.19)

indeed satisfying G0(𝟎)Gk(𝟎)Gk+1(𝟎)TX(𝟎)G_{0}^{(\mathbf{0})}\subset\cdots\subset G_{k}^{(\mathbf{0})}\subset G_{k+1}^{(\mathbf{0})}\subset\cdots\subset\mathrm{T}X^{(\mathbf{0})}.

Theorem 2.2 ([17, 15]).

System (2.1), or equivalently system (2.16), is feedback linearizable in a neighborhood of the origin of X(𝟎)X^{(\mathbf{0})} if, and only if, in this neighborhood:

  • (i)

    Gk(𝟎)G_{k}^{(\mathbf{0})} is involutive with constant dimension for all k0k\geq 0,

  • (ii)

    there exists an integer k(𝟎)nk^{(\mathbf{0})}_{\star}\leq n such that Gk(𝟎)=Gk(𝟎)(𝟎)=TX(𝟎)G_{k}^{(\mathbf{0})}=G_{k^{(\mathbf{0})}_{\star}}^{(\mathbf{0})}=\mathrm{T}X^{(\mathbf{0})} for all kk(𝟎)k\geq k^{(\mathbf{0})}_{\star}.

Note that, according to (2.17) and (2.19), G0(𝟎)={u1(0),,um(0)}G_{0}^{(\mathbf{0})}=\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{m}^{(0)}}\} is involutive with constant dimension, equal to mm, by construction.

Theorem 2.2 provides a construction of flat outputs via Frobenius theorem (see e.g. [8]) and the list of the so-called Brunovský’s controllability indices [4] as follows:

Definition 2.6.

Consider the sequence of integers

ρk(𝟎)dimGk(𝟎)/Gk1(𝟎)k1,ρ0(𝟎)dimG0(𝟎)=m.\rho_{k}^{(\mathbf{0})}\triangleq\dim G_{k}^{(\mathbf{0})}/G_{k-1}^{(\mathbf{0})}\quad\forall k\geq 1,\qquad\rho_{0}^{(\mathbf{0})}\triangleq\dim G_{0}^{(\mathbf{0})}=m.

The Brunovský controllability indices κk(𝟎)\kappa_{k}^{(\mathbf{0})}’s are defined by

κk(𝟎)#{lρl(𝟎)k},k=1,,m,\kappa_{k}^{(\mathbf{0})}\triangleq\#\{l\mid\rho_{l}^{(\mathbf{0})}\geq k\},\quad k=1,\ldots,m,

where #A\#A denotes the number of elements of an arbitrary set AA.

It can be proven (see e.g. [17, 15, 16, 28, 20]) that, for a feedback linearizable nonlinear system (2.1), or (2.16), we have:

  • the sequences {ρk(𝟎)}k0\{\rho_{k}^{(\mathbf{0})}\}_{k\geq 0} and {κk(𝟎)}k0\{\kappa_{k}^{(\mathbf{0})}\}_{k\geq 0} are non increasing,

  • ρk(𝟎)m\rho_{k}^{(\mathbf{0})}\leq m for all k,ρk(𝟎)=0k,\quad\rho_{k}^{(\mathbf{0})}=0, for all kk(𝟎)+1k\geq k^{(\mathbf{0})}_{\star}+1,

  • κ1(𝟎)=k(𝟎)+1,κm(𝟎)1\kappa_{1}^{(\mathbf{0})}=k^{(\mathbf{0})}_{\star}+1,\quad\kappa_{m}^{(\mathbf{0})}\geq 1,

  • k=0k(𝟎)ρk(𝟎)=k=1mκk(𝟎)=dimGk(𝟎)(𝟎)=n+m\displaystyle\sum_{k=0}^{k^{(\mathbf{0})}_{\star}}\rho_{k}^{(\mathbf{0})}=\sum_{k=1}^{m}\kappa_{k}^{(\mathbf{0})}=\dim G_{k^{(\mathbf{0})}_{\star}}^{(\mathbf{0})}=n+m.

The list κ1(𝟎),,κm(𝟎)\kappa_{1}^{(\mathbf{0})},\ldots,\kappa_{m}^{(\mathbf{0})} is uniquely defined up to input permutation, invariant by static state feedback and state diffeomorphism, and is indeed equal to the list of controllability indices of the associated linear system (2.15) with ri=κi(𝟎)r_{i}=\kappa_{i}^{(\mathbf{0})}, i=1,,mi=1,\ldots,m.

Moreover, for all kk and all i=1,,mi=1,\ldots,m, and possibly up to a suitable input reordering, we have

Gk(𝟎)=j=1m{adg0(𝟎)lgj(𝟎)l=0,,k(κj(𝟎)1)},Gκ1(𝟎)1(𝟎)=Gk(𝟎)(𝟎)=TX(𝟎).G_{k}^{(\mathbf{0})}=\bigoplus_{j=1}^{m}\{\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{l}g_{j}^{(\mathbf{0})}\mid l=0,\ldots,k\lor(\kappa_{j}^{(\mathbf{0})}-1)\},\quad G_{\kappa_{1}^{(\mathbf{0})}-1}^{(\mathbf{0})}=G_{k^{(\mathbf{0})}_{\star}}^{(\mathbf{0})}=\mathrm{T}X^{(\mathbf{0})}.

Then, flat outputs (y1,,ym)(y_{1},\ldots,y_{m}) are locally non trivial solutions of the system of PDE’s

Ladg0(𝟎)kgj(𝟎)yi=0,k=0,,κi(𝟎)2,j=1,,m,\displaystyle L_{\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{k}g_{j}^{(\mathbf{0})}}\;y_{i}=0,\;k=0,\ldots,\kappa_{i}^{(\mathbf{0})}-2,\;j=1,\ldots,m, (2.20)
withLadg0(𝟎)κi(𝟎)1gi(𝟎)yi0,\displaystyle\mathrm{with\leavevmode\nobreak\ }L_{\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{\kappa_{i}^{(\mathbf{0})}-1}g_{i}^{(\mathbf{0})}}\;y_{i}\neq 0,

for i=1,,mi=1,\ldots,m, where we have denoted by LηφL_{\eta}\varphi the Lie derivative of a vector function φ\varphi along the vector field η\eta. These solutions are such that the mapping

x¯(𝟎)(y1,,y1(κ1(𝟎)1),,ym,,ym(κm(𝟎)1))\overline{x}^{(\mathbf{0})}\mapsto(y_{1},\ldots,y_{1}^{(\kappa_{1}^{(\mathbf{0})}-1)},\ldots,y_{m},\ldots,y_{m}^{(\kappa_{m}^{(\mathbf{0})}-1)})

is a local diffeomorphism.

Remark 2.2.

Recall from [6] that, for single input systems, differential flatness and feedback linearizability are equivalent.

3 System Pure Prolongation

3.1 Purely Prolonged Distributions

We now introduce higher order prolongations of the vector fields defined by (2.17), called pure prolongations [7] (see also [33, 34, 2, 12]).

Given a multi-integer 𝐣(j1,,jm)m\mathbf{j}\triangleq(j_{1},\ldots,j_{m})\in\mathbb{N}^{m}, we note, as before, |𝐣|i=1mji|\mathbf{j}|\triangleq\sum_{i=1}^{m}j_{i} and the prolonged state:

x¯(𝐣)(x,u¯(𝐣))(x1,,xn,u1(0),,u1(j1),,um(0),,um(jm)),\overline{x}^{(\mathbf{j})}\triangleq(x,\overline{u}^{(\mathbf{j})})\triangleq(x_{1},\ldots,x_{n},u_{1}^{(0)},\ldots,u_{1}^{(j_{1})},\ldots,u_{m}^{(0)},\ldots,u_{m}^{(j_{m})}),

with ui(0)uiu_{i}^{(0)}\triangleq u_{i}, i=1,,mi=1,\ldots,m.

Let X(𝐣)X×m+|𝐣|X^{(\mathbf{j})}\triangleq X\times\mathbb{R}^{m+|\mathbf{j}|} be the associated 𝐣\mathbf{j}-th order jet manifold of dimension n+m+|𝐣|n+m+|\mathbf{j}| with coordinates x¯(𝐣)\overline{x}^{(\mathbf{\mathbf{j}})}.

The pure prolongation of order 𝐣\mathbf{j} of system (2.16), or otherwise said, of the vector fields (2.17), in the tangent bundle TX(𝐣)=TX×Tm+|𝐣|\mathrm{T}X^{(\mathbf{j})}=\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m+|\mathbf{j}|}, is defined by

g0(𝐣)(x¯(𝐣))\displaystyle g_{0}^{(\mathbf{j})}(\overline{x}^{(\mathbf{j})}) =i=1nfi(x,u)xi+i=1mk=0ji1ui(k+1)ui(k)\displaystyle=\sum_{i=1}^{n}f_{i}(x,u)\frac{\partial}{\partial x_{i}}+\sum_{i=1}^{m}\sum_{k=0}^{j_{i}-1}u_{i}^{(k+1)}\frac{\partial}{\partial u_{i}^{(k)}} (3.1)
gi(𝐣)(x¯(𝐣))\displaystyle g_{i}^{(\mathbf{j})}(\overline{x}^{(\mathbf{j})}) gi(ji)(x¯(𝐣))ui(ji),i=1,,m,\displaystyle\triangleq g_{i}^{(j_{i})}(\overline{x}^{(\mathbf{j})})\triangleq\frac{\partial}{\partial u_{i}^{(j_{i})}},\quad i=1,\ldots,m,

with the convention that k=0ji1ui(k+1)ui(k)0\sum_{k=0}^{j_{i}-1}u_{i}^{(k+1)}\frac{\partial}{\partial u_{i}^{(k)}}\triangleq 0 if ji=0j_{i}=0.

They are naturally associated to the adjunction of 𝐣\mathbf{j} pure integrators to u=u(0)u=u^{(0)} in (2.16) (with the same usual abuse of notations as in (2.18)):

x¯˙(𝐣)=g0(𝐣)(x¯(𝐣))+i=1mui(ji+1)gi(ji)(x¯(𝐣))\dot{\overline{x}}^{(\mathbf{j})}=g_{0}^{(\mathbf{j})}(\overline{x}^{(\mathbf{j})})+\sum_{i=1}^{m}u_{i}^{(j_{i}+1)}g_{i}^{(j_{i})}(\overline{x}^{(\mathbf{j})}) (3.2)

or

x˙=f(x,u),u˙i(k)=ui(k+1),k=0,,ji,i=1,,m,\dot{x}=f(x,u),\qquad\dot{u}_{i}^{(k)}=u_{i}^{(k+1)},\quad k=0,\ldots,j_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ i=1,\ldots,m,

u(𝐣+𝟏)=(u1(j1+1),,um(jm+1))u^{(\mathbf{j+1})}=(u_{1}^{(j_{1}+1)},\ldots,u_{m}^{(j_{m}+1)}) being the new control vector of this purely prolonged system, whose state is (x,u¯(𝐣))=x¯(𝐣)(x,\overline{u}^{(\mathbf{j})})=\overline{x}^{(\mathbf{j})}.

Remark 3.1.

Note that the state of the 𝐣\mathbf{j}-th prolonged system, x¯(𝐣)\overline{x}^{(\mathbf{j})}, coincides with the image of x¯\overline{x} by the projection p𝐣:x¯X×mp𝐣(x¯)=x¯(𝐣)X(𝐣)p_{\mathbf{j}}:\overline{x}\in X\times\mathbb{R}^{m}_{\infty}\mapsto p_{\mathbf{j}}(\overline{x})=\overline{x}^{(\mathbf{j})}\in X^{(\mathbf{j})} for all 𝐣\mathbf{j}. In addition, the family of projections p𝐢,𝐣:x¯(𝐢)X(𝐢)x¯(𝐣)=p𝐢,𝐣(x¯(𝐢))X(𝐣)p_{\mathbf{i},\mathbf{j}}:\overline{x}^{(\mathbf{i})}\in X^{(\mathbf{i})}\mapsto\overline{x}^{(\mathbf{j})}=p_{\mathbf{i},\mathbf{j}}(\overline{x}^{(\mathbf{i})})\in X^{(\mathbf{j})} for all 𝐢,𝐣\mathbf{i},\mathbf{j} such that ikjki_{k}\geq j_{k} for all k=1,,mk=1,\ldots,m, that we note 𝐢𝐣\mathbf{i}\succeq\mathbf{j}, indeed satisfies p𝐢,𝐣p𝐣,𝐤=p𝐢,𝐤p_{\mathbf{i},\mathbf{j}}\circ p_{\mathbf{j},\mathbf{k}}=p_{\mathbf{i},\mathbf{k}} for all 𝐢𝐣𝐤\mathbf{i}\succeq\mathbf{j}\succeq\mathbf{k} and thus allows us to identify the manifold X×mX\times\mathbb{R}^{m}_{\infty} with the projective limit of the family (X(𝐢),p𝐢,𝐣)(X^{(\mathbf{i})},p_{\mathbf{i},\mathbf{j}}) for all 𝐢\mathbf{i} and all 𝐣\mathbf{j} such that 𝐢𝐣\mathbf{i}\succeq\mathbf{j}, i.e. X×mlimX(𝐢)X\times\mathbb{R}^{m}_{\infty}\simeq\displaystyle\lim_{\leftarrow}X^{(\mathbf{i})} (see e.g. [3, Chap. I,§10]). A similar identification trivially holds for the associated tangent bundles, i.e. TX×TmlimTX(𝐢)\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m}_{\infty}\simeq\displaystyle\lim_{\leftarrow}\mathrm{T}X^{(\mathbf{i})} relatively to the family Tp𝐢,𝐣\mathrm{T}p_{\mathbf{i},\mathbf{j}} of tangent projections, hence the identification of the Cartan field CfC_{f}, defined by (2.4), with limg0(𝐣)\displaystyle\lim_{\leftarrow}g_{0}^{\mathbf{(j)}}, the projective limit of the vector fields g0(𝐣)g_{0}^{\mathbf{(j)}}. Nevertheless, this property does not hold for the control vector fields gi(𝐣)g_{i}^{\mathbf{(j)}} since Tp𝐣,𝐤(gi(𝐣))\mathrm{T}p_{\mathbf{j},\mathbf{k}}(g_{i}^{\mathbf{(j)}}) is not equal to gi(𝐤)g_{i}^{\mathbf{(k)}}, for i=1,,mi=1,\ldots,m and 𝐣𝐤\mathbf{j}\succeq\mathbf{k}. Moreover, the Lie bracket of vector fields is not preserved by this family of projections. This is one of the reasons why prolongations may enlarge the range of the system transformations.

Remark 3.2.

Given an arbitrary point x¯0(x0,u¯0)\overline{x}_{0}\triangleq(x_{0},\overline{u}_{0}) around which system (2.16) is defined, it is convenient to consider the shift

θ:(x,u¯)X×mθ(x,u¯)=(xx0,(uu0)¯)(z,v¯)X×m\theta:(x,\overline{u})\in X\times\mathbb{R}^{m}_{\infty}\mapsto\theta(x,\overline{u})=(x-x_{0},\overline{(u-u_{0})})\triangleq(z,\overline{v})\in X\times\mathbb{R}^{m}_{\infty}

such that x¯0\overline{x}_{0} is mapped to the origin of TX×Tm\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m}_{\infty}, denoted by 0¯\overline{0}, thus inducing the shift of vector fields:

θ(gi(𝐣))(z,v¯)gi(𝐣)(z+x0,(v+u0)¯),i=0,m,\theta_{\star}(g_{i}^{(\mathbf{j})})(z,\overline{v})\triangleq g_{i}^{(\mathbf{j})}(z+x_{0},\overline{(v+u_{0})}),\quad i=0,\ldots m, (3.3)

now defined in a neighborhood of 0¯\overline{0}. For the sake of simplicity, we will only consider such shifted vector fields in the sequel while keeping the same notation gi(𝐣)g_{i}^{(\mathbf{j})} as before, though abusive, but yet unambiguous.

We now introduce the following filtration of 𝐣\mathbf{j}-th order purely prolonged distributions of TX(𝐣)\mathrm{T}X^{(\mathbf{j})}:

G0(𝐣){u1(j1),,um(jm)},Gk+1(𝐣)Gk(𝐣)+adg0(𝐣)Gk(𝐣),k0G_{0}^{(\mathbf{j})}\triangleq\{\frac{\partial}{\partial u_{1}^{(j_{1})}},\ldots,\frac{\partial}{\partial u_{m}^{(j_{m})}}\},\qquad G_{k+1}^{(\mathbf{j})}\triangleq G_{k}^{(\mathbf{j})}+\mathrm{ad}_{g_{0}^{(\mathbf{j})}}G_{k}^{(\mathbf{j})},\quad\forall k\geq 0 (3.4)

Indeed, for 𝐣=𝟎\mathbf{j}=\mathbf{0}, i.e. j1==jm=0j_{1}=\cdots=j_{m}=0, this filtration coincides with the 𝟎\mathbf{0}th order one given by (2.19). Similarly to the 𝟎\mathbf{0}th order case, G0(𝐣)G_{0}^{(\mathbf{j})} is involutive with constant dimension, equal to mm, by construction.

Moreover, since every Gk(𝐣)TX(𝐣)G_{k}^{(\mathbf{j})}\subset\mathrm{T}X^{(\mathbf{j})}, with dimTX(𝐣)=n+m+𝐣\dim\mathrm{T}X^{(\mathbf{j})}=n+m+\mid\mathbf{j}\mid, we have

Proposition 3.1.

There exists a finite integer k(𝐣)k^{(\mathbf{j})}_{\star} such that Gk(𝐣)=Gk(𝐣)(𝐣)G_{k}^{(\mathbf{j})}=G_{k^{(\mathbf{j})}_{\star}}^{(\mathbf{j})} for all kk(𝐣)k\geq k^{(\mathbf{j})}_{\star} and

k(𝐣)n+𝐣.k^{(\mathbf{j})}_{\star}\leq n+\mid\mathbf{j}\mid. (3.5)
Proof.

Since

n+m+𝐣dimGk(𝐣)(𝐣)=k=1k(𝐣)dimGk(𝐣)/Gk1(𝐣)+dimG0(𝐣)k(𝐣)+m,n+m+\mid\mathbf{j}\mid\geq\dim G_{k^{(\mathbf{j})}_{\star}}^{(\mathbf{j})}=\sum_{k=1}^{k^{(\mathbf{j})}_{\star}}\dim G_{k}^{(\mathbf{j})}/G_{k-1}^{(\mathbf{j})}+\dim G_{0}^{(\mathbf{j})}\geq k^{(\mathbf{j})}_{\star}+m,

we get (3.5). ∎

Remark 3.3.

In full generality, k(𝐣)k^{(\mathbf{j})}_{\star} depends on the point where it is evaluated. However, if dimGk(𝐣)\dim G_{k}^{(\mathbf{j})} is constant in an open dense subset of X(𝐣)X^{(\mathbf{j})} for all large enough kk, so is k(𝐣)k^{(\mathbf{j})}_{\star}.

Let us inductively define the nn-dimensional vector functions γk,i(𝐣)\gamma_{k,i}^{(\mathbf{j})}, for k1k\geq 1, i=1,,mi=1,\ldots,m, and arbitrary 𝐣=(j1,,jm)\mathbf{j}=(j_{1},\ldots,j_{m}) as follows:

γk+1,i(𝐣)\displaystyle\gamma_{k+1,i}^{(\mathbf{j})} Lg0(𝐣)γk,i(𝐣)γk,i(𝐣)fx=Lg0(𝟎)γk,i(𝐣)+p=1ml=0jp1up(l+1)γk,i(𝐣)up(l)γk,i(𝐣)fx\displaystyle\triangleq L_{g_{0}^{(\mathbf{j})}}\gamma_{k,i}^{(\mathbf{j})}-\gamma_{k,i}^{(\mathbf{j})}\frac{\partial f}{\partial x}=L_{g_{0}^{(\mathbf{0})}}\gamma_{k,i}^{(\mathbf{j})}+\sum_{p=1}^{m}\sum_{l=0}^{j_{p}-1}u_{p}^{(l+1)}\frac{\partial\gamma_{k,i}^{(\mathbf{j})}}{\partial u_{p}^{(l)}}-\gamma_{k,i}^{(\mathbf{j})}\frac{\partial f}{\partial x} (3.6)

with

γ1,i(𝐣)=(1)(ji+1)fui(0).\gamma_{1,i}^{(\mathbf{j})}=(-1)^{(j_{i}+1)}\frac{\partial f}{\partial u_{i}^{(0)}}. (3.7)

For an arbitrary 𝐣\mathbf{j} and given i=1,,mi=1,\ldots,m, thanks to (3.7), it is readily seen that γ1,i(𝐣)\gamma_{1,i}^{(\mathbf{j})} depends at most of x¯(𝟎)\overline{x}^{(\mathbf{0})} and thus, if kji1k\leq j_{i}-1, thanks to (3.6), γk+1,i(𝐣)\gamma_{k+1,i}^{(\mathbf{j})} depends at most of x¯(𝐣k)\overline{x}^{(\mathbf{j}\bigvee k)}.


3.2 Vertical Distributions of Purely Prolonged Ones

Before stating the next Lemma, we need to recall the definition of vertical bundle. Given an arbitrary 𝐫m\mathbf{r}\in\mathbb{N}^{m} and the fiber bundle π𝐫: X(𝐫)m+𝐫\pi_{\mathbf{r}}: X^{(\mathbf{r})}\rightarrow\mathbb{R}^{m+\mid\mathbf{r}\mid}, with π𝐫(x¯(𝐫))=u¯(𝐫)\pi_{\mathbf{r}}(\overline{x}^{(\mathbf{r})})=\overline{u}^{(\mathbf{r})}, its vertical space at x¯(𝐫)\overline{x}^{(\mathbf{r})}, denoted by Vx¯(𝐫)X(𝐫)\mathrm{V}_{\overline{x}^{(\mathbf{r})}}X^{(\mathbf{r})}, is the tangent space TxX\mathrm{T}_{x}X. Its vertical bundle, denoted by VX(𝐫)\mathrm{V}X^{(\mathbf{r})}, is the vector bundle made of the vertical spaces at each x¯(𝐫)\overline{x}^{(\mathbf{r})}, i.e. the set of linear combinations i=1nαi(x¯(𝐫))xi\sum_{i=1}^{n}\alpha_{i}(\overline{x}^{(\mathbf{r})})\frac{\partial}{\partial x_{i}} whose coefficients αi\alpha_{i} are smooth functions that depend at most on x¯(𝐫)\overline{x}^{(\mathbf{r})} and where (x1,,xn)(x_{1},\ldots,x_{n}) are local coordinates of XX.

The same definition indeed holds for the vertical bundle V(X×m)\mathrm{V}(X\times\mathbb{R}^{m}_{\infty}) associated to the fiber bundle π:X×mm\pi:X\times\mathbb{R}^{m}_{\infty}\rightarrow\mathbb{R}^{m}_{\infty}, i.e. the set of linear combinations of x1,,xn\frac{\partial}{\partial x_{1}},\ldots,\frac{\partial}{\partial x_{n}} whose coefficients are smooth functions of x¯\overline{x}.

We now establish some comparison formulae between Lie brackets of the vector fields of the purely prolonged system and those of the original (non prolonged) one. More complete formulae may be found in Lemma A.1 of the Appendix A.

Lemma 3.1.

For all 𝐣=(j1,,jm)m\mathbf{j}=\left(j_{1},\ldots,j_{m}\right)\in\mathbb{N}^{m} satisfying 0j1jm0\leq j_{1}\leq\ldots\leq j_{m}, for all kjik\leq j_{i} and i=1,,mi=1,\ldots,m, we have:

adg0(𝐣)kui(ji)=(1)kui(jik)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k}\frac{\partial}{\partial u_{i}^{(j_{i})}}=(-1)^{k}\frac{\partial}{\partial u_{i}^{(j_{i}-k)}} (3.8)

and for all k1k\geq 1:

adg0(𝐣)ji+kui(ji)=(1)jiadg0(𝐣)kui(0)=γk,i(𝐣)xVX(𝐣(k1)),\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+k}\frac{\partial}{\partial u_{i}^{(j_{i})}}=(-1)^{j_{i}}\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k}\frac{\partial}{\partial u_{i}^{(0)}}=\gamma_{k,i}^{(\mathbf{j})}\frac{\partial}{\partial x}\in\mathrm{V}X^{(\mathbf{j}\bigvee(k-1))}, (3.9)

Moreover, we have

[up(jpk),adg0(𝐣)ljquq(0)]=0,k<jp,ljqs.t.k+l<jp+jq+1.\left[\frac{\partial}{\partial u_{p}^{(j_{p}-k)}},\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{l-j_{q}}\frac{\partial}{\partial u_{q}^{(0)}}\right]=0,\quad\forall k<j_{p},\;\;\forall l\geq j_{q}\leavevmode\nobreak\ \mathrm{s.t.\leavevmode\nobreak\ }k+l<j_{p}+j_{q}+1. (3.10)
Proof.

It is immediately seen that

adg0(𝐣)ui(ji)=[fx+k=1ml0uk(l+1)uk(l),ui(ji)]=ui(ji1).\mathrm{ad}_{g_{0}^{(\mathbf{j})}}\frac{\partial}{\partial u_{i}^{(j_{i})}}=\left[f\frac{\partial}{\partial x}+\sum_{k=1}^{m}\sum_{l\geq 0}u_{k}^{(l+1)}\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(j_{i})}}\right]=-\frac{\partial}{\partial u_{i}^{(j_{i-1})}}.

Iterating this computation up to k=jik=j_{i} yields (3.8):

adg0(𝐣)jiui(ji)=(1)jiui(0).\mathrm{ad}^{j_{i}}_{g_{0}^{(\mathbf{j})}}\frac{\partial}{\partial u_{i}^{(j_{i})}}=(-1)^{j_{i}}\frac{\partial}{\partial u_{i}^{(0)}}.

Then, for k=ji+1k=j_{i}+1, using the fact that [uk(l),ui(0)]=0[\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(0)}}]=0 for all ii, kk and l0l\geq 0, we have:

adg0(𝐣)ji+1ui(ji)\displaystyle\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+1}\frac{\partial}{\partial u_{i}^{(j_{i})}} =(1)ji[fx+k=1ml=0jk1uk(l+1)uk(l),ui(0)]\displaystyle=(-1)^{j_{i}}\left[f\frac{\partial}{\partial x}+\sum_{k=1}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(0)}}\right] (3.11)
=(1)jiadg0(𝟎)ui(0)=(1)(ji+1)fui(0)(x¯(𝟎))x=γ1,i(𝐣)x,\displaystyle=(-1)^{j_{i}}\mathrm{ad}_{g_{0}^{(\mathbf{0})}}\frac{\partial}{\partial u_{i}^{(0)}}=(-1)^{(j_{i}+1)}\frac{\partial f}{\partial u_{i}^{(0)}}(\overline{x}^{(\mathbf{0})})\frac{\partial}{\partial x}=\gamma_{1,i}^{(\mathbf{j})}\frac{\partial}{\partial x},

which proves that adg0(𝐣)ji+1ui(ji)VX(𝟎)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+1}\frac{\partial}{\partial u_{i}^{(j_{i})}}\in\mathrm{V}X^{(\mathbf{0})} and that (3.9) holds at the order k=1k=1.

Assuming that (3.9) holds up to k=νk=\nu, with γν,i(𝐣)\gamma_{\nu,i}^{(\mathbf{j})} depending at most on x¯(𝐣(ν1))\overline{x}^{(\mathbf{j}\lor(\nu-1))}, we have

adg0(𝐣)ji+ν+1ui(ji)=(1)ji[g0(𝐣),adg0(𝐣)νui(0)]=(Lg0(𝐣)γν,i(𝐣)γν,i(𝐣)fx)x=γν+1,i(𝐣)x,\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+\nu+1}\frac{\partial}{\partial u_{i}^{(j_{i})}}=(-1)^{j_{i}}\left[g_{0}^{(\mathbf{j})},\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{\nu}\frac{\partial}{\partial u_{i}^{(0)}}\right]=\left(L_{g_{0}^{(\mathbf{j})}}\gamma_{\nu,i}^{(\mathbf{j})}-\gamma_{\nu,i}^{(\mathbf{j})}\frac{\partial f}{\partial x}\right)\frac{\partial}{\partial x}=\gamma_{\nu+1,i}^{(\mathbf{j})}\frac{\partial}{\partial x},

and, according to (3.6)-(3.7), for all jp+1ν<jp+1j_{p}+1\leq\nu<j_{p+1}, p=1,,m1p=1,\ldots,m-1, since we differentiate γν,i(𝐣)\gamma_{\nu,i}^{(\mathbf{j})} with respect to k=1ml=0jk1uk(l+1)uk(l)\sum_{k=1}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}\frac{\partial}{\partial u_{k}^{(l)}}, it is immediate to verify that γν+1,i(𝐣)\gamma_{\nu+1,i}^{(\mathbf{j})} depends at most on

x¯(𝐣ν)=(x,u1(0),,u1(j1),,up(0),,up(jp),up+1(0),,up+1(ν),,um(0),,um(ν))\overline{x}^{(\mathbf{j}\bigvee\nu)}=\left(x,u_{1}^{(0)},\ldots,u_{1}^{(j_{1})},\ldots,u_{p}^{(0)},\ldots,u_{p}^{(j_{p})},u_{p+1}^{(0)},\ldots,u_{p+1}^{(\nu)},\ldots,u_{m}^{(0)},\ldots,u_{m}^{(\nu)}\right)

and on x¯(𝐣)=x¯(𝐣ν)\overline{x}^{(\mathbf{j})}=\overline{x}^{(\mathbf{j}\bigvee\nu)} if νjm\nu\geq j_{m}, hence (3.9).

Concerning (3.10), since adg0(𝐣)ljquq(0)=±γljq,q(𝐣)x\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{l-j_{q}}\frac{\partial}{\partial u_{q}^{(0)}}=\pm\gamma_{l-j_{q},q}^{(\mathbf{j})}\frac{\partial}{\partial x} depends at most on x¯(𝐣(ljq1))\overline{x}^{(\mathbf{j}\bigvee(l-j_{q}-1))}, the derivative of γljq,q(𝐣)\gamma_{l-j_{q},q}^{(\mathbf{j})} with respect to up(jpk)\frac{\partial}{\partial u_{p}^{(j_{p}-k)}} is indeed null if jpk>ljq1j_{p}-k>l-j_{q}-1, hence the result. ∎

3.3 Decomposition of Purely Prolonged Distributions

Let us assume, without loss of generality, that the control components have been reordered in such a way that j1j2jmj_{1}\leq j_{2}\leq\cdots\leq j_{m}. Moreover, we may suppose that j1=0j_{1}=0, as shown to be sufficient in the next section.

We now introduce two new filtrations of TX(𝐣)\mathrm{T}X^{(\mathbf{j})}, noted Γk(𝐣)\Gamma_{k}^{(\mathbf{j})} and Δk(𝐣)\Delta_{k}^{(\mathbf{j})}, for k0k\geq 0, as follows

Γk(𝐣)p=2m{up(jpl)l=0,,k(jp1)},\Gamma_{k}^{(\mathbf{j})}\triangleq\bigoplus_{p=2}^{m}\left\{\frac{\partial}{\partial u_{p}^{(j_{p}-l)}}\mid l=0,\ldots,k\vee(j_{p}-1)\right\}, (3.12)
Δk(𝐣)p=1m{adg0(𝐣)ljpup(0)l=jp,,k}.\Delta_{k}^{(\mathbf{j})}\triangleq\sum_{p=1}^{m}\left\{\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{l-j_{p}}\frac{\partial}{\partial u_{p}^{(0)}}\mid l=j_{p},\ldots,k\right\}. (3.13)

with the convention that adg0(𝐣)kjpup(0)=0\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k-j_{p}}\frac{\partial}{\partial u_{p}^{(0)}}=0 if k<jpk<j_{p}, p=1,,mp=1,\ldots,m.

We indeed have Γk(𝐣)=Γjm1(𝐣)=p=2m{up(jp),,up(1)}\Gamma_{k}^{(\mathbf{j})}=\Gamma_{j_{m}-1}^{(\mathbf{j})}=\bigoplus_{p=2}^{m}\{\frac{\partial}{\partial u_{p}^{(j_{p})}},\ldots,\frac{\partial}{\partial u_{p}^{(1)}}\} for all kjm1k\geq j_{m}-1. Thus, dimΓk(𝐣)|𝐣|\dim\Gamma_{k}^{(\mathbf{j})}\leq|\mathbf{j}| for all kk.

The definitions (3.12)–(3.13) and Lemma 3.1 readily yield, for all kk, ll and 𝐣\mathbf{j},

Γk(𝐣)¯=Γk(𝐣),[Γk(𝐣),Γl(𝐣)]Γkl(𝐣),\displaystyle\overline{\Gamma_{k}^{(\mathbf{j})}}=\Gamma_{k}^{(\mathbf{j})},\qquad[\Gamma_{k}^{(\mathbf{j})},\Gamma_{l}^{(\mathbf{j})}]\subset\Gamma_{k\wedge l}^{(\mathbf{j})}, (3.14)
Δl(𝐣)V(X×m),[Γk(𝐣),Δl(𝐣)]V(X×m),[Γk(𝐣),Δl(𝐣)]Γk(𝐣)={0}.\displaystyle\Delta_{l}^{(\mathbf{j})}\subset\mathrm{V}(X\times\mathbb{R}^{m}_{\infty}),\quad[\Gamma_{k}^{(\mathbf{j})},\Delta_{l}^{(\mathbf{j})}]\subset\mathrm{V}(X\times\mathbb{R}^{m}_{\infty}),\quad[\Gamma_{k}^{(\mathbf{j})},\Delta_{l}^{(\mathbf{j})}]\cap\Gamma_{k}^{(\mathbf{j})}=\{0\}.
Remark 3.4.

Note that, unlike the filtration {Gk(𝐣)}0kk(𝐣)\{G_{k}^{(\mathbf{j})}\}_{0\leq k\leq k^{(\mathbf{j})}_{\star}} that is increasing for kk(𝐣)k\leq k_{\star}^{(\mathbf{j})}, the mapping kdimΔk(𝐣)k\mapsto\dim\Delta_{k}^{(\mathbf{j})} is only non-decreasing in general.

Remark 3.5.

In our definition of Δk(𝐣)\Delta_{k}^{(\mathbf{j})}, we consider Lie brackets of the form adg0(𝐣)kup(0)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k}\frac{\partial}{\partial u_{p}^{(0)}}, as opposed to [2] where Lie brackets of the form adg0(𝟎)kup(0)\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{k}\frac{\partial}{\partial u_{p}^{(0)}} are used.

Proposition 3.2.

For all 𝐣\mathbf{j} such that 0=j1jm1jm0=j_{1}\leq\ldots\leq j_{m-1}\leq j_{m}, with jmj_{m} finite, if 𝐣=(j1,,ji,ji+1,,jm)\mathbf{j}=(j_{1},\ldots,j_{i},j_{i+1},\ldots,j_{m}) and 𝐣=(j1,,ji,ji+1,,jm)\mathbf{j^{\prime}}=(j_{1},\ldots,j_{i},j^{\prime}_{i+1},\ldots,j^{\prime}_{m}) for some (ji+1,,jm)(j^{\prime}_{i+1},\ldots,j^{\prime}_{m}), we have Δk(𝐣)=Δk(𝐣)\Delta_{k}^{(\mathbf{j})}=\Delta_{k}^{(\mathbf{j^{\prime}})} for all k=0,,(ji+1ji+1)1k=0,\ldots,(j_{i+1}\vee j^{\prime}_{i+1})-1.

Moreover, for all k0k\geq 0,

dimΓk(𝐣)={p=1ijp+(k+1)(mi)ifjik<ji+1,i=1,,m1𝐣ifkjm\dim\Gamma_{k}^{(\mathbf{j})}=\left\{\begin{array}[]{ll}\displaystyle\sum_{p=1}^{i}j_{p}+(k+1)(m-i)&\displaystyle\mathrm{if\leavevmode\nobreak\ }j_{i}\leq k<j_{i+1},\;i=1,\ldots,m-1\\ \displaystyle\mid\mathbf{j}\mid&\displaystyle\mathrm{if\leavevmode\nobreak\ }k\geq j_{m}\end{array}\right. (3.15)
dimΔk(𝐣){((k+1)ip=1ijp)(n+i)ifjik<ji+1,i=1,,m1((k+1)m𝐣)(n+m)ifkjm\dim\Delta_{k}^{(\mathbf{j})}\leq\left\{\begin{array}[]{ll}\displaystyle((k+1)i-\sum_{p=1}^{i}j_{p})\vee(n+i)&\displaystyle\mathrm{if\leavevmode\nobreak\ }j_{i}\leq k<j_{i+1},\;i=1,\ldots,m-1\\ \displaystyle\left((k+1)m-\mid\mathbf{j}\mid\right)\vee(n+m)&\displaystyle\mathrm{if\leavevmode\nobreak\ }k\geq j_{m}\end{array}\right. (3.16)

and we have

Gk(𝐣)=Γk(𝐣)Δk(𝐣),k0.G_{k}^{(\mathbf{j})}=\Gamma_{k}^{(\mathbf{j})}\oplus\Delta_{k}^{(\mathbf{j})},\quad\forall k\geq 0. (3.17)

Furthermore, the finite integer k(𝐣)k_{\star}^{(\mathbf{j})}, satisfying (3.5), is such that Δk(𝐣)=Δk(𝐣)(𝐣)\Delta_{k}^{(\mathbf{j})}=\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})} and Γk(𝐣)=Γk(𝐣)(𝐣)\Gamma_{k}^{(\mathbf{j})}=\Gamma_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})} for all kk(𝐣)k\geq k_{\star}^{(\mathbf{j})}.

If, in addition, dimΔk(𝐣)(𝐣)=m+n\dim\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})}=m+n, then

n+𝐣k(𝐣)jmn+𝐣m.n+\mid\mathbf{j}\mid\geq k_{\star}^{(\mathbf{j})}\geq j_{m}\wedge\frac{n+\mid\mathbf{j}\mid}{m}. (3.18)
Proof.

By definition, the generators of Γk(𝐣)\Gamma_{k}^{(\mathbf{j})} are independent for all kk and thus their number is equal to dimΓk(𝐣)\dim\Gamma_{k}^{(\mathbf{j})}, hence (3.15). The dimension of Δk(𝐣)\Delta_{k}^{(\mathbf{j})}, in turn, is lesser than, or equal to, the number of its generators, in number (k+1)++(kji+1)=(k+1)ip=1ijp(k+1)+\ldots+(k-j_{i}+1)=(k+1)i-\sum_{p=1}^{i}j_{p}, if jik<ji+1j_{i}\leq k<j_{i+1} (respectively (k+1)++(kjm+1)=(k+1)m𝐣(k+1)+\ldots+(k-j_{m}+1)=(k+1)m-\mid\mathbf{j}\mid, if kjmk\geq j_{m}), and, since, according to (3.9) of Lemma 3.1, Δk(𝐣)\Delta_{k}^{(\mathbf{j})} is contained in {u1(0),,ui(0),x1,,xn}\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{i}^{(0)}},\frac{\partial}{\partial x_{1}},\ldots,\frac{\partial}{\partial x_{n}}\} if jik<ji+1j_{i}\leq k<j_{i+1}, i=1,,m1i=1,\ldots,m-1 (respectively in {u1(0),,um(0),x1,,xn}\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{m}^{(0)}},\frac{\partial}{\partial x_{1}},\ldots,\frac{\partial}{\partial x_{n}}\} if kjmk\geq j_{m}), its dimension is bounded above by i+ni+n (resp. m+nm+n), hence (3.16)

The proof of (3.17) is by induction. For k=0k=0, by (3.12)-(3.13), we indeed have G0(𝐣)={u1(0),u2(j2),,um(jm)}={u1(0)}{u2(j2),,um(jm)}=Δ0(𝐣)Γ0(𝐣)G_{0}^{(\mathbf{j})}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(j_{2})}},\ldots,\frac{\partial}{\partial u_{m}^{(j_{m})}}\right\}=\left\{\frac{\partial}{\partial u_{1}^{(0)}}\right\}\oplus\left\{\frac{\partial}{\partial u_{2}^{(j_{2})}},\ldots,\frac{\partial}{\partial u_{m}^{(j_{m})}}\right\}=\Delta_{0}^{(\mathbf{j})}\oplus\Gamma_{0}^{(\mathbf{j})}. Thus, (3.17) is valid at the order 0.

Assume now that (3.17) holds true up to the order ν>0\nu>0 with jrν<jr+1j_{r}\leq\nu<j_{r+1} for some r{1,,m}r\in\{1,\ldots,m\}, assuming that jr<jr+1j_{r}<j_{r+1}. If jr=jr+1j_{r}=j_{r+1}, the reader may immediately go to the case ν+1=jr+1\nu+1=j_{r+1} below.

At the order ν+1\nu+1, two cases are possible: either jrν+1<jr+1j_{r}\leq\nu+1<j_{r+1} or ν+1=jr+1\nu+1=j_{r+1}. In the first case, using Lemma 3.1, we get:

Gν+1(𝐣)\displaystyle G_{\nu+1}^{(\mathbf{j})} =Gν(𝐣)+adg0(𝐣)Gν(𝐣)=Γν(𝐣)Δν(𝐣)+adg0(𝐣)Γν(𝐣)+adg0(𝐣)Δν(𝐣)\displaystyle=G_{\nu}^{(\mathbf{j})}+ad_{g_{0}^{(\mathbf{j})}}G_{\nu}^{(\mathbf{j})}=\Gamma_{\nu}^{(\mathbf{j})}\oplus\Delta_{\nu}^{(\mathbf{j})}+ad_{g_{0}^{(\mathbf{j})}}\Gamma_{\nu}^{(\mathbf{j})}+ad_{g_{0}^{(\mathbf{j})}}\Delta_{\nu}^{(\mathbf{j})}
=Γν(𝐣)Δν(𝐣)+{adg0(𝐣)ν+1u1(0),,adg0(𝐣)ν+1jrur(0),ur+1(jr+1ν1),,um(jmν1)}\displaystyle=\Gamma_{\nu}^{(\mathbf{j})}\oplus\Delta_{\nu}^{(\mathbf{j})}+\left\{ad_{g_{0}^{(\mathbf{j})}}^{\nu+1}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,ad_{g_{0}^{(\mathbf{j})}}^{\nu+1-j_{r}}\frac{\partial}{\partial u_{r}^{(0)}},\frac{\partial}{\partial u_{r+1}^{(j_{r+1}-\nu-1)}},\ldots,\frac{\partial}{\partial u_{m}^{(j_{m}-\nu-1)}}\right\}
=Γν+1(𝐣)Δν+1(𝐣).\displaystyle=\Gamma_{\nu+1}^{(\mathbf{j})}\oplus\Delta_{\nu+1}^{(\mathbf{j})}.

In the second case, namely if ν+1=jr+1\nu+1=j_{r+1},

Gν+1(𝐣)\displaystyle G_{\nu+1}^{(\mathbf{j})} =Γν(𝐣)Δν(𝐣)\displaystyle=\Gamma_{\nu}^{(\mathbf{j})}\oplus\Delta_{\nu}^{(\mathbf{j})}
+{adg0(𝐣)ν+1u1(0),,adg0(𝐣)ν+1jrur(0),ur+1(0),ur+2(jr+2ν1),,um(jmν1)}\displaystyle\hskip 28.45274pt+\left\{ad_{g_{0}^{(\mathbf{j})}}^{\nu+1}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,ad_{g_{0}^{(\mathbf{j})}}^{\nu+1-j_{r}}\frac{\partial}{\partial u_{r}^{(0)}},\frac{\partial}{\partial u_{r+1}^{(0)}},\frac{\partial}{\partial u_{r+2}^{(j_{r+2}-\nu-1)}},\ldots,\frac{\partial}{\partial u_{m}^{(j_{m}-\nu-1)}}\right\}
=Γν+1(𝐣)Δν+1(𝐣).\displaystyle=\Gamma_{\nu+1}^{(\mathbf{j})}\oplus\Delta_{\nu+1}^{(\mathbf{j})}.

The case jmν+1j_{m}\leq\nu+1 follows the same lines:

Gν+1(𝐣)=Γν(𝐣)Δν(𝐣)+{adg0(𝐣)ν+1u1(0),,adg0(𝐣)ν+1jmum(0)}=Γν+1(𝐣)Δν+1(𝐣)G_{\nu+1}^{(\mathbf{j})}=\Gamma_{\nu}^{(\mathbf{j})}\oplus\Delta_{\nu}^{(\mathbf{j})}+\left\{ad_{g_{0}^{(\mathbf{j})}}^{\nu+1}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,ad_{g_{0}^{(\mathbf{j})}}^{\nu+1-j_{m}}\frac{\partial}{\partial u_{m}^{(0)}}\right\}=\Gamma_{\nu+1}^{(\mathbf{j})}\oplus\Delta_{\nu+1}^{(\mathbf{j})}

hence (3.17) is proven and the property of the number of iterations k(𝐣)k_{\star}^{(\mathbf{j})} to simultaneously saturate the dimensions of Γk(𝐣)\Gamma_{k}^{(\mathbf{j})} and Δk(𝐣)\Delta_{k}^{(\mathbf{j})} immediately follows.

Moreover, if dimΔk(𝐣)(𝐣)=m+n\dim\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})}=m+n, we must have jmk(𝐣)j_{m}\leq k_{\star}^{(\mathbf{j})} since otherwise, using definition (3.13) for k(𝐣)<jmk_{\star}^{(\mathbf{j})}<j_{m}, Δk(𝐣)(𝐣)\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})} would not contain {um(0)}\{\frac{\partial}{\partial u_{m}^{(0)}}\} and its dimension would not exceed m1+nm-1+n. Consequently, applying once more (3.16) with dimΔk(𝐣)(𝐣)=m+n\dim\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})}=m+n, we get (k(𝐣)+1)m𝐣m+n(k_{\star}^{(\mathbf{j})}+1)m-\mid\mathbf{j}\mid\geq m+n, which, combined with (3.5), immediately yields (3.18). ∎

Remark 3.6.

The inequality (3.18) reads k(𝐣)𝐣mnmk_{\star}^{(\mathbf{j})}-\frac{\mid\mathbf{j}\mid}{m}\geq\frac{n}{m} and may thus be interpreted as an estimate of the gap between k(𝐣)k_{\star}^{(\mathbf{j})} and the average value 𝐣m\frac{\mid\mathbf{j}\mid}{m} of the prolongation lengths j1,j2,,jmj_{1},j_{2},\ldots,j_{m}, provided that the prolonged system satisfies the strong accessibility rank condition dimΔk(𝐣)(𝐣)=m+n\dim\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})}=m+n.

4 Flatness by Pure Prolongation

4.1 Equivalence by Pure Prolongation

Consider the two systems (2.5) with dimx=n\dim x=n, dimy=n\dim y=n^{\prime} and dimu=dimv=m\dim u=\dim v=m.

Given arbitrary 𝐣\mathbf{j} and 𝐥\mathbf{l}, we recall that the associated prolonged vector fields are

g(𝐣)gx+i=1mp=0jiui(p+1)ui(p),γ(𝐥)γy+i=1mp=0livi(p+1)vi(p).g^{(\mathbf{j})}\triangleq g\frac{\partial}{\partial x}+\sum_{i=1}^{m}\sum_{p=0}^{j_{i}}u_{i}^{(p+1)}\frac{\partial}{\partial u_{i}^{(p)}},\qquad\gamma^{(\mathbf{l})}\triangleq\gamma\frac{\partial}{\partial y}+\sum_{i=1}^{m}\sum_{p=0}^{l_{i}}v_{i}^{(p+1)}\frac{\partial}{\partial v_{i}^{(p)}}.

The prolonged states are, respectively, x¯(𝐣)\overline{x}^{(\mathbf{j})} and y¯(𝐥)\overline{y}^{(\mathbf{l})}, and the control inputs u(𝐣+𝟏)u^{(\mathbf{j+1})} and v(𝐥+𝟏)v^{(\mathbf{l+1})}.

Definition 4.1.

The systems (2.5) are equivalent by pure prolongation (in short P2P^{2}-equivalent) at a point x¯0X×m\overline{x}_{0}\in X\times\mathbb{R}^{m}_{\infty} if, and only if, there exist finite 𝐣\mathbf{j} and 𝐥\mathbf{l} such that the prolonged systems of order 𝐣\mathbf{j} and 𝐥\mathbf{l} respectively are equivalent by diffeomorphism and feedback, i.e. if, and only if, there exists a local diffeomorphism φ\varphi and a feedback WW:

y¯(𝐥)=φ(x¯(𝐣))v(𝐥+𝟏)=W(x¯(𝐣),u(𝐣+𝟏))\overline{y}^{(\mathbf{l})}=\varphi(\overline{x}^{(\mathbf{j})})\qquad v^{(\mathbf{l+1})}=W(\overline{x}^{(\mathbf{j})},u^{(\mathbf{j+1})})

with WW invertible with respect to u(𝐣+𝟏)u^{(\mathbf{j+1})} for all x¯(𝐣)\overline{x}^{(\mathbf{j})} in a suitable neighborhood of x¯0\overline{x}_{0}, such that

γ(𝐥)(φ×W)=φg(𝐣).\gamma^{(\mathbf{l})}\circ(\varphi\times W)=\varphi_{\star}g^{(\mathbf{j})}.

This equivalence indeed implies that n+𝐣=n+𝐥n+\mid\mathbf{j}\mid=n^{\prime}+\mid\mathbf{l}\mid.

Remark 4.1.

The equivalence relation by diffeomorphism and feedback is easily seen to be strictly finer than the P2P^{2}-equivalence (take 𝐣=𝐥=𝟎\mathbf{j}=\mathbf{l}=\mathbf{0}), which in turn is strictly finer than the Lie-Bäcklund equivalence (see e.g. Example 5.4).

Definition 4.2.

We say that system (2.16) is flat by pure prolongation (in short P2P^{2}-flat) at a point x¯0X×m\overline{x}_{0}\in X\times\mathbb{R}^{m}_{\infty} if, and only if, it is P2P^{2}-equivalent to y˙i=vi\dot{y}_{i}=v_{i}, i=1,,mi=1,\ldots,m and yy is called P2P^{2}-flat output.

It is therefore immediate to remark that a system is P2P^{2}-flat if, and only if, there exists a pure prolongation of finite order 𝐣\mathbf{j} such that the prolonged system is feedback linearizable at x¯0\overline{x}_{0}, thus recovering the definition of linearization by prolongation already introduced in [11, 12, 2].

Moreover, a P2P^{2}-flat output being obviously a flat output and pure prolongations being particular cases of Lie-Bäcklund isomorphisms, the class of P2P^{2}-flat systems is indeed contained in the class of Lie-Bäcklund equivalence to 0 (modulo the trivial field τ\tau), i.e. constitutes a subclass of differentially flat systems.

The next Lemma extends a well-know result (see e.g. [6, 33]) to our context (see also [12]).

Lemma 4.1.

We consider system (3.2), denoted by Σ(𝐣)\Sigma^{(\mathbf{j})}, with 𝐣=(j1,,jm)\mathbf{j}=(j_{1},\ldots,j_{m}), assuming, without loss of generality, that 0j1jm0\leq j_{1}\leq\ldots\leq j_{m}, possibly up to input renumbering. We denote by 𝐣=𝐣𝐣𝟏=(0,j2j1,jmj1)\mathbf{j^{\prime}}=\mathbf{j-j_{1}}=(0,j_{2}-j_{1}\ldots,j_{m}-j_{1}), and by Σ(𝐣)\Sigma^{(\mathbf{j^{\prime}})} the corresponding system. Then Σ(𝐣)\Sigma^{(\mathbf{j})} is P2P^{2}-flat at a given point (x0,u¯0)(x_{0},\overline{u}_{0}) if, and only if, Σ(𝐣)\Sigma^{(\mathbf{j^{\prime}})} is also P2P^{2}-flat at this point. Moreover, every P2P^{2}-flat output of Σ(𝐣)\Sigma^{(\mathbf{j})} at (x0,u¯0)(x_{0},\overline{u}_{0}) is a P2P^{2}-flat output of Σ(𝐣)\Sigma^{(\mathbf{j^{\prime}})} at the same point, and conversely.

Proof.

Since Σ(𝐣)\Sigma^{(\mathbf{j})} is P2P^{2}-flat, there exists 𝐣𝟎\mathbf{j_{0}} and 𝐥𝟎\mathbf{l_{0}} such that Σ(𝐣+𝐣𝟎)\Sigma^{(\mathbf{j+j_{0}})} is feedback equivalent to the linear system y(𝐥𝟎+𝟏)=wy^{(\mathbf{l_{0}+1})}=w. Therefore, since, by assumption, Σ(𝐣+𝐣𝟎)=Σ(𝐣+𝐣𝟏+𝐣𝟎)\Sigma^{(\mathbf{j+j_{0}})}=\Sigma^{(\mathbf{j^{\prime}+j_{1}+j_{0}})}, Σ(𝐣)\Sigma^{(\mathbf{j^{\prime}})} is also P2P^{2}-flat. The converse is trivial as well as the fact that Σ(𝐣+𝐣𝟎)\Sigma^{(\mathbf{j+j_{0}})} and Σ(𝐣+𝐣𝟏+𝐣𝟎)\Sigma^{(\mathbf{j^{\prime}+j_{1}+j_{0}})} have the same P2P^{2}-flat output. ∎

Remark 4.2.

In the single input case (m=1m=1), this lemma shows that a P2P^{2}-flat system is equivalent, without prolongation, to a linear system. We thus trivially recover the fact that P2P^{2}-flatness and feedback linearizability are equivalent in this case (see [6] and remark 2.2)

4.2 Necessary and Sufficient Conditions

The following Proposition is a straightforward adaptation of Theorem 2.2 for an arbitrary order 𝐣\mathbf{j}. Note that, at this stage, nothing is said about a possible choice of 𝐣\mathbf{j}, a question that will be dealt with in subsection 4.3, theorem 4.2.

Proposition 4.1.

The prolonged system of order 𝐣\mathbf{j} is feedback linearizable at 0¯\overline{0} if, and only if, Gk(𝐣)G_{k}^{(\mathbf{j})} is involutive with locally constant dimension for all kk and such that Gk(𝐣)(𝐣)=TX(𝐣)G_{k^{(\mathbf{j})}_{\ast}}^{(\mathbf{j})}=\mathrm{T}X^{(\mathbf{j})}.

Again, flat outputs can be computed via Frobenius theorem, once established the list of Brunovský’s controllability indices of order 𝐣\mathbf{j}, as follows:

Definition 4.3.

Consider the sequence of integers

ρk(𝐣)dimGk(𝐣)/Gk1(𝐣)k1,ρ0(𝐣)dimG0(𝐣)=m.\rho_{k}^{(\mathbf{j})}\triangleq\dim G_{k}^{(\mathbf{j})}/G_{k-1}^{(\mathbf{j})}\quad\forall k\geq 1,\qquad\rho_{0}^{(\mathbf{j})}\triangleq\dim G_{0}^{(\mathbf{j})}=m.

The Brunovský controllability indices of order 𝐣\mathbf{j} are defined by

κk(𝐣)#{lρl(𝐣)k},k=1,,m,\kappa_{k}^{(\mathbf{j})}\triangleq\#\{l\mid\rho_{l}^{(\mathbf{j})}\geq k\},\quad k=1,\ldots,m,

As in the 0th order case, if the prolonged system of order 𝐣\mathbf{j} is feedback linearizable at 0¯\overline{0}, we have:

  • ρk(𝐣)\rho_{k}^{(\mathbf{j})}’s and κk(𝐣)\kappa_{k}^{(\mathbf{j})}’s are non increasing sequences,

  • ρk(𝐣)m\rho_{k}^{(\mathbf{j})}\leq m for all kk and ρk(𝐣)=0\rho_{k}^{(\mathbf{j})}=0 for all kk(𝐣)+1k\geq k^{(\mathbf{j})}_{\ast}+1,

  • κ1(𝐣)=k(𝐣)+1,κm(𝐣)1\kappa_{1}^{(\mathbf{j})}=k^{(\mathbf{j})}_{\ast}+1,\quad\kappa_{m}^{(\mathbf{j})}\geq 1,

  • k=0k(𝐣)ρk(𝐣)=k=1mκk(𝐣)=dimGk(𝐣)(𝐣)=n+m+𝐣\displaystyle\sum_{k=0}^{k^{(\mathbf{j})}_{\ast}}\rho_{k}^{(\mathbf{j})}=\sum_{k=1}^{m}\kappa_{k}^{(\mathbf{j})}=\dim G_{k^{(\mathbf{j})}_{\ast}}^{(\mathbf{j})}=n+m+\mid\mathbf{j}\mid.

The list κ1(𝐣),,κm(𝐣)\kappa_{1}^{(\mathbf{j})},\ldots,\kappa_{m}^{(\mathbf{j})} is uniquely defined up to input permutation, invariant by prolonged state feedback and prolonged state diffeomorphism, and is thus equal to the list of controllability indices of the associated linear system (2.15) with ri=κi(𝐣)r_{i}=\kappa_{i}^{(\mathbf{j})}, i=1,,mi=1,\ldots,m.

Moreover, for all kk and all i=1,,mi=1,\ldots,m, and possibly up to a suitable input reordering, we have

Gk(𝐣)=p=1m{adg0(𝐣)lup(jp)l=0,k(κp(𝐣)1)},Gκ1(𝐣)1(𝐣)=Gk(𝐣)(𝐣)=TX(𝐣)G_{k}^{(\mathbf{j})}=\bigoplus_{p=1}^{m}\left\{\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{l}\frac{\partial}{\partial u_{p}^{(j_{p})}}\mid l=0,\ldots k\lor(\kappa_{p}^{(\mathbf{j})}-1)\right\},\quad G_{\kappa_{1}^{(\mathbf{j})}-1}^{(\mathbf{j})}=G_{k^{(\mathbf{j})}_{\ast}}^{(\mathbf{j})}=\mathrm{T}X^{(\mathbf{j})}

and flat outputs (y1,,ym)(y_{1},\ldots,y_{m}) are locally non trivial solutions of the system of PDE’s

Gk(𝐣),dyi=0,k=0,,κi(𝐣)2,withGκi(𝐣)1(𝐣),dyi0,i=1,,m.\left<G_{k}^{(\mathbf{j})},dy_{i}\right>=0,\;k=0,\ldots,\kappa_{i}^{(\mathbf{j})}-2,\quad\mathrm{with\leavevmode\nobreak\ }\quad\left<G_{\kappa_{i}^{(\mathbf{j})}-1}^{(\mathbf{j})},dy_{i}\right>\neq 0,\quad i=1,\ldots,m. (4.1)

Finally, the mapping

x¯(𝐣)(y1,,y1(κ1(𝐣)1),,ym,,ym(κm(𝐣)1))\overline{x}^{(\mathbf{j})}\mapsto(y_{1},\ldots,y_{1}^{(\kappa_{1}^{(\mathbf{j})}-1)},\ldots,y_{m},\ldots,y_{m}^{(\kappa_{m}^{(\mathbf{j})}-1)})

is a local diffeomorphism.


In virtue of Lemma 4.1, it suffices to restrict our analysis to prolongations of order 𝐣=(j1,,jm)\mathbf{j}=(j_{1},\ldots,j_{m}) such that 0=j1,jm0=j_{1}\leq\ldots,\leq j_{m}.


We are now ready to state our main result.

Theorem 4.1.

A necessary and sufficient condition for P2P^{2}-flatness at 0¯\overline{0} is that there exists 𝐣=(j1,,jm)m\mathbf{j}=(j_{1},\ldots,j_{m})\in\mathbb{N}^{m}, 0=j1,jm0=j_{1}\leq\ldots,\leq j_{m}, such that

  • (i)

    Δk(𝐣)¯=Δk(𝐣)\overline{\Delta_{k}^{(\mathbf{j})}}=\Delta_{k}^{(\mathbf{j})} with dimΔk(𝐣)\dim\Delta_{k}^{(\mathbf{j})} locally constant for all k0k\geq 0,

  • (ii)

    [Γk(𝐣),Δk(𝐣)]Δk(𝐣)[\Gamma_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\subset\Delta_{k}^{(\mathbf{j})} for all k0k\geq 0,

  • (iii)

    k(𝐣)k^{(\mathbf{j})}_{\ast} is such that Δk(𝐣)=TX×Tm\Delta_{k}^{(\mathbf{j})}=\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m} and Γk(𝐣)=T𝐣\Gamma_{k}^{(\mathbf{j})}=\mathrm{T}\mathbb{R}^{\mid\mathbf{j}\mid} for all kk(𝐣)k\geq k^{(\mathbf{j})}_{\ast}.

Proof.

By (3.17) of Proposition 3.2, we have Gk(𝐣)=Γk(𝐣)Δk(𝐣)G_{k}^{(\mathbf{j})}=\Gamma_{k}^{(\mathbf{j})}\oplus\Delta_{k}^{(\mathbf{j})} for all k0k\geq 0. Then, Gk(𝐣)=Gk(𝐣)¯G_{k}^{(\mathbf{j})}=\overline{G_{k}^{(\mathbf{j})}} implies that [Γk(𝐣)Δk(𝐣),Γk(𝐣)Δk(𝐣)]Γk(𝐣)Δk(𝐣)[\Gamma_{k}^{(\mathbf{j})}\oplus\Delta_{k}^{(\mathbf{j})},\Gamma_{k}^{(\mathbf{j})}\oplus\Delta_{k}^{(\mathbf{j})}]\subset\Gamma_{k}^{(\mathbf{j})}\oplus\Delta_{k}^{(\mathbf{j})}. Since Γk(𝐣)=Γk(𝐣)¯\Gamma_{k}^{(\mathbf{j})}=\overline{\Gamma_{k}^{(\mathbf{j})}} for all kk, and since [Γk(𝐣),Δk(𝐣)]Γk(𝐣)={0}[\Gamma_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\cap\Gamma_{k}^{(\mathbf{j})}=\{0\} by Lemma 3.1 and (3.14), we deduce that [Γk(𝐣),Δk(𝐣)]+[Δk(𝐣),Δk(𝐣)]Δk(𝐣)[\Gamma_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]+[\Delta_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\subset\Delta_{k}^{(\mathbf{j})}, hence [Γk(𝐣),Δk(𝐣)]Δk(𝐣)[\Gamma_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\subset\Delta_{k}^{(\mathbf{j})} and [Δk(𝐣),Δk(𝐣)]Δk(𝐣)[\Delta_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\subset\Delta_{k}^{(\mathbf{j})}, i.e. Δk(𝐣)=Δk(𝐣)¯\Delta_{k}^{(\mathbf{j})}=\overline{\Delta_{k}^{(\mathbf{j})}}, for all k0k\geq 0.

Conversely, [Γk(𝐣),Δk(𝐣)]Δk(𝐣)[\Gamma_{k}^{(\mathbf{j})},\Delta_{k}^{(\mathbf{j})}]\subset\Delta_{k}^{(\mathbf{j})} and Δk(𝐣)=Δk(𝐣)¯\Delta_{k}^{(\mathbf{j})}=\overline{\Delta_{k}^{(\mathbf{j})}} for all k0k\geq 0 trivially implies that Gk(𝐣)=Gk(𝐣)¯G_{k}^{(\mathbf{j})}=\overline{G_{k}^{(\mathbf{j})}} for all k0k\geq 0.

Moreover, since dimΓk(𝐣)\dim\Gamma_{k}^{(\mathbf{j})} is constant by construction, the fact that Gk(𝐣)G_{k}^{(\mathbf{j})} has locally constant dimension is equivalent to the fact that Δk(𝐣)\Delta_{k}^{(\mathbf{j})} has locally constant dimension too for all k0k\geq 0, hence (i).

Finally, (iii) is an immediate consequence of the condition that Gk(𝐣)=TX×Tm+𝐣G_{k}^{(\mathbf{j})}=\mathrm{T}X\times\mathrm{T}\mathbb{R}^{m+\mid\mathbf{j}\mid} for all kk(𝐣)k\geq k^{(\mathbf{j})}_{\ast}, and the theorem is proven. ∎

Remark 4.3.

Expressed in words, the above necessary and sufficient conditions read (i) involutivity with locally constant dimension of the Δk(𝐣)\Delta_{k}^{(\mathbf{j})}’s, (ii) invariance of Δk(𝐣)\Delta_{k}^{(\mathbf{j})} by Γk(𝐣)\Gamma_{k}^{(\mathbf{j})} and (iii) strong controllability rank condition (see e.g. [35, 13]).

4.3 The Pure Prolongation Algorithm

From now on, for every sequence 𝐥=(l1,,lm)m\mathbf{l}=\left(l_{1},\ldots,l_{m}\right)\in\mathbb{N}^{m}, we systematically re-order the indices {1,,m}\{1,\ldots,m\} by a suitable permutation α:{1,,m}{1,,m}\alpha:\{1,\ldots,m\}\rightarrow\{1,\ldots,m\} such that 0=lα(1)lα(m)0=l_{\alpha(1)}\leq\ldots\leq l_{\alpha(m)}. Moreover, for simplicity’s sake, the permutation α\alpha will be omitted. We will thus abusively replace lα(i)l_{\alpha(i)} by lil_{i}, i.e. 0=l1lm0=l_{1}\leq\ldots\leq l_{m}.

In the following algorithm we assume that the computations are done in a suitable open dense neighborhood of 0¯\overline{0} where all the distributions involved have constant dimension.

4.3.1 Initialization

Consider the filtration {Gk(0)k0}\{G_{k}^{(0)}\mid k\geq 0\} defined by (2.19). If every Gk(0)G_{k}^{(0)} satisfies the conditions of theorem 2.2, the system is feedback linearizable and no prolongation is needed. In particular, for m=1m=1, a case where flatness and feedback linearizability are equivalent, the results of this section are pointless.

Otherwise, we have the following alternative:

  • either there must exist k0nk_{0}\leq n such that Gk0(0)G_{k_{0}}^{(0)} is not involutive while every Gk(0)G_{k}^{(0)} is involutive for k<k0k<k_{0} (note that k01k_{0}\geq 1 since G0(𝟎)G_{0}^{(\mathbf{0})} is always involutive),

  • or the Gk(0)G_{k}^{(0)}’s are all involutive but with maxk0dimGk(0)=dimGk(𝟎)(0)<n+m\max_{k\geq 0}\dim G_{k}^{(0)}=\dim G_{k_{\star}^{(\mathbf{0})}}^{(0)}<n+m.

First Case.

There exists k0{1,,n}k_{0}\in\{1,\ldots,n\}, first index for which Gk0(0)G_{k_{0}}^{(0)} is not involutive.

Let H1(𝟎){0}H_{1}^{(\mathbf{0})}\neq\{0\} be any involutive distribution included in G1(𝟎)G_{1}^{(\mathbf{0})}, of the form

H1(𝟎){u1(0),,adg0(0)u1(0),,up0(0),,adg0(0)up0(0)}H_{1}^{(\mathbf{0})}\triangleq\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(0)}}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{p_{0}}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(0)}}\frac{\partial}{\partial u_{p_{0}}^{(0)}}\} (4.2)

for some p0mp_{0}\leq m.

By (A.3), adg0(0)up(0)=adg0(𝐥)up(0)\mathrm{ad}_{g_{0}^{(0)}}\frac{\partial}{\partial u_{p}^{(0)}}=\mathrm{ad}_{g_{0}^{(\mathbf{l})}}\frac{\partial}{\partial u_{p}^{(0)}} for all 𝐥\mathbf{l} and, thanks to (3.10), we have [uq(lqk),adg0(0)up(0)]=0[\frac{\partial}{\partial u_{q}^{(l_{q}-k)}},\mathrm{ad}_{g_{0}^{(0)}}\frac{\partial}{\partial u_{p}^{(0)}}]=0 for all lqk>1l_{q}-k>1 with q>p0q>p_{0} and all p=1,,p0p=1,\ldots,p_{0}. Therefore we can choose

Δ1(𝐥){u1(0),,adg0(𝐥)u1(0),,up0(0),,adg0(𝐥)up0(0)}=H1(𝟎)\Delta_{1}^{(\mathbf{l})}\triangleq\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(\mathbf{l})}}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{p_{0}}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(\mathbf{l})}}\frac{\partial}{\partial u_{p_{0}}^{(0)}}\}=H_{1}^{(\mathbf{0})} (4.3)

and

Γ1(𝐥){up0+1(lp0+1),up0+1(lp0+11),,um(lm),um(lm1)}\Gamma_{1}^{(\mathbf{l})}\triangleq\{\frac{\partial}{\partial u_{p_{0}+1}^{(l_{p_{0}+1})}},\frac{\partial}{\partial u_{p_{0}+1}^{(l_{p_{0}+1}-1)}},\ldots,\frac{\partial}{\partial u_{m}^{(l_{m})}},\frac{\partial}{\partial u_{m}^{(l_{m}-1)}}\} (4.4)

for any

𝐥=(0,,0p0,lp0+1,,lm),2=lp0+1lm\mathbf{l}=\left(\underbrace{0,\ldots,0}_{p_{0}},l_{p_{0}+1},\ldots,l_{m}\right),\quad 2=l_{p_{0}+1}\leq\ldots\leq l_{m} (4.5)

since they naturally satisfy Δ1(𝐥)¯=H1(𝟎)¯=H1(𝟎)=Δ1(𝐥)\overline{\Delta_{1}^{(\mathbf{l})}}=\overline{H_{1}^{(\mathbf{0})}}=H_{1}^{(\mathbf{0})}=\Delta_{1}^{(\mathbf{l})} and [Γ1(𝐥),Δ1(𝐥)]Δ1(𝐥)[\Gamma_{1}^{(\mathbf{l})},\Delta_{1}^{(\mathbf{l})}]\subset\Delta_{1}^{(\mathbf{l})} for all 𝐥\mathbf{l} given by (4.5), i.e. with lp0+1=2l_{p_{0}+1}=2.

However if, for lp0+1=1l_{p_{0}+1}=1,

Δ1(𝐥){u1(0),,adg0(𝐥)u1(0),,up0(0),,adg0(𝐥)up0(0),up0+1(0)}\Delta_{1}^{(\mathbf{l})}\triangleq\{\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(\mathbf{l})}}\frac{\partial}{\partial u_{1}^{(0)}},\ldots,\frac{\partial}{\partial u_{p_{0}}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(\mathbf{l})}}\frac{\partial}{\partial u_{p_{0}}^{(0)}},\frac{\partial}{\partial u_{p_{0}+1}^{(0)}}\} (4.6)

and

Γ1(𝐥){up0+1(1),up0+2(l2),up0+2(l21),,um(lm),um(lm1)}\Gamma_{1}^{(\mathbf{l})}\triangleq\{\frac{\partial}{\partial u_{p_{0}+1}^{(1)}},\frac{\partial}{\partial u_{p_{0}+2}^{(l_{2})}},\frac{\partial}{\partial u_{p_{0}+2}^{(l_{2}-1)}},\ldots,\frac{\partial}{\partial u_{m}^{(l_{m})}},\frac{\partial}{\partial u_{m}^{(l_{m}-1)}}\} (4.7)

also satisfy Δ1(𝐥)¯=Δ1(𝐥)\overline{\Delta_{1}^{(\mathbf{l})}}=\Delta_{1}^{(\mathbf{l})} and [Γ1(𝐥),Δ1(𝐥)]Δ1(𝐥)[\Gamma_{1}^{(\mathbf{l})},\Delta_{1}^{(\mathbf{l})}]\subset\Delta_{1}^{(\mathbf{l})}, the previous choice (4.3)-(4.4)-(4.5) maybe replaced by (4.6)-(4.7) with 1=lp0+1lm1=l_{p_{0}+1}\leq\ldots\leq l_{m}.

These choices, namely (4.3)-(4.4)-(4.5), or (4.6)-(4.7) with lp0+1=1l_{p_{0}+1}=1, for any H1(𝟎)H_{1}^{(\mathbf{0})} given by (4.2), constitutes the initialization of the algorithm.

Note that there are at most p=1m1Cmp=p=1m1m!p!(mp)!\sum_{p=1}^{m-1}{\mathrm{C}}_{m}^{p}=\sum_{p=1}^{m-1}\frac{m!}{p!(m-p)!} possibilities of such initialization.

Second Case.
Lemma 4.2.

Assume that the Gk(0)G_{k}^{(0)}’s are all involutive with maxk0dimGk(𝟎)<n+m\max_{k\geq 0}\dim G_{k}^{(\mathbf{0})}<n+m. Then Gk(𝐣)=Gk(𝐣)¯=Gk(𝟎)G_{k}^{(\mathbf{j})}=\overline{G_{k}^{(\mathbf{j})}}=G_{k}^{(\mathbf{0})} for all kk and 𝐣\mathbf{j} and maxk0dimGk(𝐣)<m+n\max_{k\geq 0}\dim G_{k}^{(\mathbf{j})}<m+n for all 𝐣\mathbf{j}.

Proof.

According to (A.3) of lemma A.1, every adg0(𝐣)kup(jp)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k}\frac{\partial}{\partial u_{p}^{(j_{p})}} is equal to ±adg0(𝟎)kjpup(0)\pm\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{k-j_{p}}\frac{\partial}{\partial u_{p}^{(0)}} + terms in Gk1(𝟎)¯=Gk1(𝟎)\overline{G_{k-1}^{(\mathbf{0})}}=G_{k-1}^{(\mathbf{0})}, thanks to the involutivity of the Gk(0)G_{k}^{(0)}’s. Therefore, Gk(𝐣)Gk(𝟎)G_{k}^{(\mathbf{j})}\subset G_{k}^{(\mathbf{0})}, the converse inclusion being immediate using (A.3). Thus dimGk(𝟎)=dimGk(𝐣)\dim G_{k}^{(\mathbf{0})}=\dim G_{k}^{(\mathbf{j})} for all kk and 𝐣\mathbf{j}, Q.E.D. ∎

We immediately conclude that, in this case, the strong controllability rank condition does not hold, which contradicts (iii) of theorem 2.2, hence the non-flatness by pure prolongation. Therefore, this case must be discarded.

4.3.2 Recursion

For all k1k\geq 1 and all 𝐥\mathbf{l} given by:

𝐥=(0,,0p0,lp0+1,,lm),1lp0+1lm,\mathbf{l}=\left(\underbrace{0,\ldots,0}_{p_{0}},l_{p_{0}+1},\ldots,l_{m}\right),\quad 1\leq l_{p_{0}+1}\leq\ldots\leq l_{m}, (4.8)

and denoting by Cmin𝐥L(lp0+1,,lm)C\!\!-\!\min_{\mathbf{l}\in L}(l_{p_{0}+1},\ldots,l_{m}) (resp. Cmax𝐥L(lp0+1,,lm)C\!\!-\!\max_{\mathbf{l}\in L}(l_{p_{0}+1},\ldots,l_{m})) the componentwise minimum (resp. maximum) of a collection of (mp0)(m-p_{0})-tuples (lp0+1,,lm)(l_{p_{0}+1},\ldots,l_{m}) in a set LL, with 1lp0+1lm1\leq l_{p_{0}+1}\leq\ldots\leq l_{m}, let us introduce the following numbers

σΔ(k)\displaystyle\sigma_{\Delta}(k) (σp0+1,Δ(k),,σm,Δ(k))\displaystyle\triangleq\left(\sigma_{p_{0}+1,\Delta}(k),\ldots,\sigma_{m,\Delta}(k)\right) (4.9)
Cmin{1lp0+1,lmΔk(𝐥)¯=Δk(𝐥)},\displaystyle\triangleq C\!\!-\!\min\{1\leq l_{p_{0}+1},\leq\ldots\leq l_{m}\mid\overline{\Delta_{k}^{(\mathbf{l})}}=\Delta_{k}^{(\mathbf{l})}\},
σΓ,Δ(k)\displaystyle\sigma_{\Gamma,\Delta}(k) (σp0+1,Γ,Δ(k),,σm,Γ,Δ(k))\displaystyle\triangleq\left(\sigma_{p_{0}+1,\Gamma,\Delta}(k),\ldots,\sigma_{m,\Gamma,\Delta}(k)\right) (4.10)
Cmin{1lp0+1lm[Γk(𝐥),Δk(𝐥)]Δk(𝐥)}\displaystyle\triangleq C\!\!-\!\min\{1\leq l_{p_{0}+1}\leq\ldots\leq l_{m}\mid[\Gamma_{k}^{(\mathbf{l})},\Delta_{k}^{(\mathbf{l})}]\subset\Delta_{k}^{(\mathbf{l})}\}

and

(jp0+1,,jm)\displaystyle\left(j_{p_{0}+1},\ldots,j_{m}\right) (4.11)
Cmaxk1(σp0+1,Δ(k)σp0+1,Γ,Δ(k),σm,Δ(k)σm,Γ,Δ(k)).\displaystyle\qquad\triangleq C\!\!-\!\max_{k\geq 1}\left(\sigma_{p_{0}+1,\Delta}(k)\wedge\sigma_{p_{0}+1,\Gamma,\Delta}(k)\leq\ldots,\leq\sigma_{m,\Delta}(k)\wedge\sigma_{m,\Gamma,\Delta}(k)\right).
Remark 4.4.

Thanks to (4.5)-(4.3)-(4.4), we get

σp0+1,Δ(1)σp0+1,Γ,Δ(1)2\sigma_{p_{0}+1,\Delta}(1)\wedge\sigma_{p_{0}+1,\Gamma,\Delta}(1)\leq 2

and if Δ1(𝐥)¯=Δ1(𝐥)\overline{\Delta_{1}^{(\mathbf{l})}}=\Delta_{1}^{(\mathbf{l})} and [Γ1(𝐥),Δ1(𝐥)]Δ1(𝐥)[\Gamma_{1}^{(\mathbf{l})},\Delta_{1}^{(\mathbf{l})}]\subset\Delta_{1}^{(\mathbf{l})}, with (4.6)-(4.7) and lp0+1,,lm1l_{p_{0}+1},\ldots,l_{m}\geq 1, we get

σp0+1,Δ(1)σp0+1,Γ,Δ(1)=1.\sigma_{p_{0}+1,\Delta}(1)\wedge\sigma_{p_{0}+1,\Gamma,\Delta}(1)=1.
Remark 4.5.

In view of the definition (3.13) of Δk(𝐥)\Delta_{k}^{(\mathbf{l})}, if lqk<lq+1l_{q}\leq k<l_{q+1} for some q{1,,m}q\in\{1,\ldots,m\} and k1k\geq 1, Δk(𝐥)\Delta_{k}^{(\mathbf{l})} does not depend on lq+1,,lml_{q+1},\ldots,l_{m}. Thus σΔ(k)=(σ1,Δ(k),,σq,Δ(k),σq,Δ(k),,σq,Δ(k)mq)\sigma_{\Delta}(k)=\left(\sigma_{1,\Delta}(k),\ldots,\sigma_{q,\Delta}(k),\underbrace{\sigma_{q,\Delta}(k),\ldots,\sigma_{q,\Delta}(k)}_{m-q}\right).

Proposition 4.2.

We have

[Γk(𝐥),Δk(𝐥)]={0},andσp0+1,Γ,Δ(k)2klp0+1>2k1,k0.[\Gamma_{k}^{(\mathbf{l})},\Delta_{k}^{(\mathbf{l})}]=\{0\},\quad\mathrm{and}\quad\sigma_{p_{0}+1,\Gamma,\Delta}(k)\leq 2k\quad\forall l_{p_{0}+1}>2k-1,\;\forall k\geq 0. (4.12)

Moreover, Δk(𝐥)\Delta_{k}^{(\mathbf{l})} is independent of lp0+1l_{p_{0}+1} if lp0+1>kl_{p_{0}+1}>k. Therefore, if Δk(𝐥)\Delta_{k}^{(\mathbf{l})} is involutive for all kk, we get

σp0+1,Δ(k)σp0+1,Γ,Δ(k)2k\sigma_{p_{0}+1,\Delta}(k)\wedge\sigma_{p_{0}+1,\Gamma,\Delta}(k)\leq 2k (4.13)

and jp0+1jm<+j_{p_{0}+1}\leq\ldots\leq j_{m}<+\infty. The minimal prolongation orders are thus obtained by minimizing (4.11) over all initializations built on (4.2).

Proof.

(4.13) is an immediate consequence of (3.10) and of the definitions of Δk(𝐥)\Delta_{k}^{(\mathbf{l})} and Γk(𝐥)\Gamma_{k}^{(\mathbf{l})}. Moreover, since, according to remark 3.4, kdimΔk(𝐥)k\mapsto\dim\Delta_{k}^{(\mathbf{l})} is non decreasing and bounded by n+mn+m, then there exists a finite KK^{\star} such that ΔK(𝐥)=Δk(𝐥)\Delta_{K^{\star}}^{(\mathbf{l})}=\Delta_{k}^{(\mathbf{l})} for all kKk\geq K^{\star} and thus the maximum with respect to kk in (4.11) is achieved for k=Kk=K^{\star}. The C-minimal prolongation orders are thus obtained by varying the initializations built on (4.2). ∎

We thus have proven:

Theorem 4.2.

We have the following alternative:

  1. 1.

    If, for every choice of initialization built on (4.2), there exists a kk for which maxp{p0+1,,m}σp,Δ(k)=+\max_{p\in\{p_{0}+1,\ldots,m\}}\sigma_{p,\Delta}(k)=+\infty or if maxkdimΔk(𝐣)<n+m\max_{k}\dim\Delta_{k}^{(\mathbf{j})}<n+m for all 𝐣\mathbf{j}, then the system is not flat by pure prolongation.

  2. 2.

    Otherwise, 𝐣\mathbf{j} is finite and given by (4.11) and the minimal prolongation orders are obtained by C-minimizing (4.11) over all initializations.

Proof.

Straightforward from what precedes. ∎

We immediately deduce the following algorithm:

Algorithm 1. flatness by pure prolongation

Input:

the vector fields g0(𝟎),g1(0)=u1(0),,gm(0)=um(0)g_{0}^{(\mathbf{0})},g_{1}^{(0)}=\frac{\partial}{\partial u_{1}^{(0)}},\ldots,g_{m}^{(0)}=\frac{\partial}{\partial u_{m}^{(0)}} (see (2.17))

output:

the minimal lengths 0=j1jm0=j_{1}\leq\ldots\leq j_{m} and k(𝐣)k_{\star}^{(\mathbf{j})} or fail if the system is not flat by pure prolongation.

Initialization.

Construct Gk(0)G_{k}^{(0)} for all knk\leq n. If k0n\exists k_{0}\leq n such that Gk0(0)G_{k_{0}}^{(0)} is not involutive, choose an involutive subdistribution H1(𝟎)Gk0(0)H_{1}^{(\mathbf{0})}\subset G_{k_{0}}^{(0)}, given by (4.2), with p0m1p_{0}\leq m-1 and compute Δ1(𝐥)\Delta_{1}^{(\mathbf{l})} and Γ1(𝐥)\Gamma_{1}^{(\mathbf{l})} by (4.3)-(4.4)-(4.5), or (4.6)-(4.7) with lp0+1=1l_{p_{0}+1}=1.

Step k1k\geq 1.

Compute Δk(𝐥)\Delta_{k}^{(\mathbf{l})}, Γk(𝐥)\Gamma_{k}^{(\mathbf{l})}, σΔ(k)\sigma_{\Delta}(k) and σΓ,Δ(k)\sigma_{\Gamma,\Delta}(k) by (4.9) and (4.10). Continue up to the first k1k_{1} such that σΓ,Δ(k1)σΔ(k1)\sigma_{\Gamma,\Delta}(k_{1})\bigwedge\sigma_{\Delta}(k_{1}) is C-maximal. Then 𝐣\mathbf{j} is given by (4.11).
If Δk(𝐥)\Delta_{k}^{(\mathbf{l})} is non involutive for some kk1k\leq k_{1}, change the initialization and restart and if Δk(𝐥)\Delta_{k}^{(\mathbf{l})} is non involutive for all initializations, then fail.

Step k(𝐣)k_{\star}^{(\mathbf{j})}.

Determine k(𝐣)k_{\star}^{(\mathbf{j})}. If dimΔk(𝐣)(𝐣)=n+m\dim\Delta_{k_{\star}^{(\mathbf{j})}}^{(\mathbf{j})}=n+m, stop. Otherwise, fail.

5 Examples

All the examples of this paper concern two input systems, i.e. m=2m=2, except example 5.5, with 33 inputs, resuming the example 2 of [18].

In the two input examples, the prolongation index 𝐣=(j1,j2)\mathbf{j}=(j_{1},j_{2}) is supposed to be such that 0=j1j20=j_{1}\leq j_{2}, up to a suitable input permutation. However, for the sake of readability, we will keep the original input numbering unchanged and thus consider that 𝐣=(0,j2)\mathbf{j}=(0,j_{2}) or (j1,0)(j_{1},0) depending on the context. At the exception of this modification, we strictly apply algorithm 4.2 in all the examples.

The first example gives a detailed presentation of the application of this algorithm, in particular the role played by σΓ,Δ(k)\sigma_{\Gamma,\Delta}(k) and σΔ(k)\sigma_{\Delta}(k). The second one shows the importance of the sole number σ1,Δ(k)\sigma_{1,\Delta}(k) to determine the prolongation length, and the third one, borrowed from [7], and carried on again in [2, Section 3.1], is reported here to compare our approach with the one of [2]. Then, the pendulum example is presented to show that non flat systems by pure prolongation can be detected in a finite number of steps. To conclude this section, the last example, with 3 inputs, compares the constuction proposed in [18, Example 2] with our algorithm leading to a minimal prolongation, strictly smaller than the one of [18].

5.1 Chained System [25]

x1(3)=u1\displaystyle x_{1}^{(3)}=u_{1} (5.1)
x¨2=u2\displaystyle\ddot{x}_{2}=u_{2}
x˙3=u1u2\displaystyle\dot{x}_{3}=u_{1}u_{2}

This system has been proven to be flat in [25, section 3.1.1] with the following flat output

y1=x3x¨1u2+x˙1u˙2x1u¨2,y2=x2.y_{1}=x_{3}-\ddot{x}_{1}u_{2}+\dot{x}_{1}\dot{u}_{2}-x_{1}\ddot{u}_{2},\quad y_{2}=x_{2}. (5.2)

5.1.1 Associated Non Prolonged Distributions

Let us start this section by showing that system (5.1) is not static feedback linearizable. We denote the state coordinates by (x1,1,x1,2,x1,3,x2,1,x2,2,x3,u1(0),u2(0))(x_{1,1},x_{1,2},x_{1,3},x_{2,1},x_{2,2},x_{3},u_{1}^{(0)},u_{2}^{(0)}) (n=6n=6 and m=2m=2), with x1,jx1(j)x_{1,j}\triangleq x_{1}^{(j)}, j=1,2,3j=1,2,3, x2,jx2(j)x_{2,j}\triangleq x_{2}^{(j)}, j=1,2j=1,2 and ui(0)=uiu_{i}^{(0)}=u_{i}, i=1,2i=1,2.
The two input variables are ui(1)u˙iu_{i}^{(1)}\triangleq\dot{u}_{i}, i=1,2i=1,2.

The system vector fields are

g0(0)\displaystyle g_{0}^{(0)} x1,2x1,1+x1,3x1,2+u1(0)x1,3+x2,2x2,1+u2(0)x2,2+u1(0)u2(0)x3\displaystyle\triangleq x_{1,2}\frac{\partial}{\partial x_{1,1}}+x_{1,3}\frac{\partial}{\partial x_{1,2}}+u_{1}^{(0)}\frac{\partial}{\partial x_{1,3}}+x_{2,2}\frac{\partial}{\partial x_{2,1}}+u_{2}^{(0)}\frac{\partial}{\partial x_{2,2}}+u_{1}^{(0)}u_{2}^{(0)}\frac{\partial}{\partial x_{3}} (5.3)
g1(0)\displaystyle g_{1}^{(0)} u1(0),g2(0)u2(0)\displaystyle\triangleq\frac{\partial}{\partial u_{1}^{(0)}},\qquad g_{2}^{(0)}\triangleq\frac{\partial}{\partial u_{2}^{(0)}}

One can verify that

adg0(0)g1(0)=x1,3u2(0)x3,adg0(0)2g1(0)=x1,2,adg0(0)3g1(0)=x1,1,adg0(0)4g1(0)=0\displaystyle\mathrm{ad}_{g_{0}^{(0)}}g_{1}^{(0)}=-\frac{\partial}{\partial x_{1,3}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}},\quad\mathrm{ad}_{g_{0}^{(0)}}^{2}g_{1}^{(0)}=\frac{\partial}{\partial x_{1,2}},\quad\mathrm{ad}_{g_{0}^{(0)}}^{3}g_{1}^{(0)}=-\frac{\partial}{\partial x_{1,1}},\quad\mathrm{ad}_{g_{0}^{(0)}}^{4}g_{1}^{(0)}=0

and

adg0(0)g2(0)=x2,2u1(0)x3,adg0(0)2g2(0)=x2,1,adg0(0)3g2(0)=0.\displaystyle\mathrm{ad}_{g_{0}^{(0)}}g_{2}^{(0)}=-\frac{\partial}{\partial x_{2,2}}-u_{1}^{(0)}\frac{\partial}{\partial x_{3}},\quad\mathrm{ad}_{g_{0}^{(0)}}^{2}g_{2}^{(0)}=\frac{\partial}{\partial x_{2,1}},\quad\mathrm{ad}_{g_{0}^{(0)}}^{3}g_{2}^{(0)}=0.

Therefore

G0(0)={u1(0),u2(0)}=G0(0)¯,G_{0}^{(0)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\overline{G_{0}^{(0)}},
G1(0)=G0(0)+{x1,3u2(0)x3,x2,2u1(0)x3}G1(0)¯,G_{1}^{(0)}=G_{0}^{(0)}+\left\{-\frac{\partial}{\partial x_{1,3}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}},-\frac{\partial}{\partial x_{2,2}}-u_{1}^{(0)}\frac{\partial}{\partial x_{3}}\right\}\neq\overline{G_{1}^{(0)}},

since, e.g. , [g2(0),adg0(0)g1(0)]=x3G1(0)[g_{2}^{(0)},\mathrm{ad}_{g_{0}^{(0)}}g_{1}^{(0)}]=-\frac{\partial}{\partial x_{3}}\not\in G_{1}^{(0)}, and dimG1(0)=4\dim G_{1}^{(0)}=4 whereas dimG1(0)¯=5\dim\overline{G_{1}^{(0)}}=5,

G2(0)=G1(0)+{x1,2,x2,1}G2(0)¯,dimG2(0)¯=7,G_{2}^{(0)}=G_{1}^{(0)}+\left\{\frac{\partial}{\partial x_{1,2}},\frac{\partial}{\partial x_{2,1}}\right\}\neq\overline{G_{2}^{(0)}},\quad\dim\overline{G_{2}^{(0)}}=7,
G3(0)=G2(0)+{x1,1}G3(0)¯G_{3}^{(0)}=G_{2}^{(0)}+\left\{\frac{\partial}{\partial x_{1,1}}\right\}\neq\overline{G_{3}^{(0)}}

and Gk(0)=G3(0)G_{k}^{(0)}=G_{3}^{(0)} for all k3k\geq 3. Moreover, dimG3(0)=7<dimG3(0)¯=n+m=8\dim{G_{3}^{(0)}}=7<\dim{\overline{G_{3}^{(0)}}}=n+m=8. We conclude that the system is not feedback linearizable.

Since G1(0)G_{1}^{(0)} is not involutive, there are two possible initializations:

H1,1(𝟎)={g1(0),adg0(0)g1(0)}={u1(0),x1,3u2(0)x3}=H1,1(𝟎)¯H_{1,1}^{(\mathbf{0})}=\{g_{1}^{(0)},\mathrm{ad}_{g_{0}^{(0)}}g_{1}^{(0)}\}=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{1,3}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}}\}=\overline{H_{1,1}^{(\mathbf{0})}} (5.4)

or

H1,2(𝟎)={g2(0),adg0(0)g2(0)}={u2(0),x2,2u1(0)x3}=H1,2(𝟎)¯.H_{1,2}^{(\mathbf{0})}=\{g_{2}^{(0)},\mathrm{ad}_{g_{0}^{(0)}}g_{2}^{(0)}\}=\{\frac{\partial}{\partial u_{2}^{(0)}},-\frac{\partial}{\partial x_{2,2}}-u_{1}^{(0)}\frac{\partial}{\partial x_{3}}\}=\overline{H_{1,2}^{(\mathbf{0})}}. (5.5)

We use the second possibility, which amounts to prolonging the first input.

5.1.2 Flatness by Pure Prolongation of the First Input

Let us now apply theorem 4.1 and algorithm 4.2 with j2=0j_{2}=0, i.e. g2(j2)g2(0)=u2(0)g_{2}^{(j_{2})}\triangleq g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}} to determine if this system is flat by pure prolongation and compute j11j_{1}\geq 1. Recall that we have set g0(l1,0)=g0(0)+p=0l11u1(p+1)u1(p)g_{0}^{(l_{1},0)}=g_{0}^{(0)}+\sum_{p=0}^{l_{1}-1}u_{1}^{(p+1)}\frac{\partial}{\partial u_{1}^{(p)}}.

𝐤=𝟎.\bullet\leavevmode\nobreak\ \mathbf{k=0.}

We have Γ0(l1,0)={u1(l1)}\Gamma_{0}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}}\right\}, Δ0(l1,0)={u2(0)}=Δ0(l1,0)¯\Delta_{0}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\overline{\Delta_{0}^{(l_{1},0)}} and [Γ0(l1,0),Δ0(l1,0)]Δ0(l1,0)\left[\Gamma_{0}^{(l_{1},0)},\Delta_{0}^{(l_{1},0)}\right]\subset\Delta_{0}^{(l_{1},0)} for all l11l_{1}\geq 1.

𝐤=𝟏.\bullet\leavevmode\nobreak\ \mathbf{k=1.}

If l12l_{1}\geq 2,

Γ1(l1,0)={u1(l1),u1(l11)}\Gamma_{1}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}}\right\}

and

Δ1(l1,0)={u2(0),adg0(l1,0)u2(0)}={u2(0),x2,2+u1(0)x3}=Δ1(l1,0)¯.\Delta_{1}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}}+u_{1}^{(0)}\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{1}^{(l_{1},0)}}.

Moreover, [Γ1(l1,0),Δ1(l1,0)]={0}Δ1(l1,0)\left[\Gamma_{1}^{(l_{1},0)},\Delta_{1}^{(l_{1},0)}\right]=\{0\}\subset\Delta_{1}^{(l_{1},0)}.

Now, for l1=1l_{1}=1, we have Γ1(1,0)={u1(1)}\Gamma_{1}^{(1,0)}=\left\{\frac{\partial}{\partial u_{1}^{(1)}}\right\} and Δ1(1,0)={u1(0),u2(0),adg0(1,0)u2(0)}={u1(0),u2(0),x2,2+u1(0)x3}\Delta_{1}^{(1,0)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(1,0)}}\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}}+u_{1}^{(0)}\frac{\partial}{\partial x_{3}}\right\} which is not involutive.

Thus σ1,Δ(1)=2\sigma_{1,\Delta}(1)=2 and σ1,Γ,Δ(1)=0\sigma_{1,\Gamma,\Delta}(1)=0 which implies that

j1=maxk0σ1,Δ(k)σ1,Γ,Δ(k)σ1,Δ(1)σ1,Γ,Δ(1)=2.j_{1}=\max_{k\geq 0}\sigma_{1,\Delta}(k)\wedge\sigma_{1,\Gamma,\Delta}(k)\geq\sigma_{1,\Delta}(1)\wedge\sigma_{1,\Gamma,\Delta}(1)=2.
𝐤=𝟐.\bullet\leavevmode\nobreak\ \mathbf{k=2.}

If l13l_{1}\geq 3, we have

Γ2(l1,0)={u1(l1),u1(l11),u1(l12)}\Gamma_{2}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}},\frac{\partial}{\partial u_{1}^{(l_{1}-2)}}\right\}

and

Δ2(l1,0)\displaystyle\Delta_{2}^{(l_{1},0)} ={u2(0),adg0(l1,0)u2(0),adg0(l1,0)2u2(0)}\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}^{2}\frac{\partial}{\partial u_{2}^{(0)}}\right\}
={u2(0),x2,2+u1(0)x3,x2,1u1(1)x3}=Δ2(l1,0)¯.\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}}+u_{1}^{(0)}\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2,1}}-u_{1}^{(1)}\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{2}^{(l_{1},0)}}.

Moreover, it is readily verified that [Γ2(l1,0),Δ2(l1,0)]Δ2(l1,0)\left[\Gamma_{2}^{(l_{1},0)},\Delta_{2}^{(l_{1},0)}\right]\subset\Delta_{2}^{(l_{1},0)} only if l14l_{1}\geq 4, condition (ii) of theorem 4.1 being violated if l1=3l_{1}=3 and we have σ1,Δ(2)=0\sigma_{1,\Delta}(2)=0 and σ1,Γ,Δ(2)=4\sigma_{1,\Gamma,\Delta}(2)=4 which implies that j1=maxk0σ1,Δ(k)σ1,Γ,Δ(k)maxr=1,2σ1,Δ(r)σ1,Γ,Δ(r)4j_{1}=\max_{k\geq 0}\sigma_{1,\Delta}(k)\wedge\sigma_{1,\Gamma,\Delta}(k)\geq\max_{r=1,2}\sigma_{1,\Delta}(r)\wedge\sigma_{1,\Gamma,\Delta}(r)\geq 4.

𝐤=𝟑.\bullet\leavevmode\nobreak\ \mathbf{k=3.}

Again, if l14l_{1}\geq 4, we have:

Γ3(l1,0)={u1(l1),u1(l11),u1(l12)),u1(l13)}\Gamma_{3}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}},\frac{\partial}{\partial u_{1}^{(l_{1}-2))}},\frac{\partial}{\partial u_{1}^{(l_{1}-3)}}\right\}

and

Δ3(l1,0)\displaystyle\Delta_{3}^{(l_{1},0)} ={u2(0),adg0(l1,0)u2(0),adg0(l1,0)2u2(0),adg0(l1,0)3u2(0)}\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}^{2}\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(l_{1},0)}}^{3}\frac{\partial}{\partial u_{2}^{(0)}}\right\}
={u2(0),x2,2+u1(0)x3,x2,1u1(1)x3,u1(2)x3}\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}}+u_{1}^{(0)}\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2,1}}-u_{1}^{(1)}\frac{\partial}{\partial x_{3}},u_{1}^{(2)}\frac{\partial}{\partial x_{3}}\right\}
={u2(0),x2,2,x2,1,x3}=Δ3(l1,0)¯\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}},\frac{\partial}{\partial x_{2,1}},\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{3}^{(l_{1},0)}}

provided that u1(2)0u_{1}^{(2)}\neq 0. We also indeed have [Γ3(l1,0),Δ3(l1,0)]Δ3(l1,0)\left[\Gamma_{3}^{(l_{1},0)},\Delta_{3}^{(l_{1},0)}\right]\subset\Delta_{3}^{(l_{1},0)} for l14l_{1}\geq 4 hence σ1,Γ,Δ(3)=4\sigma_{1,\Gamma,\Delta}(3)=4 and j1maxr=1,2,3σ1,Δ(r)σ1,Γ,Δ(r)4j_{1}\geq\max_{r=1,2,3}\sigma_{1,\Delta}(r)\wedge\sigma_{1,\Gamma,\Delta}(r)\geq 4.

𝐤=𝟒\bullet\leavevmode\nobreak\ \mathbf{k=4}

If l15l_{1}\geq 5, we have Γ4(l1,0)={u1(l1),,u1(l14)}\Gamma_{4}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\ldots,\frac{\partial}{\partial u_{1}^{(l_{1}-4)}}\right\} and Δ4(l1,0)=Δ3(l1,0)\Delta_{4}^{(l_{1},0)}=\Delta_{3}^{(l_{1},0)} since adg0(l1,0)4u2(0)=u1(3)x3Δ3(l1,0)\mathrm{ad}_{g_{0}^{(l_{1},0)}}^{4}\frac{\partial}{\partial u_{2}^{(0)}}=-u_{1}^{(3)}\frac{\partial}{\partial x_{3}}\in\Delta_{3}^{(l_{1},0)}.

If now l1=4l_{1}=4, Γ4(4,0)={u1(4),,u1(1)}\Gamma_{4}^{(4,0)}=\left\{\frac{\partial}{\partial u_{1}^{(4)}},\ldots,\frac{\partial}{\partial u_{1}^{(1)}}\right\} and

Δ4(4,0)\displaystyle\Delta_{4}^{(4,0)} ={u1(0),u2(0),adg0(4,0)u2(0),,adg0(4,0)4u2(0)}\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\mathrm{ad}_{g_{0}^{(4,0)}}\frac{\partial}{\partial u_{2}^{(0)}},\ldots,\mathrm{ad}_{g_{0}^{(4,0)}}^{4}\frac{\partial}{\partial u_{2}^{(0)}}\right\}
={u1(0),u2(0),x2,2,x2,1,x3}=Δ4(4,0)¯.\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}},\frac{\partial}{\partial x_{2,1}},\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{4}^{(4,0)}}.

We thus immediately get [Γ4(l1,0),Δ4(l1,0)]Δ4(l1,0)\left[\Gamma_{4}^{(l_{1},0)},\Delta_{4}^{(l_{1},0)}\right]\subset\Delta_{4}^{(l_{1},0)} for all l14l_{1}\geq 4.

𝐤𝟓\bullet\leavevmode\nobreak\ \mathbf{k\geq 5}

Finally, the reader may easily check that Γk(4,0)=Γ4(4,0)\Gamma_{k}^{(4,0)}=\Gamma_{4}^{(4,0)} for all k5k\geq 5 and

Δ5(4,0)\displaystyle\Delta_{5}^{(4,0)} =Δ4(4,0)+{adg0(4,0)u1(0)}={u1(0),x1,3,u2(0),x2,2,x2,1,x3}\displaystyle=\Delta_{4}^{(4,0)}+\left\{\mathrm{ad}_{g_{0}^{(4,0)}}\frac{\partial}{\partial u_{1}^{(0)}}\right\}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{1,3}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}},\frac{\partial}{\partial x_{2,1}},\frac{\partial}{\partial x_{3}}\right\}
=Δ5(4,0)¯\displaystyle=\overline{\Delta_{5}^{(4,0)}}
Δ6(4,0)\displaystyle\Delta_{6}^{(4,0)} =Δ5(4,0)+{adg0(4,0)2u1(0)}={u1(0),x1,3,x1,2,u2(0),x2,2,x2,1,x3}\displaystyle=\Delta_{5}^{(4,0)}+\left\{\mathrm{ad}_{g_{0}^{(4,0)}}^{2}\frac{\partial}{\partial u_{1}^{(0)}}\right\}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{1,3}},\frac{\partial}{\partial x_{1,2}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{2,2}},\frac{\partial}{\partial x_{2,1}},\frac{\partial}{\partial x_{3}}\right\}
=Δ6(4,0)¯\displaystyle=\overline{\Delta_{6}^{(4,0)}}
Δ7(4,0)\displaystyle\Delta_{7}^{(4,0)} =Δ6(4,0)+{adg0(4,0)3u1(0)}=T8.\displaystyle=\Delta_{6}^{(4,0)}+\left\{\mathrm{ad}_{g_{0}^{(4,0)}}^{3}\frac{\partial}{\partial u_{1}^{(0)}}\right\}=\mathrm{T}\mathbb{R}^{8}.

We also indeed have [Γk(4,0)),Δk(4,0))]Δk(4,0)\left[\Gamma_{k}^{(4,0))},\Delta_{k}^{(4,0))}\right]\subset\Delta_{k}^{(4,0)} for all k5k\geq 5.

Hence, for all k0k\geq 0, the minimal 𝐣\mathbf{j} is equal to (4,0)(4,0) and we conclude that system (5.1), with the first input channel prolonged up to j1=4j_{1}=4, i.e. controlled by u1(5)u_{1}^{(5)}, is feedback linearizable.

Note that the bounds (3.5) and (3.18) are indeed satisfied. They read n+j1=10k(4,0)=7n+j12j1=5n+j_{1}=10\geq k_{\star}^{(4,0)}=7\geq\frac{n+j_{1}}{2}\wedge j_{1}=5, but they are not tight.

Accordingly, evaluating the bound on the number of integrators needed to linearize the system, proposed by [34] for m=2m=2, we find 2n3=92n-3=9, and the one proposed by [12] gives 2n16(8+2414)=2n3=92n-\frac{1}{6}(8+24-14)=2n-3=9 as well.

5.1.3 Flat output computation

The prolonged system is now expressed in the state coordinates

x¯(4,0)(x1,1,x1,2,x1,3,x2,1,x2,2,x3,u1(0),u1(1),u1(2),u1(3),u1(4),u2(0))\overline{x}^{(4,0)}\triangleq(x_{1,1},x_{1,2},x_{1,3},x_{2,1},x_{2,2},x_{3},u_{1}^{(0)},u_{1}^{(1)},u_{1}^{(2)},u_{1}^{(3)},u_{1}^{(4)},u_{2}^{(0)})

still with x1,jx1(j1)x_{1,j}\triangleq x_{1}^{(j-1)}, j=1,2,3j=1,2,3, x2,jx2(j1)x_{2,j}\triangleq x_{2}^{(j-1)}, j=1,2j=1,2 and ui(0)=uiu_{i}^{(0)}=u_{i}, i=1,2i=1,2. We indeed still have n=6n=6 but the prolonged state dimension is now equal to n+m+|j|=12n+m+|j|=12 with the two input variables u1(5)u_{1}^{(5)} and u2(1)u_{2}^{(1)}.

The prolonged system vector fields are

g0(4,0)g0(0)+j=03u1(j+1)u1(j),g1(4)u1(4),g2(0)u2(0)g_{0}^{(4,0)}\triangleq g_{0}^{(0)}+\sum_{j=0}^{3}u_{1}^{(j+1)}\frac{\partial}{\partial u_{1}^{(j)}},\qquad g_{1}^{(4)}\triangleq\frac{\partial}{\partial u_{1}^{(4)}},\qquad g_{2}^{(0)}\triangleq\frac{\partial}{\partial u_{2}^{(0)}}

and the corresponding distributions are, according to (3.17),

G0(4,0)\displaystyle G_{0}^{(4,0)} ={u1(4),u2(0)}=G0(4,0)¯,\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(4)}},\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\overline{G_{0}^{(4,0)}},
G1(4,0)\displaystyle G_{1}^{(4,0)} =G0(4,0)+{u1(3),x2,2u1(0)x3}=G1(4,0)¯,\displaystyle=G_{0}^{(4,0)}+\left\{\frac{\partial}{\partial u_{1}^{(3)}},-\frac{\partial}{\partial x_{2,2}}-u_{1}^{(0)}\frac{\partial}{\partial x_{3}}\right\}=\overline{G_{1}^{(4,0)}},
G2(4,0)\displaystyle G_{2}^{(4,0)} =G1(4,0)+{u1(2),x2,1u1(1)x3}=G2(4,0)¯,\displaystyle=G_{1}^{(4,0)}+\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial x_{2,1}}-u_{1}^{(1)}\frac{\partial}{\partial x_{3}}\right\}=\overline{G_{2}^{(4,0)}},
G3(4,0)\displaystyle G_{3}^{(4,0)} =G2(4,0)+{u1(1),u1(2)x3}=G3(4,0)¯,\displaystyle=G_{2}^{(4,0)}+\left\{\frac{\partial}{\partial u_{1}^{(1)}},-u_{1}^{(2)}\frac{\partial}{\partial x_{3}}\right\}=\overline{G_{3}^{(4,0)}},
G4(4,0)\displaystyle G_{4}^{(4,0)} =G3(4,0)+{u1(0)}=G4(4,0)¯,\displaystyle=G_{3}^{(4,0)}+\left\{\frac{\partial}{\partial u_{1}^{(0)}}\right\}=\overline{G_{4}^{(4,0)}},
G5(4,0)\displaystyle G_{5}^{(4,0)} =G4(4,0)+{x1,3}=G5(4,0)¯,\displaystyle=G_{4}^{(4,0)}+\left\{\frac{\partial}{\partial x_{1,3}}\right\}=\overline{G_{5}^{(4,0)}},
G6(4,0)\displaystyle G_{6}^{(4,0)} =G5(4,0)+{x1,2}=G6(4,0)¯,\displaystyle=G_{5}^{(4,0)}+\left\{\frac{\partial}{\partial x_{1,2}}\right\}=\overline{G_{6}^{(4,0)}},
G7(4,0)\displaystyle G_{7}^{(4,0)} =G6(4,0)+{x1,1}=G7(4,0)¯=T12.\displaystyle=G_{6}^{(4,0)}+\left\{\frac{\partial}{\partial x_{1,1}}\right\}=\overline{G_{7}^{(4,0)}}=\mathrm{T}\mathbb{R}^{12}.

This confirms that the system (5.1) is flat by pure prolongation in any neighborhood excluding u1(2)=0u_{1}^{(2)}=0, with k(4,0)=7k^{(4,0)}_{\ast}=7. The reader may easily verify that ρk(4,0)=2\rho_{k}^{(4,0)}=2 for k=0,,3k=0,\ldots,3 and ρk(4,0)=1\rho_{k}^{(4,0)}=1 for k=4,,7k=4,\ldots,7, which yields κ1(4,0)=8\kappa_{1}^{(4,0)}=8 and κ2(4,0)=4\kappa_{2}^{(4,0)}=4.

The corresponding flat outputs are thus obtained by solving the set of P.D.E.’s

Gk(4,0),dy1=0,k=0,,6,G7(4,0),dy10\displaystyle\left<G_{k}^{(4,0)},dy_{1}\right>=0,\quad k=0,\ldots,6,\qquad\left<G_{7}^{(4,0)},dy_{1}\right>\neq 0 (5.6)
Gk(4,0),dy2=0,k=0,,2,G3(4,0),dy20\displaystyle\left<G_{k}^{(4,0)},dy_{2}\right>=0,\quad k=0,\ldots,2,\qquad\left<G_{3}^{(4,0)},dy_{2}\right>\neq 0

whose solution is

y1=x1,y2=x3x2,2u1(0)+x2,1u1(1)=x3x˙2u1+x2u˙1.y_{1}=x_{1},\quad y_{2}=x_{3}-x_{2,2}u_{1}^{(0)}+x_{2,1}u_{1}^{(1)}=x_{3}-\dot{x}_{2}u_{1}+x_{2}\dot{u}_{1}. (5.7)
Remark 5.1.

Using the first intialization (5.4), which amounts to prolonging the second input, we leave to the reader the verification that one obtains a prolongation of order 6, thus non minimal, with flat outputs given by (5.2).

Remark 5.2.

In [25, section 3.1.1], the authors consider a dual notion of minimality, called rr-flatness, where rr is the minimal number over all possible flat outputs of the maximal number of derivatives of the inputs that appear in the flat outputs, i.e.

r=minY:flatoutputmaxi=1,,m{si𝐬=(s1,,sm),y=Y(x,u¯(𝐬))}.r=\min_{Y:\leavevmode\nobreak\ \mathrm{flat\leavevmode\nobreak\ output}}\leavevmode\nobreak\ \max_{i=1,\ldots,m}\{s_{i}\mid\mathbf{s}=(s_{1},\ldots,s_{m}),\leavevmode\nobreak\ y=Y(x,\overline{u}^{(\mathbf{s})})\}.

They conjectured that rr should be equal to 1 in the present case (with their notations, α1=3\alpha_{1}=3, α2=2\alpha_{2}=2 and min(α1,α2)1=1\min(\alpha_{1},\alpha_{2})-1=1). As the reader may easily verify, it is indeed minimal since the minimal 𝐣\mathbf{j} is (4,0)(4,0) and moreover since, by the equations of the first line of (5.6), y1y_{1} neither can depend on u¯1(4)\overline{u}_{1}^{(4)} nor on u2(0)u_{2}^{(0)}, and, by the second line, y2y_{2} cannot depend on u1(4),u1(3),u1(2)u_{1}^{(4)},u_{1}^{(3)},u_{1}^{(2)} but explicitly depends on u1(1)u_{1}^{(1)} by the definition of G2(4,0)G_{2}^{(4,0)}.

5.2 4-dimensional Driftless Bilinear System [25, 26, 27]

x˙1\displaystyle\dot{x}_{1} =u1\displaystyle=u_{1} (5.8)
x˙2\displaystyle\dot{x}_{2} =x3u1\displaystyle=x_{3}u_{1}
x˙3\displaystyle\dot{x}_{3} =x4u1\displaystyle=x_{4}u_{1}
x˙4\displaystyle\dot{x}_{4} =u2\displaystyle=u_{2}

It is immediate to verify that this system is flat with flat output

y1=x1,y2=x2y_{1}=x_{1},\quad y_{2}=x_{2} (5.9)

(see [26, 27] and [25, theorems 4 and 5]) but not static feedback linearizable.

According to our formalism, we consider the state (x1,x2,x3,x4,u1(0),u2(0))(x_{1},x_{2},x_{3},x_{4},u_{1}^{(0)},u_{2}^{(0)}) of dimension 6, with n=4n=4 and m=2m=2, and the new inputs (u1(1),u2(1))(u_{1}^{(1)},u_{2}^{(1)}). The associated vector fields are

g0(0)=u1(0)(x1+x3x2+x4x3)+u2(0)x4,g1(0)=u1(0),g2(0)=u2(0).g_{0}^{(0)}=u_{1}^{(0)}\left(\frac{\partial}{\partial x_{1}}+x_{3}\frac{\partial}{\partial x_{2}}+x_{4}\frac{\partial}{\partial x_{3}}\right)+u_{2}^{(0)}\frac{\partial}{\partial x_{4}},\quad g_{1}^{(0)}=\frac{\partial}{\partial u_{1}^{(0)}},\quad g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}}.

The reader may easily verify that G1(𝟎)G_{1}^{(\mathbf{0})} is not involutive and that there are two possible initializations:

H1,1(𝟎)={g1(0),adg0(𝟎)g1(0)}={u1(0),x1+x3x2+x4x3}=H1,1(𝟎)¯H_{1,1}^{(\mathbf{0})}=\{g_{1}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{1}^{(0)}\}=\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{1}}+x_{3}\frac{\partial}{\partial x_{2}}+x_{4}\frac{\partial}{\partial x_{3}}\}=\overline{H_{1,1}^{(\mathbf{0})}}

or

H1,2(𝟎)={g2(0),adg0(𝟎)g2(0)}={u2(0),x4}=H1,2(𝟎)¯.H_{1,2}^{(\mathbf{0})}=\{g_{2}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{2}^{(0)}\}=\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}}\}=\overline{H_{1,2}^{(\mathbf{0})}}.

We use the second one, which amounts to prolonging the first input u1(1)u_{1}^{(1)}.

5.2.1 the distributions Γk(𝐥)\Gamma_{k}^{(\mathbf{l})} and Δk(𝐥)\Delta_{k}^{(\mathbf{l})}

We set 𝐥=(l1,0)\mathbf{l}=(l_{1},0), l11l_{1}\geq 1. We thus have

g0(l1,0)=g0(0)+p=0l11u1(p+1)u1(p),g1(l1)=u1(l1),g2(0)=u2(0)g_{0}^{(l_{1},0)}=g_{0}^{(0)}+\sum_{p=0}^{l_{1}-1}u_{1}^{(p+1)}\frac{\partial}{\partial u_{1}^{(p)}},\quad g_{1}^{(l_{1})}=\frac{\partial}{\partial u_{1}^{(l_{1})}},\quad\quad g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}}

and

Γ0(l1,0)={u1(l1)},Δ0(l1,0)={u2(0)}\Gamma_{0}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}}\right\},\quad\Delta_{0}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}}\right\}

for all l11l_{1}\geq 1.

𝐤=𝟏.\bullet\leavevmode\nobreak\ \mathbf{k=1.}

If l12l_{1}\geq 2, we have

Γ1(l1,0)={u1(l1),u1(l11)},Δ1(l1,0)=Δ0(l1,0)+{adg0(l1,0)u2(0)}={u2(0),x4}\Gamma_{1}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}}\right\},\quad\Delta_{1}^{(l_{1},0)}=\Delta_{0}^{(l_{1},0)}+\{\mathrm{ad}_{g_{0}^{(l_{1},0)}}\frac{\partial}{\partial u_{2}^{(0)}}\}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}}\right\}

and if l1=1l_{1}=1,

Γ1(1,0)={u1(1)},Δ1(1,0)=Δ0(1,0)+{u1(0),adg0(1,0)u2(0)}={u1(0),u2(0),x4}.\Gamma_{1}^{(1,0)}=\left\{\frac{\partial}{\partial u_{1}^{(1)}}\right\},\quad\Delta_{1}^{(1,0)}=\Delta_{0}^{(1,0)}+\{\frac{\partial}{\partial u_{1}^{(0)}},\mathrm{ad}_{g_{0}^{(1,0)}}\frac{\partial}{\partial u_{2}^{(0)}}\}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}}\right\}.

We thus have Δ1(l1,0)=Δ1(l1,0)¯\Delta_{1}^{(l_{1},0)}=\overline{\Delta_{1}^{(l_{1},0)}} and [Γ1(l1,0),Δ1(l1,0)]={0}Δ1(l1,0)\left[\Gamma_{1}^{(l_{1},0)},\Delta_{1}^{(l_{1},0)}\right]=\{0\}\subset\Delta_{1}^{(l_{1},0)} for all l11l_{1}\geq 1, hence σ1,Γ,Δ(1)=σ1,Δ(1)=0\sigma_{1,\Gamma,\Delta}(1)=\sigma_{1,\Delta}(1)=0.

𝐤=𝟐.\bullet\leavevmode\nobreak\ \mathbf{k=2.}

If l13l_{1}\geq 3, we have

Γ2(l1,0)={u1(l1),u1(l11),u1(l12)},Δ2(l1,0)={u2(0),x4,x3},[Γ2(l1,0),Δ2(l1,0)]={0}\begin{array}[]{c}\Gamma_{2}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}},\frac{\partial}{\partial u_{1}^{(l_{1}-2)}}\right\},\quad\Delta_{2}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}}\right\},\vspace{10pt}\\ \left[\Gamma_{2}^{(l_{1},0)},\Delta_{2}^{(l_{1},0)}\right]=\{0\}\end{array}

and if l1=2l_{1}=2,

Γ2(2,0)={u1(2),u1(1)},Δ2(2,0)={u1(0),u2(0),x4,x3},[Γ2(2,0),Δ2(2,0)]={0}.\begin{array}[]{c}\Gamma_{2}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}}\right\},\quad\Delta_{2}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}}\right\},\vspace{10pt}\\ \left[\Gamma_{2}^{(2,0)},\Delta_{2}^{(2,0)}\right]=\{0\}.\end{array}

Finally, if l1=1l_{1}=1,

Γ2(1,0)={u1(1)},Δ2(1,0)\displaystyle\Gamma_{2}^{(1,0)}=\left\{\frac{\partial}{\partial u_{1}^{(1)}}\right\},\quad\Delta_{2}^{(1,0)} ={u1(0),x1x3x2x4x3,u2(0),x4,x3}\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{1}}-x_{3}\frac{\partial}{\partial x_{2}}-x_{4}\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}}\right\}
Δ2(1,0)¯.\displaystyle\neq\overline{\Delta_{2}^{(1,0)}}.

We thus have Δ2(l1,0)=Δ2(l1,0)¯\Delta_{2}^{(l_{1},0)}=\overline{\Delta_{2}^{(l_{1},0)}} and [Γ2(l1,0),Δ2(l1,0)]Δ2(l1,0)\left[\Gamma_{2}^{(l_{1},0)},\Delta_{2}^{(l_{1},0)}\right]\subset\Delta_{2}^{(l_{1},0)} for all l12l_{1}\geq 2 but, for l1=1l_{1}=1, Δ2(1,0)\Delta_{2}^{(1,0)} is not involutive. Therefore, σ1,Δ(2)=2\sigma_{1,\Delta}(2)=2 and our search may be restricted to l12l_{1}\geq 2.

𝐤=𝟑.\bullet\leavevmode\nobreak\ \mathbf{k=3.}

If l14l_{1}\geq 4, we have

Γ3(l1,0)={u1(l1),u1(l11),u1(l12),u1(l13)},Δ3(l1,0)={u2(0),x4,x3,x2}.\Gamma_{3}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{1}^{(l_{1})}},\frac{\partial}{\partial u_{1}^{(l_{1}-1)}},\frac{\partial}{\partial u_{1}^{(l_{1}-2)}},\frac{\partial}{\partial u_{1}^{(l_{1}-3)}}\right\},\quad\Delta_{3}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}}\right\}.

If l1=3l_{1}=3,

Γ3(3,0)={u1(3),u1(2),u1(1)},Δ3(3,0)={u1(0),u2(0),x4,x3,x2}.\Gamma_{3}^{(3,0)}=\left\{\frac{\partial}{\partial u_{1}^{(3)}},\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}}\right\},\quad\Delta_{3}^{(3,0)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}}\right\}.

If l1=2l_{1}=2,

Γ3(2,0)={u1(2),u1(1)},Δ3(2,0)={u1(0),u2(0),x4,x3,x2,x1}=T6\Gamma_{3}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}}\right\},\quad\Delta_{3}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial x_{1}}\right\}=\mathrm{T}\mathbb{R}^{6}

Therefore, Δk(2,0)=Δk(2,0)¯\Delta_{k}^{(2,0)}=\overline{\Delta_{k}^{(2,0)}} and [Γk(2,0),Δk(2,0)]Δk(2,0)\left[\Gamma_{k}^{(2,0)},\Delta_{k}^{(2,0)}\right]\subset\Delta_{k}^{(2,0)} for all k0k\geq 0 and Δ3(2,0)=T6\Delta_{3}^{(2,0)}=\mathrm{T}\mathbb{R}^{6}.

We conclude that the conditions of theorem 4.1 hold true whenever j12j_{1}\geq 2, which proves that system (5.8) with the pure prolongation of order 𝐣=(2,0)\mathbf{j}=(2,0) is feedback linearizable.

The reader may easily check that, using the first initialization H1,1(𝟎)H_{1,1}^{(\mathbf{0})}, since Δ2(0,l2)\Delta_{2}^{(0,l_{2})} is not involutive for all l23l_{2}\geq 3, no pure prolongation of the second input channel, u2(1)u_{2}^{(1)}, can lead to the linearizability conditions. Therefore the minimal prolongation is 𝐣=(2,0)\mathbf{j}=(2,0).

5.2.2 Flat output computation

The prolonged state is now x¯(2,0)(x1,x2,x3,x4,u1(0),u1(1),u1(2),u2(0))\overline{x}^{(2,0)}\triangleq(x_{1},x_{2},x_{3},x_{4},u_{1}^{(0)},u_{1}^{(1)},u_{1}^{(2)},u_{2}^{(0)}) of dimension 8, and the new inputs are (u1(3),u2(1))(u_{1}^{(3)},u_{2}^{(1)}).

G0(2,0)={u1(2),u2(0)}=G0(2,0)¯,G_{0}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{2}^{(0)}}\right\}=\overline{G_{0}^{(2,0)}},
G1(2,0)={u1(2),u1(1),u2(0),x4}=G1(2,0)¯,G_{1}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}}\right\}=\overline{G_{1}^{(2,0)}},
G2(2,0)={u1(2),u1(1),u1(0),u2(0),x4,x3}=G2(2,0)¯,G_{2}^{(2,0)}=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}},\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}}\right\}=\overline{G_{2}^{(2,0)}},
G3(2,0)\displaystyle G_{3}^{(2,0)} ={u1(2),u1(1),u1(0),x1+x3x2+x4x3,u2(0),x4,x3,x2}\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(2)}},\frac{\partial}{\partial u_{1}^{(1)}},\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{1}}+x_{3}\frac{\partial}{\partial x_{2}}+x_{4}\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}}\right\}
=G3(2,0)¯=T8\displaystyle=\overline{G_{3}^{(2,0)}}=\mathrm{T}\mathbb{R}^{8}

hence the feedback linearizability of the purely prolonged system with ρ0(2,0)=ρ1(2,0)=ρ2(2,0)=ρ3(2,0)=2\rho_{0}^{(2,0)}=\rho_{1}^{(2,0)}=\rho_{2}^{(2,0)}=\rho_{3}^{(2,0)}=2 and κ1(2,0)=κ2(2,0)=4\kappa_{1}^{(2,0)}=\kappa_{2}^{(2,0)}=4.

Flat outputs (y1,y2)(y_{1},y_{2}) are locally non trivial solutions of the system (5.10), i.e. :

Gk(2,0),dyi=0,k=0,1,2,withG3(2,0),dyi0,i=1,2.\left<G_{k}^{(2,0)},dy_{i}\right>=0,\;k=0,1,2,\quad\mathrm{with\leavevmode\nobreak\ }\quad\left<G_{3}^{(2,0)},dy_{i}\right>\neq 0,\quad i=1,2. (5.10)

It is immediate to verify that

y1=x1,y2=x2y_{1}=x_{1},\qquad y_{2}=x_{2}

is a solution of (5.10) and that the mapping

x¯(2,0)(y1,,y1(3),y2,,y2(3))\overline{x}^{(2,0)}\mapsto\left(y_{1},\ldots,y_{1}^{(3)},y_{2},\ldots,y_{2}^{(3)}\right)

is a local diffeomorphism.

5.3 An Example from [7]

In our formalism, considering the inputs (u1,u2)(u1(0),u2(0))(u_{1},u_{2})\triangleq(u_{1}^{(0)},u_{2}^{(0)}) as part of the state, with n=4n=4 and m=2m=2, this example from [7, Example 2] reads:

x˙1=x2+x3u2(0)\displaystyle\dot{x}_{1}=x_{2}+x_{3}u_{2}^{(0)} (5.11)
x˙2=x3+x1u2(0)\displaystyle\dot{x}_{2}=x_{3}+x_{1}u_{2}^{(0)}
x˙3=u1(0)+x2u2(0)\displaystyle\dot{x}_{3}=u_{1}^{(0)}+x_{2}u_{2}^{(0)}
x˙4=u2(0)\displaystyle\dot{x}_{4}=u_{2}^{(0)}
u˙1(0)=u1(1)\displaystyle\dot{u}_{1}^{(0)}=u_{1}^{(1)}
u˙2(0)=u2(1).\displaystyle\dot{u}_{2}^{(0)}=u_{2}^{(1)}.

It is shown in [7] that this sytem does not satisfy the sufficient, but otherwise not necessary, condition for dynamic linearization of Theorem 4.2 of this paper. Nevertheless, it satisfies the algorithm of [2, section 3.1], that constitutes a sufficient condition for flatness by pure prolongation, without proof of minimality of the obtained prolongation. We show here that it is linearizable by pure prolongation by application of our algorithm, thus providing the minimal prolongation.

The non prolonged vctor fields are

g0(𝟎)=(x2+x3u2(0))x1+(x3+x1u2(0))x2+(u1(0)+x2u2(0))x3+u2(0)x4\displaystyle g_{0}^{(\mathbf{0})}=(x_{2}+x_{3}u_{2}^{(0)})\frac{\partial}{\partial x_{1}}+(x_{3}+x_{1}u_{2}^{(0)})\frac{\partial}{\partial x_{2}}+(u_{1}^{(0)}+x_{2}u_{2}^{(0)})\frac{\partial}{\partial x_{3}}+u_{2}^{(0)}\frac{\partial}{\partial x_{4}}
g1(𝟎)=u1(0),g2(0)=u2(0).\displaystyle g_{1}^{(\mathbf{0})}=\frac{\partial}{\partial u_{1}^{(0)}},\qquad g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}}.

and it is easily seen that

G1(𝟎)={u1(0),u2(0),x3,x3x1x1x2x2x3x4}G1(𝟎)¯G_{1}^{(\mathbf{0})}=\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial x_{3}},-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}}\}\neq\overline{G_{1}^{(\mathbf{0})}}

and that the subdistributions

H1,10={g1(𝟎),adg0(𝟎)g1(𝟎)}={u1(0),x3}=H1,10¯H_{1,1}^{0}=\{g_{1}^{(\mathbf{0})},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{1}^{(\mathbf{0})}\}=\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{3}}\}=\overline{H_{1,1}^{0}}

or

H1,20={g2(𝟎),adg0(𝟎)g2(𝟎)}={u2(0),x3x1x1x2x2x3x4}=H1,20¯H_{1,2}^{0}=\{g_{2}^{(\mathbf{0})},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{2}^{(\mathbf{0})}\}=\{\frac{\partial}{\partial u_{2}^{(0)}},-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}}\}=\overline{H_{1,2}^{0}}

can be taken as possible initializations.

Choosing H1,10H_{1,1}^{0} amounts to prolonging the second input at an arbitrary order l21l_{2}\geq 1 and set 𝐥=(0,l2)\mathbf{l}=(0,l_{2}). For l21l_{2}\geq 1, we denote, as before,

g0(0,l2)=(x2+x3u2(0))x1+(x3+x1u2(0))x2+(u1(0)+x2u2(0))x3\displaystyle g_{0}^{(0,l_{2})}=(x_{2}+x_{3}u_{2}^{(0)})\frac{\partial}{\partial x_{1}}+(x_{3}+x_{1}u_{2}^{(0)})\frac{\partial}{\partial x_{2}}+(u_{1}^{(0)}+x_{2}u_{2}^{(0)})\frac{\partial}{\partial x_{3}}
+u2(0)x4+p=0l21u2(p+1)u2(p)\displaystyle\hskip 170.71652pt+u_{2}^{(0)}\frac{\partial}{\partial x_{4}}+\sum_{p=0}^{l_{2}-1}u_{2}^{(p+1)}\frac{\partial}{\partial u_{2}^{(p)}}
g1(𝟎)=u1(0),g2(l2)=u2(l2)\displaystyle g_{1}^{(\mathbf{0})}=\frac{\partial}{\partial u_{1}^{(0)}},\qquad g_{2}^{(l_{2})}=\frac{\partial}{\partial u_{2}^{(l_{2})}}

and we indeed have

Γ0(0,l2)={u2(l2)},Δ0(0,l2)={u1(0)}=Δ0(0,l2)¯,[Γ0(0,l2),Δ0(0,l2)]Δ0(0,l2),l21.\Gamma_{0}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{2}^{(l_{2})}}\right\},\quad\Delta_{0}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{1}^{(0)}}\right\}=\overline{\Delta_{0}^{(0,l_{2})}},\quad\left[\Gamma_{0}^{(0,l_{2})},\Delta_{0}^{(0,l_{2})}\right]\subset\Delta_{0}^{(0,l_{2})},\quad\forall l_{2}\geq 1.
𝐤=𝟏.\bullet\leavevmode\nobreak\ \mathbf{k=1.}

For all l22l_{2}\geq 2, adg0(0,l2)u1(0)=x3\displaystyle\mathrm{ad}_{g_{0}^{(0,l_{2})}}\frac{\partial}{\partial u_{1}^{(0)}}=-\frac{\partial}{\partial x_{3}} and

Γ1(0,l2)={u2(l2),u2(l21)},Δ1(0,l2)={u1((0)),x3}=Δ1(0,l2)¯,[Γ1(0,l2),Δ1(0,l2)]Δ1(0,l2).\begin{array}[]{c}\displaystyle\Gamma_{1}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{2}^{(l_{2})}},\frac{\partial}{\partial u_{2}^{(l_{2}-1)}}\right\},\qquad\Delta_{1}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{1}^{((0))}},-\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{1}^{(0,l_{2})}},\vspace{1em}\\ \displaystyle\left[\Gamma_{1}^{(0,l_{2})},\Delta_{1}^{(0,l_{2})}\right]\subset\Delta_{1}^{(0,l_{2})}.\end{array}

For l2=1l_{2}=1:

Γ1(0,1)={u2(1)},Δ1(0,1)={u2(0),u1(0),x3}=Δ1(0,1)¯,[Γ1(0,1),Δ1(0,1)]Δ1(0,1).\begin{array}[]{c}\displaystyle\Gamma_{1}^{(0,1)}=\left\{\frac{\partial}{\partial u_{2}^{(1)}}\right\},\quad\Delta_{1}^{(0,1)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{3}}\right\}=\overline{\Delta_{1}^{(0,1)}},\vspace{1em}\\ \displaystyle\left[\Gamma_{1}^{(0,1)},\Delta_{1}^{(0,1)}\right]\subset\Delta_{1}^{(0,1)}.\end{array}
𝐤=𝟐.\bullet\leavevmode\nobreak\ \mathbf{k=2.}

For all l23l_{2}\geq 3, adg0(0,l2)2u1(0)=u2(0)x1+x2\displaystyle\mathrm{ad}_{g_{0}^{(0,l_{2})}}^{2}\frac{\partial}{\partial u_{1}^{(0)}}=u_{2}^{(0)}\frac{\partial}{\partial x_{1}}+\frac{\partial}{\partial x_{2}} and

Γ2(0,l2)={u2(l2),u2(l21),u2(l22)},Δ2(0,l2)={u1(0),x3,u2(0)x1+x2}=Δ2(0,l2)¯,[Γ2(0,l2),Δ2(𝟎,l2)]Δ2(0,l2).\begin{array}[]{c}\displaystyle\Gamma_{2}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{2}^{(l_{2})}},\frac{\partial}{\partial u_{2}^{(l_{2}-1)}},\frac{\partial}{\partial u_{2}^{(l_{2}-2)}}\right\},\vspace{1em}\\ \Delta_{2}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{3}},u_{2}^{(0)}\frac{\partial}{\partial x_{1}}+\frac{\partial}{\partial x_{2}}\right\}=\overline{\Delta_{2}^{(0,l_{2})}},\vspace{1em}\\ \displaystyle\left[\Gamma_{2}^{(0,l_{2})},\Delta_{2}^{(\mathbf{0},l_{2})}\right]\subset\Delta_{2}^{(0,l_{2})}.\end{array}

But for l2=2l_{2}=2:

Γ2(0,2)={u2(2),u2(1)},Δ2(0,2)={u2(0),u1(0),x3,u2(0)x1+x2}Δ2(0,2)¯\Gamma_{2}^{(0,2)}=\left\{\frac{\partial}{\partial u_{2}^{(2)}},\frac{\partial}{\partial u_{2}^{(1)}}\right\},\quad\Delta_{2}^{(0,2)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{3}},u_{2}^{(0)}\frac{\partial}{\partial x_{1}}+\frac{\partial}{\partial x_{2}}\right\}\neq\overline{\Delta_{2}^{(0,2)}}

therefore, σ2,Δ(2)=2\sigma_{2,\Delta}(2)=2 and we must exclude j2=2j_{2}=2.

𝐤=𝟑.\bullet\leavevmode\nobreak\ \mathbf{k=3.}

For all l24l_{2}\geq 4, adg0(0,l2)3u1(0)=(u2(1)1)x1(u2(0))2x2u2(0)x3\displaystyle\mathrm{ad}_{g_{0}^{(0,l_{2})}}^{3}\frac{\partial}{\partial u_{1}^{(0)}}=(u_{2}^{(1)}-1)\frac{\partial}{\partial x_{1}}-\left(u_{2}^{(0)}\right)^{2}\frac{\partial}{\partial x_{2}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}} and, if we exclude the points where u2(0)=0u_{2}^{(0)}=0 and u2(1)=1u_{2}^{(1)}=1,

Γ3(0,l2)={u2(l2),u2(l21),u2(l22),u2(l23)},\Gamma_{3}^{(0,l_{2})}=\left\{\frac{\partial}{\partial u_{2}^{(l_{2})}},\frac{\partial}{\partial u_{2}^{(l_{2}-1)}},\frac{\partial}{\partial u_{2}^{(l_{2}-2)}},\frac{\partial}{\partial u_{2}^{(l_{2}-3)}}\right\},
Δ3(𝟎,l2)\displaystyle\Delta_{3}^{(\mathbf{0},l_{2})} ={u1(0),x3,u2(0)x1+x2,(u2(1)1)x1+(u2(0))2x2u2(0)x3}\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{3}},u_{2}^{(0)}\frac{\partial}{\partial x_{1}}+\frac{\partial}{\partial x_{2}},(u_{2}^{(1)}-1)\frac{\partial}{\partial x_{1}}+\left(u_{2}^{(0)}\right)^{2}\frac{\partial}{\partial x_{2}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}}\right\}
={u1((0)),x3,x2,x1}=Δ3(0,l2)¯,\displaystyle=\left\{\frac{\partial}{\partial u_{1}^{((0))}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial x_{1}}\right\}=\overline{\Delta_{3}^{(0,l_{2})}},
[Γ3(0,l2),Δ3(0,l2)]Δ3(0,l2).\left[\Gamma_{3}^{(0,l_{2})},\Delta_{3}^{(0,l_{2})}\right]\subset\Delta_{3}^{(0,l_{2})}.

The reader may then easily check that the same holds for l2=3l_{2}=3:

Γ3(0,3)={u2(3),u2(2),u2(1)},\Gamma_{3}^{(0,3)}=\left\{\frac{\partial}{\partial u_{2}^{(3)}},\frac{\partial}{\partial u_{2}^{(2)}},\frac{\partial}{\partial u_{2}^{(1)}}\right\},
Δ3(0,3)={u2(0),u1(0),x3,x2,x1}=Δ3(0,3)¯.\Delta_{3}^{(0,3)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial x_{1}}\right\}=\overline{\Delta_{3}^{(0,3)}}.
𝐅𝐨𝐫𝐚𝐥𝐥𝐤𝟒.\bullet\leavevmode\nobreak\ \mathbf{For\leavevmode\nobreak\ all\leavevmode\nobreak\ k\geq 4.}

Since adg0(0,l2)ku1(0)\mathrm{ad}_{g_{0}^{(0,l_{2})}}^{k}\frac{\partial}{\partial u_{1}^{(0)}} and adg0(0,l2)k3u2(0)\mathrm{ad}_{g_{0}^{(0,l_{2})}}^{k-3}\frac{\partial}{\partial u_{2}^{(0)}} are linear combinations of x1\frac{\partial}{\partial x_{1}}, x2\frac{\partial}{\partial x_{2}} and x3\frac{\partial}{\partial x_{3}} only, we have, for all l24l_{2}\geq 4:

Δk(0,l2)=Δk(0,l2)¯,[Γk(0,l2),Δk(0,l2)]Δk(0,l2),\Delta_{k}^{(0,l_{2})}=\overline{\Delta_{k}^{(0,l_{2})}},\quad\left[\Gamma_{k}^{(0,l_{2})},\Delta_{k}^{(0,l_{2})}\right]\subset\Delta_{k}^{(0,l_{2})},

and, for l2=3l_{2}=3, using the fact that adg0(0,3)u2(0)=x3x1x1x2x2x3x4\mathrm{ad}_{g_{0}^{(0,3)}}\frac{\partial}{\partial u_{2}^{(0)}}=-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}}, we have

Γ4(0,3)={u2(3),u2(2),u2(1)},\Gamma_{4}^{(0,3)}=\left\{\frac{\partial}{\partial u_{2}^{(3)}},\frac{\partial}{\partial u_{2}^{(2)}},\frac{\partial}{\partial u_{2}^{(1)}}\right\},
Δ4(0,3)\displaystyle\Delta_{4}^{(0,3)} ={u2(0),x3x1x1x2x2x3x4,u1(0),x3,x2,x1}\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial x_{1}}\right\}
={u2(0),u1(0),x4,x3,x2,x1}=Δ4(0,3)¯=T6\displaystyle=\left\{\frac{\partial}{\partial u_{2}^{(0)}},\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial x_{4}},\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial x_{1}}\right\}=\overline{\Delta_{4}^{(0,3)}}=\mathrm{T}\mathbb{R}^{6}

and

Δk(0,3)=T6k4,\Delta_{k}^{(0,3)}=\mathrm{T}\mathbb{R}^{6}\quad\forall k\geq 4,

hence j2=3j_{2}=3 and k(0,3)=4k_{\ast}^{(0,3)}=4.

We conclude that the conditions of theorem 4.1 are satisfied for all kk provided that j2=3j_{2}=3.

On the contrary, initializing the algorithm by H1,20H_{1,2}^{0}, which amounts to prolonging the first input u1(1)u_{1}^{(1)}, we may easily check that, for all l13l_{1}\geq 3,

Δ2(l1,0)={u2(0),x3x1x1x2x2x3x4,(x1u1(0))x1x3x3}\Delta_{2}^{(l_{1},0)}=\left\{\frac{\partial}{\partial u_{2}^{(0)}},-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}},(x_{1}-u_{1}^{(0)})\frac{\partial}{\partial x_{1}}-x_{3}\frac{\partial}{\partial x_{3}}\right\}

is not involutive, thus contradicting condition (i) of Theorem 4.1, which proves that the minimal prolongation of the second input is equal to 3.

Let us finally give the construction of the flat output and prolonged state diffeomorphism. The prolonged state is (x1,x2,x3,x4,u1(0),u2(0),u2(1),u2(2),u2(3))(x_{1},x_{2},x_{3},x_{4},u_{1}^{(0)},u_{2}^{(0)},u_{2}^{(1)},u_{2}^{(2)},u_{2}^{(3)}) of dimension 9=n+m+𝐣9=n+m+\mid\mathbf{j}\mid.

We get

G0(0,3)={u1(0),u2(3)}=G0(0,3)¯,G1(0,3)={x3,u2(2)}G0(0,3)=G1(0,3)¯,\displaystyle G_{0}^{(0,3)}=\left\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(3)}}\right\}=\overline{G_{0}^{(0,3)}},\quad G_{1}^{(0,3)}=\left\{\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial u_{2}^{(2)}}\right\}\oplus G_{0}^{(0,3)}=\overline{G_{1}^{(0,3)}},
G2(0,3)={u2(0)x1+x2,u2(1)}G1(0,3)=G2(0,3)¯,\displaystyle G_{2}^{(0,3)}=\left\{u_{2}^{(0)}\frac{\partial}{\partial x_{1}}+\frac{\partial}{\partial x_{2}},\frac{\partial}{\partial u_{2}^{(1)}}\right\}\oplus G_{1}^{(0,3)}=\overline{G_{2}^{(0,3)}},
G3(0,3)={(u2(1)1)x1u2(0)x3,u2(0)}G2(0,3)=G3(0,3)¯,\displaystyle G_{3}^{(0,3)}=\left\{(u_{2}^{(1)}-1)\frac{\partial}{\partial x_{1}}-u_{2}^{(0)}\frac{\partial}{\partial x_{3}},\frac{\partial}{\partial u_{2}^{(0)}}\right\}\oplus G_{2}^{(0,3)}=\overline{G_{3}^{(0,3)}},
G4(0,3){x3x1x1x2x2x3x4}G4(0,3)=T9\displaystyle G_{4}^{(0,3)}\left\{-x_{3}\frac{\partial}{\partial x_{1}}-x_{1}\frac{\partial}{\partial x_{2}}-x_{2}\frac{\partial}{\partial x_{3}}-\frac{\partial}{\partial x_{4}}\right\}\oplus G_{4}^{(0,3)}=\mathrm{T}\mathbb{R}^{9}

The Brunovský’s controllability indices are κ1(0,3)=5\kappa_{1}^{(0,3)}=5 and κ2(0,3)=4\kappa_{2}^{(0,3)}=4 and the system of PDE’s that the flat outputs must satisfy is :

Gk(0,3),dy1=0,k=0,,3,G4(0,3),dy10\displaystyle\left<G_{k}^{(0,3)},dy_{1}\right>=0,\quad k=0,\ldots,3,\qquad\left<G_{4}^{(0,3)},dy_{1}\right>\neq 0
Gk(0,3),dy2=0,k=0,,2,G3(0,3),dy20.\displaystyle\left<G_{k}^{(0,3)},dy_{2}\right>=0,\quad k=0,\ldots,2,\qquad\left<G_{3}^{(0,3)},dy_{2}\right>\neq 0.

Its solution is given by y1=x4y_{1}=x_{4}, y2=x1u2(0)x2y_{2}=x_{1}-u_{2}^{(0)}x_{2}.

5.4 The Pendulum Example [10, section II. C]

This model of pendulum in the vertical plane has been studied in [10, section II. C], [21, section 6.2.3],[22, section 5.3] where it is shown to be flat. We prove here that it is not flat by pure prolongation.

Though naturally control-affine, it is presented here in its prolonged form (2.16):

x˙1=x2x˙2=u1(0)y˙1=y2y˙2=u2(0)θ˙1=θ2θ˙2=u1(0)εcosθ1+u2(0)+1εsinθ1u˙1(0)=u1(1)u˙2(0)=u2(1).\begin{array}[]{cclcccl}\displaystyle\dot{x}_{1}&=&\displaystyle x_{2}&\nobreak\leavevmode\hfil&\displaystyle\dot{x}_{2}&=&\displaystyle u_{1}^{(0)}\\ \displaystyle\dot{y}_{1}&=&\displaystyle y_{2}&\nobreak\leavevmode\hfil&\displaystyle\dot{y}_{2}&=&\displaystyle u_{2}^{(0)}\\ \displaystyle\dot{\theta}_{1}&=&\displaystyle\theta_{2}&\nobreak\leavevmode\hfil&\displaystyle\dot{\theta}_{2}&=&\displaystyle-\frac{u_{1}^{(0)}}{\varepsilon}\cos\theta_{1}+\frac{u_{2}^{(0)}+1}{\varepsilon}\sin\theta_{1}\\ \displaystyle\dot{u}_{1}^{(0)}&=&\displaystyle u_{1}^{(1)}&\nobreak\leavevmode\hfil&\displaystyle\dot{u}_{2}^{(0)}&=&\displaystyle u_{2}^{(1)}.\end{array} (5.12)

The state is (x1,x2,y1,y2,θ1,θ2,u1(0),u2(0))(x_{1},x_{2},y_{1},y_{2},\theta_{1},\theta_{2},u_{1}^{(0)},u_{2}^{(0)}), of dimension n+m=6+2=8n+m=6+2=8. The associated non prolonged vector fields are

g0(0)\displaystyle g_{0}^{(0)} =x2x1+y2y1+θ2θ1+1εsinθ1θ2\displaystyle=x_{2}\frac{\partial}{\partial x_{1}}+y_{2}\frac{\partial}{\partial y_{1}}+\theta_{2}\frac{\partial}{\partial\theta_{1}}+\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}} (5.13)
+u1(0)(x21εcosθ1θ2)+u2(0)(y2+1εsinθ1θ2)\displaystyle+u_{1}^{(0)}\left(\frac{\partial}{\partial x_{2}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}}\right)+u_{2}^{(0)}\left(\frac{\partial}{\partial y_{2}}+\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\right)
g1(0)\displaystyle g_{1}^{(0)} =u1(0)g2(0)=u2(0)\displaystyle=\frac{\partial}{\partial u_{1}^{(0)}}\qquad g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}}

We have

G0(𝟎)={g1(0),g2(0)}={u1(0),u2(0)}G_{0}^{(\mathbf{0})}=\{g_{1}^{(0)},g_{2}^{(0)}\}=\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}}\}

and

G1(𝟎)\displaystyle G_{1}^{(\mathbf{0})} ={g1(0),g2(0),adg0(𝟎)g10,adg0(𝟎)g20}\displaystyle=\{g_{1}^{(0)},g_{2}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{1}^{0},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{2}^{0}\}
={u1(0),u2(0),x2+1εcosθ1θ2,y21εsinθ1θ2}=G1(𝟎)¯\displaystyle=\{\frac{\partial}{\partial u_{1}^{(0)}},\frac{\partial}{\partial u_{2}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},-\frac{\partial}{\partial y_{2}}-\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\}=\overline{G_{1}^{(\mathbf{0})}}

but

G2(𝟎)\displaystyle G_{2}^{(\mathbf{0})} =G1(𝟎)+{x11εcosθ1θ1θ2εsinθ1θ2,y1+1εsinθ1θ1θ2εcosθ1θ2}\displaystyle=G_{1}^{(\mathbf{0})}+\{\frac{\partial}{\partial x_{1}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{\theta_{2}}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial y_{1}}+\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{\theta_{2}}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}}\}
G2(𝟎)¯.\displaystyle\neq\overline{G_{2}^{(\mathbf{0})}}.

As before, we may initialize the pure prolongation algorithm by choosing

H1,10={g1(0),adg0(𝟎)g10}={u1(0),x2+1εcosθ1θ2}=H1,10¯H_{1,1}^{0}=\{g_{1}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{1}^{0}\}=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}}\}=\overline{H_{1,1}^{0}}

which amounts to prolonging u2(1)u_{2}^{(1)}, or

H1,20={g2(0),adg0(𝟎)g20}={u2(0),y21εsinθ1θ2}=H1,20¯H_{1,2}^{0}=\{g_{2}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{2}^{0}\}=\{\frac{\partial}{\partial u_{2}^{(0)}},-\frac{\partial}{\partial y_{2}}-\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\}=\overline{H_{1,2}^{0}}

which amounts to prolonging u1(1)u_{1}^{(1)}.

It is noticeable that the inputs u1(1)u_{1}^{(1)} and u2(1)u_{2}^{(1)} play a symmetric role . Thus, one may choose indifferently one of them as the non prolonged input. Let us choose u1(1)u_{1}^{(1)} as non prolonged input, with the initialization H1,10H_{1,1}^{0}. Thus, the vector fields associated to a prolongation of length l2l_{2} on the second input read:

g0(0,l2)\displaystyle g_{0}^{(0,l_{2})} =x2x1+y2y1+θ2θ1+1εsinθ1θ2+u1(0)(x21εcosθ1θ2)\displaystyle=x_{2}\frac{\partial}{\partial x_{1}}+y_{2}\frac{\partial}{\partial y_{1}}+\theta_{2}\frac{\partial}{\partial\theta_{1}}+\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}+u_{1}^{(0)}\left(\frac{\partial}{\partial x_{2}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}}\right) (5.14)
+u2(0)(y2+1εsinθ1θ2)+p=0l21u2(p+1)u2(p)\displaystyle+u_{2}^{(0)}\left(\frac{\partial}{\partial y_{2}}+\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\right)+\sum_{p=0}^{l_{2}-1}u_{2}^{(p+1)}\frac{\partial}{\partial u_{2}^{(p)}}
g1(0)\displaystyle g_{1}^{(0)} =u1(0)g2(l2)=u2(l2).\displaystyle=\frac{\partial}{\partial u_{1}^{(0)}}\qquad g_{2}^{(l_{2})}=\frac{\partial}{\partial u_{2}^{(l_{2})}}.
𝐤=𝟏.\bullet\leavevmode\nobreak\ \mathbf{k=1.}

For all l22l_{2}\geq 2,

Γ1(0,l2)={u2(l2),u2(l21)},Δ1(0,l2)={u1(0),x2+1εcosθ1θ2},\Gamma_{1}^{(0,l_{2})}=\{\frac{\partial}{\partial u_{2}^{(l_{2})}},\frac{\partial}{\partial u_{2}^{(l_{2}-1)}}\},\quad\Delta_{1}^{(0,l_{2})}=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}}\},

and if l2=1l_{2}=1,

Γ1(0,1)={u2(1)},Δ1(0,l2)={u1(0),x2+1εcosθ1θ2,u2(0)}.\Gamma_{1}^{(0,1)}=\{\frac{\partial}{\partial u_{2}^{(1)}}\},\quad\Delta_{1}^{(0,l_{2})}=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial u_{2}^{(0)}}\}.

Thus, we have Δ1(0,l2)¯=Δ1(0,l2)\overline{\Delta_{1}^{(0,l_{2})}}=\Delta_{1}^{(0,l_{2})} and [Γ1(0,l2),Δ1(0,l2)]Δ1(0,l2)[\Gamma_{1}^{(0,l_{2})},\Delta_{1}^{(0,l_{2})}]\subset\Delta_{1}^{(0,l_{2})} for all l21l_{2}\geq 1.


𝐤=𝟐.\bullet\leavevmode\nobreak\ \mathbf{k=2.}

For all l23l_{2}\geq 3,

Γ2(0,l2)\displaystyle\Gamma_{2}^{(0,l_{2})} ={u2(l2),u2(l21),u2(l22)},\displaystyle=\{\frac{\partial}{\partial u_{2}^{(l_{2})}},\frac{\partial}{\partial u_{2}^{(l_{2}-1)}},\frac{\partial}{\partial u_{2}^{(l_{2}-2)}}\},
Δ2(0,l2)\displaystyle\Delta_{2}^{(0,l_{2})} ={u1(0),x2+1εcosθ1θ2,x11εcosθ1θ11εθ2sinθ1θ2}.\displaystyle=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial x_{1}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{1}{\varepsilon}\theta_{2}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\}.

If l2=2l_{2}=2,

Γ2(0,2)\displaystyle\Gamma_{2}^{(0,2)} ={u2(2),u2(1)},\displaystyle=\{\frac{\partial}{\partial u_{2}^{(2)}},\frac{\partial}{\partial u_{2}^{(1)}}\},
Δ2(0,2)\displaystyle\Delta_{2}^{(0,2)} ={u1(0),x2+1εcosθ1θ2,x11εcosθ1θ11εθ2sinθ1θ2,u2(0)}.\displaystyle=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial x_{1}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{1}{\varepsilon}\theta_{2}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial u_{2}^{(0)}}\}.

If l2=1l_{2}=1,

Γ2(0,1)\displaystyle\Gamma_{2}^{(0,1)} ={u2(1)},\displaystyle=\{\frac{\partial}{\partial u_{2}^{(1)}}\},
Δ2(0,1)\displaystyle\Delta_{2}^{(0,1)} ={u1(0),x2+1εcosθ1θ2,x11εcosθ1θ11εθ2sinθ1θ2,\displaystyle=\{\frac{\partial}{\partial u_{1}^{(0)}},-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial x_{1}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{1}{\varepsilon}\theta_{2}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}},
u2(0),y21εsinθ1θ2}.\displaystyle\hskip 113.81102pt\frac{\partial}{\partial u_{2}^{(0)}},-\frac{\partial}{\partial y_{2}}-\frac{1}{\varepsilon}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}\}.

Again, we have [Γ2(0,l2),Δ2(0,l2)]Δ2(0,l2)[\Gamma_{2}^{(0,l_{2})},\Delta_{2}^{(0,l_{2})}]\subset\Delta_{2}^{(0,l_{2})} for all l21l_{2}\geq 1, but Δ2(0,l2)¯Δ2(0,l2)\overline{\Delta_{2}^{(0,l_{2})}}\neq\Delta_{2}^{(0,l_{2})} for all l21l_{2}\geq 1 since, e.g.

[x2+1εcosθ1θ2,x11εcosθ1θ11εθ2sinθ1θ2]=1ε2sin2θ1θ2Δ2(0,l2)[-\frac{\partial}{\partial x_{2}}+\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{2}},\frac{\partial}{\partial x_{1}}-\frac{1}{\varepsilon}\cos\theta_{1}\frac{\partial}{\partial\theta_{1}}-\frac{1}{\varepsilon}\theta_{2}\sin\theta_{1}\frac{\partial}{\partial\theta_{2}}]=-\frac{1}{\varepsilon^{2}}\sin 2\theta_{1}\frac{\partial}{\partial\theta_{2}}\not\in\Delta_{2}^{(0,l_{2})}

for all l21l_{2}\geq 1. Changing the non prolonged input u1(0)u_{1}^{(0)} in u2(0)u_{2}^{(0)}, as previously announced, a similar calculation, left to the reader, shows that Δ2(l1,0)\Delta_{2}^{(l_{1},0)} is not involutive for all l11l_{1}\geq 1. Thus, according to the first item of theorem 4.2, system (5.12) is not flat by pure prolongation, though differentially flat, as shown in [10, section C].

5.5 An example with 3 inputs [18]

x˙1=u1(0)x˙2=x3u1(0)x˙3=x4u1(0)+x1u3(0)x˙4=u2(0).\begin{array}[]{ccl}\dot{x}_{1}&=&u_{1}^{(0)}\\ \dot{x}_{2}&=&x_{3}u_{1}^{(0)}\\ \dot{x}_{3}&=&x_{4}u_{1}^{(0)}+x_{1}u_{3}^{(0)}\\ \dot{x}_{4}&=&u_{2}^{(0)}.\end{array} (5.15)

The state is (x1,,x4,u1(0),u2(0),u3(0))(x_{1},\ldots,x_{4},u_{1}^{(0)},u_{2}^{(0)},u_{3}^{(0)}) (of dimension n+m=4+3=7n+m=4+3=7) and the control inputs are (u1(1),u2(1),u3(1))(u_{1}^{(1)},u_{2}^{(1)},u_{3}^{(1)}). The reader may easily see that G1(𝟎)G_{1}^{(\mathbf{0})} is not involutive and thus that this system is not static feedback linearizable.

This example may be found in [18] in the context of control affine systems with nn states and n1n-1 inputs with the property that there exists i=1,2,3i=1,2,3 such that

dim{f1,f2,f3,adgfi}=4,\dim\{f_{1},f_{2},f_{3},\mathrm{ad}{g}f_{i}\}=4, (5.16)

with

f1=x1+x3x2+x4x3,f2=x4,f3=x1x3,g=u1(0)f1+u2(0)f2+u3(0)f3,f_{1}=\frac{\partial}{\partial x_{1}}+x_{3}\frac{\partial}{\partial x_{2}}+x_{4}\frac{\partial}{\partial x_{3}},\quad f_{2}=\frac{\partial}{\partial x_{4}},\quad f_{3}=x_{1}\frac{\partial}{\partial x_{3}},\quad g=u_{1}^{(0)}f_{1}+u_{2}^{(0)}f_{2}+u_{3}^{(0)}f_{3},

a property satisfied for i=3i=3 since

adgf3=u1(0)[f1,f3]=u1(0)(x3x1x2){f1,f2,f3}.\mathrm{ad}{g}f_{3}=u_{1}^{(0)}[f_{1},f_{3}]=u_{1}^{(0)}\left(\frac{\partial}{\partial x_{3}}-x_{1}\frac{\partial}{\partial x_{2}}\right)\not\in\{f_{1},f_{2},f_{3}\}.

Therefore, in [18, Theorem 3], flat outputs are exhibited as 3 independent first integrals of f3f_{3}, yielding the prolongation u˙1(1)=u1(2)\dot{u}_{1}^{(1)}=u_{1}^{(2)}, u˙2(1)=u2(2)\dot{u}_{2}^{(1)}=u_{2}^{(2)}, with 𝐣=(1,1,0)\mathbf{j}=(1,1,0) and thus 𝐣=2\mid\mathbf{j}\mid=2.

We show here that this prolongation is not minimal and we compute the minimal one.

We initialize the algorithm by remarking that {f2,f3}\{f_{2},f_{3}\} is the unique maximal involutive subdistribution of {f1,f2,f3}\{f_{1},f_{2},f_{3}\}, thus indicating that u2(1)u_{2}^{(1)} and u3(1)u_{3}^{(1)} are suitable candidates of non prolonged inputs. We thus, consider the following vector fields, corresponding to prolongations of u1(1)u_{1}^{(1)}, for all l1l\geq 1:

g0(l,0,0)u1(0)(x1+x3x2+x4x3)+u2(0)x4+u3(0)x1x3+p=0l1u1(p+1)u1(p),g_{0}^{(l,0,0)}\triangleq u_{1}^{(0)}\left(\frac{\partial}{\partial x_{1}}+x_{3}\frac{\partial}{\partial x_{2}}+x_{4}\frac{\partial}{\partial x_{3}}\right)+u_{2}^{(0)}\frac{\partial}{\partial x_{4}}+u_{3}^{(0)}x_{1}\frac{\partial}{\partial x_{3}}+\sum_{p=0}^{l-1}u_{1}^{(p+1)}\frac{\partial}{\partial u_{1}^{(p)}},
g1(l)=u1(l),g2(0)=u2(0),g3(0)=u3(0).g_{1}^{(l)}=\frac{\partial}{\partial u_{1}^{(l)}},\qquad g_{2}^{(0)}=\frac{\partial}{\partial u_{2}^{(0)}},\qquad g_{3}^{(0)}=\frac{\partial}{\partial u_{3}^{(0)}}.

The reader may easily check that Δk(l,0,0)=Δk(l,0,0)¯\Delta_{k}^{(l,0,0)}=\overline{\Delta_{k}^{(l,0,0)}} and [Γk(l,0,0),Δk(l,0,0)]Δk(l,0,0)[\Gamma_{k}^{(l,0,0)},\Delta_{k}^{(l,0,0)}]\subset\Delta_{k}^{(l,0,0)} for all l1l\geq 1 and k=0,1,2k=0,1,2, and that dimΔ2(1,0,0)=n+m=7\dim\Delta_{2}^{(1,0,0)}=n+m=7. Therefore, the prolongation 𝐣=(1,0,0)\mathbf{j}=(1,0,0), with 𝐣=1\mid\mathbf{j}\mid=1, is minimal and the minimally purely prolonged system is equivalent by diffeomorphism and feedback to

y1(3)=u1(2),y¨2=u2(1)y3(3)=W(y1,,y1(3),y2,,y¨2,y3,,y¨3,u3(1))y_{1}^{(3)}=u_{1}^{(2)},\qquad\ddot{y}_{2}=u_{2}^{(1)}\qquad y_{3}^{(3)}=W(y_{1},\ldots,y_{1}^{(3)},y_{2},\ldots,\ddot{y}_{2},y_{3},\ldots,\ddot{y}_{3},u_{3}^{(1)})

with flat outputs

y1=x1,y2=x4,y3=x2y_{1}=x_{1},\quad y_{2}=x_{4},\quad y_{3}=x_{2}

and with

W(y1,,y1(3),y2,,y¨2,y3,,y¨3,u3(1))y2y˙1(y¨1(y˙1)2y1)+y˙2(y˙1)2+y3y1(3)y˙1y˙3y¨1(1y1+2y¨1(y˙1)2)+y¨3(y˙1y1+2y¨1y˙1)+y1y˙1u3(1).\begin{array}[]{l}\displaystyle W(y_{1},\ldots,y_{1}^{(3)},y_{2},\ldots,\ddot{y}_{2},y_{3},\ldots,\ddot{y}_{3},u_{3}^{(1)})\triangleq y_{2}\dot{y}_{1}\left(\ddot{y}_{1}-\frac{\left(\dot{y}_{1}\right)^{2}}{y_{1}}\right)+\dot{y}_{2}\left(\dot{y}_{1}\right)^{2}\\ \displaystyle\hskip 56.9055pt+y_{3}\frac{y_{1}^{(3)}}{\dot{y}_{1}}-\dot{y}_{3}\ddot{y}_{1}\left(\frac{1}{y_{1}}+2\frac{\ddot{y}_{1}}{(\dot{y}_{1})^{2}}\right)+\ddot{y}_{3}\left(\frac{\dot{y}_{1}}{y_{1}}+2\frac{\ddot{y}_{1}}{\dot{y}_{1}}\right)+y_{1}\dot{y}_{1}u_{3}^{(1)}.\end{array}

6 Concluding Remarks

We have established necessary and sufficient conditions for a system to be flat by pure prolongation, i.e. belonging to the equivalence class of 0 with respect to the equivalence by pure prolongation relation. These conditions extend preliminary results of [6, 7, 33, 34, 2, 12] thanks to a thorough study of purely prolonged vector fields. We then deduce a computationally tractable algorithm giving the minimal prolongation in a finite number of steps using only Lie brackets and linear algebra.

Possible extensions of this work towards general flatness necessary and sufficient conditions are under study.

Acknowledgements— The author wishes to express his warm thanks to Ph. Martin and Y. Kaminski for many fruitful discussions.

References

  • [1] E. Aranda-Bricaire, C.H. Moog, and J.-B. Pomet, A linear algebraic framework for dynamic feedback linearization, IEEE Trans. Automat. Contr. 40 (1995), no. 1, 127–132.
  • [2] S. Battilotti and C. Califano, A constructive condition for dynamic feedback linearization, Systems & Control Letters 52 (2004), 329–338.
  • [3] N. Bourbaki, Algèbre, Éléments de Mathématiques, Chapitres 1-3, Hermann, Paris, 1970.
  • [4] P. Brunovský, A classification of linear controllable systems, Kybernetika 6 (1970), 173–178.
  • [5] É. Cartan, Sur l’équivalence absolue de certains systèmes d’équations différentielles et sur certaines familles de courbes, Bull. Soc. Math. France 42 (1914), 12–48, in Oeuvres Complètes, part II, vol 2, pages 1133–1168, CNRS, Paris, 1984.
  • [6] B. Charlet, J. Lévine, and R. Marino, On dynamic feedback linearization, Systems & Control Letters 13 (1989), 143–151.
  • [7]  , Sufficient conditions for dynamic state feedback linearization, SIAM J. on Control and Optimization 29 (1991), no. 1, 38–57.
  • [8] C. Chevalley, Theory of Lie Groups, Princeton University Press, 1946.
  • [9] M. Fliess, J. Lévine, Ph. Martin, and P. Rouchon, Flatness and defect of nonlinear systems: introductory theory and examples, Int. J. Control 61 (1995), no. 6, 1327–1361.
  • [10]  , A Lie-Bäcklund approach to equivalence and flatness of nonlinear systems, IEEE Trans. Automat. Contr. 44 (1999), no. 5, 922–937.
  • [11] J. Franch, Flatness, tangent systems and flat outputs, Ph.D. thesis, Universitat Politecnica de Catalunya, 1999.
  • [12] J. Franch and E. Fossas, Linearization by prolongations: New bound on the number of integrators, European Journal of Control 11 (2005), 171–179.
  • [13] R. Hermann and A.J. Krener, Nonlinear controllability and observability, IEEE Trans. Automat. Contr. 22 (1977), 728–740.
  • [14] L.R. Hunt, R. Su, and G. Meyer, Design for multi-input nonlinear systems, Differential Geometric Control Theory (R.W. Brockett, R.S. Millman, and H.J. Sussmann, eds.), Birkhäuser, Boston, 1983, pp. 268–298.
  • [15]  , Global transformations of nonlinear systems, IEEE Trans. Automat. Contr. 28 (1983), 24–31.
  • [16] A. Isidori, Nonlinear Control Systems. An Introduction, 1st ed., Springer, Berlin, New York, 1985.
  • [17] B. Jakubczyk and W. Respondek, On linearization of control systems, Bull. Acad. Pol. Sci. Ser. Sci. Math. 28 (1980), no. 9–10, 517–522.
  • [18] Y. J. Kaminski, J. Lévine, and F. Ollivier, On singularities of flat affine systems with nn states and n1n-1 controls, International Journal of Robust and Nonlinear Control 30 (2020), no. 9, 3547–3565.
  • [19] I. S. Krasil’shchik, V. V. Lychagin, and A. M. Vinogradov, Geometry of Jet Spaces and Nonlinear Partial Differential Equations, Gordon and Breach, New York, 1986.
  • [20] J. Lévine, Static and dynamic state feedback linearization, Nonlinear Systems (A. Fossard and D. Normand-Cyrot, eds.), vol. 3, Chapman & Hall, 1997, pp. 93–126.
  • [21]  , Analysis and Control of Nonlinear Systems: A Flatness-based Approach, Mathematical Engineering, Springer, 2009.
  • [22]  , On necessary and sufficient conditions for differential flatness, Applicable Algebra in Engineering, Communication and Computing 22 (2011), 47–90.
  • [23] R. Marino, On the largest feedback linearizable subsystem, Systems & Control Letters 6 (1986), 345–351.
  • [24] Ph. Martin, Contribution à l’étude des systèmes diffèrentiellement plats, Ph.D. thesis, École des Mines de Paris, 1992.
  • [25] Ph. Martin, R.M. Murray, and P. Rouchon, Flat systems, Plenary Lectures and Minicourses, Proc. ECC 97, Brussels (G. Bastin and M. Gevers, eds.), 1997, pp. 211–264.
  • [26] Ph. Martin and P. Rouchon, Any (controllable) driftless system with 3 inputs and 5 states is flat, Systems & Control Letters 25 (1995), 167–173.
  • [27]  , Any (controllable) driftless system with m inputs and m+2 states is flat, Proc. of the 34th IEEE Conf. on Decision and Control, New Orleans, 1995, pp. 2886–2891.
  • [28] H. Nijmeijer and A.J. van der Schaft, Nonlinear Dynamical Control Systems, Springer, New York, 1990.
  • [29] J.-B. Pomet, A differential geometric setting for dynamic equivalence and dynamic linearization, Geometry in Nonlinear Control and Differential Inclusions (B. Jakubczyk, W. Respondek, and T. Rzeżuchowski, eds.), Banach Center Publications, Warsaw, 1993, pp. 319–339.
  • [30]  , A necessary condition for dynamic equivalence, SIAM J. on Control and Optimization 48 (2009), 925–940.
  • [31] A.J. van der Schaft, Linearization and input-output decoupling for general nonlinear systems, Systems & Control Letters 5 (1984), 27–33.
  • [32] W.F. Shadwick, Absolute equivalence and dynamic feedback linearization, Systems & Control Letters 15 (1990), 35–39.
  • [33] W.M. Sluis, A necessary condition for dynamic feedback linearization, Systems & Control Letters 21 (1993), 277–283.
  • [34] W.M. Sluis and D.M. Tilbury, A bound on the number of integrators needed to linearize a control system, Systems & Control Letters 29 (1996), no. 1, 43–50.
  • [35] H.J. Sussmann and V. Jurdjevic, Controllability of nonlinear systems, J. Differential Equations 12 (1972), 95–116.
  • [36] V.V. Zharinov, Geometrical Aspect of Partial Differential Equations, World Scientific, Singapore, 1992.

Appendix A A Comparison Formula

We establish a comparison formula expressing the iterated Lie brackets adg0(𝐣)kgi(ji)\mathrm{ad}^{k}_{g_{0}^{(\mathbf{j})}}g_{i}^{(j_{i})} for all k0k\geq 0, in terms of a combination of Lie brackets of the vector fields g0(𝟎)=fxg_{0}^{(\mathbf{0})}=f\frac{\partial}{\partial x}, g1(0)=u1(0)g_{1}^{(0)}=\frac{\partial}{\partial u_{1}^{(0)}}, …, gm(0)=um(0)g_{m}^{(0)}=\frac{\partial}{\partial u_{m}^{(0)}} of the original system (2.16).

For simplicity’s sake, we introduce a fictitious input u0u0(0)tu_{0}\triangleq u_{0}^{(0)}\triangleq t so that u0(1)=1u_{0}^{(1)}=1, j0=1j_{0}=1, and u0(k)=0u_{0}^{(k)}=0 for all k>1k>1. Thus, l=0j01u0(l+1)g0(𝐥)=u0(1)g0(𝟎)=g0(𝟎)\sum_{l=0}^{j_{0}-1}u_{0}^{(l+1)}g_{0}^{(\mathbf{l})}=u_{0}^{(1)}g_{0}^{(\mathbf{0})}=g_{0}^{(\mathbf{0})}, and

g0(𝐣)=g0(𝟎)+k=1ml=0jk1uk(l+1)gk(l)=k=0ml=0jk1uk(l+1)gk(l).g_{0}^{(\mathbf{j})}=g_{0}^{(\mathbf{0})}+\sum_{k=1}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}g_{k}^{(l)}=\sum_{k=0}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}g_{k}^{(l)}.
Lemma A.1.

For every i=1,,mi=1,\ldots,m, we have:

adg0(𝐣)kgi(ji)=(1)kui(jik)=(1)kgi(jik)ifkji,\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{k}g_{i}^{(j_{i})}=(-1)^{k}\frac{\partial}{\partial u_{i}^{(j_{i}-k)}}=(-1)^{k}g_{i}^{(j_{i}-k)}\quad\mathrm{if\leavevmode\nobreak\ }k\leq j_{i}, (A.1)

and, for k=ji+νk=j_{i}+\nu, for all ν1\nu\geq 1,

adg0(𝐣)ji+νgi(ji)=(1)ji(r,𝐥,𝐩)𝕁i,ν(𝐣)cν,𝐥𝐩upr(lr)up1(l1)[gpr(0),,[gp1(0),gi(0)]]\displaystyle\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+\nu}g_{i}^{(j_{i})}=(-1)^{j_{i}}\sum_{(r,\mathbf{l},\mathbf{p})\in\mathbb{J}^{(\mathbf{j})}_{i,\nu}}c_{\nu,\mathbf{l}}^{\mathbf{p}}u_{p_{r}}^{(l_{r})}\cdots u_{p_{1}}^{(l_{1})}[g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots] (A.2)
=(1)jiadg0(𝟎)νgi(0)+(q,r,𝐥,𝐩)𝕀i,ν(𝐣)cν,q,𝐥𝐩upr(lr)up1(l1)[gpr(0),,[gp1(0),adg0(𝟎)qgi(0)]]\displaystyle=(-1)^{j_{i}}\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{\nu}g_{i}^{(0)}+\sum_{(q,r,\mathbf{l},\mathbf{p})\in\mathbb{I}^{(\mathbf{j})}_{i,\nu}}c_{\nu,q,\mathbf{l}}^{\mathbf{p}}u_{p_{r}}^{(l_{r})}\cdots u_{p_{1}}^{(l_{1})}[g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{q}g_{i}^{(0)}]\ldots] (A.3)

where 𝕁i,ν(𝐣)\mathbb{J}^{(\mathbf{j})}_{i,\nu} (resp. 𝕀i,ν\mathbb{I}_{i,\nu}) is the set of multi-integers (r,𝐥,𝐩)(r,l1,,lr,p1,,pr)(r,\mathbf{l},\mathbf{p})\triangleq(r,l_{1},\ldots,l_{r},p_{1},\ldots,p_{r}) (resp. (q,r,𝐥,𝐩)(q,r,l1,,lr,p1,,pr)(q,r,\mathbf{l},\mathbf{p})\triangleq(q,r,l_{1},\ldots,l_{r},p_{1},\ldots,p_{r})) defined by

𝕁i,ν(𝐣){(r,𝐥,𝐩)1rν,0pαm, 1lαjα,α=1,,r,𝐥=ν},\displaystyle\mathbb{J}^{(\mathbf{j})}_{i,\nu}\triangleq\{(r,\mathbf{l},\mathbf{p})\mid 1\leq r\leq\nu,\quad 0\leq p_{\alpha}\leq m,\leavevmode\nobreak\ 1\leq l_{\alpha}\leq j_{\alpha},\leavevmode\nobreak\ \alpha=1,\ldots,r,\leavevmode\nobreak\ \;\mid\mathbf{l}\mid=\nu\}, (A.4)
(resp.\displaystyle(resp.
𝕀i,ν(𝐣){q,r1, 0pαm, 1lαjα,α=1,,r,pr0,r𝐥=νq})\displaystyle\mathbb{I}^{(\mathbf{j})}_{i,\nu}\triangleq\left\{\displaystyle q,r\geq 1,\leavevmode\nobreak\ 0\leq p_{\alpha}\leq m,\leavevmode\nobreak\ 1\leq l_{\alpha}\leq j_{\alpha},\leavevmode\nobreak\ \alpha=1,\ldots,r,\leavevmode\nobreak\ p_{r}\neq 0,\leavevmode\nobreak\ r\leq\leavevmode\nobreak\ \mid\mathbf{l}\mid\leavevmode\nobreak\ =\nu-q\right\}) (A.5)

with 𝐥α=1rlα\mid\mathbf{l}\mid\triangleq\sum_{\alpha=1}^{r}l_{\alpha}, cν,𝐥𝐩cν,l1,,lrp1,,prc_{\nu,\mathbf{l}}^{\mathbf{p}}\triangleq c_{\nu,l_{1},\ldots,l_{r}}^{p_{1},\ldots,p_{r}}\in\mathbb{Z} (resp. cν,q,𝐥𝐩cν,q,l1,,lrp1,,prc_{\nu,q,\mathbf{l}}^{\mathbf{p}}\triangleq c_{\nu,q,l_{1},\ldots,l_{r}}^{p_{1},\ldots,p_{r}}\in\mathbb{Z}) and with the convention that the summation of the right-hand side of (A.2) (resp. (A.3)) vanishes if 𝕁i,ν(𝐣)=\mathbb{J}^{(\mathbf{j})}_{i,\nu}=\emptyset (resp. 𝕀i,ν(𝐣)=\mathbb{I}^{(\mathbf{j})}_{i,\nu}=\emptyset), each bracket [gpr(0),,[gp1(0),gi(0)]][g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots] in (A.2)(resp. [gpr(0),,[gp1(0),adg0(𝟎)qgi(0)]][g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{q}g_{i}^{(0)}]\ldots] in (A.3)) depending at most on x¯(𝟎)\overline{x}^{(\mathbf{0})}.

Proof.

By induction. It is immediately seen that

adg0(𝐣)gi(ji)=[fx+k=1ml=0jk1uk(l+1)uk(l),ui(ji)]=ui(ji1)=gi(ji1)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}g_{i}^{(j_{i})}=[f\frac{\partial}{\partial x}+\sum_{k=1}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(j_{i})}}]=-\frac{\partial}{\partial u_{i}^{(j_{i-1})}}=-g_{i}^{(j_{i}-1)}

Iterating this computation up to k=jik=j_{i} yields (A.1). In particular:

adg0(𝐣)jigi(ji)=(1)jiui(0)=(1)jigi(0).\mathrm{ad}^{j_{i}}_{g_{0}^{(\mathbf{j})}}g_{i}^{(j_{i})}=(-1)^{j_{i}}\frac{\partial}{\partial u_{i}^{(0)}}=(-1)^{j_{i}}g_{i}^{(0)}.

Then, for k=ji+1k=j_{i}+1, using the fact that [uk(l),ui(0)]=0[\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(0)}}]=0 for all ii, kk and l0l\geq 0, we have:

adg0(𝐣)ji+1gi(ji)\displaystyle\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+1}g_{i}^{(j_{i})} =[g0(𝐣),adg0(𝐣)jigi(ji)]=(1)ji[g0(𝐣),gi(𝟎)]\displaystyle=[g_{0}^{(\mathbf{j})},\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}}g_{i}^{(j_{i})}]=(-1)^{j_{i}}[g_{0}^{(\mathbf{j})},g_{i}^{(\mathbf{0})}] (A.6)
=(1)ji[fx+k=1ml=0jk1uk(l+1)uk(l),ui(0)]\displaystyle=(-1)^{j_{i}}[f\frac{\partial}{\partial x}+\sum_{k=1}^{m}\sum_{l=0}^{j_{k}-1}u_{k}^{(l+1)}\frac{\partial}{\partial u_{k}^{(l)}},\frac{\partial}{\partial u_{i}^{(0)}}]
=(1)ji[fx,ui(0)]=(1)jiadg0(𝟎)gi(0)\displaystyle=(-1)^{j_{i}}[f\frac{\partial}{\partial x},\frac{\partial}{\partial u_{i}^{(0)}}]=(-1)^{j_{i}}\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{i}^{(0)}

which proves that (A.2) and (A.3) hold at the order k=ji+1k=j_{i}+1, i.e. ν=1\nu=1, the summation of the right-hand side of (A.3) being equal to 0 since 𝕀i,1(𝐣)=\mathbb{I}^{(\mathbf{j})}_{i,1}=\emptyset. Furthermore, a direct calculation shows that

(1)jiadg0(𝟎)gi(0)=(1)(ji+1)fui(0)(x¯(𝟎))x,(-1)^{j_{i}}\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{i}^{(0)}=(-1)^{(j_{i}+1)}\frac{\partial f}{\partial u_{i}^{(0)}}(\overline{x}^{(\mathbf{0})})\frac{\partial}{\partial x}, (A.7)

which proves that adg0(𝐣)ji+1gi(ji)VX(𝟎)\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+1}g_{i}^{(j_{i})}\in\mathrm{V}X^{(\mathbf{0})}, where VX(𝟎)\mathrm{V}X^{(\mathbf{0})} is the vertical bundle of X(𝟎)X^{(\mathbf{0})}, i.e. the set of vector fields that are linear combinations of x1,,xn\frac{\partial}{\partial x_{1}},\ldots,\frac{\partial}{\partial x_{n}}, and whose coefficients are smooth functions that depend at most on x¯(𝟎)\overline{x}^{(\mathbf{0})} (but not of u(l)u^{(l)} for all l1l\geq 1). It results that [up(l),adg0(𝟎)gi(0)]=[gp(l),adg0(𝟎)gi(0)]=0[\frac{\partial}{\partial u_{p}^{(l)}},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{i}^{(0)}]=[g_{p}^{(l)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}g_{i}^{(0)}]=0 for all l1l\geq 1, all p=1,,mp=1,\ldots,m, and all i=1,,mi=1,\ldots,m.

Assume now that (A.2) and (A.3) hold up to k=ji+νk=j_{i}+\nu, with ν>1\nu>1, and that all the brackets [gpr(0),,[gp1(0),gi(0)]][g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots] and [gpr(0),[,[gp1(0),adg0(𝟎)qgi(0)]]][g_{p_{r}}^{(0)},[\ldots,[g_{p_{1}}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{q}g_{i}^{(0)}]\ldots]] depend at most on x¯(𝟎)\overline{x}^{(\mathbf{0})}. We immediately deduce that

[upr+1(lr+1),[gpr(0),[[gp1(0),adg0(𝟎)qgi(0)]]]]=[gpr+1(lr+1),[gp1(0),[[gpr(0),adg0(𝟎)qgi(0)]]]]=0[\frac{\partial}{\partial u_{p_{r+1}}^{(l_{r+1})}},[g_{p_{r}}^{(0)},[\ldots[g_{p_{1}}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{q}g_{i}^{(0)}]\ldots]]]=[g_{p_{r+1}}^{(l_{r+1})},[g_{p_{1}}^{(0)},[\ldots[g_{p_{r}}^{(0)},\mathrm{ad}_{g_{0}^{(\mathbf{0})}}^{q}g_{i}^{(0)}]\ldots]]]=0

for all lr+11l_{r+1}\geq 1, pr+1{1,,m}p_{r+1}\in\{1,\ldots,m\}, q{0,,ν1}q\in\{0,\ldots,\nu-1\}, r1r\geq 1, p1,,pr{0,,m}p_{1},\ldots,p_{r}\in\{0,\ldots,m\}, and i{1,,m}i\in\{1,\ldots,m\}. Thus:

adg0(𝐣)ji+ν+1gi(ji)=[g0(𝐣),adg0(𝐣)ji+νgi(ji)]\displaystyle\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+\nu+1}g_{i}^{(j_{i})}=[g_{0}^{(\mathbf{j})},\mathrm{ad}_{g_{0}^{(\mathbf{j})}}^{j_{i}+\nu}g_{i}^{(j_{i})}]
=(1)ji(r,𝐥,𝐩)𝕁i,ν(𝐣)[g0(𝐣),cν,𝐥𝐩upr(lr)up1(l1)[gpr(0),,[gp1(0),gi(0)]]]\displaystyle=(-1)^{j_{i}}\sum_{(r,\mathbf{l},\mathbf{p})\in\mathbb{J}_{i,\nu}^{(\mathbf{j})}}[g_{0}^{(\mathbf{j})},c_{\nu,\mathbf{l}}^{\mathbf{p}}u_{p_{r}}^{(l_{r})}\cdots u_{p_{1}}^{(l_{1})}[g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots]]
=(1)ji(r,𝐥,𝐩)𝕁i,ν(𝐣)pr+1=0mcν,𝐥𝐩upr+1(1)upr(lr)up1(l1)[gpr+1(0),[gpr(0),,[gp1(0),gi(0)]]]\displaystyle=(-1)^{j_{i}}\sum_{(r,\mathbf{l},\mathbf{p})\in\mathbb{J}_{i,\nu}^{(\mathbf{j})}}\sum_{p_{r+1}=0}^{m}c_{\nu,\mathbf{l}}^{\mathbf{p}}u_{p_{r+1}}^{(1)}u_{p_{r}}^{(l_{r})}\cdots u_{p_{1}}^{(l_{1})}[g_{p_{r+1}}^{(0)},[g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots]]
+(1)ji(r,𝐥,𝐩)𝕁i,ν(𝐣)pr+1=1mlr+1=0jpr+1cν,𝐥𝐩upr+1(lr+1+1)upr+1(lr+1)(upr(lr)up1(l1))[gpr(0),,[gp1(0),gi(0)]]\displaystyle\qquad+(-1)^{j_{i}}\sum_{(r,\mathbf{l},\mathbf{p})\in\mathbb{J}_{i,\nu}^{(\mathbf{j})}}\sum_{p_{r+1}=1}^{m}\sum_{l_{r+1}=0}^{j_{p_{r+1}}}c_{\nu,\mathbf{l}}^{\mathbf{p}}u_{p_{r+1}}^{(l_{r+1}+1)}\frac{\partial}{\partial u_{p_{r+1}}^{(l_{r+1})}}\left(u_{p_{r}}^{(l_{r})}\cdots u_{p_{1}}^{(l_{1})}\right)[g_{p_{r}}^{(0)},\ldots,[g_{p_{1}}^{(0)},g_{i}^{(0)}]\ldots]
(r,λ,μ)𝕂i,ν+1(𝐣)cν+1,λμuμr+1(λr+1)uμ1(λ1)[gμr+1(0),,[gμ1(0),gi(0)]].\displaystyle\triangleq\sum_{(r,\mathbf{\lambda},\mathbf{\mu})\in\mathbb{K}_{i,\nu+1}^{(\mathbf{j)}}}c_{\nu+1,\mathbf{\lambda}}^{\mathbf{\mu}}u_{\mu_{r+1}}^{(\lambda_{r+1})}\cdots u_{\mu_{1}}^{(\lambda_{1})}[g_{\mu_{r+1}}^{(0)},\ldots,[g_{\mu_{1}}^{(0)},g_{i}^{(0)}]\ldots].

Applying the Leibnitz rule to the penultimate line, using the fact that 𝐥=α=1rlα=ν\mid\mathbf{l}\mid=\sum_{\alpha=1}^{r}l_{\alpha}=\nu by the induction assumption, we get 1+𝐥=α=1r+1λα=ν+11+\mid\mathbf{l}\mid=\sum_{\alpha=1}^{r+1}\lambda_{\alpha}=\nu+1, with λα1,α=1,,r+1\lambda_{\alpha}\geq 1,\;\alpha=1,\ldots,r+1, and 0μαm0\leq\mu_{\alpha}\leq m, which proves that the latter formula holds for all (r,λ,μ)𝕂i,ν+1(𝐣)𝕁i,ν+1(𝐣)(r,\mathbf{\lambda},\mathbf{\mu})\in\mathbb{K}_{i,\nu+1}^{(\mathbf{j})}\subset\mathbb{J}_{i,\nu+1}^{(\mathbf{j})}. We leave as an exercise to the reader the proof of the converse inclusion 𝕁i,ν+1(𝐣)𝕂i,ν+1(𝐣)\mathbb{J}_{i,\nu+1}^{(\mathbf{j})}\subset\mathbb{K}_{i,\nu+1}^{(\mathbf{j})}. Therefore (A.2) is valid for all ν\nu. The proof of (A.3) follows exactly the same lines. The lemma is proven. ∎