984
Views
7
CrossRef citations to date
0
Altmetric
Research Article

A structurally flat triangular form based on the extended chained form

, &
Pages 1144-1163 | Received 20 Jul 2020, Accepted 19 Oct 2020, Published online: 11 Nov 2020

ABSTRACT

In this paper, we present a structurally flat triangular form which is based on the extended chained form. We provide a complete geometric characterisation of the proposed triangular form in terms of necessary and sufficient conditions for an affine input system with two inputs to be static feedback equivalent to this triangular form. This yields a sufficient condition for an affine input system to be flat.

1. Introduction

The concept of flatness was introduced in control theory by Fliess, Lévine, Martin and Rouchon, see, e.g. Fliess et al. (Citation1992Citation1995) and has attracted a lot of interest in the control systems theory community. The flatness property allows an elegant systematic solution of feed-forward and feedback problems, see, e.g. Fliess et al. (Citation1995). Roughly speaking, a nonlinear control system (1) x˙=f(x,u)(1) with dim(x)=n states and dim(u)=m inputs is flat, if there exist m differentially independent functions yj=φj(x,u,u1,,uq), uk denoting the kth time derivative of u, such that x and u can locally be parameterised by y and its time derivatives. Up to now, there do not exist verifiable necessary and sufficient conditions for testing a system of the form (Equation1) for flatness, only for certain subclasses of systems, the flatness problem has been solved. Recent research in the field of flatness can be found in e.g. Schöberl et al. (Citation2010), Schlacher and Schöberl (Citation2013), Li et al. (Citation2013), Schöberl and Schlacher (Citation2014), Kolar et al. (Citation2015), Nicolau and Respondek (Citation2017).

Structurally flat triangular forms are of special interest in the problem of deriving flat outputs for nonlinear control systems. In Bououden et al. (Citation2011), a structurally flat triangular form for a class of 0-flat systems is proposed and geometric necessary and sufficient conditions for the existence of a transformation of a nonlinear control system into this triangular form are provided. A structurally flat implicit triangular form for 1-flat systems, together with a constructive scheme for transforming a system into the proposed triangular form, can be found in Schöberl and Schlacher (Citation2014). A complete solution of the flatness problem of systems that become static feedback linearisable after a one-fold prolongation of a suitably chosen control is presented in Nicolau and Respondek (Citation2017). Normal forms for this class of systems can be found in Nicolau and Respondek (Citation2019). Another class of systems for which the flatness problem has been solved are two-input driftless systems, see Martin and Rouchon (Citation1994). Flat two-input driftless systems are static feedback equivalent to a structurally flat triangular form, referred to as chained form. In Li et al. (Citation2013), an extension of the chained form for systems with drift, the so-called extended chained form, is considered. Geometric necessary and sufficient conditions for a two-input affine input system (AI-system) to be static feedback equivalent to this extended chained form can be found in Silveira et al. (Citation2015). Conditions for the case with m2 inputs are provided in Li et al. (Citation2016), Nicolau et al. (Citation2014) and Nicolau (Citation2014).

In Gstöttner et al. (Citation2020), a triangular form which generalises the extended chained form is considered and necessary and sufficient conditions for a two-input AI-system to be static feedback equivalent to this triangular form are provided. The proposed triangular form consists of three subsystems. It generalises the extended chained form by augmenting it with two subsystems in Brunovsky normal form. To be precise, two equally lengthened integrator chains are attached to the inputs of a subsystem in extended chained form and furthermore, the top variables (flat outputs) of this subsystem in extended chained form act as inputs for two arbitrary lengthened integrator chains. The structurally flat triangular form obtained this way contains the (extended) chained form as a special case. In this contribution, we further develop the ideas presented in Gstöttner et al. (Citation2020). We again augment the extended chained form with integrator chains, but here, the integrator chains attached to the inputs of the subsystem in extended chained form differ in length by one integrator. As a consequence, the (extended) chained form is not contained as a special case. The top variables of the subsystem in extended chained form again act as inputs for two arbitrary lengthened integrator chains. It turns out that a broad variety of practical and academic examples is static feedback equivalent to this triangular form. Among others, e.g. the planar VTOL aircraft, also considered in e.g. Fliess et al. (Citation1999) and Schöberl et al. (Citation2010) and the model of a gantry crane, considered in e.g. Fliess et al. (Citation1995). These systems cannot be handled with the triangular form presented in Gstöttner et al. (Citation2020). We again provide necessary and sufficient conditions for an AI-system to be static feedback equivalent to this triangular form. This again provides a sufficient condition for an AI-system to be flat. In contrast to Gstöttner et al. (Citation2020), where proofs are only sketched, detailed proofs are provided in this contribution.

2. Notation

Let X be an n-dimensional smooth manifold, equipped with local coordinates xi, i=1,,n. Its tangent bundle and cotangent bundle are denoted by (T(X),τX,X) and (T(X),τX,X). For these bundles, we have the induced local coordinates (xi,x˙i) and (xi,x˙i) with respect to the bases {xi} and {dxi}, respectively. Throughout, the Einstein summation convention is used. The exterior derivative of a p-form ω is denoted by dω. By Lvkφ, we denote the k-fold Lie derivative of a function φ along a vector field v. Let v and w be two vector fields. Their Lie bracket is denoted by [v,w], for the repeated application of the Lie bracket, we use the common notation advkw=[v,advk1w], k1 and adv0w=w. Let furthermore D1 and D2 be two distributions. By [v,D1], we denote the distribution spanned by the Lie bracket of v with all basis vector fields of D1, and by [D1,D2] the distribution spanned by the Lie brackets of all possible pairs of basis vector fields of D1 and D2. The ith derived flag of a distribution D is denoted by D(i) and defined by D(0)=D and D(i+1)=D(i)+[D(i),D(i)] for i0. The ith Lie flag of a distribution D is denoted by D(i) and defined by D(0)=D and D(i+1)=D(i)+[D,D(i)] for i0. The involutive closure of D is denoted by D¯, it is the smallest involutive distribution which contains D. It can be determined via the derived flag. We denote the Cauchy characteristic distribution of D by C(D). It is spanned by all vector fields c which belong to D and satisfy [c,D]D. Cauchy characteristic distributions are always involutive. They allow us to find a basis for a distribution which is independent of certain coordinates. Since C(D) is involutive, it can be straightened out such that C(D)=span{x1,,xnc}, with nc=dim(C(D)). From [C(D),D]D, it follows that in these coordinates, a basis for D which does not depend on the coordinates (x1,,xnc) can be constructed. Consider an AI-system with m-inputs (2) x˙=a(x)+bj(x)uj,j=1,,m.(2) Geometrically, such a system is represented by the drift vector field a=ai(x)xi and the input vector fields bj=bji(x)xi, j=1,,m, i=1,,n, on the state manifold X. Throughout, we assume that all vector fields and functions we deal with are smooth. We call two AI-systems static feedback equivalent, if they are locally equivalent via a diffeomorphism x~=Φ(x) on the state space and an invertible feedback transformation u~j=gj(x)+mkj(x)uk. The equivalent system reads x~˙=a~(x~)+b~j(x~)u~j,j=1,,m,where a~i(x~)=(xlΦi(x)(al(x)bkl(x)mˆjk(x)gj(x)))Φˆ(x~) and b~ji(x~)=(xlΦi(x)bkl(x)mˆjk(x))Φˆ(x~), with the inverse x=Φˆ(x~) of the state transformation x~=Φ(x) and the inverse matrix [mˆjk] of the matrix [mkj] from the input transformation.

3. Known results

In this section, we summarise some known results from the literature which are of particular importance for characterising our triangular form. Throughout, we assume all distributions to have locally constant dimension, we consider generic points only. In particular, we call a system static feedback equivalent to a certain normal form, even though the transformation into this form may exhibit singularities. Consider again an m-input AI-system (Equation2). Such a system is called static feedback linearisable, if it is static feedback equivalent to a linear controllable system, in particular to the Brunovsky normal form. The static feedback linearisation problem has been solved in Jakubczyk and Respondek (Citation1980) and Hunt and Su (Citation1981). The geometric necessary and sufficient conditions read as follows. For (Equation2), we define the distributions Di+1=Di+[a,Di], i1, where D1=span{b1,,bm}.

Theorem 3.1

The m-input AI-system (Equation2) is static feedback linearisable if and only if all the distributions Di, i1 are involutive and Dn1=T(X).

In Martin and Rouchon (Citation1994), it is shown that a two-input driftless system of the form (3) x˙=b1(x)u1+b2(x)u2(3) is flat, if and only if it is static feedback equivalent to the structurally flat triangular form (4) x˙1=u2,x˙2=x3u2,x˙n1=xnu2,x˙n=u1,(4) referred to as chained form. The input vector fields of a system in chained form read (5) b1=xn,b2=x1+x3x2++xnxn1.(5) The geometric necessary and sufficient conditions for a driftless system (Equation3) to be static feedback equivalent to the chained form (Equation4) are summarised in the following theorem.

Theorem 3.2

The driftless system (Equation3) is static feedback equivalent to the chained form (Equation4) if and only if D=span{b1,b2} satisfies dim(D(i))=2+i, i=0,,n2.

In Murray (Citation1994), it is shown that locally around a point of the state space at which the additional regularity condition dim(D(i))=2+i, i=0,,n2 on the Lie flag of D holds, the transformation into chained form does not exhibit singularities. A system in chained form is flat with the pair of top variables (x1,x2) forming a possible flat output. For a comprehensive analysis of the flatness of systems static feedback equivalent to the chained form, a characterisation of all their x-flat outputs and their singularities, we refer to Li and Respondek (Citation2012). The structurally flat triangular form (6) x˙1=u2x˙2=x3u2+a2(x1,x2,x3)x˙3=x4u2+a3(x1,x2,x3,x4)x˙n1=xnu2+an1(x1,,xn)x˙n=u1,(6) referred to as extended chained form, was first considered in Li et al. (Citation2013). In Nicolau (Citation2014) and Silveira et al. (Citation2015), geometric necessary and sufficient conditions for an AI-system with two inputs to be static feedback equivalent to (Equation6) are provided. Those are summarised in the following theorem.

Theorem 3.3

An AI-system (Equation2) with two inputs (m=2) is static feedback equivalent to the extended chained form (Equation6) if and only if

  1. D=span{b1,b2} satisfies dim(D(i))=2+i, i=0,,n2.

  2. The drift of the system meets the compatibility condition (7) [a,C(D(i))]D(i),i=1,,n3.(7)

These conditions can be interpreted as follows. The first condition assures that the driftless system obtained by setting a(x)=0 is static feedback equivalent to the chained form (Equation4). Indeed, this condition coincides with that of Theorem 3.2. The same regularity condition dim(D(i))=2+i, i=0,,n2 on the Lie flag of D applies for excluding singularities. The second condition assures that the drift is compatible with the chained form, i.e. that in coordinates in which the input vector fields are in chained form, the drift takes the desired triangular structure. Thus, a flat output of the driftless system obtained by setting a(x)=0 is also a flat output of (Equation6). For a comprehensive analysis of the flatness of systems, static feedback equivalent to the extended chained form, a characterisation of their flat outputs and their singularities, we refer to Li et al. (Citation2016) or Nicolau (Citation2014).

Remark 3.1

Note that a system in chained form (Equation4) becomes static feedback linearisable by (n2)-fold prolonging the input u2. The same holds for a system in extended chained form (Equation6), see also Nicolau (Citation2014).

4. A structurally flat triangular form based on the extended chained form

In the following, we consider the structurally flat triangular form (8) x˙1=f1(x1,x21,x22)x˙2=f2(x1,x2,x3,11,x3,21)x˙3=f3(x3,u1,u2),(8) which consists of three subsystems. Throughout, the first subscript of the variables, which is either 1, 2 or 3, always denotes to which subsystem the variables belong. The x1-subsystem is in Brunovsky normal form, it consists of two integrator chains. Its states are denoted by x1,ji, where j = 1, 2 refers to the integrator chain and 1in1,j refers to the position within the jth integrator chain (9) x˙1,11=x1,12x˙1,21=x1,22x˙1,12=x1,13x˙1,22=x1,23x˙1,1n1,1=x21x˙1,2n1,2=x22.(9) The x2-subsystem is essentially in extended chained form (10) x˙21=x3,21x˙22=x23x3,21+a22(x1,x21,x22,x23)x˙2n21=x2n2x3,21+a2n21(x1,x2)x˙2n2=x3,11+g(x1,x2)x3,21.(10) The x3-subsystem is again in Brunovsky normal form, it consists of two integrator chains which differ in length by one integrator. Its states are denoted by x3,ji, where j = 1, 2 refers to the integrator chain and the superscript i again refers to the position within the corresponding integrator chain (11) x˙3,11=x3,12x˙3,21=x3,22x˙3,12=x3,13x˙3,22=x3,23x˙3,1n31=x3,1n3x˙3,2n31=u2x˙3,1n3=u1.(11) The triangular form (Equation8) consists of three subsystems. The x1-subsystem is in Brunovsky normal form, it consists of two integrator chains of arbitrary lengths n1,10 and n1,20. In total, it consists of n1=n1,1+n1,20 states. The x2-subsystem is essentially in extended chained form (we assume n23) and the top variables x21 and x22 of this subsystem act as inputs for the x1-subsystem.

Remark 4.1

The x2-subsystem differs from the extended chained form in two minor ways. First, the functions a2i, i=2,,n21, which represent the drift of the x2-subsystem may also depend on the stats x1. Second, in the last equation of the x2-subsystem, besides x3,11, there may also occur the term g(x1,x2)x3,21. The function g(x1,x2) can be zero, however, due to the differing lengths of the integrator chains in the x3-subsystem, the normalisation x¯3,11=x3,11+g(x1,x2)x3,21 to eliminate the term g(x1,x2)x3,21 from the x2-subsystem is not possible. Nevertheless, the x2-subsystem has analogous structural properties as a system in extended chained form.

The x3-subsystem is again in Brunovsky normal form, it consists of two integrator chains which differ in length by one integrator. The top variables x3,11 and x3,21 act as inputs for the x2-subsystemFootnote1 (we assume n31, for n3=1, the x3-subsystem only consists of a single integrator, namely x˙3,11=u1, and in the x2-subsystem, x3,21 is replaced by u2). In conclusion, the x3-subsystem and the x2-subsystem form an endogenous dynamic feedback for the x1-subsystem. The x3-subsystem in turn is an endogenous dynamic feedback for the x2-subsystem. The total number of states of (Equation8) is given by n=n1+n2+2n31.

Remark 4.2

In conclusion, the restriction on the dimensions of the subsystems in (Equation8) are n23 and n31. However, it turns out that a system of the form (Equation8) with n2=3 and n1=0 meets the conditions of Theorem 3.1 and thus, it is static feedback linearisable. Therefore, n2=3 only makes sense if n11, which we always assume if n2=3.

Remark 4.3

A system of the form (Equation8) becomes static feedback linearisable after an (n21)-fold prolongation of u2 (one prolongation accounts for the differing lengths of the integrator chains in the x3-subsystem, the remaining (n22) prolongations correspond to those in Remark 3.1). In particular, a system of the form (Equation8) with n2=3 becomes static feedback linearisable after a two-fold prolongation of u2. A geometric characterisation of systems that become static feedback linearisable after a two-fold prolongation of a suitably chosen control can be found in Nicolau and Respondek (Citation2016a). However, due to Assumption 2 in Nicolau and Respondek (Citation2016a), the geometric necessary and sufficient conditions for linearisability via a two-fold prolongation provided therein do not apply to a system of the form (Equation8) if n1,11 and n1,21. The special case n2=3 and n1,12 or n1,22 is indeed fully covered by Nicolau and Respondek (Citation2016a). Our geometric characterisation of (Equation8) provided in the following section is not subject to restrictions on n1, n2 and n3.

As a motivating example, consider the planar VTOL aircraft, also treated e.g. in Fliess et al. (Citation1999), Schöberl et al. (Citation2010) or Schöberl and Schlacher (Citation2011), and given by (12) x˙=vxz˙=vzθ˙=ωv˙x=ϵcos(θ)u2sin(θ)u1v˙z=cos(θ)u1+ϵsin(θ)u21ω˙=u2.(12) This system is not static feedback linearisable, but it is known to be flat. It is not static feedback equivalent to the triangular form proposed in Gstöttner et al. (Citation2020), but it is static feedback equivalent to the triangular form (Equation8). In Section 5, we will systematically derive a state and input transformation, which brings (Equation12) into the form (13) x˙1,11=x21x˙1,21=x22x˙21=u~2x˙22=x23u~2+x23x˙23=x3,11x˙3,11=u~1,(13) which is of the form (Equation8) with n1,1=n1,2=1, n2=3 and n3=1.

Remark 4.4

In Nicolau and Respondek (Citation2020), normal forms for systems that become static feedback linearisable after a two-fold prolongation of a suitably chosen control are presented. Therein, based on a given suitable flat output a representation of the VTOL analogous to (Equation13) is derived. Note however that the geometric necessary and sufficient conditions for linearisability via a twofold prolongation provided in Nicolau and Respondek (Citation2016a) do not apply to the VTOL.

The triangular form (Equation8) is similar to the triangular form presented in Gstöttner et al. (Citation2020). The difference between the triangular form considered here and the triangular form in Gstöttner et al. (Citation2020) is that in (Equation8) the integrator chains in the x3-subsystem (Equation11) differ in length by one integrator, whereas in the triangular form in Gstöttner et al. (Citation2020) those have the same length, i.e. for the triangular form in Gstöttner et al. (Citation2020), we would have an x3-subsystem of the form (14) x˙3,11=x3,12x˙3,21=x3,22x˙3,12=x3,13x˙3,22=x3,23x˙3,1n3=u1x˙3,2n3=u2,(14) instead of the form (Equation11) (and g = 0 in the x2-subsystem (Equation10)).

4.1. Characterisation of the triangular form

In this section, we provide necessary and sufficient conditions for a two-input AI-system to be static feedback equivalent to the triangular form (Equation8) and thus provide a sufficient condition for such a system to be flat. Consider a two-input AI-system (15) x˙=a(x)+b1(x)u1+b2(x)u2.(15) We define the distributions Di, i=1,,n3+1 where D1=span{b1,b2} and Di+1=Di+[a,Di], with the smallest integer n3 such that Dn3+1 is not involutive. (If the considered AI-system is indeed static feedback equivalent to (Equation8), this integer n3 is the length of the first integrator chain in the x3-subsystem.) We again assume all distributions to have locally constant dimension and we omit discussing singularities coming along with flat outputs of (Equation8) or singularities in the problem of transforming a given system into the form (Equation8). We consider generic points only, regularity conditions are omitted.

Theorem 4.1

The AI-system (Equation15) is static feedback equivalent to the triangular form (Equation8) if and only if dim(Di)=2i, i=1,,n3+1, C(Dn3+1)Dn3 and there exists a vector field bp=α1b1+α2b2 such that with the distributions Δ0=Dn31+span{adan31bp} and Δ1=Dn3+span{adan3bp}, the following conditions are satisfied:

  1. C(Δ1)=Δ0.

  2. The derived flags of the non involutive distribution Δ1 satisfy dim(Δ1(i))=dim(Δ1)+i,i=1,,n22,with the smallest integer n2 such that Δ1(n22)=Δ¯1.

  3. The drift satisfies the compatibility conditionsFootnote2 (16) [a,C(Δ1(i))]Δ1(i),i=1,,n23,(16) (17) dim(Δ¯1+[a,Δ1(n23)])=dim(Δ¯1)+1.(17)

  4. The distributions Gi+1, i0 are involutive, where Gi+1=Gi+[a,Gi] and G0=Δ¯1.

  5. Gs=T(X) holds for some integer s.

All these conditions are easily verifiable and require differentiation and algebraic operations only. Also the construction of a vector field bp, which is needed for verifying the conditions, requires differentiation and algebraic operations only. The construction is discussed in the next section, where we will see that there exist at most two candidates for a vector field bp and we are able to compute them without solving PDEs. Let us outline the meaning of the individual conditions of the theorem and of the vector field bp. Consider a system of the form (Equation8). The main idea of Theorem 4.1 is to characterise the three subsystems of (Equation8) on their own and have separate conditions which take into account their coupling. Since the individual subsystems are either in Brunovsky normal form or essentially in extended chained form, they are in fact characterised by Theorems 3.1 and 3.3 in Section 3. The x3-subsystem (Equation11) consists of two integrator chains with the lengths n3 and n31. The vector field bp corresponds to the input vector field of the longer integrator chain, i.e. in (Equation11) this would be bp=x3,1n3. The involutive distributions D1Dn31Δ0 characterise the x3-subsystem, i.e. Δ0=span{x3}. The top variables x3,11 and x3,21 of the x3-subsystem act as inputs for the x2-subsystem. The input vector fields of the x2-subsystem with respect to the variables x3,11 and x3,21 read b2,1=b1c and b2,2=b2c+gb1c with b1c=x2n2 and b2c=x21+x23x22++x2n2x2n21. The vector fields b1c and b2c are structurally of the form (Equation5). Since the shorter integrator chain of the x3-subsystem has only a length of n31, the distribution Dn3 already contains one of these input vector field of the x2-subsystem, namely b2,2. That is why the vector field bp is needed, it allows us to separate the x3-subsystem from the x2-subsystem despite the differing lengths of the integrator chains. Constructing bp roughly speaking means identifying the longer integrator chain of the x3-subsystem. Given bp, we can calculate the distribution Δ0 and thus explicitly identify the states which belong to the x3-subsystem, i.e. Δ0=span{x3}.

The distribution Δ1 is spanned by x3 and both input vector fields b2,1 and b2,2 of the x2-subsystem. Item (a) is crucial for the coupling of the x2-subsystem with the x3-subsystem, it assures that the x2-subsystem indeed allows an AI representation with respect to its inputs x3,11 and x3,21. Item (b) is in fact a condition on the distribution spanned by the input vector fields of the x2-subsystem, the condition in fact matches that of Theorem 3.2 for the normal chained form. Item (b) therefore guarantees that the input vector fields of the x2-subsystem can indeed be transformed into a chained structure.

The triangular dependence of the functions a2i, i=2,,n21, i.e. the drift of the x2-subsystem, on the states x2 is assured by the condition (Equation16), which essentially matches the compatibility condition (Equation7) in Theorem 3.3 for the extended chained form. The drift vector field of a system of the form (Equation8) of course does not only consist of the drift vector field a2=a22x22++a2n21x2n21 of the x2-subsystem, it also contains the input vector fields of the x2-subsystem and has additional components which belong to the x1-subsystem and the x3-subsystem, the drift vector field of (Equation8) actually reads a=f1+x3,21(b2c+gb1c)+x3,11b1c+a2+a3 where f1=x1,12x1,11++x21x1,1n1,1+x1,22x1,21++x22x1,2n1,2 and a3=x3,12x3,11++x3,1n3x3,1n31+x3,22x3,21++x3,2n31x3,2n32. However, as we will see in the proof of Theorem 4.1, the compatibility condition of the extended chained form still analogously applies, i.e. when testing the compatibility of the drift of the x2-subsystem, all the additional components of the drift do not matter.

The involutive closure Δ¯1 of Δ1 allows us to separate the x1-subsystem from the x2-subsystem and the x3-subsystem, i.e. Δ¯1=span{x3,x2}. Condition (Equation17) is crucial for the coupling of the x1-subsystem with the x2-subsystem. The static feedback linearisable x1-subsystem is characterised by the involutive distributions of item (d) and by item (e).

As already mentioned, the difference between the triangular form considered here and the triangular form in Gstöttner et al. (Citation2020) is that for the triangular form in Gstöttner et al. (Citation2020), we would have an x3-subsystem of the form (Equation14) instead of the form (Equation11) (and g = 0 in the x2-subsystem (Equation10)). An AI-system (Equation15) is static feedback equivalent to this simpler triangular form presented in Gstöttner et al. (Citation2020) if and only if the items (a) to (e) of Theorem 4.1 are met with Dn3 and Dn3+1 instead of Δ0 and Δ1. So in this case, because of the equal length of the integrator chains, the x3-subsystem is simply characterised by the involutive distributions D1Dn3 and the distribution Dn3+1 is spanned by x3 and the input vector fields of the x2-subsystem. No extra effort is needed for identifying the states which belong to the x3-subsystem. Despite the similarity of the two triangular forms, the triangular form with differing lengths of the integrator chains in the x3-subsystem, i.e. the triangular form covered in this paper, is applicable to many practical and academic examples which cannot be handled with the triangular form presented in Gstöttner et al. (Citation2020). Among others, e.g. the already mentioned planar VTOL aircraft and the model of a gantry crane and some academic examples considered in Schöberl (Citation2014), see Section 5. Although the proof of Theorem 4.1 is also somewhat similar to the proof of the main theorem in Gstöttner et al. (Citation2020), the differing lengths of the integrator chains in the x3-subsystem greatly increase the complexity of the proof. In contrast to Gstöttner et al. (Citation2020), where proofs are only sketched, detailed proofs are provided in this paper.

4.1.1. Determining a vector field bp

According to Theorem 4.1, the test for static feedback equivalence of an AI-system (Equation15) to the triangular form (Equation8) involves finding a certain linear combination bp=α1b1+α2b2 of its input vector fields b1 and b2. It should be noted that in Theorem 4.1, only the direction of bp matters, i.e. if the conditions of Theorem 4.1 are met with bp, then they are also met with b~p=λbp with an arbitrary non-zero function λ of the state of the system. A detailed proof of this property is provided in Appendix A.1. As already mentioned, provided that the system under consideration is indeed static feedback equivalent to the triangular form (Equation8), such a vector field bp corresponds to the input vector field of the longer integrator chain in the corresponding x3-subsystem. In the following, we explain the construction of such a vector field bp. We will start by deriving a necessary condition, which every such vector field bp has to meet. This will allow us to determine candidates for the vector field bp. We will then show that this necessary condition yields only at most two non-collinear candidates for bp. Thus, if the system is indeed static feedback equivalent to (Equation8), at least with one of those candidates, the conditions of Theorem 4.1 must indeed be met. Let us introduce the abbreviations v1=adan31b1, v2=adan31b2 and vp=α1v1+α2v2. With those, because of adan31bp=α1adan31b1+α2adan31b2 mod Dn31 and analogously adan3bp=α1adan3b1+α2adan3b2 mod Dn3, we have adan31bp=vp mod Dn31 and adan3bp=[a,vp] mod Dn3. Thus, for the distributions Δ0 and Δ1 of Theorem 4.1, we have Δ0=Dn31+span{vp} and Δ1=Dn3+span{[a,vp]}. Item (a) of Theorem 4.1 requires that C(Δ1)=Δ0. Since vpΔ0, this implies that [vp,Δ1]Δ1 must hold and in particular [vp,[a,vp]]Δ1 must hold. A necessary condition on vp=α1v1+α2v2 (i.e. a necessary condition on the coefficients α1 and α2 in this linear combination) is thus (18) [vp,[a,vp]]=(α1)2[v1,[a,v1]]+α1Lv1α1[a,v1]+α1α2[v1,[a,v2]]+α1Lv1α2[a,v2]+α2α1[v2,[a,v1]]+α2Lv2α1[a,v1]+(α2)2[v2,[a,v2]]+α2Lv2α2[a,v2] mod Dn3!Δ1,(18) which is a system of non-linear PDEs in α1 and α2. From solutions, we obtain candidates bp=α1b1+α2b2 (or vp=α1v1+α2v2). Since by construction we have Δ1Dn3+1, a solution of (Equation18) also meets [vp,[a,vp]]Dn3+1. Thus, we have the weaker necessary condition [vp,[a,vp]]!Dn3+1, which because of [a,v1],[a,v2]Dn3+1, is purely algebraic and readsFootnote3 (19) (α1)2[v1,[a,v1]]+2α1α2[v1,[a,v2]]+(α2)2[v2,[a,v2]]!Dn3+1.(19) Although (Equation19) is again only a necessary condition which the coefficients α1 and α2 must fulfill, this necessary condition yields at most two non-collinear candidates for the vector field bp=α1b1+α2b2. This is shown in detail in Appendix A.3. Thus, if the system under consideration is indeed static feedback equivalent to (Equation8), at least with one of those two non-collinear candidates for bp, the conditions of Theorem 4.1 must indeed be met. The conditions of Theorem 4.1 may even be met with both candidates. In this case, there exist two different triangular representation (which in turn lead to different flat outputs), since the distributions involved in the conditions of Theorem 4.1 depend on bp.

Remark 4.5

There exists a simpler method for determining a vector field bp in certain cases. If for an AI-system (Equation15adan3+1b1H or adan3+1b2H, with the distribution H=Dn3+1+[Dn3,Dn3+1] holds, the direction of bp=α1b1+α2b2 is uniquely determined by the condition adan3+1bpH. Since adan3+1bp=α1adan3+1b1+α2adan3+1b2 mod Dn3+1 and Dn3+1H, the condition adan3+1bpH yields a system of linear equations for determining α1 and α2. If adan3+1b1, adan3+1b2H, this method for determining bp is of course not applicable, since adan3+1bpH would be met for any linear combination bp=α1b1+α2b2 of the input vector fields of the system. A proof of this property can be found in Appendix A.2.

4.2. Determining flat outputs

For determining flat outputs of a system which is static feedback equivalent to (Equation8), there is no need to actually transform the system into the form (Equation8). A system which is static feedback equivalent to (Equation8) meets the conditions of Theorem 4.1. Flat outputs can be derived directly from the distributions Δ1 and Gi, which are involved in Theorem 4.1, items (b) and (d). All what follows in this section is actually contained in the sufficiency part of the proof of Theorem 4.1. Here, we only summarise the computation of compatible flat outputs, for details, we refer to the proof of Theorem 4.1. Consider a system of the form (Equation8). Depending on the length of the integrator chains in the x1-subsystem (Equation9), flat outputs are determined differently. In particular, we have to distinguish between the cases that

  1. both integrator chains have at least length one, i.e. n1,1,n1,21,

  2. one of the integrator chains has length zero, and the other one at least a length of one,

  3. both integrator chains have length zero, i.e. n1,1=n1,2=0, an x1-subsystem does not exist at all.

Given a system which meets the conditions of Theorem 4.1, we can easily test which case applies. If we have dim(G1)=dim(Δ¯1)+2, in the corresponding triangular form (Equation8), both chains have at least length one. If dim(G1)=dim(Δ¯1)+1, one chain has length zero, if Δ¯1=T(X), both have length zero. In the following, we discuss these three cases in more detail.

Case 1: If n1,1,n1,21, i.e. both integrator chains of the x1-subsystem (Equation9) have at least length one, flat outputs are all pairs of functions (φ1,φ2), which form a linearising output of the x1-subsystem. The x1-subsystem is characterised by the distributions Gi of Theorem 4.1 item (d). So in this case, flat outputs are determined from the sequence of involutive distributions Gi of Theorem 4.1 item (d), in the same way as linearising outputs are determined from the sequence of involutive distributions involved in the test for static feedback linearisability (see, e.g. Jakubczyk & Respondek, Citation1980; Nijmeijer & van der Schaft, Citation1990).

Case 2: Here, the x1-subsystem determines one component φ1 of a flat output. This function is obtained by integrating Gs1, i.e. by finding a function φ1 such that span{dφ1}=Gs1 with Gs1 of Theorem 4.1, item (d) and the smallest integer s such that item (e) holds. A possible second component φ2 is obtained by integrating the integrable codistribution L=(Δ1(n23))+span{dLasφ1}. A function φ2 whose differential dφ2 together with {dφ1,dLaφ1,,dLasφ1} spans the codistribution L, is a possible second component.

Case 3: If both chains have length zero (the x1-subsystem does not exist at all), the problem of finding flat outputs, is in fact the same as finding flat outputs of a system that is static feedback equivalent to the chained form. This problem is addressed in Li and Respondek (Citation2012). In this case, flat outputs are all pairs of functions (φ1,φ2), which meet L=(span{dφ1,dφ2})Δ1(n23), with Δ1(n23) of Theorem 4.1 item (b). In Li and Respondek (Citation2012), Theorem 2.10, a method for constructing such a distribution L is provided. The distribution L is not unique, one has to choose one function φ1 whose differential dφ10 annihilates C(Δ1(n23)). Once such a function has been chosen, the distribution L can be calculated, and in turn a possible second function φ2, which together with φ1 forms a possible flat output, can be calculated. Equivalent to the method for determining L provided in Li and Respondek (Citation2012), once a function φ1 whose differential annihilates C(Δ1(n23)) has been chosen, the annihilator of L can also be calculated via L=(Δ1(n23))+span{dφ1}.

Note that Cases 2 and 3 are in fact similar. The function Lasφ1 in Case 2 corresponds to φ1 from Case 3. In Case 3, we have to choose the function φ1. Once this function has been chosen, the distribution L is uniquely determined by this function. In Case 2, the distribution L is uniquely determined by the function Lasφ1, which is imposed by the x1-subsystem.

4.3. Proof of Theorem 4.1

To keep the proof reasonably compact, parts of it are condensed into propositions and small facts, which are proven in Appendix A.1. The following two lemmas are of particular importance for the sufficiency part of the proof. Proofs of these lemmas are also provided in the appendix.

Lemma 4.2

Let D be a distribution. Every characteristic vector field of D, i.e. every vector field cC(D) is also characteristic for its derived flag D(1), i.e. C(D)C(D(1)).

An immediate consequence of Lemma 4.2 is that the Cauchy characteristic distributions C(D(i)), i0 form the sequence of nested involutive distributions C(D)C(D(1))C(D(2))

Lemma 4.3

If a d-dimensional distribution D satisfies dim(C(D))=d2 and dim(D(i))=d+i, i=1,,l with l such that D(l)=D¯, then the Cauchy characteristics C(D(i)) satisfy dim(C(D(i)))=d2+i and C(D(i))D(i1), i=1,,l1.

Lemma 4.3 is based on a similar one in Cartan (Citation1914), see also Martin and Rouchon (Citation1994, Lemma 2).

Necessity. We have to show that a system of the form (Equation8) meets the conditions of Theorem 4.1. We assume n32, nothing substantially changes if n3=1. Recall that the drift vector field of a system of the form (Equation8) reads a=f1+x3,21(b2c+gb1c)+x3,11b1c+a2+a3 where f1=x1,12x1,11++x21x1,1n1,1+x1,22x1,21++x22x1,2n1,2, b1c=x2n2, b2c=x21+x23x22++x2n2x2n21, a2=a22x22++a2n21x2n21 and a3=x3,12x3,11++x3,1n3x3,1n31+x3,22x3,21++x3,2n31x3,2n32. The input vector fields of (Equation8) are given by b1=x3,1n3 and b2=x3,2n31. The distributions defined right before Theorem 4.1 are thus given by D1=span{x3,1n3,x3,2n31}D2=span{x3,1n3,x3,2n31,x3,1n31,x3,2n32}Dn31=span{x3,1n3,x3,2n31,,x3,12,x3,21}Dn3=span{x3,1n3,x3,2n31,,x3,12,x3,21,x3,11x3,b2c+gb1c}Dn3+1=span{x3,b1c,b2c,[a,b2c+gb1c]}.The distributions D1,,Dn3 are involutive and they meet dim(Di)=2i.

Fact 4.1

The distribution Dn3+1 is not involutive, it meets dim(Dn3+1)=2n3+2 and C(Dn3+1)Dn3.

The input vector field belonging to the longer integrator chain in the x3-subsystem is b1=x3,1n3. With bp=b1, we obtain adan31bp=(1)n31x3,11 and adan3bp=(1)n3b1c, and thus Δ0=Dn31+span{adan31bp}=span{x3} and Δ1=Dn3+span{adan3bp}=span{x3,b1c,b2c}. With these two distributions, the items (a) to (e) of Theorem 4.1 are met. Item (a) is met since the vector fields b1c and b2c are independent of the variables x3. We have Δ1(i)=span{x3,x21+x23x22++x2n2ix2n2i1,x2n2,,x2n2i},i=0,,n22,where the (n22)th derived flag of Δ1 is actually the involutive closure of Δ1, i.e. Δ1(n22)=Δ¯1=span{x3,x2}, the last noninvolutive distribution of this sequence is Δ1(n23)=span{x3,x21+x23x22,x2n2,,x23}. Their Cauchy characteristic distributions are given by C(Δ1(i))=span{x3,x2n2,,x2n2i+1},i=1,,n23.The distributions Δ1(i) meet the condition dim(Δ1(i))=dim(Δ1)+i, i=1,,n22 of item (b). To show that the condition (Equation16) of item (c), i.e. [a,C(Δ1(i))]Δ1(i), i=1,,n23 is met, note that we haveFootnote4 [x2n2i+1,a]=(x3,21+x2n2i+1a2n2i)x2n2i mod C(Δ1(i)),i=1,,n23,i.e. [x2n2i+1,a]Δ1(i), i=1,,n23. Condition (Equation17) of item (c) is met since Δ¯1+[a,Δ1(n23)]=Δ¯1+span{[a,x2n2],,[a,x23]Δ¯1,[a,x21+x23x22]Δ¯1},i.e. [a,Δ1(n23)] yields only one new direction with respect to Δ¯1. For the distributions Gi of item (d), we obtain G0=Δ¯1=span{x3,x2}Gi=span{x3,x2,x1,1n1,1,,x1,1n1,1i+1,x1,2n1,2,,x1,2n1,2i+1},i1,where any x1,jk, j = 1, 2 with k0 has to be omitted. These distributions are obviously involutive. Thus, item (d) is met. Item (e) is met since we have Gs=T(X) for s=max{n1,1,n1,2}.

Sufficiency. We have to show that an AI-system which meets the conditions of Theorem 4.1 can be transformed into the triangular form (Equation8). Item (a) implies that Δ0 is involutive. Because of Lemma 4.2, the Cauchy characteristics of the derived flags of Δ1 form the sequence of nested involutive distributions C(Δ1)C(Δ1(1))C(Δ1(n23)),where C(Δ1)=Δ0. We thus have the sequence of nested involutive distributions (20) D1Dn31Δ0C(Δ1(1))C(Δ1(n23))Δ¯1G1Gs=T(X).(20) The transformation of (Equation15) into the form (Equation8) is done in the following six steps.

Step 1: Straighten out all the distributions (Equation20) simultaneously.

Proposition 4.1

In coordinates in which the distributions (Equation20) are straightened out, the system (Equation15) takes the form (21) x˙1=f1(x1,x21,x22,x23)x˙2=f2(x1,x2,x31,x32,x33)x˙3=f3(x1,x2,x3,u1,u2).(21)

Here, the variables of the x1-subsystem are denoted by x1i, 1in1 and the variables of the x3-subsystem are denoted by x3i, 1i2n31. (Both of these subsystems are not yet decomposed into two separate integrator chains and therefore, a second subscript as in (Equation9) and (Equation11) is not yet needed.) The top three variables x21, x22 and x23 of the x2-subsystem act as inputs for the x1-subsystem and the top three variables x31, x32 and x33 of the x3-subsystem act as inputs for the x2-subsystem. However, we have redundancy among these inputs as the following proposition asserts.

Proposition 4.2

The subsystems in (Equation21) meet the rank conditions rank((x21,x22,x23)f1)2, rank((x31,x32,x33)f2)=2 and rank((x32,x33)f2)=1Footnote5.

Note that by only straightening out the distributions Δ0 and Δ¯1, i.e. applying a change of coordinates such that Δ0=span{x3} and Δ¯1=span{x3,x2}, we already obtain a decomposition of the system (Equation15) into three subsystems. Simultaneously straightening out the remaining distributions of (Equation20) only affects the structure of these three subsystems. In particular, by straightening out D1Dn31 and G1Gs1, the x1-subsystem and the x3-subsystem take a triangular structure, known from the static feedback linearisation problem (see, e.g. Nijmeijer & van der Schaft, Citation1990). The inputs u1 and u2 of course occur affine in f3, since we started with an AI-system and only applied a state transformation, which of course preserves the AI structure.

Step 2: Transform the subsystem x˙1=f1(x1,x21,x22,x23) into Brunovsky normal form, i.e. separate it into two integrator chains, by successively introducing new coordinates from top to bottom. In the prior to last step, we then have (22) x˙1,11=x1,12x˙1,21=x1,22x˙1,12=x1,13x˙1,22=x1,23x˙1,1n1,1=φ1(x¯1,x21,x22,x23)x˙1,2n1,2=φ2(x¯1,x21,x22,x23)x˙2=f¯2(x¯1,x2,x31,x32,x33)x˙3=f¯3(x¯1,x2,x3,u1,u2),(22) where x¯1=(x1,11,,x1,1n1,1,x1,21,,x1,2n1,2) are the newly introduced states in which the x1-subsystem is in Brunovsky normal form. In the following, we have to distinguish between the three possible cases regarding the actual number of integrator chains in the x1-subsystem, which were already mentioned in Section 4.2. The rank of the Jacobian matrix (x21,x22,x23)f1 corresponds to the actual number of non-redundant inputs of the x1-subsystem and thus to the actual number of integrator chains in the x1-subsystem, which can either be 2, 1 or 0.

Case 1: If rank((x21,x22,x23)f1)=2 holds, the functions φj(x¯1,x21,x22,x23), j = 1, 2 in (Equation22) determine the desired top variables for the x2-subsystem. (Because of rank((x21,x22,x23)f1)=2, these functions meet dx¯1dφ1dφ20 and thus, they can indeed serve as states for the x2-subsystem.) These functions have to meet L=(span{dx¯1,dφ1,dφ2})Δ1(n23), the distribution L is of importance in the problem of transforming the x2-subsystem into (extended) chained form.

Proposition 4.3

If rank((x21,x22,x23)f1)=2 holds, the functions φj(x¯1,x21,x22,x23), j = 1, 2 in (Equation22) satisfy L=(span{dx¯1,dφ1,dφ2})Δ1(n23).

Case 2: If rank((x21,x22,x23)f1)=1 holds, the x1-subsystem only consists of one integrator chain (one chain in (Equation22) is missing), so it determines only one function φ1(x¯1,x21,x22,x23) which we want as top variable in the x2-subsystem. In this case, there always exists a second function φ2(x¯1,x21,x22,x23) which together with φ1 fulfils L=(span{dx¯1,dφ1,dφ2})Δ1(n23).

Proposition 4.4

If rank((x21,x22,x23)f1)=1 holds, there always exists a function φ2(x¯1,x21,x22,x23) which together with φ1 fulfils L=(span{dx¯1,dφ1,dφ2})Δ1(n23).

Case 3: Finally, in the case that the x1-subsystem does not exist at all, i.e. Δ¯1=T(X), two functions φj(x21,x22,x23), which fulfill L=(span{dφ1,dφ2})Δ1(n23), have to be found. Such functions always exist.

Proposition 4.5

If Δ¯1=T(X), there always exist two functions φj(x21,x22,x23), which fulfill L=(span{dφ1,dφ2})Δ1(n23).

Step 3: Introduce the functions φj(x¯1,x21,x22,x23), as the top variables of the x2-subsystem, i.e. apply the state transformation x~2j=φj(x¯1,x21,x22,x23), j = 1, 2. (When introducing these function as new states of the x2-subsystem, it may be necessary to also introduce x~23=φ3(x¯1,x21,x22,x23) with a suitably chosen function φ3, since it may happen that dx¯1dφ1dφ2dx23=0.) This completes the transformation of the x1-subsystem into Brunovsky normal form, i.e.  x˙1,11=x1,12x˙1,21=x1,22x˙1,1n1,1=x~21x˙1,2n1,2=x~22x~˙2=f~2(x¯1,x~2,x31,x32,x33)x˙3=f~3(x¯1,x~2,x3,u1,u2),where x~2=(x~21,x~22,x23,,x2n2). Furthermore, this transformation straightens out the distribution L=(span{dx¯1,dφ1,dφ2}) simultaneously with the distributions (Equation20), i.e. L=span{x3,x2n2,,x23}.

The following two steps deal with the transformation of the x2-subsystem into essentially (extended) chained form, see also Remark 4.1.

Step 4: Because of the rank condition rank((x32,x33)f2)=1 of Step 1, we also have rank((x32,x33)f~2)=1. Therefore, without loss of generality, we can assume that f~2 explicitly depends on x32, if not, swap x32 and x33. Let us assume that the first component f~21 of f~2 explicitly depends on x32 (after eventually swapping x32 and x33). This enables us to replace the state x32 of the x3-subsystem by the new stateFootnote6 (23) x3,21=f~21(x¯1,x~2,x31,x32,x33),(23) i.e. with this new state, we replace x32 and leave all the other coordinates unchanged. This transformation normalises the first equation of the x2-subsystem, i.e. x~˙21=x3,21. The following fact guarantees that this transformation is indeed a regular transformation.

Fact 4.2

The function f~21 in (Equation23) indeed explicitly depends on x32 (after eventually swapping x32 and x33).

Proposition 4.6

After applying the transformation (Equation23), the x2-subsystem reads (24) x~˙21=x3,21x~˙22=b22(x¯1,x~21,x~22,x23)x3,21+a22(x¯1,x~21,x~22,x23)x˙23=b23(x¯1,x~21,x~22,x23,x24)x3,21+a23(x¯1,x~21,x~22,x23,x24)x˙2n21=b2n21(x¯1,x~2)x3,21+a2n21(x¯1,x~2)x˙2n2=g(x¯1,x~2,x31,x3,21).(24)

The triangular dependence of the functions b2i, i=2,,n21 on the states (x23,,x2n2) is in fact a consequence of item (b). The triangular dependence of the functions a2i, i=2,,n21 on these states is guaranteed by item (c) condition (Equation16), i.e. [a,C(Δ1(i))]Δ1(i), i=1,,n23.

Step 5: Successively introduce the functions b2i as new states in the x2-subsystem from top to bottom. After n22 such steps, the x2-subsystem reads (25) x~˙21=x3,21x~˙22=x~23x3,21+a~22(x¯1,x~21,x~22,x~23)x~˙23=x~24x3,21+a~23(x¯1,x~21,x~22,x~23,x~24)x~˙2n21=x~2n2x3,21+a~2n21(x¯1,x~2)x~˙2n2=g~(x¯1,x~2,x31,x3,21).(25)

Remark 4.6

Introducing x3,11=g~(x¯1,x~2,x31,x3,21) would complete the transformation of the x2-subsystem to extended chained form (except for the dependence of the drift a~2 on the states x¯1, see also Remark 4.1). However, after this transformation, the distribution Dn31 would in general no longer be straightened out.

From the involutivity of Dn3, it follows that in the last line of (Equation25), we actually have x~˙2n2=g~1(x¯1,x~2,x31)+g~2(x¯1,x~2)x3,21, i.e. x3,212g~=0 and x31x3,21g~=0Footnote7. Introducing x3,11=g~1(x¯1,x~2,x31), i.e. replacing x31 by the new state x3,11 and leaving all the other coordinates unchanged, keeps all the distributions (Equation20) straightened out and results in x~˙2n2=x3,11+g~2(x¯1,x~2)x3,21. (For n3=1, it can analogously be shown that x~˙2n2=g~1(x¯1,x~2,x31)+g~2(x¯1,x~2)u~2, which after introducing x3,11=g~1(x¯1,x~2,x31) yields x~˙2n2=x3,11+g~2(x¯1,x~2)u~2.)

Step 6: Transform the x3-subsystem into Brunovsky normal form, by successively introducing new states from top to bottom and finally applying a suitable static feedback.

The transformation of an AI-system into the triangular form (Equation8) by following these six steps is demonstrated on two different examples in the following section.

5. Examples

5.1. Planar VTOL aircraft

Consider again our motivating example, the planar VTOL aircraft (Equation12). In the following, we apply Theorem 4.1 to show that this system is indeed static feedback equivalent to the triangular form (Equation8). Based on that, we will derive a flat output compatible with the triangular form. Finally, we will explicitly transform the system into the form (Equation8). The input vector fields of (Equation12) are given by b1=sin(θ)vx+cos(θ)vz and b2=ϵcos(θ)vx+ϵsin(θ)vz+ω. The drift is given by a=vxx+vzz+ωθvz. The distribution D1=span{b1,b2} is involutive, the distribution D2=D1+[a,D1]=span{sin(θ)vx+cos(θ)vz,ϵcos(θ)vx+ϵsin(θ)vz+ω,sin(θ)xcos(θ)zωcos(θ)vxωsin(θ)vz,ϵcos(θ)x+ϵsin(θ)z+θ+ϵωsin(θ)vxϵωcos(θ)vz}is not involutive, so we have n3=1 and the conditions dim(Di)=2i, i = 1, 2 hold. The condition C(D2)D1 of Theorem 4.1 is also met. Before we can evaluate the remaining conditions of Theorem 4.1, we have to construct a vector field bp for this system. For the distribution H=D2+[D1,D2], see Remark 4.5, we obtain H=span{sin(θ)xcos(θ)z,ϵx+cos(θ)θ,vx,vz,ω}and for the vector fields adan3+1b1 and adan3+1b2 we have ada2b1=2ωcos(θ)x+2ωsin(θ)z+ω2sin(θ)vxω2cos(θ)vz,ada2b2=2ϵωsin(θ)x2ϵωcos(θ)zϵω2cos(θ)vxϵω2sin(θ)vz.The vector field ada2b2 is contained in H, the vector field ada2b1 is not contained in H, so we have bp=b2 (see again Remark 4.5). Thus, for the distributions Δ0=Dn31+span{adan31bp} and Δ1=Dn3+span{adan3bp} we obtain Δ0=span{b2} and Δ1=span{b1,b2,[a,b2]}, respectively. With these distributions, the items (a)–(e) of Theorem 4.1 are met. We have Δ1(1)=span{ϵcos(θ)x+ϵsin(θ)z+θ,vx,vz,ω}=Δ¯1,thus item (b) is met and we have n2=3. Furthermore, we have G1=Δ¯1+[a,Δ¯1]=T(X), therefore dim(G1)=dim(Δ¯1)+2 holds and thus the x1-subsystem in the corresponding triangular form (Equation8) consists of two integrator chains, G1=T(X) furthermore implies that both of these chains are of length one. Thus, according to Section 4.2, Case 1, flat outputs compatible with the triangular form are all pairs of functions (φ1,φ2), which satisfy span{dφ1,dφ2}=(Δ¯1). From (Δ¯1)=span{dzϵsin(θ)dθ,dxϵcos(θ)dθ}, a possible pair of such functions follows as e.g. φ1=z+ϵcos(θ), φ2=xϵsin(θ). The same flat output has been derived in Martin et al. (Citation1996) based on physical considerations, or in Schöberl (Citation2014) based on an implicit triangular form. In conclusion, the planar VTOL (Equation12) is static feedback equivalent to the triangular form (Equation8) with n1,1=n1,2=1, n2=3 and n3=1. Let us demonstrate the transformation of the VTOL (Equation12) into the form (Equation8), such that the components of the flat output φ1=z+ϵcos(θ), φ2=xϵsin(θ) appear as top variables, by following the six steps of the sufficiency part of the proof of Theorem 4.1.

Step 1: In this example, the sequence of involutive distributions (Equation20) reduces to (26) Δ0Δ¯1G1=T(X).(26) These distributions can be straightened out by the state transformation (27) x1,11=φ1=z+ϵcos(θ)x22=Laφ2=vxϵωcos(θ)x1,21=φ2=xϵsin(θ)x23=θx21=Laφ1=vzϵωsin(θ)x31=ω.(27)

Remark 5.1

Choosing x1,11=φ1, x1,21=φ2, x21=Laφ1 and x22=Laφ2 is not mandatory for straightening out the distributions (Equation26). It is a short cut which immediately transforms the x1-subsystem into Brunovsky normal form, i.e. it joins together the Steps 1–3 of the proof. Alternatively, we could choose a transformation which just straightens out the distributions (Equation26), then introduce the components of the flat output as the states of the x1-subsystem (which here corresponds to applying Step 2) and then apply the Step 3.

Applying the transformation (Equation27) to (Equation12) results in (28) x˙1,11=x21x˙1,21=x22x˙21=cos(x23)(u1ϵ(x31)2)1x˙22=sin(x23)(u1ϵ(x31)2)x˙23=x31x˙31=u2.(28) Step 4: Since we have n3=1, in order to normalise the first equation of the x2-subsystem, we have to apply the input transformation explained in Note 6, instead of the state transformation (Equation23). The input transformation reads u~2=cos(x23)(u1ϵ(x31)2)1, u~1=u2. Applying this transformation to (Equation28) results in x˙1,11=x21x˙1,21=x22x˙21=u~2x˙22=tan(x23)(u~2+1)x˙23=x31x˙31=u~1.Step 5: We have to successively introduce the components of the input vector field associated with the input u~2 of the x2-subsystem as new states. Since n2=3, in this example we have only one such step, namely introducing x~23=tan(x23), which results in x˙1,11=x21x˙1,21=x22x˙21=u~2x˙22=x~23(u~2+1)x~˙23=(1+(x~23)2)x31x˙31=u~1.The last equation of the x2-subsystem is normalised by introducing x3,11=(1+(x~23)2)x31, which results in x˙1,11=x21x˙1,21=x22x˙21=u~2x˙22=x~23(u~2+1)x~˙23=x3,11x˙3,11=2x~231+(x~23)2(x3,11)2(1+(x~23)2)u~1.Step 6: By introducing u~~1=2x~231+(x~23)2(x3,11)2(1+(x~23)2)u~1, the x3-subsystem takes Brunovsky normal form, i.e. x3,11=u~~1, which completes the transformation into the form (Equation8). The individual transformation steps summarised to one state and input transformation read (29) x1,11=z+ϵcos(θ)x~23=tan(θ)x1,21=xϵsin(θ)x3,11=ωcos2(θ)x21=vzϵωsin(θ)u~~1=2ω2sin(θ)+cos(θ)u2cos3(θ)x22=vxϵωcos(θ)u~2=(u1ϵω2)cos(θ)1.(29)

5.2. Academic example

Consider the system (30) x˙1=u1x˙2=u2x˙3=sin(u1u2),(30) also considered in Lévine (Citation2009) and Schöberl (Citation2014). This system is not an AI-system, however, every non-linear system of the general form x˙=f(x,u) becomes an AI-system with according properties regarding flatness, by onefold prolonging every control. By prolonging both controls of (Equation30), we obtain the AI-system (31) x˙1=u1x˙2=u2x˙3=sin(u1u2)u˙1=u11u˙2=u12,(31) with the new state z=(x1,x2,x3,u1,u2), the new inputs u11 and u12, the input vector fields b1=u1 and b2=u2 and the drift a=u1x1+u2x2+sin(u1u2)x3. In the following, we show that (Equation31) is static feedback equivalent to the triangular form (Equation8) by applying Theorem 4.1. The distribution D1=span{b1,b2} is involutive, the distribution D2=span{u1,u2,u2x1+cos(u1u2)x3,u1x1+u2x2} is not involutive, so we have n3=1 and the conditions dim(Di)=2i, i = 1, 2 are met. The condition C(D2)D1 of Theorem 4.1 is also met. Before we can evaluate the remaining conditions of Theorem 4.1, we have to construct a vector field bp for this system. For the distribution H=D2+[D1,D2], see Remark 4.5, we obtain H=T(X). Thus, a vector field bp cannot be determined via the method described in Remark 4.5. Instead, we determine a vector field bp via (Equation19). In this example, we have v1=adan31b1=b1=u1v2=adan31b2=b2=u2.By inserting those vector fields into (Equation19), we obtain (32) (α1)2sin(u1u2)(u2)2+2α1α2(cos(u1u2)u2sin(u1u2)u1)u2+(α2)2(sin(u1u2)u12cos(u1u2)u2)u1x3=!0 mod D2.(32) The condition (Equation32) admits two independent non-trivial solutions, namely α1=λu1, α2=λu2 and α1=λ(u1tan(u1u2)2u2), α2=λu2tan(u1u2), both solutions with an arbitrary non-zero function λ(z). Thus, with the choice λ=1, we obtain the candidates bp=u1u1+u2u2 and bp=(u1tan(u1u2)2u2)u1+u2tan(u1u2)u2. If (Equation31) is indeed static feedback equivalent to (Equation8), the remaining conditions of Theorem 4.1 must be met with at least one of these candidates. This is indeed the case, namely with the vector field bp=u1u1+u2u2, i.e. the vector field constructed from the first solution. For the distributions Δ0=Dn31+span{adan31bp} and Δ1=Dn3+span{adan3bp} we obtain Δ0=span{bp}=span{u1u1+u2u2}Δ1=span{b1,b2,[a,bp]}=span{u1,u2,u1x1+u2x2}.With these distributions, the items (a) to (e) of Theorem 4.1 are met. We have Δ1(1)=span{u1,u2,x1,x2}=Δ¯1,thus item (b) is met and we have n2=3. Furthermore, we have G1=Δ¯1+[a,Δ¯1]=T(X), therefore dim(G1)=dim(Δ¯1)+1 holds and thus, the x1-subsystem in the corresponding triangular form (Equation8) only consists of one integrator chain, G1=T(X), i.e. s = 1, furthermore implies that this chain is of length one. Thus, according to Section 4.2, Case 2, flat outputs compatible with the triangular form are all pairs of functions (φ1,φ2), which satisfy L=span{dφ1,dLaφ1,dφ2}=Δ1+span{dLaφ1} with span{dφ1}=(Δ¯1). We have (Δ¯1)=span{dx3}, thus φ1=φ1(x3). Furthermore, we have Laφ1(x3)=sin(u1u2)x3φ1(x3) and thus L=Δ1+span{dLaφ1}=span{u2dx1u1dx2,u2du1u1du2,dx3}. Therefore, φ2=φ2(u1/u2,x1x2u1/u2,x3), chosen such that dφ1dLaφ1dφ20. A possible flat output is thus e.g. φ1=x3, φ2=x1x2u1/u2. In conclusion, (Equation31) is static feedback equivalent to the triangular form (Equation8) with n1=1, n2=3 and n3=1. Indeed, by applying a suitable state and input transformation to (Equation31), which again can be derived systematically following the six steps of the sufficiency part of the proof of Theorem 4.1, the system (Equation31) takes the form z˙1,11=z21z˙21=u~12z˙22=z23u~12z˙23=z3,11+z21z231(z21)2u~12z˙3,11=u~11,which is of the form (Equation8).

5.3. Further academic examples

Consider the following two academic examples: x˙1=u1x˙2=u2x˙3=u1u2x˙1=u1x˙2=u2x˙3=u1u2,which are similar to (Equation30) in the previous section and are also treated in e.g. Schöberl (Citation2014). Also these two systems are static feedback equivalent to (Equation8) (after turning them into AI-systems by prolonging each of their controls, as demonstrated on the previous example) and thus, can be transformed into the form (Equation8) systematically. For these systems, the dimensions of the individual subsystems in a corresponding triangular form (Equation8) would be n1=1, n2=3, n3=1 and n1=0, n2=4, n3=1, respectively. Therefore, these systems become static feedback linearisable by prolonging a suitably chose control twofold (as n2=3) or threefold (as n2=4), respectively. Flat outputs for these systems can again be derived systematically as described in Section 4.2, without actually transforming the systems into the form (Equation8). For these two systems as well as for (Equation30), flat outputs can also be derived using the results in Pomet (Citation1997). However, in order to apply these results, the systems have to be converted to AI-systems with four states first. For that, an input transformation such that one of the new inputs occurs affine has to be found. For the first of the above examples, this is already the case (either u1 or u2 can be interpreted as the affine occurring input), but for the other two examples it is not obvious how to find such an input transformation.

5.4. Explicit transformation into the triangular form

Based on the following academic example, we once more demonstrate the transformation into the triangular form (Equation8) by following the six steps of the sufficiency part of the proof of Theorem 4.1. Consider the system (33) x˙1=x2x˙6=x7(x9x8x10)x˙2=x4+sin(x6)x˙7=x1(x8x10x9)+sin(x8)x˙3=x2+x5x˙8=x9+x10x˙4=(x9x8x10)(1cos(x6)x7)x˙9=u1x˙5=x6(x9x8x10)x˙10=u2.(33) The input vector fields are given by b1=x9 and b2=x10, the drift is given by a=x2x1+(x4+sin(x6))x2+(x2+x5)x3+(x9x8x10)(1cos(x6)x7)x4+x6(x9x8x10)x5+x7(x9x8x10)x6+(x1(x8x10x9)+sin(x8))x7+(x9+x10)x8.The distributions D1=span{b1,b2}=span{x10,x9}D2=D1+[a,D1]=span{x10,x9,x8,(x7cos(x6)1)x4x6x5x7x6+x1x7}are involutive, the distribution D3=D2+[a,D2]=span{x10,x9,x8,x7,x2+x6x3+cos(x6)sin(x8)x4sin(x8)x6,x7x2+x6x7x3+sin(x8)x4+x6sin(x8)x5},is not invoultive, so we have n3=2. The conditions dim(Di)=2i, i=1,,3 and C(D3)D2 are met. For the distribution H=D3+[D2,D3], see Remark 4.5, we obtain H=span{x10,x9,x8,x7,x7cos(x6)x2+sin(x8)x6,x7x2x6sin(x8)x5,x2+x6x3,x7x2+sin(x8)x4}.For the vector fields adan3+1b1 and adan3+1b2 we have ada3b1=x1+1x6x2 mod Hada3b2=x8x1x8x6x2 mod H.The linear combination x8ada3b1+ada3b2=0 mod H is obviously contained in H. Thus, we have bp=x8b1+b2=x8x9+x10 and for the distributions Δ0=Dn31+span{adan31bp} and Δ1=Dn3+span{adan3bp} we obtain Δ0=span{x10,x9,x8}Δ1=span{x10,x9,x8,x7,(1x7cos(x6))x4+x6x5+x7x6}.With these distributions the items (a) to (e) of Theorem 4.1 are met. We have Δ1(1)=span{x10,x9,x8,x7,cos(x6)x4x6,x4+x6x5}Δ1(2)=span{x10,,x4}=Δ¯1,thus item (b) is met and we have n2=4. Furthermore, we have G1=span{x10,,x2}, G2=T(X) and dim(G1)=dim(Δ¯1)+2 hold. Thus, in the corresponding triangular form (Equation8), the x1-subsystem consists of two integrator chains with the lengths one and two. Thus, according to Section 4.2, Case 1, flat outputs compatible with the triangular form are all pairs of functions (φ1,φ2), which satisfy span{dφ1}=G1 and span{dφ1,dLaφ1,dφ2}=(Δ¯1). From G1=span{dx1}, φ1=φ1(x1) follows. From Laφ1(x1)=x2x1φ1(x1) and (Δ¯1)=span{dx1,dx2,dx3}, it follows that φ2=φ2(x1,x2,x3), chosen such that dφ1dLaφ1φ20. A possible flat output is thus, e.g. φ1=x1, φ2=x3. In the following, we transform (Equation33) into the triangular form (Equation8), such that the components of the flat output φ1=x1, φ2=x3 appear as top variables in the triangular form.

Step 1: In this example, the distributions D1Δ0C(Δ1(1))Δ¯1G1G2=T(X),corresponding to the sequence (Equation20), are already straightened out. Therefore, (Equation33) is structurally already in the form (Equation21). Indeed, by renaming the states according to x11=x1x12=x2x13=x3x21=x4x22=x5x23=x6x24=x7x31=x8x32=x9x33=x10,we obtain x˙11=x12x˙12=x21+sin(x23)x˙13=x12+x22x˙21=(x32x31x33)(1cos(x23)x24)x˙22=x23(x32x31x33)x˙23=x24(x32x31x33)x˙24=x11(x31x33x32)+sin(x31)x˙31=x32+x33x˙32=u1x˙33=u2,which is exactly the form (Equation21). The rank conditions rank((x21,x22,x23)f1)=2, rank((x31,x32,x33)f2)=2 and rank((x32,x33)f2)=1 hold.

Step 2: The x1-subsystem is already in Brunovsky normal form except for a normalisation of the ‘inputs’ of the integrator chains. To obtain exactly the representation (Equation22), we only have to rename the states of the x1-subsystem according to x1,11=x11, x1,12=x12 and x1,21=x13. This results in x˙1,11=x1,12x˙1,12=x21+sin(x23)x˙1,21=x1,12+x22.Step 3: The transformation of the x1-subsystem into Brunovsky normal form is completed by normalising the last two equations of the x1-subsystem, i.e. by introducing x~21=x21+sin(x23) and x~22=x1,12+x22, resulting in x˙1,11=x1,12x˙1,12=x~21x˙1,21=x~22x~˙21=x32x31x33x~˙22=x23(x32x31x33)+x~21x˙23=x24(x32x31x33)x˙24=x1,11(x31x33x32)+sin(x31).Step 4: Next, we normalise the first equation of the x2-subsystem, by introducing x3,21=x32x31x33. This leads to x~˙21=x3,21x~˙22=x23x3,21+x~21x˙23=x24x3,21x˙24=sin(x31)x1,11x3,21x˙31=(1+x31)x33+x3,21x˙3,21=h(x31,x3,21,x33,u1,u2)x˙33=u2,the x2-subsystem is indeed of the form (Equation24).

Step 5: We have to successively introduce the components of the input vector field associated with the input x3,21 of the x2-subsystem as new states (which here is actually already the case, we only have to rename the states according to x~23=x23 and x~24=x24 to be consistent with the notation in the proof of Theorem 4.1). Normalising the last equation of the x2-subsystem, i.e. introducing x3,11=sin(x31)x1,11x3,21, would complete the transformation of the x2-subsystem to extended chained form. However, this transformation would result in x~˙21=x3,21x~˙22=x~23x3,21+x~21x~˙23=x~24x3,21x~˙24=x3,11x˙3,11=h1(x1,11,x1,12,x3,11,x3,21,x33,u1,u2)x˙3,21=h2(x1,11,x3,11,x3,21,x33,u1,u2)x˙33=u2,preventing us from transforming the x3-subsystem into Brunovsky normal from by successively introducing new coordinates from top to bottom, since the inputs u1 and u2 occur in all three equations of the x3-subsystem (the distribution D1 is not straightened out anymore, see also Remark 4.6). Instead, we only introduce x3,11=sin(x31), which results in x~˙21=x3,21x~˙22=x~23x3,21+x~21x~˙23=x~24x3,21x~˙24=x3,11x1,11x3,21x˙3,11=1(x3,11)2x3,21+1+arcsin(x3,11)x33x˙3,21=x3,21+1+arcsin(x3,11)x33x33arcsin(x3,11)u2+u1x˙33=u2and keeps D1 straightened out, so the inputs u1 and u2 still only occur in the last two equations of the x3-subsystem.

Step 6: The last step is to transform the x3-subsystem into Brunovsky normal form. For that, we first introduce x3,12=1(x3,11)2(x3,21+(1+arcsin(x3,11))x33), to obtain x˙3,11=x3,12x˙3,12=h1(x3,11,x3,12,u1,u2)x˙3,21=h2(x3,11,x3,12,x3,21,u1,u2).Finally, we complete the transformation by introducing u~1=h1 and u~2=h2. After applying this input transformation, the complete system reads x˙1,11=x1,12x˙1,12=x~21x˙1,21=x~22x~˙21=x3,21x~˙22=x~23x3,21+x~21x~˙23=x~24x3,21x~˙24=x3,11x1,11x3,21.x˙3,11=x3,12x˙3,21=u~2x˙3,12=u~1.which is of the form (Equation8).

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The first author and the second author have been supported by the Austrian Science Fund (FWF) under grant number P 32151 and P 29964.

Notes

1 Note that the top variable x3,11 of the longer integrator chain corresponds to the input u1 in (Equation6), i.e. the input which only occurs in the very last equation of the extended chained form. This is crucial, if x3,11 and x3,21 would be swapped, the system would be static feedback equivalent to a system of the form (Equation8) with equally lengthened integrator chains in the x3-subsystem, i.e. it would be static feedback equivalent to the triangular form proposed in Gstöttner et al. (2020). In this case, the last equation of the x2-subsystem would belong to the shorter integrator chain, which would compensate the length difference.

2 If Δ¯1=T(X), the condition (Equation17) and the items (d) and (e) have to be omitted. In this case, the system is static feedback equivalent to (Equation8) with n1,1=n1,2=0 if and only if all the other conditions of the theorem are met.

3 Here, we used [v2,[a,v1]]=[v1,[a,v2]] mod Dn3+1, which follows from the Jacobi identity [v2,[a,v1]]+[v1,[v2,a]]=[v1,[a,v2]]+[a,[v1,v2]Dn3]Dn3+1=0.

4 Evaluated for e.g. i = 1, we obtain [x2n2,a]=(x3,21+x2n2a2n21)x2n21 mod C(Δ1(1)).Because of Δ1(1)=span{x3,x21+x23x22++x2n21x2n22,x2n2,x2n21} and C(Δ1(1))=span{x3,x2n2}, it indeed follows that [a,C(Δ1(1))]Δ1(1).

5 For n3=1, we have x˙2=f2(x1,x2,x31,u1,u2), i.e. f2 can depend on all the states and both inputs. The latter two rank conditions then read rank((u1,u2,x31)f2)=2 and rank((u1,u2)f2)=1.

6 In case that n3=1 holds, instead of the states x32 or x33 at least one of the inputs u1 or u2 occurs in f~2 (the inputs u1 and u2 would of course occur affine in f~2). Instead of the state transformation (Equation23), we then have the input transformation u~2=f~21(x¯1,x~2,x31,u1,u2) and u1 left unchanged. Crucial for this transformation to be regular is that f~21 indeed explicitly depends on u2 (after eventually swapping u1 and u2).

7 Calculating Dn3 in the coordinates obtained so far, we obtain Dn3=span{x3,x~21+x~23x~22++x~2n2x~2n21+x3,21g~x~2n2}. If the function g~ would be of the general nonlinear form g~(x¯1,x~2,x31,x3,21) with x3,212g~0 or x31x3,21g~0, the distribution Dn3 would not be involutive, which contradicts with Dn3 being indeed involutive.

8 The following proof of the feedback invariance of Dn3+1 is a replication of a part of the proof of Proposition 7.1 in Nicolau and Respondek (Citation2016b), adapted to our notation.

9 For n3=1, we have f2=f2(x1,x2,x31,u1,u2), i.e. f2 can depend on all the states and both inputs. The relation Δ1=Δ0+[a,Δ0] actually only holds for n32 (for n3=1, we have dim(Δ0)=1 and thus, the distribution Δ1=Δ0+[a,Δ0] would be of dimension 2). However, by an analogous reasoning based on the system obtained by onefold prolonging both inputs of the system, for f2=f2(x1,x2,x31,u1,u2) the rank conditions rank((x31,u1,u2)f2)=2 and rank((u1,u2)f2)=1 can be shown.

10 For n2=3, we would have Δ1(1)=Δ¯1 and thus C(Δ1(1))=Δ¯1. In this case, replace C(Δ1(1)) by L=span{x3,x2n2}. Because of LΔ1, there again exists a linear combination of v1 and v2 which is contained in span{x3,x2n2}, where x2n2=x23 in this case.

11 If n3=1, the state transformation (Equation23) is replaced by the input transformation u~2=f~21(x¯1,x~2,x31,u1,u2), see also Note 6. In (A3), x3,21 and x33 would then be replaced by u~2 and u1. By an analogous reasoning as above based on the prolonged system (see also Note 9), we would then find that the x2-subsystem is actually independent of u1 and that x31 again occurs only in the very last equation of the x2-subsystem, i.e. the x2-subsystem would again be of the form (A4), but with x3,21 replaced by u~2. That the input u~2 occurs affine in the x2-subsystem would follow directly from the fact that we started with an AI-system and only applied transformations which preserve the AI structure.

12 Note that we have L=span{x3,x2n2,,x23}. Thus, the vector field b2, by having a component in the x¯21-direction cannot be contained in L.

13 Note that because of bj,v~jDn3 and the involutivity of Dn3, we have [a~,v~j]=[a+γ1b1+γ2b2,v~j]=[a,v~j] mod Dn3.

14 All non-trivial solutions of the linear homogeneous equation α1β12+α2β22=0 are of the form α1=λβ22, α2=λβ12 with arbitrary λ0. We have at least β220 or β120, otherwise, the input transformation from above would not be invertible (i.e. for β11β22β12β21=0, the new input vector fields b~1 and b~2 would be linearly dependent). Inserting this solution into the second factor (α1β11+α2β21) yields λ(β11β22β12β21). This term can only vanish for λ=0 since β11β22β12β210 for a regular transformation. However, λ=0 is the trivial solution α1=α2=0.

15 Otherwise, we would have C(Dn3+1)=Dn3 and thus, the condition C(Dn3+1)Dn3 of Theorem 4.1 would be violated.

16 We have [v1,[a,v2]]=[b1c,b2c] mod Dn3+1 and [v2,[a,v2]]=[b2c+gb1c,[a,b2c+gb1c]]. Recall that we have Dn3+1=span{x3,b1c,b2c,[a,b2c+gb1c]} and H=span{x3,b1c,b2c,[a,b2c+gb1c],[b2c+gb1c,[a,b2c+gb1c]],[b1c,b2c]}, see (A5). Thus, we actually have H=Dn3+1+span{[v1,[a,v2]],[v2,[a,v2]]}. For dim(H)=dim(Dn3+1)+2, there neither [v1,[a,v2]] nor [v2,[a,v2]] can already be contained in Dn3+1, nor they can be collinear mod Dn3+1.

References

Appendix

Supplements

In this section, details omitted in the proof of our main theorem and proofs concerning the construction of a vector field bp are provided.

Proof

Proof of Lemma 4.2

Let {c1,,cnc} be a basis for C(D) and {v1,,vd} a basis for D. We obviously have [ci,vj]D and [vj,vk]D(1). From the Jacobi identity [vj,[ci,vk]D]D(1)+[vk,[vj,ci]D]D(1)+[ci,[vk,vj]D(1)]=0,i{1,,nc}, j,k{1,,d},it follows that [ci,[vk,vj]]D(1). Thus, every vector field ciC(D) is also characteristic for D(1), i.e. C(D)C(D(1)).

Proof

Proof of Lemma 4.3

Let us construct a special basis for the distribution D, namely D=span{c1,,cd2,v1,v2}, with cjC(D). Because of dim(D(i))=d+i, bases for D(1) and D(2) are then given by D(1)=span{c1,,cd2,v1,v2,v3} and D(2)=span{c1,,cd2,v1,v2,v3,v4} (with v3=[v1,v2] and v4=[v1,v3] or v4=[v2,v3] if [v1,v3]=0 mod D(1)). We obviously have [v1,v2]=0 mod D(1). Furthermore, we have [v1,v3]=α1v4 mod D(1) and [v2,v3]=α2v4 mod D(1), where α1 and α2 are some functions and at least α10 or α20. The vector field v~1=α2v1α1v2D satisfies [v~1,vi]=0 mod D(1), i=1,,3 and [v~1,cj]=0 mod DD(1), j=1,,d2. Thus, v~1 is a characteristic vector field of D(1). Because of Lemma 4.2, we furthermore have C(D)C(D(1)). Thus, we have C(D(1))=C(D)+span{v~1}D. The rest of the proof follows the same line. A basis for D(1) is given by D(1)=span{c1,,cd1,w1,w2}, where cd1=v~1, w1=v1 (or w1=v2 if v1 and v~1 are collinear mod C(D)) and w2=v3. This way, we formally obtained the same problem as before. Thus, by essentially the same argumentation as before, it follows that C(D(2))=C(D(1))+span{w~1}D(1), with w~1 being a suitable linear combination of w1 and w2. Continuing this argumentation, dim(C(D(i)))=d2+i and C(D(i))D(i1), i=1,,l1 follows.

A.1. Details omitted in the proof of Theorem 4.1

A.1.1. Feedback invariance of certain distributions

Right before Theorem 4.1, for an AI-system (Equation15), the distributions Di, i=1,,n3+1, D1=span{b1,b2} and Di+1=Di+[a,Di], with the smallest integer n3 such that Dn3+1 is not involutive, are defined. Since they are of importance for Theorem 4.1, let us discuss their invariance with respect to invertible static feedback transformations.Footnote8 It is well known that the involutive distributions Di, i=1,,n3 are feedback invariant, see, e.g. Jakubczyk and Respondek (Citation1980) or Nijmeijer and van der Schaft (Citation1990). To show that also the first non-involutive distribution Dn3+1=Dn3+[a,Dn3] is feedback invariant, we calculate Dn3+1 after applying an invertible static feedback. After applying an invertible static feedback to an AI-system (Equation15), its input vector fields and drift read b~1=β11b1+β12b2, b~2=β21b1+β22b2, with β11β22β12β210 and a~=a+γ1b1+γ2b2, where βij and γj are functions of the state of the system. Since D~n3=Dn3 is involutive, we obtain D~n3+1=Dn3+[a~,Dn3]=Dn3+[a+γ1b1+γ2b2,Dn3]=Dn3+[a,Dn3]=Dn3+1+[γ1b1+γ2b2,Dn3]Dn3,i.e. D~n3+1=Dn3+1 indeed holds. Also the distributions Δ0=Dn31+span{adan31bp} and Δ1=Dn3+span{adan3bp} play a crucial role in Theorem 4.1. These are also feedback invariant, i.e. for a fixed vector field bpD1, calculating these distribution with the feedback modified drift a~=a+γ1b1+γ2b2, yields the same distributions Δ0 and Δ1. Furthermore, only the direction of bp matters, i.e. with b~p=λbp with an arbitrary non-zero function λ of the state of the system, we again obtain the same distributions Δ0 and Δ1. To show this, note that because of the involutivity of Di, i=1,,n3 and adai1bjDi, we have ada~1b~p=[a+γ1b1+γ2b2,λbp]=λada1bp mod D1ada~2b~p=[a+γ1b1+γ2b2,λada1bp mod D1]=λada2bp mod D2ada~n31b~p=λadan31bp mod Dn31ada~n3b~p=λadan3bp mod Dn3and thus, the distribution Δ~0=Dn31+span{ada~n31b~p} indeed coincides with Δ0=Dn31+span{adan31bp}, and Δ~1=Dn3+span{ada~n3b~p} indeed coincides with Δ1=Dn3+span{adan3bp}.

A.1.2. Details necessity

Proof

Proof of Fact 4.1

In the necessity part of the proof of Theorem 4.1, we claimed that the distribution Dn3+1=span{x3,b1c,b2c,[a,b2c+gb1c]}, with b1c=x2n2, b2c=x21+x23x22++x2n2x2n21 and a=f1+x3,21(b2c+gb1c)+x3,11b1c+a2+a3 is not involutive, meets dim(Dn3+1)=2n3+2 and C(Dn3+1)Dn3. To show this, note that the vector field [a,b2c+gb1c] is of the form [a,b2c+gb1c]=[f1,b2c]+x3,11[b1c,b2c]+[a2,b2c]+g[a2,b1c]h(x1,x2)[b1c,b2c] mod span{x3,b1c,b2c}=[f1,b2c]+h~(x1,x2,x3,11)[b1c,b2c]+[a2,b2c] mod span{x3,b1c,b2c}.Since [b1c,b2c]=x2n21span{x3,b1c,b2c}, the condition dim(Dn3+1)=2n3+2 indeed holds, it holds independently of the dimension n1=n1,1+n1,2 of the x1-subsystem and the actual form of the drift vector field a2 of the x2-subsystem. Next, let us show the non-involutivity of Dn3+1 by contradiction. Assume that Dn3+1 would be involutive. The distribution Dn3+1 contains the vector fields b1c and b2c. Since [b1c,b2c]=x2n21span{x3,b1c,b2c}, in order for Dn3+1 to be involutive, there must hold [a,b2c+gb1c]=h(x1,x2,x3,11)[b1c,b2c] mod span{x3,b1c,b2c} (otherwise, the Lie bracket [b1c,b2c] would not be contained in Dn3+1). This is only possible if [f1,b2c]=0 mod span{x3,b1c,b2c,[b1c,b2c]} and thus n1=0 (for n1>0, [f1,b2c] has non-zero components in the x1-direction and those are certainly not contained in span{x3,b1c,b2c,[b1c,b2c]}). This shows the non-involutivity of Dn3+1 for the case n1>0. For n1=0, the case [a,b2c+gb1c]=h(x1,x2,x3,11)[b1c,b2c] mod span{x3,b1c,b2c} can indeed occur (it occurs when n1=0 and [a2,b2c]=0 mod span{x3,b1c,b2c,[b1c,b2c]}) and results in Dn3+1=span{x3,b1c,b2c,[b1c,b2c]}. This distribution is only involutive, if [[b1c,b2c],b2c]=x2n22Dn3+1, which only holds for n2=3 (we do not consider the degenerated case n22, see also Remark 4.2). However, for n2=3, we have Dn3+1=T(X) and in turn, the system is static feedback linearisable. In the case that Dn3+1=span{x3,b1c,b2c,[b1c,b2c]} holds, we have C(Dn3+1)=span{x3,b1c}Dn3. For n11 or [a2,b2c]0 mod span{x3,b1c,b2c,[b1c,b2c]}, the distribution Dn3+1 explicitly depends on x3,11, i.e. [x3,11,Dn3+1]Dn3+1. In this case, because of x3,11Dn3, the condition C(Dn3+1)Dn3 also holds.

A.1.3. Details sufficiency

Details step 1. Consider again the sequence of nested involutive distributions (Equation20). By assumption, we have dim(Di)=2i for i=1,,n3+1 and by construction, we have dim(Δ0)=dim(Dn31)+1=2n31. Furthermore, by construction we have dim(Δ1)=dim(Dn3)+1=2n3+1=dim(Δ0)+2. Because of items (a) and (b), Lemma 4.3 applies to Δ1. Thus, for the Cauchy characteristic distributions C(Δ1(i)), we have dim(C(Δ1(i)))=dim(Δ0)+i for i=1,,n23. For the involutive closure of Δ1, we have dim(Δ¯1)=dim(Δ0)+n2. In Step 1 of the sufficiency part of the proof of Theorem 4.1, we claimed that after straightening out the distributions (Equation20), i.e. applying a change of coordinates such that (A1) D1=span{x32n31,x32n32}Dn31=span{x32n31,,x32}Δ0=span{x3}C(Δ1(1))=span{x3,x2n2}C(Δ1(n23))=span{x3,x2n2,,x24}G0=Δ¯1=span{x3,x2}Gs=span{x3,x2,x1}=T(X),(A1) the system is decomposed into the form (A2) x˙1=f1(x1,x21,x22,x23)x˙2=f2(x1,x2,x31,x32,x33)x˙3=f3(x1,x2,x3,u1,u2),(A2) with rank((x21,x22,x23)f1)2, rank((x31,x32,x33)f2)=2 and rank((x32,x33)f2)=1. In the following, we explain why this is indeed the case.

Proof

Proof of Proposition 4.1

That f1 only depends on the states (x1,x21,x22,x23) and is independent of the states (x24,,x2n2,x3) is implied by item (c) condition (Equation16) evaluated for i=n23, i.e. [a,C(Δ1(n23))]Δ1(n23). We have C(Δ1(n23))=span{x3,x2n2,,x24}. If f1 would depend on any of the states (x24,,x2n2,x3), then [a,C(Δ1(n23))] would contain vector fields with a x1-component and thus, [a,C(Δ1(n23))] would not be contained in Δ1(n23)Δ¯1=span{x3,x2}.

If n2=3, condition (Equation16) does not exist and we have x2=(x21,x22,x23). In this case, f1 can depend on all the states x2. To show that also in this case f1 cannot depend on the stats x3, note that we have Δ0=span{x3} and by construction [a,Δ0]Δ1Δ¯1=span{x3,x2}. If f1 would depend on x3, then [a,Δ0] would contain vector fields with a x1-components and thus, [a,Δ0] would not be contained in Δ¯1=span{x3,x2}.

Next, let us show that f2=f2(x1,x2,x31,x32,x33). We have Dn32=span{x32n31,,x34} and Dn31=span{x32n31,,x32}. Since by construction [a,Dn32]Dn31 and Dn31 contains no vector fields which have a x2-component, it follows that f2 is indeed independent of the states (x34,,x32n31).

Proof

Proof of Proposition 4.2

We have to show several rank conditions. First, let us show that rank((x21,x22,x23)f1)2 holds, or equivalently that dim(G1)dim(Δ¯1)+2 holds. Recall that because of items (a) and (b), Lemma 4.3 applies to Δ1 and thus, we have dim(C(Δ1(n23)))=dim(Δ¯1)3 and dim(Δ1(n23))=dim(Δ¯1)1. Therefore, there exist three vector fields v1,,v3 such that Δ1(n23)=C(Δ1(n23))+span{v1,v2} and Δ¯1(n23)=C(Δ1(n23))+span{v1,v2,v3}. Due to item (c) condition (Equation17), i.e. dim(Δ¯1+[a,Δ1(n23)])=dim(Δ¯1)+1, the vector fields [a,v1] and [a,v2] are collinear mod Δ¯1. Therefore, the dimension of G1=Δ¯1+span{[a,v1],[a,v2],[a,v3]} exceeds that of Δ¯1 at most by two and thus rank((x21,x22,x23)f1)2 holds.

Next, let us show that rank((x31,x32,x33)f2)=2 holds. Note that because of f2=f2(x1,x2,x31,x32,x33) we have rank((x31,x32,x33)f2)=rank(x3f2), i.e. since f2 is independent of (x34,,x32n31), the Jacobian matrices of f2 with respect to (x31,x32,x33) and with respect to all the states x3 have the same rank. We have Δ0=span{x3}. From Δ1=Δ0+[a,Δ0] and dim(Δ1)=dim(Δ0)+2 it follows that rank(x3f2)=2 holds and thus, also rank((x31,x32,x33)f2)=2 indeed holds .Footnote9

The last rank condition, namely rank((x32,x33)f2)=1, follows from Dn31Δ0Dn3, dim(Dn3)=dim(Δ0)+1 and Dn3=Dn31+[a,Dn31]. Since Dn31Δ0Dn3, we also have Dn3=Δ0+[a,Dn31] and since dim(Dn3)=dim(Δ0)+1, [a,Dn31] yields exactly one direction which is not already contained in Δ0=span{x3}. Because of Dn31=span{x32n31,,x33,x32}, this implies rank((x32,x33)f2)=1.

Details step 2. In the following, we show Propositions 4.3–4.5, i.e. we show that in any case regarding the actual form of the x1-subsystem, there always exists a certain involutive distribution L. The distribution L is of importance in the problem of transforming the x2-subsystem into (extended) chained form.

Proof

Proof of Proposition 4.3

For rank((x21,x22,x23)f1)=2, the x1-subsystem consists of two integrator chains and we have to show that the functions φj(x¯1,x21,x22,x23), j = 1, 2 in (Equation22) meet L=(span{dx¯1,dφ1,dφ2})Δ1(n23). To show this, let us first construct a special basis for the distribution Δ1(n23). Recall that Lemma 4.3 applies to Δ1 and thus, we have dim(C(Δ1(n23)))=dim(Δ1(n23))2. Therefore, the distribution Δ1(n23) can be represented as Δ1(n23)=C(Δ1(n23))+span{v1,v2}, with suitable vector fields v1,v2Δ1(n23). Because of item (c) condition (Equation17), i.e. dim(Δ¯1+[a,Δ1(n23)])=dim(Δ¯1)+1, the vector fields [a,v1] and [a,v2] are collinear mod Δ¯1, i.e. (permute v1 and v2 if necessary) [a,v2]=λ[a,v1] mod Δ¯1. The vector field v~2=v2λv1 therefore satisfies [a,v~2]=λ[a,v1]λ[a,v1]Laλv1=0 mod Δ¯1and thus Lv~2φj=0. Therefore, by choosing a basis of C(Δ1(n23)) together with v1 and v~2 as basis for Δ1(n23), i.e. Δ1(n23)=C(Δ1(n23))+span{v1,v~2}, we obtain a basis of which all basis vector fields, except for v1, are annihilated by dφ1 and dφ2. The 1-form ω=(dφ2v1)dφ1(dφ1v1)dφ2,because of ωv1=(dφ2v1)(dφ1v1)(dφ1v1)(dφ2v1)=0annihilates all basis vector fields of Δ1(n23). This 1-form together with dx¯1 therefore spans the annihilator of Δ1(n23). This shows that the annihilator of Δ1(n23) is indeed a sub-codistribution of L=span{dx¯1,dφ1,dφ2}, i.e. (Δ1(n23))span{dx¯1,dφ1,dφ2}=L, or, equivalently L=(span{dx¯1,dφ1,dφ2})Δ1(n23) indeed holds.

Proof

Proof of Proposition 4.4

For rank((x21,x22,x23)f1)=1, the x1-subsystem consists only of one integrator chain. In this case, the x1-subsystem determines one function φ1(x¯1,x21,x22,x23), i.e. the ‘input’ of the single integrator chain. We have to show that in this case, there always exists a second function φ2(x¯1,x21,x22,x23) such that L=(span{dx¯1,dφ1,dφ2})Δ1(n23). For that, we again make use of a special basis for the distribution Δ1(n23), namely Δ1(n23)=C(Δ1(n23))+span{v1,v2} with vj=vji(x21,x22,x23)x2i, i=1,,3, and thus [c,vj]=0 mod C(Δ1(n23)) for any cC(Δ1(n23)). The non-zero vector field v~=(dφ1v2)v1(dφ1v1)v2 annihilates dφ1, i.e. Lv~φ1=0. Together with a basis of C(Δ1(n23)), this vector field v~ spans the involutive distribution L=C(Δ1(n23))+span{v~}Δ1(n23). We obviously have dφ1L and since L is involutive, there exists a second function φ2(x21,x22,x23) such that L=(span{dx¯1,dφ1,dφ2})Δ1(n23). The distribution L is uniquely determined by the function φ1(x21,x22,x23), the particular choice of the basis vector fields v1 and v2 does not matter. In fact, for any pair of vector fields w1,w2, which together with a basis of C(Δ1(n23)) spans Δ1(n23), the distribution L~=C(Δ1(n23))+span{(dφ1w2)w1(dφ1w1)w2} coincides with L=C(Δ1(n23))+span{(dφ1v2)v1(dφ1v1)v2} from above, i.e. L~=L. To show this, note that any vector fields w1 and w2, which together with C(Δ1(n23)) span the distribution Δ1(n23), can be written as a linear combination wj=βj1v1+βj2v2 mod C(Δ1(n23)), j = 1, 2, with the vector fields vj=vji(x21,x22,x23)x2i, i=1,,3, from above and functions βji=βji(x¯1,x2,x3) and β11β22β21β120. Because of (dφ1w2)w1(dφ1w1)w2=(dφ1v1)β21+(dφ1v2)β22(β11v1+β12v2)(dφ1v1)β11+(dφ1v2)β12(β21v1+β22v2) mod C(Δ1(n23))=(β11β22β21β12)0(dφ1v2)v1(dφ1v1)v2 mod C(Δ1(n23)),the distribution L~=C(Δ1(n23))+span{(dφ1w2)w1(dφ1w1)w2} indeed coincides with L=C(Δ1(n23))+span{(dφ1v2)v1(dφ1v1)v2}. Therefore, we obtain the unique distribution L via L=C(Δ1(n23))+span{(dφ1w2)w1(dφ1w1)w2} by choosing arbitrary vector fields w1 and w2, which together with C(Δ1(n23)) span the distribution Δ1(n23).

Proof

Proof of Proposition 4.5

For Δ¯1=T(X), there does not exist an x1-subsystem. We have to show that in this case, there always exist two functions φj(x21,x22,x23), j = 1, 2 which fulfill L=(span{dφ1,dφ2})Δ1(n23). The construction of L is essentially the same as in the proof of Proposition 4.4. We just have to omit x¯1 and since there is no x1-subsystem which determines a function φ1, we have to choose one. A valid choice is any function φ1(x21,x22,x23), dφ10. The distribution L is then obtained via L=C(Δ1(n23))+span{(dφ1w2)w1(dφ2w1)w2}, again with arbitrary vector fields w1 and w2, which together with C(Δ1(n23)) span the distribution Δ1(n23). The distribution L is again independent of the particular choice of w1 and w2, it is uniquely determined by the choice of φ1(x21,x22,x23). This construction in fact coincides with the construction of L provided in Li Respondek (Citation2012), Theorem 2.10, i.e. in case that no x1-subsystem exists, the results from Li Respondek (Citation2012) directly apply.

There is another way to calculate the distribution L to a given function φ1 (either determined by the x1-subsystem or chosen if no x1-subsystem exists). As we have seen above, the distribution L is uniquely determined by the function φ1 and there always exists a second function φ2 such that L=(span{dx¯1,dφ1,dφ2})Δ1(n23) (omit x¯1 in case that there is no x1-subsystem). Recall that we have dim(Δ1(n23))=dim(Δ¯1)1 and Δ¯1=span{x3,x2}. The annihilator of Δ1(n23) is thus of the form (Δ1(n23))=span{dx¯1,ω} and because of L=(span{dx¯1,dφ1,dφ2})Δ1(n23), the annihilator of Δ1(n23) is a sub-codistribution of L=span{dx¯1,dφ1,dφ2}, i.e. (Δ1(n23))=span{dx¯1,ω}span{dx¯1,dφ1,dφ2}. Thus, the 1-form ω is a linear combination of the differentials dφ1 and dφ2 and thus, we have L=span{dx¯1,dφ1,ω}. Therefore, given φ1, we immediately obtain the annihilator of the associated distribution L via L=(Δ1(n23))+span{dφ1}. By integrating this codistribution, we obtain a possible second function φ2.

Details step 4. In the following, we show that after applying the transformation (Equation23), the x2-subsystem takes the form (Equation24), as asserted in Proposition 4.6. The Fact 4.2 is shown subsequently.

Proof

Proof of Proposition 4.6

The transformation (Equation23) normalises the first equation of the x2-subsystem, i.e. applying this transformation immediately yields an x2-subsystem of the form (A3) x~˙21=x3,21x~˙22=f~22(x¯1,x~2,x31,x3,21,x33)x˙2n2=f~2n2(x¯1,x~2,x31,x3,21,x33).(A3) The same way as in the proof of Proposition 4.2 the rank condition rank((x32,x33)f2)=1 is shown, for (A3) the rank condition rank((x3,21,x33)f~2)=1 can be shown, which implies that the functions f~2j, j=2,,n2 are actually independent of x33. In other words, the transformation (Equation23) eliminates the redundancy among the inputs of the x2-subsystem. For n32, we have Δ1=Δ0+[a,Δ0] and thus Δ1=span{x3,v1,v2} with the vector fields v1=x~21+x3,21f~22x~22++x3,21f~2n2x2n2 and v2=x31f~22x~22++x31f~2n2x2n2. Furthermore, Lemma 4.3 applies to Δ1. Therefore, there exists a linear combination of v1 and v2 which is contained in C(Δ1(1))=span{x3,x2n2} Footnote10. The vector field v1 has a non-zero component in the x~21-direction, the vector field v2 has not. Therefore, linear combinations of v1 and v2 which are contained in span{x3,x2n2} consist of v2 only, and thus, we actually have v2C(Δ1(1)). Since v2=x31f~22x~22++x31f~2n2x2n2, it follows that x31 can only occur in the function f~2n2. Furthermore, in order for C(Δ1)=Δ0 to hold, x3,21 must occur affine in the functions f~2j, j=2,,n21, i.e. the x2-subsystem is actually of the form Footnote11 (A4) x~˙21=x3,21x~˙22=b22(x¯1,x~2)x3,21+a22(x¯1,x~2)x˙2n21=b2n21(x¯1,x~2)x3,21+a2n21(x¯1,x~2)x˙2n2=g(x¯1,x~2,x31,x3,21).(A4) Next, we show that the functions b2i in (A4) depend on the states of the x2-subsystem in a triangular manner. Lemma 4.3 applies to Δ1. Based on that, we will first show that Δ1(i)=span{b2}+C(Δ1(i+1)), i=0,,n24, with b2=x~21+b22x~22+b23x23++b2n21x2n21, i.e. that the derived flags Δ1(i) are composed of the one-dimensional distribution spanned by the vector field b2 and the Cauchy characteristic distributions of their next derived flags. To show this, note that the vector field b2 has a component in the x~21-direction. Thus, it cannot belong to any of the Cauchy characteristics C(Δ1(i+1))=span{x3,x2n2,,x2n2i}, i=0,,n24. However, because of b2Δ1, the vector field b2 also belongs to all the derived flags of Δ1. Furthermore, because of Lemma 4.3, we have dim(C(Δ1(i+1)))=dim(Δ1(i))1, i=0,,n24. Thus, span{b2} completes C(Δ1(i+1)) to Δ1(i). By Proposition 4.3, 4.4 or 4.5, depending on which case actually applies, we furthermore have the distribution L=(span{dx¯1,dφ1,dφ2})Δ1(n23) and it follows that b2L Footnote12. Thus, Δ1(n23)=span{b2}+L holds, i.e. Δ1(n23)=span{x3,x2n2,,x23,b2}. In conclusion, we have Δ1(i)=span{x3,x2n2,,x2n2i,b2}, i=0,,n23, from which x2jb2k=0 for k+2jn2, k=2,,n22, and x2jb2k0 for j = k + 1, k=2,,n21 follows. This exactly describes the triangular dependence of the functions b2i, i=2,,n21 on the states (x23,,x2n2) in (Equation24).

The triangular dependence of the functions a2i, i=2,,n21 on the states (x23,,x2n2) in (Equation24) is implied by item (c) condition (Equation16), i.e. [a,C(Δ1(i))]Δ1(i), i=1,,n23. We have C(Δ1(i))=span{x3,x2n2,,x2n2i+1}, i=1,,n23 and Δ1(i)=span{x3,x2n2,,x2n2i,b2}, i=0,,n23. Evaluating [a,C(Δ1(i))]Δ1(i), i=1,,n23 therefore yields x2ja2k=0 for k+2jn2, k=2,,n22. The condition [a,C(Δ1(i))]Δ1(i), i=1,,n23 in fact coincides with the compatibility condition (Equation7) in Theorem 3.3 for the extended chained form.

Proof

Proof of Fact 4.2

It follows from the construction of the x2-subsystem via the derived flags of Δ1 and Δ1=Δ0+[a,Δ0] that every component of the right-hand side of the x2-subsystem, i.e. every function f~2i, i=1,,n2 in x~˙2=f~2(x¯1,x~2,x31,x32,x33) explicitly depends on at least one of the inputs (x31,x32,x33) of the x2-subsystem. Under the assumption that f~21 indeed explicitly depends on x32 (after eventually swapping x32 and x33), in Step 4 of the proof, x32 is replaced by x3,21=f~21(x¯1,x~2,x31,x32,x33), which results in an x2-subsystem of the form (Equation24) and Dn31=span{x32n31,,x33,x3,21}, i.e. this transformation certainly keeps the distribution Dn31 straightened out. To show that f~21 indeed depends on x32 or x33, let us instead replace the state x31 by x3,21=f~21(x¯1,x~2,x31,x32,x33) and keep x32 and x33 as coordinate (if this transformation is not regular, i.e. if f~21 is independent of x31, then we can stop here, since then we readily have that f~21 explicitly depends on x32 or x33). By an in fact analogous reasoning as in the proof of Proposition 4.6, it then follows that after this transformation, the x2-subsystem takes the form x~˙21=x3,21x~˙22=b22(x¯1,x~21,x~22,x23)x3,21+a22(x¯1,x~21,x~22,x23)x˙23=b23(x¯1,x~21,x~22,x23,x24)x3,21+a23(x¯1,x~21,x~22,x23,x24)x˙2n21=b2n21(x¯1,x~2)x3,21+a2n21(x¯1,x~2)x˙2n2=g(x¯1,x~2,x32,x33,x3,21).Furthermore, in the new coordinates, the distribution Dn31 takes the form Dn31=span{x32n31,,x34,x33+x33f~21x3,21,x32+x32f~21x3,21}. Assume that f~21 is independent of x32 and x33. Then, we have Dn31=span{x32n31,,x34,x33,x32}, i.e. Dn31 is still straightened out. However, this leads to Dn3=Dn31+[a,Dn31]=span{x3,x2n2} and Dn3+1=Dn3+[a,Dn3]=span{x3,x2n2,x2n21,x~21+b22x~22++b2n22x2n22} and in turn, C(Dn3+1)=Dn3 would hold, or, if n2=3, Dn3+1 would be involutive. (In fact, this would lead exactly to the case mentioned in Note 1, where the longer integrator chain of the x3-subsystem is attached to the ‘wrong’ input of the x2-subsystem.) For n3=1, we have the input transformation u~2=f~21(x¯1,x~2,x31,u1,u2) instead of the state transformation x3,21=f~21(x¯1,x~2,x31,x32,x33), see also Note 6. Based on the system obtained by onefold prolonging both inputs of the system it can then analogously be shown that an f~21 which does not depend on an input u1 or u2 leads to the same contradiction.

A.2. Proof of the simple method for determining bp

In the following, we show why in the case adan3+1b1H or adan3+1b2H, with the distribution H=Dn3+1+[Dn3,Dn3+1], a vector field bp can indeed be determined from the criterion adan3+1bpH, as proposed in Remark 4.5. For a system of the form (Equation8), we obtain (recall that we have Dn3=span{x3,b2c+gb1c} and Dn3+1=span{x3,b1c,b2c,[a,b2c+gb1c]}) (A5) H=Dn3+1+[Dn3,Dn3+1]=span{x3,b1c,b2c,[a,b2c+gb1c],[b2c+gb1c,[a,b2c+gb1c]],[b1c,b2c]}.(A5) The input vector fields of (Equation8) are b1=x3,1n3 and b2=x3,2n31 (for n3=1, we actually have b2=b2c+gb1c), where b1 is the input vector field belonging to the longer integrator chain of the x3-subsystem. The conditions of Theorem 4.1 are met with any non-zero vector field bp which is collinear with b1. For any vector field bp which is collinear with b1=x3,1n3, i.e. bp=λx3,1n3 with an arbitrary non-zero function λ of the state of the system, we obtain adan3+1bp=λ(1)n3[a,b1c] mod Dn3+1 and thus, because of [a,b1c]span{x3,b1c,[b1c,b2c]}H and Dn3+1H, we have adan3+1bpH. Whereas adan3+1b2=(1)n31[a,[a,b2c+gb1c]] may or may not be contained in H. If adan3+1b2H, then adan3+1bpH is indeed only met for vector fields bp which are collinear with b1, i.e. collinear with the input vector field of the longer integrator chain in the x3-subsystem. However, if also adan3+1b2H, this criterion for determining a vector field bp is not applicable, since adan3+1bpH would be met for every linear combination bp of the input vector fields of the system.

A.3. Analysis of the necessary condition (19)

In the following, we analyse the necessary condition (Equation19) in terms of uniqueness of the direction of candidates for bp=α1b1+α2b2. For a system of the form (Equation8), we have b1=x3,1n3 and b2=x3,2n31 (for n3=1, we actually have b2=b2c+gb1c) and thus v1=adan31b1=(1)n31x3,11 and v2=adan31b2=(1)n31(b2c+gb1c), again with b1c=x2n2 and b2c=x21+x23x22++x2n2x2n21. The conditions of Theorem 4.1 are met with any non-zero vector field bp which is collinear with b1=x3,1n3. Assume we apply a regular input transformation on the system. Then, we have b~1=β11b1+β12b2 and b~2=β21b1+β22b2, with β11β22β12β210, and a~=a+γ1b1+γ2b2 (βij and γj being functions of the state of the system, i.e. βij=βij(x) and γj=γj(x)). For v~j=ada~n31b~j, we accordingly obtain v~1=β11v1+β12v2 mod Dn31 and v~2=β21v1+β22v2 mod Dn31. In the following, we show that by solving the necessary condition (Equation19), we obtain at most two non-collinear candidates for the vector field bp, and that one of these candidates is collinear with b1=x3,1n3. We start by inserting v~1 and v~2 into (Equation19), i.e.  (A6) (α1)2[v~1,[a~,v~1]]+2α1α2[v~1,[a~,v~2]]+(α2)2[v~2,[a~,v~2]]!Dn3+1.(A6) By inserting the corresponding expressions for v~1, v~2 and a~ from above, we obtainFootnote13 (α1)2[β11v1+β12v2,[a,β11v1+β12v2]]+2α1α2[β11v1+β12v2,[a,β21v1+β22v2]]+(α2)2[β21v1+β22v2,[a,β21v1+β22v2]]!Dn3+1.Expanding yields (α1)2(β11)2[v1,[a,v1]]+β11β12[v1,[a,v2]]+β12β11[v2,[a,v1]]+(β12)2[v2,[a,v2]]+2α1α2β11β21[v1,[a,v1]]+β11β22[v1,[a,v2]]+β12β21[v2,[a,v1]]+β12β22[v2,[a,v2]]+(α2)2(β21)2[v1,[a,v1]]+β21β22[v1,[a,v2]]+β22β21[v2,[a,v1]]+(β22)2[v2,[a,v2]]!Dn3+1.With [v2,[a,v1]]=[v1,[a,v2]] mod Dn3+1 (following from the Jacobi identity), and [v1,[a,v1]]Dn3+1 (actually [v1,[a,v1]]=0), we obtain (α1)22β11β12[v1,[a,v2]]+(β12)2[v2,[a,v2]]+2α1α2(β11β22+β12β21)[v1,[a,v2]]+β12β22[v2,[a,v2]]+(α2)22β21β22[v1,[a,v2]]+(β22)2[v2,[a,v2]]!Dn3+1,and after some rearranging 2(α1)2β11β12+α1α2(β11β22+β12β21)+(α2)2β21β22[v1,[a,v2]]+(α1)2(β12)2+2α1α2β12β22+(α2)2(β22)2[v2,[a,v2]]!Dn3+1,and finally (A7) (α1β12+α2β22)2(α1β11+α2β21)[v1,[a,v2]]+(α1β12+α2β22)[v2,[a,v2]]!Dn3+1.(A7) In the following, we have to distinguish between two cases, namely between [v1,[a,v2]] and [v2,[a,v2]] being collinear mod Dn3+1 or not.

Case 1: Let us first consider the case that [v1,[a,v2]] and [v2,[a,v2]] are not collinear mod Dn3+1. In this case, there does not exist a non-trivial linear combination of the vector fields [v1,[a,v2]] and [v2,[a,v2]] which is contained in Dn3+1. Furthermore, the factors (α1β11+α2β21) and (α1β12+α2β22) cannot vanish simultaneously for α10 or α20 Footnote14, or in other words, we cannot chose α1 and α2 such that in (A7) there occurs a trivial linear combination of [v1,[a,v2]] and [v2,[a,v2]]. Thus, in this case, in order for (A7) to hold, the factor (α1β12+α2β22) must vanish and thus α1=λβ22 and α2=λβ12 with arbitrary λ0, i.e. in this case the solution of the necessary condition (A6) is unique up to a multiplication with arbitrary λ0. With this solution, for bp we obtain bp=α1b~1+α2b~2=λ(β22(β11b1+β12b2)β12(β21b1+β22b2))=λ(β11β22β12β21)0b1,i.e. we indeed recover the direction of b1=x3,1n3.

Case 2: In this case, there exists a non-trivial linear combination of the vector fields [v1,[a,v2]] and [v2,[a,v2]] which is contained in Dn3+1, i.e. there exist functions κ1 and κ2 such that κ1[v1,[a,v2]]+κ2[v2,[a,v2]]Dn3+1with at least κ10 or κ20. At least [v1,[a,v2]]Dn3+1 or [v2,[a,v2]]Dn3+1 holds.Footnote15 Therefore, either [v2,[a,v2]]=κ[v1,[a,v2]] mod Dn3+1 or [v1,[a,v2]]=0 mod Dn3+1 and thus (A7) simplifies to either (A8) (α1β12+α2β22)2(α1β11+α2β21)+κ(α1β12+α2β22)[v1,[a,v2]]!Dn3+1(A8) or (A9) (α1β12+α2β22)2[v2,[a,v2]]!Dn3+1.(A9) For (A8), we have the solutions α1=λβ22, α2=λβ12 and α1=λ(2β21+κβ22), α2=λ(2β11+κβ12), both with arbitrary λ0. For (A9), we have the solution α1=λβ22, α2=λβ12, again with arbitrary λ0. Therefore, in any case, with one of the solutions, namely α1=λβ22 and α2=λβ12 with arbitrary λ0, we recover the direction of b1=x3,1n3. For the possibly existing second solution (α1=λ(2β21+κβ22) and α2=λ(2β11+κβ12)), which yields the candidate bp=λ(β11β22β21β12)(κb12b2), the conditions of Theorem 4.1 may or may not be met.

(If we are in Case 1, i.e. if [v1,[a,v2]] and [v2,[a,v2]] are not collinear mod Dn3+1, and thus if (A6) certainly yields only one candidate for bp, can be deduced from the dimension of the distribution H=Dn3+1+[Dn3,Dn3+1]. To be precise, if dim(H)=dim(Dn3+1)+2, [v1,[a,v2]] and [v2,[a,v2]] are not collinear mod Dn3+1 Footnote16. The Case 2 occurs if dim(H)=dim(Dn3+1)+1.)

In conclusion, in any case the necessary condition (A6) yields at most two non-collinear candidates for the vector field bp.