Feedback equivalence of nonlinear control systems - European ...

21 downloads 233041 Views 489KB Size Report
4 Canonical form for single-input systems with controllable linearization. 24. 5 Dual normal ... 13 Analytic normal forms: a class of strict feedforward systems. 88.
Feedback equivalence of nonlinear control systems: a survey on formal approach Witold Respondek Laboratoire de Math´ematiques, INSA de Rouen, Pl. Emile Blondel, 76131 Mont Saint Aignan, France [email protected]

and Issa Amadou Tall Department of Mathematics, Natural Science Division, Tougaloo College, 500 W. County Line Road, MS, 39174, USA [email protected]

Keywords: feedback equivalence, normal form, canonical form, invariants, symmetries, feedforward form AMS subject classification 2000: 93B11, 93B17, 93B27. Abstract This paper is a survey devoted to the formal approach to the feedback equivalence problem for nonlinear control systems. We show how classical Poincar´e’s approach, developed for dynamical systems, generalizes to control systems for continuous and discrete-time. We present normal forms and canonical forms for nonlinear control systems (single-input and multi-input, control-affine and general). We use the formal approach to study symmetries of nonlinear control systems as well as discuss special forms: linear and feedforward. We illustrate presented forms by various examples in dimensions 3 and 4.

1

Contents 1 Introduction

4

2 Equivalence of dynamical systems: Poincar´ e theorem 3 Normal forms for single-input systems with 3.1 Introduction . . . . . . . . . . . . . . . . . . 3.2 Notations . . . . . . . . . . . . . . . . . . . 3.3 Normal form and m-invariants . . . . . . . . 3.4 Normal form for non-affine systems . . . . .

10

controllable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

13 13 14 15 22

4 Canonical form for single-input systems with controllable linearization

24

5 Dual normal form and dual m-invariants

29

6 Dual canonical form

32

7 Normal forms for single-input systems with uncontrollable linearization 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 Taylor series expansions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3 Linear part and resonances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Notations and definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5 Weighted homogeneous systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6 Weighted homogeneous invariants . . . . . . . . . . . . . . . . . . . . . . . . . . 7.7 Explicit normalizing transformations . . . . . . . . . . . . . . . . . . . . . . . . 7.8 Weighted normal form for single-input systems with uncontrollable linearization 7.9 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 Normal forms for multi-input 8.1 Non-affine normal forms . . 8.2 Affine normal forms . . . . . 8.3 Examples . . . . . . . . . .

. . . . . . . . .

34 34 34 36 37 38 42 44 44 46

nonlinear control systems 48 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

9 Feedback linearization

56

10 Normal forms for discrete-time control systems 60 10.1 Example: Bressan and Rampazzo Pendulum . . . . . . . . . . . . . . . . . . . . . 63 11 Symmetries of control systems 11.1 Symmetries . . . . . . . . . . . . . . . . . . . . . . 11.2 Symmetries of single-input nonlinearizable systems 11.3 Symmetries of the canonical form . . . . . . . . . . 11.4 Formal symmetries . . . . . . . . . . . . . . . . . . 11.5 Symmetries of feedback linearizable systems . . . .

2

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

64 65 66 67 68 69

12 Feedforward and strict feedforward forms 12.1 Introduction and notations . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2 Feedforward and strict feedforward normal forms . . . . . . . . . . . . . 12.3 Feedforward and strict feedforward form: first nonlinearizable term . . . 12.4 Feedforward and strict feedforward forms: the general step . . . . . . . . 12.5 Feedforward and strict feedforward systems on R4 . . . . . . . . . . . . . 12.6 Geometric characterization of feedforward and strict feedforward systems 12.7 Symmetries and strict feedforward form . . . . . . . . . . . . . . . . . . . 12.8 Strict feedforward form: affine versus general . . . . . . . . . . . . . . . . 12.9 Strict feedforward systems on the plane . . . . . . . . . . . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

72 73 75 76 78 80 83 83 87 87

13 Analytic normal forms: a class of strict feedforward systems.

88

14 Conclusion

90

3

1

Introduction

In this survey we will be dealing with nonlinear control systems of the form Π : x˙ = F (x, u), where x ∈ X, an open subset of Rn and u ∈ U ⊂ Rm , and F (x, u) is a family of vector fields, C ∞ -smooth with respect to (x, u). The variables x = (x1 , . . . , xn )T represent the state of the system and the variables u = (u1 , . . . , um )T represent the control, that is, an external influence on the system. Π can be understood as un underdetermined system of ordinary differential equations: n equations for n + m variables. We will be interested in equivalence problems for the system Π. Consider another system of the same form ˜ : x˜˙ = F˜ (˜ Π x, u˜), ˜ an open subset of Rn and u˜ ∈ U˜ ⊂ Rm . A natural equivalence can be defined where x˜ ∈ X, ˜ are state-space equivalent, shortly as follows. Assume that U = U˜ . We say that Π and Π S-equivalent, if there exist a diffeomorphism x˜ = φ(x) u˜ = u transforming solutions into solutions. More precisely, if (x(t), u(t)) is a solution of Π, then ˜ which is equivalent to (φ(x(t)), u(t)) is a solution of Π, Dφ(x) · F (x, u) = F˜ (φ(x), u), for any u ∈ U , where Dφ(x) denotes the derivative of φ at x. It means that the state-space equivalence establishes a diffeomorphic correspondence of the right hand sides of differential equations corresponding to the constant controls, which can be expressed as (φ∗ F )(˜ x, u) = F˜ (˜ x, u),

u ∈ U,

where for any vector field f and any diffeomorphism x˜ = φ(x), we denote (φ∗ f )(˜ x) = Dφ(φ−1 (˜ x))· −1 f (φ (˜ x)). S-equivalence is well understood. It establishes a one-to-one smooth correspondence between the trajectories of equivalent systems (corresponding to the same measurable, not necessarily constant, controls). For accessible systems (see, e.g., [36] and [64] for definition) the set of complete invariants for the local S-equivalence is formed by all iterative Lie-brackets evaluated at a nominal point (in the analytic case) or in its neighborhood (in the smooth case), see, e.g., [38]. Since the system Π has state and control variables, another natural transformation is to apply ˜ × U˜ that changes both x and u, that is, to Π a diffeomorphism Υ = (φ, ψ)T of X × U onto X x˜ = φ(x, u) u˜ = ψ(x, u) ˜ Taking a C 1 -solution (x(t), u(t)) of Π and and transforms the solutions of Π into those of Π. ˜ we using the fact that its image (φ(x(t), u(t)), ψ(x(t), u(t)) is assumed to be a solution of Π, 4

conclude that ∂φ F (x, u) + ∂φ u˙ = F˜ (φ(x, u), ψ(x, u)). Now it is easy to see that, since F and ∂x ∂u F˜ do not depend, respectively, on u˙ and u˜˙ , the map φ cannot depend on u. This implies that any Υ preserving the system solutions must actually be a triangular diffeomorphism Υ:

x˜ = φ(x) u˜ = ψ(x, u),

satisfying

Dφ(x) · F (x, u) = F˜ (φ(x), ψ(x, u)), ˜ equivalent via Υ, are called feedback which is called a feedback transformation. Systems Π and Π, equivalent, shortly F-equivalent. The states x and x˜ of two feedback equivalent systems are ˜ while the thus related by a diffeomorphism φ between the corresponding state spaces X and X ˜ controls u and u˜ are related by a diffeomorphism ψ between U and U which depends on the state ˜ locally feedback equivalent at (x0 , u0 ) and (˜ x. We will call Π and Π x0 , u˜0 ), respectively, if (φ, ψ) is a local diffeomorphism satisfying (φ, ψ)(x0 , u0 ) = (˜ x0 , u˜0 ).) It is the feedback equivalence and its local counterpart, which are the main topic of our survey. Notice that the diffeomorphism φ establishes a one-to-one correspondence of x-trajectories of two feedback equivalent systems although equivalent trajectories are differently parameterized by controls. Indeed, a trajectory x(t) of the first system corresponding to a control u(t) is mapped ˜ corresponding to u˜(t) = ψ(x(t), u(t)). Based into the curve φ(x(t)), which is the trajectory of Π ˜ asking that there on this observation, one can define a weaker notion of equivalence of Π and Π exists a one-to-one correspondence between trajectories (corresponding to, say, C ∞ -controls) and omitting the assumption that the correspondence is given by a diffeomorphism. This leads to the important notion of dynamic feedback equivalence (see, e.g., [17], [18], [39], [66]), which we, however, will not study in this survey. The main subject of this paper is feedback equivalence, which has been extensively studied during the last twenty years. Although being natural, this problem is very involved (mainly because of functional parameters appearing in the classification that we will describe briefly in a moment). Many existing results are devoted to systems that are affine with respect to controls, that is, are of the form Σ : x˙ = f (x) +

m X

gi (x)ui = f (x) + g(x)u,

i=1

where x ∈ X, f and gi are C ∞ -smooth control vector fields on X, u = (u1 , . . . , um )T ∈ U = Rm and g = (g1 , . . . , gm ). When studying feedback equivalence of control-affine systems, we will apply feedback transformations that are affine with respect to controls: Γ:

x˜ = φ(x) u = α(x) + β(x)˜ u,

where u = ψ −1 (x, u˜) = α(x) + β(x)˜ u, with α and β being C ∞ -smooth functions with values in Rm and Gl(m, R), respectively. Consider another control-affine system ˜ : x˜˙ = f˜(˜ Σ x) +

m X

g˜i (x)˜ ui = f˜(˜ x) + g˜(˜ x)˜ u,

i=1

5

where u˜ = (˜ u1 , . . . , u˜m )T ∈ U˜ = Rm and g˜ = (˜ g1 , . . . , g˜m ). ˜ are feedback equivalent The general definition implies that the control-affine systems Σ and Σ if and only if φ∗ (f + gα) = f˜ and φ∗ (gβ) = g˜, which we will write shortly

˜ Γ∗ (Σ) = Σ. ˜ are locally feedback equivalent at x0 and x˜0 , We will say that the control-affine systems Σ and Σ respectively, if φ is a local diffeomorphism satisfying φ(x0 ) = x˜0 and α and β are defined locally around x0 . Notice that local feedback equivalence is local in the state-space X but global in the control space U = Rm . ˜ the problem of their (local) feedback equivalence Given two control-affine systems Σ and S, st amounts thus to solving the system of 1 -order partial differential equations

(CDE)

∂φ (x)(f (x) + g(x)α(x)) = f˜(φ(x)) ∂x ∂φ (x)(g(x)β(x)) = g˜(φ(x)). ∂x

It turns out that feedback equivalence of general systems Π under Υ and of control-affine systems Σ under Γ are very closely related. Consider a general nonlinear control system Π : x˙ = F (x, u), where x ∈ X, an open subset of Rn , u ∈ U , an open subset of Rm . Together with Π, we consider its extension (preintegration) Σe : x˙ e = f e (xe ) + g e (xe )ue , where xe = (x, u) ∈ X e = X × U , ue ∈ U e = Rm , and the dynamics are given by x˙ = F (x, u) u˙ = ue , that is f e (xe ) = (F (x, u), 0)T and g e (xe ) = (0, Id)T . Notice that Σe is a control-affine system controlled by the derivatives u˙ i = uei of the original controls ui , for 1 ≤ i ≤ m. ˜ are equivalent (resp. locally equivalent at (x0 , u0 ) Proposition 1.1 Two control systems Π and Π and (˜ x0 , u˜0 )) under a general feedback transformation Υ if and only if their respective extensions e ˜ e are equivalent (resp. locally equivalent at xe0 = (x0 , u0 ) and x˜e0 = (˜ x0 , u˜0 )) under an Σ and Σ affine feedback Γ. As a consequence, many problems concerning feedback equivalence are studied and solved for control-affine systems and their extension to the general case can be done by an appopriate application of Proposition 1.1. In order to geometrize the problem of feedback equivalence, we associate to the system Π its field of admissible velocities F(x) = {F (x, u) : u ∈ U } ⊂ Tx X. 6

The field of admissible velocities of the control-affine system Σ is the following field of affine subspaces (equivalently, an affine distribution): A(x) = {f (x) +

m X

gi (x)ui : ui ∈ R} = f (x) + G(x) ⊂ Tx X,

i=1

where G denotes the distribution spanned by the vector fields g1 , . . . , gm . Now it is easy to see ˜ are feedback equivalent, then the corresponding affine that if two control affine-systems Σ and Σ distributions are equivalent, that is, ˜ φ∗ A = A. Moreover, the converse holds if the distributions G and G˜ are of constant ranks. Analogous implications (the converse under the constant rank assumption) are true for local feedback equivalence. Notice that attaching the field of admissible velocities to an affine system results in eliminating controls from the description: what remains is a geometric object, which is the affine distribution A while the choice of controls (equivalently, the choice of sections of A) becomes irrelevant. Example 1.2 To illustrate the notion of feedback equivalence, let us recall the first (historically) studied feedback classification problem, which is that for linear control systems of the form Λ : x˙ = Ax + Bu = Ax +

m X

ui b i ,

i=1

where x ∈ Rn , Ax and b1 , . . . , bm are, respectively, linear and constant vector fields on Rn , and u = (u1 , . . . , um )T ∈ Rm . In order to preserve the linear form of the system, we apply to it the linear feedback transformation. x˜ = T x u = Kx + L˜ u, where T , K, and L are matrices of appropriate sizes, T and L being invertible. The system Λ is transformed into ˜ : x˜˙ = A˜ ˜x + B ˜ u˜ = T (A + BK)T −1 x˜ + T BL˜ Λ u. It is a classical result of the linear control theory (see, e.g., [47]) that any linear controllable system is feedback equivalent to the following system (called Brunovsk´ y canonical form): x˜˙ i,j = x˜i,j+1 ˙x˜i,ρi = u˜i ,

1 ≤ j ≤ ρi − 1 1 ≤ i ≤ m,

where x˜ = (˜ x1,1 , x˜1,2 , . . . , x˜1,ρ1 , . . . ,P x˜m,ρm )T . m y or Kronecker The integers ρ1 ≥ · · · ≥ ρm , i=1 ρi = n (called controllability, Brunovsk´ indices) form a set of complete feedback invariants of the linear feedback linear group action on controllable systems and are defined as follows ρi = max{rj | rj ≥ i},

(1.1)

where d0 = 0, di = rank (B, . . . , Ai−1 B) and ri = di − di−1 for 1 ≤ i ≤ n. Notice that for the linear control system Λ, the field of admissible velocities is given by the field of m-dimensional affine subspaces A(x) = Ax + B of Rn , where B is the image of Rm under the linear map B : Rm → Rn . The Brunovsk´ y canonical form thus gives a canonical form for the field A under linear invertible transformations x˜ = T x. ¤ 7

Observe that the dimension of the space of linear systems (of pairs (A, B)) is n2 + nm and the dimension of the group of linear feedback (of the triples (T, K, L)) is n2 + nm + m2 . We can thus expect open orbits to exist and, indeed, open orbits exist (those of systems with the maximal vector of di ’s). The picture gets completely different for nonlinear control systems under the action of nonlinear feedback. Although both are infinite dimensional, the group of (local) feedback transformations is much ”smaller” than the space of all (local) control systems and, as a consequence, functional parameters must necessarily appear in the feedback classification. To see this, notice that the space of general systems Π is parameterized by n functions of n + m variables (components of F ) while the group of feedback transformations by m functions of n + m variables (components of ψ) and n functions of n variables (components of φ). Thus functional parameters are to be expected if m < n, that is, in all interesting cases. To make this argument precise, we will follow [38] and compute the dimension dΠ (k) of the space of k-jets of the system Π and the dimension dΥ (k) of the corresponding jet-space of the feedback group acting on the space of k-jets of the systems. To this end, recall that the dimension of the space of polynomials of n variables, of degree not greater than k, is equal to (k + n)! (k + 1)(k + 2) · · · (k + n) = , k!n! n! which is a polynomial of k of degree n starting with k n /n!. The dynamics of the system are represented by n components of F (x, u), each being a function of n + m variables. The feedback group is represented by n components of the diffeomorphism φ(x), each being a function of n variables and m components of the map ψ(x, u), each being a function of n + m variables. Notice that (k + 1)-jet of a diffeomorphism act on the k-jet of the system. Thus dΠ (k) = n

(k + n + m)! , k!(n + m)!

dΥ (k) = n

(k + 1 + n)! (k + n + m)! +m . (k + 1)!n! k!(n + m)!

The codimension of any orbit, of the feedback group action on the space of systems, is bounded from bellow by the difference dΠ (k) − dΥ (k), which is a polynomial of k of degree n + m, whose coefficient multiplying k n+m is n−m . (n + m)! This coefficient is positive when m < n and thus the polynomial and the codimension of any orbit tend to infinity when k tends to infinity. As a consequence, functional moduli must appear in the feedback classification if m < n (which exhausts all interesting cases). Observe that a control-affine system is defined by m + 1 vector fields f, g1 , . . . , gm and the feedback group by the diffeomorphism φ and m + m2 components of the pair (α, β). Therefore in the case of control-affine systems the corresponding dimensions are: dΣ = n(m + 1)

(k + n)! , k!(n)!

dΓ = n

(k + 1 + n)! (k + n)! + m(m + 1) . (k + 1)!n! k!(n)!

The codimension of any orbit in the k-jets space is bounded below by the difference dΣ − dΓ , which is a polynomial of k of degree n, whose coefficient multiplying k n is m(n − m − 1) . n! 8

If m < n − 1, then this coefficient is positive and thus the polynomial and the codimension of any orbit under the feedback group action tend to infinity as k tends to infinity. As a consequence, functional moduli must appear in feedback classification of control-affine systems if m < n−1. In the case m = n − 1 we can hope, however, for normal forms without functional parameters and, indeed, such normal forms have been obtained by Respondek and Zhitomirskii for m = 2, n = 3 in [73] and in the general case in [94]. It is the existence of functional moduli, which causes one of the main difficulties of the feedback equivalence problem. Four basic methods have been proposed to study various aspects of feedback equivalence. The first method, used for control-affine systems, is based on studying invariant properties of two geometric objects attached to the system: the distribution G and the affine distribution A. Notice that feedback equivalence of control-linear systems (that is control-affine system Σ with f ≡ 0) coincides with equivalence, under a diffeomorphism, of the ˜ So this approach is linked, in a natural way, with the corresponding distributions G and G. classification and with singularities of vector fields and distributions, and their invariants. Using that method a large variety of feedback classification problems have been solved, see e.g. [7], [12], [35], [38], [43], [44], [54], [67], [73], [94]. The second approach, proposed by Gardner [19], uses Cartan’s method of equivalence [11]. To the control system Π, we can associate the Pfaffian system given by the differential forms ˜ is analyzed dxi −Fi (x, u)dt, for 1 ≤ i ≤ n, on X ×U ×R and the feedback equivalence of Π and Π by studying the equivalence of the corresponding Pfaffian systems and their geometry, see [23], [21], [22], [60]. The third method, inspired by the hamiltonian formalism for optimal control problems, has been developed by Bonnard and Jakubczyk [6], [7], [42], [40] and has led to a very nice description of feedback invariants in terms of singular extremals. Another approach based also on the hamiltonian formalism for optimal control has been proposed by Agrachev [?] and has led to a construction of a fundamental geometric invariant of feedback equivalence: the curvature of control systems. Finally, a very fruitful approach was proposed by Kang and Krener [52] and then followed by Kang [48], [49]. Their idea, which is closely related with classical Poincar´e’s technique for linearization of dynamical systems (see e.g. [1]), is to analyze the system Π and the feedback transformation Υ (the system Σ and the transformation Γ, respectively, in the control-affine ˜ also step case) step by step and, as a consequence, to produce a simpler equivalent system Π by step. It is this approach, and various classification results obtained using it, which form the subject of this survey. This survey is organized as follows. We will present in Section 2 classical Poincar´e’s approach to the problem of formal equivalence of dynamical systems. In Section 3 we will generalize, following Kang and Krener, that formal approach to nonlinear control systems. We will present a normal form for homogeneous systems, their invariants, explicit normalizing transformations and, finally, a normal form under a formal feedback. We will also extend the normal form to general (non-affine systems). Then, in Section 4, we will propose a canonical form for nonlinear control systems. In the two following Sections 5 and 6 we will dualize results of preceding sections and give a dual normal form (together with dual invariants and explicit normalizing transformations) and a dual canonical form. Then in Section 7 we will pass to systems with uncontrollable linearization, introduce weighted homogeneity and we will give: a normal form, invariants, explicit normalizing transformations, and a formal normal form. This section gen9

eralizes, from the one hand side, results for systems with controllable linearization (presented in earlier sections) and, on the other hand side, results on dynamical systems from Section 2. Section 8 will be devoted to multi-input normal forms (for space related reasons we just treat the controllable case): it generalizes results on normal forms obtained in Section 3. A discrete-time version of Section 3 will be given in Section 10. In Section 9 we compare well known results devoted to feedback linearization with their counterpart obtained via the formal feedback. We will also discuss systems that are feedback equivalent to linear uncontrollable systems. Then the two following Sections present applications of the formal approach to the classification of control systems. We discuss symmetries of control systems in Section 11 and show an enormous difference between the group symmetries of feedback linearizable and nonlinearizable systems. In Section 12 we characterize, using the formal approach, systems that are feedback equivalent to feedforward and strict feedforward forms. Finally, in Section 13 we give a class of analytic strict feedforward forms than can be transformed to a normal form via constructive analytic transformations. Because of space limits, this survey does not touch many important results. To mention just a few: we do not discuss analysis of bifurcations based on formal approach (e.g., [50], [51], [57], [58]), bifurcations of discrete-time systems, normal forms for observed dynamics. Each of those subjects requires its own survey proving efficiency of the formal approach.

2

Equivalence of dynamical systems: Poincar´ e theorem

In this section we will summarize very briefly Poincar´e’s approach to the problem of (formal) equivalence of dynamical systems. The goal of this section is three-folds. Firstly, to make our survey complete and self-contained. Secondly, to show how the formal approach to the equivalence of dynamical systems generalizes to the formal approach to feedback equivalence of control systems. Thirdly, some of results on formal normal forms for dynamical systems and of formal linearization (Theorems 2.3 and 2.4 stated at the end of this section) will be used in Sections 7 and 9 of the survey. Consider the uncontrolled dynamical system x˙ = f (x), where x ∈ X, an open subset of Rn and f is a C ∞ -smooth vector field on X. A C ∞ -smooth diffeomorphism x˜ = φ(x) brings the considered dynamical system into x˜˙ = f˜(˜ x) = (φ∗ f )(˜ x), where

∂φ −1 (φ (˜ x)) · f (φ−1 (˜ x)). ∂x Now given two dynamical systems x˙ = f (x) and x˜˙ = f˜(˜ x), the problem of establishing their equivalence is to find a diffeomorphism x˜ = φ(x) satisfying (φ∗ f )(˜ x) =

(DE)

∂φ (x)f (x) = f˜(φ(x)), ∂x 10

which is a system of n first order partial differential equations for the components of φ(x). Notice that in the most interesting case of f (x0 ) = f˜(˜ x0 ) = 0, this is a system of singular partial differential equations. Consider the infinite Taylor series expansion of our dynamical system x˙ = f (x) = Jx +

∞ X

f [m] (x)

m=2

around an equilibrium, which is assumed to be x0 = 0 ∈ Rn , where f [m] denotes a polynomial vector field, whose all components are homogenous polynomials of degree m. Apply to it a formal change of coordinates given by an invertible formal transformation of the form x˜ = φ(x) = x +

∞ X

φ[m] (x),

m=2

which preserves 0 ∈ Rn and starts with the identity, where all components of φ[m] are homogeneous polynomials of degree m. In order to study the action of φ(x) on f (x), we will see how its homogenous part of degree m acts on terms of degree m of f . To this end, apply to x˙ = Jx + f [m] (x) the transformation x˜ = x + φ[m] (x), where m ≥ 2. We have, modulo terms of higher degree, ∂φ[m] x˜˙ = Jx + f [m] (x) + (x)(Jx + f [m] (x)) ∂x ∂φ[m] = J x˜ − Jφ[m] (x) + f [m] (x) + (x)Jx ∂x = J x˜ + f [m] (x) + [Jx, φ[m] (x)] = J x˜ + f˜[m] (˜ x), ∂v where [v, w](x) = ∂w (x)v(x) − ∂x w(x) is the Lie bracket of two vector fields v and w. Using the ∂x notation adv w = [v, w], we obtain

(HE)

adJx φ[m] (x) = f˜[m] (x) − f [m] (x),

which we will call a homological equation. Consider the action of adJx on the space P [m] of polynomial vector fields whose all components are homogeneous polynomials of degree m. For a multi-index k = (k1 , . . . , kn ), denote xk = xk11 · · · xknn . Lemma 2.1 Assume that J is diagonal, say J = diag(λ1 , . . . , λn ). Then adJx is a diagonal operator on the space P [m] in the eigenbasis formed by the eigenvectors xk ∂x∂ i , for all multiindices k such that k1 + · · · + kn = m and 1 ≤ i ≤ n. The eigenvalues of adJx depend linearly on the eigenvalues of J, more precisely, we have adJx (xk

∂ ∂ ) = (k, λ)(xk ), ∂xi ∂xi

where λ = (λ1 , . . . , λn ), and (k, λ) = k1 λ1 + · · · + kn λn . 11

Corollary 2.2 The operator adJx is invertible on the space P [m] if there does not hold any relation of the form n X ks λs = λj , s=1

where ks are nonnegative integers, |k| = k1 + · · · + kn ≥ 2 and 1 ≤ j ≤ n. P For any relation λj = ns=1 ks λs , called resonance, we define Rj = {k = (k1 , . . . , κn ) : λj = k1 λ1 + · · · + kn λn , ki ∈ N ∪ {0}, |k| ≥ 2}, which will be called the resonant set associated with λj . Theorem 2.3 Consider the differential equation ∞ X

x˙ = f (x) = Jx +

f [m] (x).

m=2

and assume that all eigenvalues are real and distinct, and that the spectrum of J is nonresonant. (i) For each m ≥ 2 and any homogenous vector fields f [m] and f˜[m] of degree m, the homological equation (HE) is solvable within the class of Rn -valued polynomials φ[m] of degree m. ∞ ∞ P P f [m] (x) and x˜˙ = f˜(˜ x) = J x˜ + f˜[m] (˜ x) (ii) The differential equations x˙ = f (x) = Jx + m=2

are equivalent via an invertible formal transformation of the form x˜ = x + (iii) The differential equation x˙ = f (x) = Jx +

∞ P

∞ P

m=2

φ[m] (x).

m=2

f [m] (x) can be formally linearized, that is,

m=2

can be brought to the form x˜˙ = J x˜ via an invertible formal transformation of the form ∞ P φ[m] (x). x˜ = x + m=2

Item (i) is a direct consequence of Corollary 2.2. Item (ii) follows by a successive application of (i) for m = 2, 3 and so on. Finally, (iii) is an immediate consequence of (ii), applied for f˜ = J x˜. If the spectrum of J is resonant, then using the adJx -operator we can get rid of all nonresonant terms, which leads to the following: Theorem 2.4 Consider the differential equation x˙ = f (x) = Jx +

∞ X

f [m] (x).

m=2

Assume that J is diagonal, that is, J = diag(λ1 , . . . , λn ). There exits a formal invertible trans∞ P formation of the form x˜ = x + φ[m] (x) bringing x˙ = f (x) into x˜˙ = f˜(˜ x) of the form m=2

f˜j (˜ x) = λj x˜j +

X

γjk x˜k11 · · · x˜knn ,

k∈Rj

where γjk ∈ R and the summation is taken over all resonances (k1 , . . . , kn ) forming the resonant set Rj associated with λj . 12

If the eigenvalues of J are distinct but not necessarily real, then an analogous results holds; we will state it in Section 7. Theorems 2.3 and 2.4 summarize Poincar´e’s approach in the formal category. The idea of this approach is very natural: in order to establish the equivalence of two dynamical systems, we replace the singular partial differential equation (DE) by an infinite sequence of homological equations (HE), which are simply linear equations with respect to the unknown components of the homogenous part φ[m] of φ. A much more delicate and difficult issues of constructing C ∞ -smooth or real analytic transformations that linearize the equation (in the non-resonant case) or annihilate all nonresonant terms (in the general case) are discussed very briefly in Section 9.

3 3.1

Normal forms for single-input systems with controllable linearization Introduction

In this section we will be studying nonlinear single-input control-affine systems of the form Σ : ξ˙ = f (ξ) + g(ξ)u, where ξ ∈ X, an open subset of Rn , u ∈ R, and f and g are C ∞ -smooth vector fields on X. Throughout this section we will study the system Σ around a point ξ0 at which f (ξ0 ) = 0 and g(ξ0 ) 6= 0. Without loss of generality, we will assume that ξ0 = 0. We will also assume throughout this section that the linear part (F, G) of the system is controllable, where F = ∂f (0) ∂ξ and G = g(0). The goal of this section is to obtain a normal form of Σ under the action of the feedback group consisting of feedback transformation of the form Γ :

x = φ(ξ) u = α(ξ) + β(ξ)v.

Together with the system Σ and the feedback transformation Γ, we will consider their Taylor series expansions Σ∞ and Γ∞ , respectively, and we will study the action of Γ∞ on Σ∞ step-by-step, that is, the action of the homogeneous part Γm of Γ∞ on the homogeneous part Σ[m] of Σ∞ . In other words, we will generalize to control systems the approach that Poincar´e has developed for dynamical systems (and which we recalled in Section 2). It was Kang and Krener [52], [48], [49] who proposed that approach in the context of control systems and who have obtained fundamental results. Their pioneering work has inspired the authors who have obtained further results and all of them form now-a-days a relatively complete theory of formal feedback classification of nonlinear control systems. The first results of Kang and Krener were devoted to obtaining a normal form for single-input control-affine systems with controllable linear approximation and we will also start our systematic presentation in this section by discussing that case. A generalization to non-affine systems will be given at the end of this section while further developments (uncontrollable linear approximation and the problem of canonical forms) will be discussed in next sections. 13

The section is organized as follows. In Section 3.2 we will say a few words about the notation used in the whole section as well as in Sections 4, 5, and 6. Main results are given in Section 3.3: a normal forms for homogeneous systems, explicit transformations bringing to it, m-invariants, and normal form under formal feedback. Finally, in Section 3.4 we will generalize the normal form to non-affine systems.

3.2

Notations

We will denote by P [m] (ξ) the space of homogeneous polynomials of degree m of the variables ξ1 , . . . , ξn , by P ≤m (ξ) the space of polynomials of degree m of the variables ξ1 , . . . , ξn , and by P ≥m (ξ) the space of formal power series of the variables ξ1 , . . . , ξn starting from terms of degree m. Analogously, we will denote by V [m] (ξ) the space of homogeneous vector fields whose components are in P [m] (ξ), by V ≤m the space of polynomial vector fields whose components are in P ≤m (ξ), and by V ≥m (ξ) the space of vector fields formal power series whose components are in P ≥m (ξ). Similar notations P [m] (ξ, u), V [m] (ξ, u) etc. will also appear and they stand, respectively, for homogeneous polynomials and homogeneous polynomial vector fields depending on the state variables ξ = (ξ1 , . . . , ξn ) and control variable u, with homogeneity being understood with respect to the all variables (ξ, u). Because of various normal forms and various transformations that are used throughout the paper, we will keep the following notation. Together with Σ, we will also consider its Taylor series expansion Σ∞ and its homogeneous part Σ[m] of degree m given, respectively, by the following systems Σ[m] : ξ˙ = Aξ + Bu + f [m] (ξ) + g [m−1] (ξ)u, ∞ ¡ ¢ P Σ∞ : ξ˙ = Aξ + Bu + f [k] (ξ) + g [k−1] (ξ)u . k=2

The systems Σ, Σ[m] , and Σ∞ will stand for the systems under consideration. Their state vector will be denoted by ξ and their control by u (x and v being reserved, respectively, for the state and control of various normal forms). The system Σ[m] (resp. the system Σ∞ ) transformed via ˜ [m] (resp. by Σ ˜ ∞ ). Its state vector will be denoted by x, its control feedback will be denoted by Σ by v, and the vector fields, defining its dynamics, by f˜[k] and g˜[k−1] . Feedback equivalence of ˜ [m] will be established via a smooth feedback, that is precisely, homogeneous systems Σ[m] and Σ via a homogeneous feedback Γm . On the other hand, feedback equivalence of systems Σ∞ and ˜ ∞ will be established via a formal feedback Γ∞ . Σ We will introduce two kinds of normal forms, Kang normal forms and dual normal forms (Sections 3 and 5), as well as canonical form and dual canonical form (Sections 4 and 6). The [m] symbol “bar” will correspond to the vector field f¯[m] defining the Kang normal forms ΣN F and ∞ Σ∞ ¯[m−1] defining the dual normal N F and the canonical form ΣCF as well as to the vector field g [m] ∞ forms ΣDN F and Σ∞ DN F and the dual canonical form ΣDCF . Analogously, the m-invariants (resp. [m−1] dual m-invariants) of the system Σ[m] will be denoted by a[m]j,i+2 (resp. by bj ) and the m[m] [m] invariants (resp. dual m-invariants) of the normal form ΣN F (resp. dual normal form ΣDN F ) by [m−1] a ¯[m]j,i+2 (resp. by ¯bj ). Other normal forms will appear in Section 12.

14

3.3

Normal form and m-invariants

All objects, that is, functions, maps, vector fields, control systems, etc., are considered in a neighborhood of 0 ∈ Rn and assumed to be C ∞ -smooth. Let h be a smooth R-valued function. By ∞ X [0] [1] [2] h(ξ) = h (ξ) + h (ξ) + h (ξ) + · · · = h[m] (ξ) m=0

we denote its Taylor series expansion at 0 ∈ Rn , where h[m] (ξ) stands for a homogeneous polynomial of degree m. Similarly, for a map φ of an open subset of Rn to Rn (resp. for a vector field f on an open subset of Rn ) we will denote by φ[m] (resp. by f [m] ) the homogeneous term of degree m of its [m] [m] Taylor series expansion at 0 ∈ Rn , that is, each component φj of φ[m] (resp. fj of f [m] ) is a homogeneous polynomial of degree m in ξ. Consider the Taylor series expansion of the system Σ given by ∞

Σ

: ξ˙ = F ξ + Gu +

∞ X ¡ [m] ¢ f (ξ) + g [m−1] (ξ)u ,

(3.1)

m=2

(0) and G = g(0). Recall that we assume in this section that f (0) = 0 and where F = ∂f ∂ξ g(0) 6= 0. Consider also the Taylor series expansion Γ∞ of the feedback transformation Γ given by x = Tξ +

∞ P

φ[m] (ξ)

m=2

Γ∞ :

u = Kξ + Lv +

∞ ¡ P

α

[m]

(ξ) + β

[m−1]

¢

(3.2)

(ξ)v ,

m=2

where T is an invertible matrix and L 6= 0. Analogously to the Poincar´e’s approach presented in Section 2, we analyze the action of Γ∞ on the system Σ∞ step by step. To start with, consider the linear system ξ˙ = F ξ + Gu. Throughout the section we will assume that it is controllable. It can be thus transformed by a linear feedback transformation of the form Γ1 :

x = Tξ u = Kξ + Lv

into the Brunovsk´ y canonical form (A, B), see  0 1 ···  ..  . A= 0 0 · · · 0 0 ···

e.g. [47] and Example 1.2 in Section 1:    0 0  ..       , B = . . 0 1 1 0

15

Assuming that the linear part (F, G), of the system Σ∞ given by (3.1), has been transformed to the Brunovsk´ y canonical form (A, B), we follow an idea of Kang and Krener [52], [48] and apply successively a series of transformations Γm :

x = ξ + φ[m] (ξ) u = v + α[m] (ξ) + β [m−1] (ξ)v,

(3.3)

for m = 2, 3, . . . . A feedback transformation defined as an infinite series of successive compositions of Γm , m = 1, 2, . . . is also denoted by Γ∞ (that is, Γ∞ = · · · Γm ◦ Γm−1 ◦ · · · ◦ Γ1 ) because, as a formal power series, it is of the form (3.2). We will not address the problem of convergence in general (see Sections 9 and 13 for some comments on this issue, and for a convergent class of analytic systems) and we will call such a series of successive compositions a formal feedback transformation. Observe that each transformation Γm , for m ≥ 2, leaves invariant all homogeneous terms of degree smaller than m of the system Σ∞ and we will call Γm a homogeneous feedback transformation of degree m. We will study the action of Γm on the following homogeneous system Σ[m] : ξ˙ = Aξ + Bu + f [m] (ξ) + g [m−1] (ξ)u.

(3.4)

˜ [m] given by Consider another homogeneous system Σ ˜ [m] : x˙ = Ax + Bv + f˜[m] (x) + g˜[m−1] (x)v. Σ

(3.5)

We will say that the homogeneous system Σ[m] is feedback equivalent to the homogenous system ˜ [m] if there exists a homogeneous feedback transformation of the form (3.3), which brings Σ[m] Σ ˜ [m] modulo terms in V ≥m+1 (x, v). into Σ The starting point for formal classification of single-input control systems is the following result, proved by Kang [48]. Proposition 3.1 The homogeneous feedback transformation Γm , defined by (3.3), brings the ˜ [m] , given by (3.5), if and only if the following relations system Σ[m] , given by (3.4), into Σ  [m] [m] [m] [m]  LAξ φj − φj+1 (ξ) = f˜j (ξ) − fj (ξ)        [m] [m−1] [m−1]  LB φj (ξ) = g˜j (ξ) − gj (ξ)  (3.6)  [m] [m] [m]  [m] ˜  LAξ φn + α (ξ) = fn (ξ) − fn (ξ)        [m] [m−1] [m−1] LB φn (ξ) + β [m−1] (ξ) = g˜n (ξ) − gn (ξ). [m]

hold for any 1 ≤ j ≤ n − 1, where φj

are the components of φ[m] .

This proposition represents the essence of the method developed by Kang and Krener and used for many results in this survey paper. The problem of studying the feedback equivalence of ˜ requires, in general, solving the system (CDE) of 1-st order two control-affine systems Σ and Σ partial differential equations (as we have already explained in Section 1). On the other hand, if we perform the analysis step by step, then the problem of establishing the feedback equivalence 16

˜ [m] reduces to solving the algebraic system (3.6), called sometimes of two systems Σ[m] and Σ the control homological equation by its analogy with Poincar´e’s homological equation (HE) of Section 2. Indeed, (3.6) can be re-written in the following compact from adAξ φ[m] (ξ) = f˜[m] (ξ) − f (ξ) − Bα[m] (ξ) (CHE) [m]

adB (ξ) = g˜[m−1] (ξ) − g [m−1] (ξ) − Bβ [m−1] (ξ), which reduces to (HE) if the control vector field B + g [m−1] (ξ) is not present, with A playing the role of J. Therefore for control systems, solving the differential equation (CDE) is replaced by an infinite sequence of algebraic homological equations (CHE) exactly like for dynamical systems, where the differential equation (CDE) is replaced by an infinite sequence of homological equations (HE) (compare Section 2). Using the above proposition, Kang [48] proved the following result: Theorem 3.2 The homogeneous system Σ[m] can be transformed, via a homogeneous feedback transformation Γm , into the following normal form  n P [m−2]  x2i P1,i (x1 , . . . , xi ) x ˙ = x +  1 2   i=3   ..    .   n  P  [m−2]  x2i Pj,i (x1 , . . . , xi ) x˙ j = xj+1 +    i=j+2  [m] . . (3.7) ΣN F : .   [m−2]  2  x˙ n−2 = xn−1 + xn Pn−2,n (x1 , . . . , xn )         x˙ n−1 = xn       x˙ n = v , [m−2]

where Pj,i variables.

(x1 , . . . , xi ) are homogeneous polynomials of degree m−2 depending on the indicated

In order to illustrate this result, consider the case m = 2, which actually was for Kang and Krener [52] the starting point for the formal approach to feedback equivalence. Applying the above theorem to m = 2 yields that the homogeneous system Σ[2] can be transformed, via a homogeneous feedback transformation Γ2 , into the following normal form  x˙ 1 = x2 + a1,3 x23 + a1,4 x24 + · · · + a1,n x2n          x˙ 2 = x3 + a2,4 x24 + · · · + a2,n x2n        .. [2] . ΣN F :  x˙ n−2 = xn−1 + an−2,n x2n         x˙ n−1 = xn        x˙ = u, n 17

[m]

where aj,i ∈ R. Notice that the general normal form ΣN F exhibits the same triangular structure [2] as ΣN F , the only difference being the replacement of the constants aj,i by the polynomials [m−2] Pj,i (x1 , . . . , xi ). Now we will make a few comments on the number of constants aj,i (for m = 2) and that of the polynomials (in the general case) present in the normal forms. Compare the analysis given below (performed for the homogeneous system Σ[m] ) with a similar analysis given for general multi-input systems Π and control-affine systems Σ in Section 1. Recall that the dimension of the space of polynomials P [m] of degree m of n variables and of the space V [m] of polynomial vector fields on Rn , whose all components belong to P [m] , are, respectively (n + m − 1)! (n + m − 1)! and n . m!(n − 1)! m!(n − 1)! Homogeneous systems Σ[m] are given by two vector fields f [m] ∈ V [m] and g [m−1] ∈ V [m−1] . So the dimension of the space of single-input systems, homogenous of degree m, is dΣ[m] = n

(n + m − 1)! (n + m − 2)! +n . m!(n − 1)! (m − 1)!(n − 1)!

The feedback group Γm is given by n components of the diffeomorphism φ[m] , each in P [m] , and two functions α[m] ∈ P [m] and β [m−1] ∈ P [m−1] . Hence the dimension of Γm is dΓm = n

(n + m − 2)! (n + m − 1)! (n + m − 1)! + + . m!(n − 1)! m!(n − 1)! (m − 1)!(n − 1)!

Both dimensions are polynomials of degree n−1 of m and their difference is thus also a polynomial of degree n − 1 of m starting with dΣ[m] − dΓm =

n − 2 n−1 m + ··· , (n − 1)!

where dots stand for lower order terms. Now observe that the dimension of the space of n − 2 functions, each belonging to P [m] , is also a polynomial of degree n − 1 of m starting with (n − 2)

(n + m − 1)! n − 2 n−1 = m + ··· , m!(n − 1)! (n − 1)! [m]

which explains why in the normal form ΣN F we have n − 2 polynomials of n variables. Since dΣ[m] −dΓm < (n−2) (n+m−1)! , it follows that polynomials of fewer variables show up in the normal m!(n−1)! [m]

form ΣN F . Analogous argument applied to m tending to infinity explains the appearance of n−2 functions of n variables in the normal form Σ∞ N F (see Theorem 3.6 below). In order to calculate the exact number of invariants in the form Σ[m] (which is bounded from below by dΣ[m] − dΓm ), we have to study the action of Γm on the space of homogeneous systems of degree m. This action is not free, the isotropy group being of dimension 1 (see [48], [81] and Proposition 4.4 for a detailed calculation). This can be illustrated by the homogeneous system Σ[2] of degree 2, + nn (we have n components of f [2] and n components of g [1] ) and for which dΣ[2] = n n(n+1) 2 dΓ2 = n n(n+1) + n(n+1) + n (we have n components of φ[2] and the function α[2] and β [1] ). It 2 2 18

2

follows that dΣ[2] −dΓ2 = n −3n while the number of parameters aj,i (which is actually the number 2 2 [2] of invariants of Σ , see the next section) is (n−1)(n−2) . The difference (n−1)(n−2) − n −3n = 1 is 2 2 2 2 actually the dimension of the isotropy subgroup of Γ , which is the dimension of the group of symmetries of any Σ[2] (see Section 11). [m]

The two following questions concerning the normal form ΣN F are important and arise naturally. [m−2] (Q1) Are the polynomials Pj,i invariant, that is, unique under feedback Γm ? [m] (Q2) How to bring a given system Σ[m] into its normal form ΣN F ? The answer to question (Q1) is positive, and to construct invariants under homogeneous feedback transformations, let us define the vector fields Xim−1 (ξ) = (−1)i adiAξ+f [m] (ξ) (B + g [m−1] (ξ)) [m−1]

and let Xi the subspace

be its homogeneous part of degree m − 1. By πi we will denote the projection on © ª Wi = ξ = (ξ1 , . . . , ξn )T ∈ Rn : ξi+1 = · · · = ξn = 0

that is πi (ξ) = (ξ1 , . . . , ξi , 0, . . . , 0). Following Kang [48], we denote by a[m]j,i+2 (ξ) the homogeneous part of degree m − 2 of the polynomials ³ ´ £ m−1 m−1 ¤ [m−1] j−1 j−1 Ai+1 B [m−1] CA Xi , Xi+1 (πn−i (ξ)) = CA adAi B Xi+1 − ad Xi (πn−i (ξ)), where C = (1, 0, . . . , 0)T ∈ Rn , and (j, i) ∈ ∆ ⊂ N × N, defined by ∆ = {(j, i) ∈ N × N : 1 ≤ j ≤ n − 2, 0 ≤ i ≤ n − j − 2} . The homogeneous polynomials a[m]j,i+2 , for (j, i) ∈ ∆, will be called m-invariants of Σ[m] , under the action of Γm . The following result of Kang [48] asserts that m-invariants a[m]j,i+2 , for (j, i) ∈ ∆, are complete invariants of homogeneous feedback and, moreover, illustrates their meaning for the homogeneous [m] normal form ΣN F . ˜ [m] and let Consider two homogeneous systems Σ[m] and Σ { a[m]j,i+2 : (j, i) ∈ ∆ },

and

{a ˜[m]j,i+2 : (j, i) ∈ ∆ }

denote, respectively, their m-invariants. The following result was proved by Kang [48]: Theorem 3.3 The m-invariants have the following properties: ˜ [m] are equivalent via a homogeneous feedback trans(i) Two homogeneous systems Σ[m] and Σ m formation Γ if and only if a[m]j,i+2 = a ˜[m]j,i+2 ,

19

for any (j, i) ∈ ∆.

[m]

(ii) The m-invariants a ¯[m]j,i+2 of the normal form ΣN F , defined by (3.7), are given by a ¯[m]j,i+2 (x) =

∂2 2 [m−2] xn−i Pj,n−i (x1 , . . . , xn−i ), 2 ∂xn−i

for any (j, i) ∈ ∆.

(3.8)

In order to answer to the second question (Q2) we will construct an explicit feedback trans[m] formation that brings the homogeneous system Σ[m] to its normal form ΣN F . Define the homo[m−1] [m−1] [m−1] geneous polynomials ψj,i (ξ) by setting ψj,0 (ξ) = ψ1,1 (ξ) = 0, Ã ! n−i X [m−1] [m−1] [m] ψj,i (ξ) = −CAj−1 adn−i + (−1)t adt−1 , Aξ g Aξ adAn−i−t B f t=1

if 1 ≤ j < i ≤ n and [m−1]

ψj,i

[m]

[m−1]

(ξ) = LAn−i B fj−1 (πi (ξ)) + LAξ ψj−1,i (πi (ξ)) Z [m−1] +ψj−1,i−1 (πi−1 (ξ))

ξi

+ 0

(3.9) [m−1] LAn−i+1 B ψj−1,i (πi (ξ))dξi ,

[m−1] ψj,i (πi (ξ)) [m]

if 1 ≤ i ≤ j, where [m] components φj of φ

[m−1]

is the restriction of ψj,i (ξ) to the subspace Wi . Define the , for 1 ≤ j ≤ n, and the feedback (α[m] , β [m−1] ) by n R P ξi

[m]

φj (ξ) =

0

i=1

[m−1]

ψj,i

(πi (ξ))dξi ,

[m]

[m]

1 ≤ j ≤ n − 1,

[m]

φn (ξ) = fn−1 (ξ) + LAξ φn−1 (ξ), α

[m]

³ (ξ) = −

[m] fn (ξ)

(3.10)

´

+

[m] LAξ φn (ξ)

,

´ ³ [m] [m−1] (ξ) + LB φn (ξ) . β [m−1] (ξ) = − gn We have the following result of the authors [81]: Theorem 3.4 The homogeneous feedback transformation Γm :

x = ξ + φ[m] (ξ) u = v + α[m] (ξ) + β [m−1] (ξ)v, [m]

where α[m] , β [m−1] , and the components φj of φ[m] are defined by (3.10), brings the homogeneous [m] system Σ[m] into its normal form ΣN F given by (3.7). Example 3.5 To illustrate results of this section, we consider the system Σ[m] , given by (3.4) on R3 . Theorem 3.2 implies that the system Σ[m] is equivalent, via a homogeneous feedback [m] transformation Γm defined by (3.10), to its normal form ΣN F (see (3.7)) x˙ 1 = x2 + x23 P [m−2] (x1 , x2 , x3 ) x˙ 2 = x3 x˙ 3 = v, 20

where P [m−2] (x1 , x2 , x3 ) is a homogeneous polynomial of degree m − 2 of the variables x1 , x2 , x3 . ¤ We would like now to discuss the interest of Theorem 3.4. As we have already mentioned, Poincar´e’s method allows to replace the partial differential equation (CDE) (given in Section 1) by solving successively linear algebraic equations defined by the control homological equation (CHE), see [52] and [48], and Proposition 3.1. The solvability of this equation was proved in [52] and [48] while Theorem 3.4 provides an explicit solution (in the form of the transformations (3.10), that are easily computable via differentiation and integration of homogeneous polynomials) to the control homological equation. As a consequence, for any given control system, Theorem 3.4 gives transformations bringing the homogeneous part of the system into its normal form. For example, if the system is feedback linearizable, up to order m0 − 1 (see [54]), then a diffeomorphism and a feedback compensating all nonlinearities of degree lower than m0 can be calculated explicitly without solving partial differential equations (compare Section 9). More generally, by a successive application of transformations given by (3.10) we can bring the system, without solving partial differential equations, to its normal form given in Theorem 3.6 below. Consider the system Σ∞ of the form (3.1) and recall that we assume the linear part (F, G) to be controllable. Apply successively to Σ∞ a series of transformations Γm , m = 1, 2, 3, . . . , such that [m] each Γm brings Σ[m] to its normal form ΣN F . More precisely, bring (F, G) into the Brunonvsk´ y 1 ∞,1 1 ∞ canonical form (A, B) via a linear feedback Γ and denote Σ = Γ∗ (Σ ). Assume that a system Σ∞,m−1 has been defined. Let Γm be a homogeneous feedback transformation transforming Σ[m] , [m] which is the homogeneous part of degree m of Σ∞,m−1 , to the normal form ΣN F (Γm can be ∞,m−1 taken, for instance, as the transformations defined by (3.10)). Define Σ∞,m = Γm ). ∗ (Σ m ∞,m−1 Notice that we apply Γ to the whole system Σ (and not only to its homogeneous part [m] Σ ). Successive repeating of Theorem 3.2 gives the following result of Kang [48]. Theorem 3.6 There exists a formal feedback transformation Γ∞ which brings the system Σ∞ to a normal form Σ∞ N F given by  n P   x2i P1,i (x1 , . . . , xi ) x ˙ = x + 1 2   i=3    ..   .    n P    x˙ j = xj+1 + x2i Pj,i (x1 , . . . , xi )    i=j+2 .. Σ∞ : (3.11) NF .     x˙ n−2 = xn−1 + x2n Pn−2,n (x1 , . . . , xn )         x˙ n−1 = xn       x˙ n = v , where Pj,i (x1 , . . . , xi ) are formal power series depending on the indicated variables. Example 3.7 Consider a system Σ defined on R3 whose linear part is controllable (compare Example 3.5). Theorem 3.6 implies that the system Σ is equivalent, via a formal feedback

21

transformation Γ∞ , to its normal form Σ∞ NF x˙ 1 = x2 + x23 P (x1 , x2 , x3 ) x˙ 2 = x3 x˙ 3 = v, where P (x1 , x2 , x3 ) is a formal power series of the variables x1 , x2 , x3 .

3.4

¤

Normal form for non-affine systems [m]

In this section we will generalize normal forms ΣN F and Σ∞ N F to general control systems. As we explained in Section 1, such a generalization can be performed using Proposition 1.1. Consider a general control system of the form Π : ξ˙ = F (ξ, u), around an equilibrium point (ξ0 , u0 ), that is F (ξ0 , u0 ) = 0. Without loss of generality, we can assume that (ξ0 , u0 ) = (0, 0). Together with Π we will consider its Taylor series expansion ∞

Π

: ξ˙ = F ξ + Gu +

∞ X

F [m] (ξ, u),

m=2

where F [m] (ξ, u) stands for homogeneous terms of degree m and homogeneity is understood in this section with respect to the state and control variables together. Consider the feedback transformation Υ (compare Section 1) Υ :

x = φ(ξ) v = ψ(ξ, u)

and its Taylor series expansion Υ∞ given by x = Tξ +

∞ P

φ[m] (ξ)

m=2

Υ∞ :

v = Kξ + Lu +

∞ P

ψ [m] (ξ, u),

m=2

where T is an invertible matrix and L 6= 0. We will assume throughout this section that the pair (F, G) is controllable and so we can suppose that it is in the Brunovsk´ y canonical form (A, B). Like in the control-affine case, we will consider the action of the homogenous part Υm of Υ∞ given by Υm :

x = ξ + φ[m] (ξ) v = u + ψ [m] (ξ, u)

on the homogeneous part Π[m] of Π∞ given by Π[m] : ξ˙ = Aξ + Bu + F [m] (ξ, u). Combining Theorem 3.2 with Proposition 1.1 leads to the following result: 22

Theorem 3.8 The general homogeneous system Π[m] can be transformed, via a homogeneous feedback transformation Υm , into the following normal form  n P [m−2] [m−2]   x ˙ = x + x2i P1,i (x1 , . . . , xi ) + v 2 P1 (x1 , . . . , xn , v)  1 2   i=3   ..   .    n  P [m−2] [m−2]   x ˙ = x + x2i Pj,i (x1 , . . . , xi ) + v 2 Pj (x1 , . . . , xn , v)  j j+1   i=j+2  [m] .. ΠN F : (3.12) .   [m−2] [m−2]   x˙ n−2 = xn−1 + x2n Pn−2,n (x1 , . . . , xn ) + v 2 Pn−1 (x1 , . . . , xn , v)       [m−2]   x˙ n−1 = xn + v 2 Pn (x1 , . . . , xn , v)        x˙ n = v , [m−2]

[m−2]

where Pj,i (x1 , . . . , xi ) and Pj (x1 , . . . , xn , v) are homogeneous polynomials of degree m − 2 depending on the indicated variables. Notice that formally the above normal form can be obtained as follows. Consider the affine [m] normal form ΣN F and apply the reduction defined as the inverse of the extension (preintegration) [m] described just before Proposition 1.1. More precisely, assume that ΣN F is controlled by xn , so skip the last equation x˙ n = v and denote xn by v. What we obtain is an (n − 1)-dimensional [m] system, nonlinear with respect to v, which gives actually the (n − 1)-dimensional form ΠN F . Like in the control-affine case, a successive application of Theorem 3.8 gives a formal normal [m] ∞ form for Π∞ N F under Υ . It has the same structure as ΠN F , the only difference being that the [m−2] [m−2] polynomials Pj,i (x1 , . . . , xi ) and Pj (x1 , . . . , xn , v) are replaced by formal power series of the same variables. We will end up with a simple example, which, actually, is a non-affine version of Examples 3.5 and 3.7. Example 3.9 Consider the general system Π[m] on R2 . Theorem 3.6 implies that the system [m] Π[m] is equivalent, via a homogeneous feedback transformation Υm to its normal form ΠN F , see (3.12): [m−2] x˙ 1 = x2 + v 2 P1 (x1 , x2 , v) x˙ 2 = v, [m−2]

where P1 (x1 , x2 , v) is a homogeneous polynomial of degree m−2 of the variables x1 , x2 and v. As a consequence, the general system Π∞ on R2 is equivalent, via a formal feedback transformation Υ∞ to its normal form Π∞ NF : x˙ 1 = x2 + v 2 P1 (x1 , x2 , v) x˙ 2 = v, where P1 (x1 , x2 , v) is a formal power series of the variables x1 , x2 and v.

23

¤

4

Canonical form for single-input systems with controllable linearization [m]

As proved by Kang and recalled in Theorem 3.3, the normal form ΣN F is unique under homogeneous feedback transformation Γm . The normal form Σ∞ N F is constructed by a successive m application of homogeneous transformations Γ , for m ≥ 1, which bring the corresponding ho[m] mogeneous systems Σ[m] into their normal forms ΣN F . Therefore a natural and fundamental question which arises is whether the system Σ∞ can admit two different normal forms, that is, whether the normal forms given by Theorem 3.6 are in fact canonical forms under a general formal feedback transformations of the form Γ∞ . It turns out that a given system can admit different normal forms, as shows the following example of Kang [48]. The main reason for the nonunique[m] ness of the normal form Σ∞ N F is that, although the normal form ΣN F is unique, homogeneous [m] feedback transformation Γm bringing Σ[m] into ΣN F is not. It is this small group of homogeneous [m] feedback transformations of order m that preserve ΣN F (described by Proposition 4.4), which causes the nonuniqueness of Σ∞ NF . Example 4.1 Consider the following system ξ˙1 = ξ2 + ξ32 − 2ξ1 ξ32 ξ˙2 = ξ3 ξ˙3 = u,

(4.1)

on R3 . Clearly, this system is in Kang normal form (compare with Theorem 3.6), say Σ∞ 1,N F . The feedback transformation

Γ≤3 :

x1 = ξ1 − ξ12 − 43 ξ23 x2 = ξ2 − 2ξ1 ξ2 x3 = ξ3 − 2(ξ22 + ξ1 ξ3 ) − 2ξ2 ξ32 u

= v + 6ξ2 ξ3 + 12ξ1 ξ2 ξ3 − 4ξ33 + 2(ξ1 + 2ξ12 + 2ξ2 ξ3 )v

brings the system (4.1) into the form x˙ 1 = x2 + x23 x˙ 2 = x3 x˙ 3 = v modulo terms in V ≥4 (x, v). Applying successively homogeneous feedback transformations Γm given, for any m ≥ 4, by (3.10), we transform the above system into the following normal form Σ∞ 2,N F : x˙ 1 = x2 + x23 + x23 P (x) x˙ 2 = x3 (4.2) x˙ 3 = v, where P is a formal power series whose 1-jet at 0 ∈ R3 vanishes. The systems (4.1) and (4.2) are ∞ in their normal forms (Σ∞ 1,N F and Σ2,N F , respectively) and, moreover, the systems are feedback equivalent, but the system (4.2) does not contain any term of degree 3. As a consequence, the normal form Σ∞ ¤ N F is not unique under feedback transformations. 24

A natural and important problem is thus to construct a canonical form and the aim of this section is indeed to construct a canonical form for Σ∞ under feedback transformation Γ∞ . Consider the system Σ∞ of the form ∞

Σ

: ξ˙ = F ξ + Gu +

∞ X ¡ [m] ¢ f (ξ) + g [m−1] (ξ)u .

(4.3)

m=2

Since its linear part (F, G) is assumed to be controllable, we bring it, via a linear transformation and linear feedback, to the Brunovsk´ y canonical form (A, B). Let the first homogeneous term of Σ∞ which cannot be annihilated by a feedback transformation be of degree m0 . As proved by Krener [54], the degree m0 is given by the largest integer such that all distributions Dk = span {g, . . . , adk−1 f g}, for 1 ≤ k ≤ n − 1, are involutive modulo terms of order m0 − 2. We can thus, due to Theorems 3.2 and 3.3, assume that, after applying a suitable feedback Γ≤m0 , the system Σ∞ takes the form ∞ X ¡

ξ˙ = Aξ + Bu + f¯[m0 ] (ξ) +

¢ f [m] (ξ) + g [m−1] (ξ)u ,

m=m0 +1

where (A, B) is in Brunovsk´ y canonical form and the first nonvanishing homogeneous vector field f¯[m0 ] is in the normal form (by Theorem 3.2) with components given by  n P 2 [m0 −2]   ξi Pj,i (ξ1 , . . . , ξi ), 1 ≤ j ≤ n − 2,  [m0 ] i=j+2 ¯ fj (ξ) =    0, n − 1 ≤ j ≤ n. Let (i1 , . . . , in−s ), where i1 + · · · + in−s = m0 and in−s ≥ 2, be the largest, in the lexicographic ordering, (n − s)-tuple of nonnegative integers such that for some 1 ≤ j ≤ n − 2, we have ∂ m0 f¯j

[m0 ] i

n−s ∂ξ1i1 · · · ∂ξn−s

Define

6= 0.

(

∂ m0 f¯j

[m0 ]

j ∗ = sup j = 1, . . . , n − 2 :

i

n−s ∂ξ1i1 · · · ∂ξn−s

) 6= 0 .

The following results, whose proofs are detailed in [81], describe the canonical form obtained by the authors. Theorem 4.2 The system Σ∞ given by (4.3) is equivalent by a formal feedback Γ∞ to a system of the form ∞ X ∞ f¯[m] (x), ΣCF : x˙ = Ax + Bv + m=m0 [m] f¯j (x)

¯[m]

where, for any m ≥ m0 , the components of f (x) are given by  n P 2 [m−2]   xi Pj,i (x1 , . . . , xi ), 1 ≤ j ≤ n − 2,  [m] i=j+2 ¯ fj (x) =    0, n − 1 ≤ j ≤ n, 25

(4.4)

additionally, we have

∂ m0 f¯j ∗ 0

[m ] i

n−s ∂xi11 · · · ∂xn−s

= ±1

(4.5)

and, moreover, for any m ≥ m0 + 1, ∂ m0 f¯j ∗

[m] i

n−s ∂xi11 · · · ∂xn−s

(x1 , 0, . . . , 0) = 0.

(4.6)

∞ The form Σ∞ CF satisfying (4.4), (4.5) and (4.6) will be called the canonical form of Σ . The name is justified by the following ∞ Theorem 4.3 Two systems Σ∞ 1 and Σ2 are formally feedback equivalent if and only if their ∞ ∞ canonical forms Σ1,CF and Σ2,CF coincide.

Kang [48], generalizing [52], proved that any system Σ∞ can be brought by a formal feedback into the normal form Σ∞ N F , for which (4.4) is satisfied. He also observed that his normal forms are not unique, see Example 4.1. Our results, Theorems 4.2 and 4.3, complete his study. We show that for each degree m of homogeneity we can use a 1-dimensional subgroup of feedback transformations which preserves the “triangular” structure of (4.4) and at the same time allows us to normalize one higher order term. The form of (4.5) and (4.6) is a result of this normalization. These 1-dimensional subgroups of feedback transformations are given by the following proposition. Proposition 4.4 The transformation Γm given by (3.3) leaves invariant the system Σ[m] defined by (3.4) if and only if [m] m φj = am Lj−1 1 ≤ j ≤ n, Aξ ξ1 , α[m]

= −am LnAξ ξ1m ,

(4.7)

n−1 m β [m−1] = −am LB LAξ ξ1 ,

where am is an arbitrary real parameter. Theorem 4.2 establishes an equivalence of the system Σ∞ with its canonical form Σ∞ CF via a formal feedback. Its direct corollary yields the following result for equivalence under a smooth feedback of the form x = φ(ξ) Γ : u = α(ξ) + β(ξ)v, up to an arbitrary order. Indeed, we have the following: Corollary 4.5 Consider a smooth control system Σ : ξ˙ = f (ξ) + g(ξ)u. For any positive integer k we have:

26

(i) There exists a smooth feedback Γ transforming Σ, locally around 0 ∈ Rn , into its canonical form Σ≤k CF given by k X Σ≤k : x˙ = Ax + Bv + f¯[m] (x), CF

m=m0 [m] modulo terms in V ≥k+1 (x, v), where the components f¯j (x) of f¯[m] (x), for any m0 ≤ m ≤ k, satisfy (4.4),(4.5), (4.6). ≥k+1 (ii) Feedback equivalence of Σ and Σ≤k (x, v), can be established via a CF , modulo terms in V ≤k polynomial feedback transformation Γ of degree k. (iii) Two smooth systems Σ1 and Σ2 are feedback equivalent modulo terms in V ≥k+1 (x, v) if and ≤k only if their canonical forms Σ≤k 1,CF and Σ2,CF coincide.

This Corollary follows directly from Theorems 4.2 and 4.3. To end this section will illustrate our results by two examples. Example 4.6 Let us reconsider the system Σ given by Example 3.7. It is equivalent, via a formal feedback, to the normal form x˙ 1 = x2 + x23 P (x1 , x2 , x3 ) x˙ 2 = x3 x˙ 3 = v, where P (x1 , x2 , x3 ) is a formal power series. Assume, for simplicity, that m0 = 2, which is equivalent to the following generic condition : g, adf g, and [g, adf g] are linearly independent at 0 ∈ R3 . This implies that we can express P = P (x1 , x2 , x3 ) as P = c + P1 (x1 ) + x2 P2 (x1 , x2 ) + x3 P3 (x1 , x2 , x3 ), where c 6= 0 and P1 (0) = 0. Observe that any P (x1 , x2 , x3 ), of the above form, gives a normal ∞ form Σ∞ N F . In order to get the canonical form ΣCF , we use Theorem 4.2 which assures the existence of a feedback transformation Γ∞ of the form x˜ = φ(x) v = α(x) + β(x)˜ v, which normalizes the constant c and annihilates the formal power series P1 (x1 ). More precisely, Γ∞ transforms Σ into its canonical form Σ∞ CF x1 , x˜2 , x˜3 ) x˜˙ 1 = x˜2 + x˜23 P˜ (˜ ˙x˜2 = x˜3 x˜˙ 3 = v˜, where the formal power series P˜ (˜ x1 , x˜2 , x˜3 ) is of the form P˜ (˜ x1 , x˜2 , x˜3 ) = 1 + x˜2 P˜2 (˜ x1 , x˜2 ) + x˜3 P˜3 (˜ x1 , x˜2 , x˜3 ), showing clearly a difference between the normal and canonical form: in the latter, the free term is normalized and the term depending on x1 is annihilated. ¤ 27

Now, we give an example of constructing the canonical form for a physical model of variable length pendulum. Example 4.7 Consider the variable length pendulum of Bressan and Rampazzo [8] (see also [17]). We denote by ξ1 the length of the pendulum, by ξ2 its velocity, by ξ3 the angle with respect to the horizontal, and by ξ4 the angular velocity. The control u = ξ˙4 = ξ¨3 is the angular acceleration. The mass is normalized to 1. The equations are (compare [8] and [17]) ξ˙1 ξ˙2 ξ˙3 ξ˙4

= = = =

ξ2 −g sin ξ3 + ξ1 ξ42 ξ4 u,

where g denotes the gravity. Notice that if we suppose to control the angular velocity ξ4 = ξ˙3 , which is the case of [8] and [17], then the system is three-dimensional but the control enters nonlinearly. At any equilibrium point ξ0 = (ξ10 , ξ20 , ξ30 , ξ40 )T =(ξ10 , 0, 0, 0)T , the linear part of the system is controllable. Our goal is to produce, for variable length pendulum, a normal form and the canonical form as well as to answer the question whether the systems corresponding to various values of the gravity constant g are feedback equivalent. In order to get a normal form, put x1 x2 x3 x4 v

= = = = =

ξ1 ξ2 −g sin ξ3 −gξ4 cos ξ3 gξ42 sin ξ3 − ug cos ξ3 .

x˙ 1 x˙ 2 x˙ 3 x˙ 4

= = = =

The system becomes x2 1 x3 + x24 g2x−x 2 3 x4 v,

which gives a normal form. Indeed, we rediscover Σ∞ N F , given by (3.11), with P1,3 = 0, P1,4 = 0, and x1 . P2,4 = 2 g − x23 To bring the system to its canonical form Σ∞ CF , firstly observe that m0 = 3. Indeed, the 2 x1 function x4 g2 −x2 starts with third order terms, which corresponds to the fact that the invariants 3 a[2]j,i+2 vanish for any 1 ≤ j ≤ 2 and any 0 ≤ i ≤ 2 − j. The only non zero component of f¯[3] [1] [3] is f¯2 = x24 P2,4 . Hence j ∗ = 2 and the only, and thus the largest, quadruplet (i1 , i2 , i3 , i4 ) of nonnegative integers, satisfying i1 + i2 + i3 + i4 = 3 and such that ∂ 3 f¯2 6= 0, ∂xi11 . . . ∂xi44 [3]

[3]

is (i1 , i2 , i3 , i4 ) = (1, 0, 0, 2). In order to normalize f2 , put x˜i = a1 xi , 1 ≤ i ≤ 4, v˜ = a1 v, 28

where a1 = 1/g. We get the following canonical form for the variable length pendulum x˜˙ 1 x˜˙ 2 x˜˙ 3 x˜˙ 4

= = = =

x˜2 x ˜1 x˜3 + x˜24 1−˜ x23 x˜4 v˜.

Independently of the value of the gravity constant g, all systems are feedback equivalent to each other. ¤

5

Dual normal form and dual m-invariants [m]

In the normal form ΣN F given by (3.7), all the components of the control vector field g [m−1] are annihilated and all non removable nonlinearities are grouped in f [m] . Kang and Krener in their pioneering paper [52] have shown that it is possible to transform, via a homogeneous transformation Γ2 of degree 2, the homogeneous system Σ[2] : ξ˙ = Aξ + Bu + f [2] (ξ) + g [1] (ξ)u to a dual normal form. In that form the components of the drift f [2] are annihilated while all non removable nonlinearities are, this time, present in g [1] . The aim of this section is to propose, for an arbitrary m, a dual normal form for the system Σ[m] and a dual normal form for the system Σ∞ . Our dual normal form, on the one hand, generalizes, for higher order terms, that given in [52] [m] for second order terms, and, on the other hand, dualizes, the normal form ΣN F . The structure of this section will follow that of Section 3: we will give the dual normal form, then we define and study dual m-invariants, and, finally, we give an explicit construction of transformations bringing the system into its dual normal form. For the proofs of all results contained in this section the reader is sent to [81]. Our first result asserts that we can always bring the homogeneous system Σ[m] , given by (3.4), into a dual normal form. Theorem 5.1 The homogeneous system Σ[m] is equivalent, via a homogeneous feedback trans[m] formations Γm , to the dual normal form ΣDN F given by  x˙ 1 = x2    [m−2]  x˙ 2 = x3 + vxn Q2,n (x1 , . . . , xn )     ..   .   n  P  [m−2]  xi Qj,i (x1 , . . . , xi )  x˙ j = xj+1 + v [m] i=n−j+2 ΣDN F : (5.1) .  .  .    n P  [m−2]   x ˙ = x + v xi Qj,i (x1 , . . . , xi ) n−1 n    i=3      x˙ n = v , [m−2]

where Qj,i variables.

(x1 , . . . , xi ) are homogeneous polynomials of degree m−2 depending on the indicated

29

[2]

We will give the dual normal form ΣDN F for homogeneous systems of degree two: x˙ 1 = x2 x˙ 2 = x3 + .. .

+ vxn q2,n

x˙ n−1 = xn + vx3 qn−1,3 + · · · + vxn qn−1,n x˙ n = v, where qj,i ∈ R. The following example is a particular case of the system above and help illustrate Theorem 5.1. Example 5.2 Consider the system Σ[2] defined in R3 by ξ˙1 = ξ2 + ξ32 ξ˙2 = ξ3 ξ˙3 = u. It is easy to check that the change of coordinates x1 = ξ1 , x2 = ξ2 + ξ32 , x3 = ξ3 , and x4 = ξ4 yields the dual normal form (n = 3, q2,3 = 2, and v = u) x˙ 1 = x2 x˙ 2 = x3 + 2x3 v x˙ 3 = v. Now we will define dual m-invariants. To start with, recall that the homogeneous vector field [m−1] Xi is defined by taking the homogeneous part of degree m − 1 of the vector field Xim−1 = (−1)i adiAξ+f [m] (B + g [m−1] ). [m−1]

[m−1]

By Xi (πi (ξ)) we will denote Xi evaluated at the point πi (ξ) = (ξ1 , . . . , ξi , 0, . . . , 0)T of the subspace © ª Wi = ξ = (ξ1 , . . . , ξn )T ∈ Rn : ξi+1 = · · · = ξn = 0 . [m−1]

Consider the system Σ[m] and for any j, such that 2 ≤ j ≤ n − 1, define the polynomial bj by setting [m−1] bj

=

[m−1] gj

+

j−1 X

[m] LB Lj−k−1 fk Aξ



n X

Z LB Lj−1 Aξ

i=1

k=1

ξi 0

[m−1]

CXn−i (πi (ξ))dξi .

[m−1]

The homogeneous polynomials bj , for 2 ≤ j ≤ n − 1, will be called the dual m-invariants of [m] the homogeneous system Σ . ˜ [m] of the form (3.4) and (3.5). Let Consider two systems Σ[m] and Σ [m−1]

{ bj

: 2 ≤ j ≤ n − 1 },

and

{ ˜bj

[m−1]

: 2 ≤ j ≤ n − 1}

denote, respectively, their dual m-invariants. The following result dualizes that of Theorem 3.3. Theorem 5.3 The dual m-invariants have the following properties: 30

˜ [m] are equivalent via a homogeneous feedback transformation Γm if (i) two systems Σ[m] and Σ and only if [m−1] [m−1] bj = ˜bj , for any 2 ≤ j ≤ n − 1; [m−1] [m] (ii) the dual m-invariants ¯bj of the dual normal form ΣDN F , defined by (5.1), are given by n X

¯b[m−1] (x) = j

[m−2]

xi Qj,i

(x1 , . . . , xi ),

for any 2 ≤ j ≤ n − 1.

i=n−j+2

The above result asserts that the dual m-invariants, similarly like m-invariants, form a set of complete invariants of the homogeneous feedback transformation. Notice, however, that the same information is encoded in both sets of invariants in different ways. [m]

Like Theorem 3.3 for normal form ΣN F , Theorem 5.3 shows that the polynomial functions [m−2] [m] Qj,i defining the dual normal form ΣDN F are unique under feedback transformation Γm . The [m] remaining question is how to bring a given system into its dual normal form ΣDN F . To this end, define the following homogeneous polynomials [m] φ1

= −

n Z X i=1

[m]

ξi 0

[m]

φj+1

[m]

= fj

α[m] β [m−1]

[m−1]

CXn−i (πi (ξ))dξi

+ LAξ φj , 1 ≤ j ≤ n − 1, ´ ³ [m] [m] = − fn + LAξ φn ´ ³ [m] [m−1] + LB φn . = − gn

(5.2)

Theorem 5.4 The feedback transformation m

Γ

:

x = ξ + φ[m] (ξ) u = v + α[m] (ξ) + β [m−1] (ξ)v, [m]

where α[m] , β [m−1] , and the components φj of φ[m] are defined by (5.2), brings the system Σ[m] [m] into its dual normal form ΣDN F given by (5.1). ∞ Now our aim is to dualize the normal form Σ∞ of the form (4.3) N F . Consider the system Σ and assume that its linear part (F, G) is controllable. Consider the system Σ∞ of the form (3.1) and recall that we assume the linear part (F, G) to be controllable. Apply successively to Σ∞ a series of transformations Γm , m = 1, 2, 3, . . . , [m] such that each Γm brings Σ[m] to its normal form ΣDN F . More precisely, bring (F, G) into the Brunonvsk´ y canonical form (A, B) via a linear feedback Γ1 and denote Σ∞,1 = Γ1∗ (Σ∞ ). Assume that a system Σ∞,m−1 has been defined. Let Γm be a homogeneous feedback transformation transforming Σ[m] , which is the homogeneous part of degree m of Σ∞,m−1 , to the dual normal [m] form ΣDN F (Γm can be taken, for instance, as the transformations defined by (5.2)). Define ∞,m−1 ). Notice that we apply Γm to the whole system Σ∞,m−1 (and not only to its Σ∞,m = Γm ∗ (Σ homogeneous part Σ[m] ). Successive repeating of Theorem 5.4 gives the following dual normal form.

31

Theorem 5.5 The system Σ∞ can be transformed via a formal feedback transformation Γ∞ , into the dual normal form Σ∞ DN F given by  x˙ 1 = x2     x˙ 2 = x3 + vxn Q2,n (x1 , . . . , xn )    ..   .    n  P   xi Qj,i (x1 , . . . , xi )  x˙ j = xj+1 + v i=n−j+2 : Σ∞ (5.3) DN F ..   .   n  P    xi Qj,i (x1 , · · · , xi ) x ˙ = x + v n−1 n   i=3       x˙ n = v, where Qj,i (x1 , . . . , xi ) are formal power series depending on the indicated variables.

6

Dual canonical form

Naturally, similarly to normal forms, a given system can admit different dual normal forms. We are thus interested in constructing a dual canonical form (which would dualize the canonical form ∞ ∞ Σ∞ CF in the same way as ΣDN F dualizes ΣN F ). Assuming that the linear part (F, G) of the system ∞ Σ , of the form (4.3), is controllable, we denote by m0 the degree of the first homogeneous term of the system Σ∞ which cannot be annihilated by a feedback transformation. Thus by Theorems 5.3 and 5.4 (using transformations (5.2)), we can assume, after applying a suitable feedback, that Σ∞ takes the form ∞

Σ

∞ X ¡

: ξ˙ = Aξ + Bu + g¯[m0 −1] (ξ)u +

¢ f [m] (ξ) + g [m−1] (ξ)u ,

m=m0 +1

where (A, B) is in Brunovsk´ y canonical form and the first nonvanishing homogeneous vector [m0 −1] field g¯ is in the dual normal form, compare (5.1) with components given by  n P [m −2]   ξi Qj,i 0 (ξ1 , . . . , ξi ), 2≤j ≤n−1  [m0 −1] i=n−j+2 g¯j (ξ) =    0, j = 1 or j = n. Define

o n [m −1] j∗ = inf j = 2, . . . , n − 1 : g¯j 0 (ξ) 6= 0

and let (i1 , . . . , in ), such that i1 + · · · + in = m0 − 1, be the largest, in the lexicographic ordering, n-tuple of nonnegative integers such that [m −1]

∂ m0 −1 g¯j∗ 0

∂ξ1i1 · · · ∂ξnin We have the following result: 32

6= 0.

Theorem 6.1 There exists a formal feedback transformation Γ∞ which brings the system Σ∞ into the following form ∞ X ∞ ΣDCF : x˙ = Ax + Bv + g¯[m−1] (x)v, m=m0 [m−1]

where for any m ≥ m0 , the components g¯j of g¯[m−1] are given by  n P [m−2]   xi Qj,i (x1 , . . . , xi ), 2 ≤ j ≤ n − 1  [m−1] i=n−j+2 = g¯j    0, j = 1 or j = n.

(6.1)

Moreover, [m −1]

∂ m0 −1 g¯j∗ 0

∂xi11 · · · ∂xinn and for any m ≥ m0 + 1

= ±1

(6.2)

[m−1]

∂ m0 −1 g¯j∗

∂xi11 · · · ∂xinn

(x1 , 0, . . . , 0) = 0 .

(6.3)

∞ The form Σ∞ DCF , which satisfies (6.1), (6.2) and (6.3), will be called dual canonical form of Σ . The name is justified by the following. ∞ Theorem 6.2 Two systems Σ∞ 1 and Σ2 are formally feedback equivalent if and only if their ∞ ∞ dual canonical forms Σ1,DCF and Σ2,DCF coincide.

Example 6.3 Let us consider the system Σ : ξ˙ = f (ξ) + g(ξ)u,

ξ ∈ R3 , u ∈ R,

whose linear part is assumed to be controllable. Theorem 5.5 assures that the system Σ is formally feedback equivalent to the dual normal form Σ∞ DN F given by x˙ 1 = x2 x˙ 2 = x3 + vx3 Q(x1 , x2 , x3 ) x˙ 3 = v, where Q(x1 , x2 , x3 ) is a formal power series of the variables x1 , x2 , x3 . Assume for simplicity that m0 = 2, which is equivalent to the condition: g, adf g, and [g, adf g] linearly independent at 0 ∈ R3 . This implies that we can represent Q = Q(x1 , x2 , x3 ), as Q = c + x1 Q1 (x1 ) + x2 Q2 (x1 , x2 ) + x3 Q3 (x1 , x2 , x3 ), where c ∈ R, c 6= 0. Observe that any Q of the above form gives a dual normal form Σ∞ DN F . In order to get the dual canonical form we use Theorem 6.1, which assures that the system Σ is formally feedback equivalent to its dual canonical form Σ∞ DCF defined by x˜˙ 1 = x˜2 ˜ x1 , x˜2 , x˜3 ) x˜˙ 2 = x˜3 + v˜x˜3 Q(˜ ˙x˜3 = v˜, 33

˜ x1 , x˜2 , x˜3 ) is a formal power series such that where Q(˜ ˜ x1 , x˜2 , x˜3 ) = 1 + x˜2 Q ˜ 2 (˜ ˜ 3 (˜ Q(˜ x1 , x˜2 ) + x˜3 Q x1 , x˜2 , x˜3 ). This ends the example.

7 7.1

¤

Normal forms for single-input systems with uncontrollable linearization Introduction [m]

In Section 3 we presented the normal forms ΣN F and Σ∞ N F of the system Σ whose linearization (that is, linear approximation) is controllable. In this section we will deal with system with uncontrollable linearization. A normal form for homogeneous systems of degree 2, with uncontrollable linearization, was proposed by Kang [49]. The normal form presented in this section was obtained by the authors ([79] and [83] and it generalizes, on the one hand side, the normal form of Kang [49] (uncontrollable linearization) and, on the other hand side, the normal form [m] ΣN F (controllable linearization) also obtained by Kang and presented in Section 3. Another normal form for single-input systems with uncontrollable linearization has also been proposed by Krener, Kang, and Chang [56], [57] and [58], and by the authors [76], which differ from ours by another definition of homogeneity (we consider the latter with respect to the linearly controllable variables while theirs is respect to all variables); see Example 7.5 illustrating this.

7.2

Taylor series expansions

All objects, that is, functions, maps, vector fields, control systems, etc., are considered in a neighborhood of 0 ∈ Rn and assumed to be C ∞ -smooth. Consider the single-input system Σ : ξ˙ = f (ξ) + g(ξ)u,

ξ ∈ Rn , u ∈ R.

We assume throughout this section that f (0) = 0 and g(0) 6= 0. Let Σ[1] : ξ˙ = F ξ + Gu, where F = ∂f (0) and G = g(0), be the linear approximation of the system around the equilibrium ∂ξ point 0 ∈ Rn . If the linear approximation is not controllable, which is the case studied in this section, then there exists a positive integer r ∈ N such that rank (G, F G, . . . , F n−1 G) = n − r. Moreover, there exist coordinates ξ = (ξ1 , . . . , ξr , ξr+1 , . . . , ξn )T of Rr × Rn−r in which the pair (F, G) admits the following Kalman decomposition µ ¶ µ ¶ F1 0 0 A= and B = , F3 F2 n×n G2 n×1 where the pair (F2 , G2 ) is controllable. Throughout this section r will stand for the dimension of the uncontrollable part of the linear approximation of the system and ξ = (ξ1 , . . . , ξr , ξr+1 , . . . , ξn )T will denote coordinates defining the Kalman decomposition. 34

We will use the notation C0∞ (Rr ) for the space of germs at 0 ∈ Rr of smooth R-valued functions of ξ = (ξ1 , . . . , ξr )T ∈ Rr . By R[[ξ1 , . . . , ξr ]] we will denote the space of formal power series of ξ1 , . . . , ξr with coefficients in R. Let h be a smooth R-valued function defined in a neighborhood of 0 × 0 ∈ Rr × Rn−r . By h(ξ) = h[0] (ξ) + h[1] (ξ) + h[2] (ξ) + · · · =

∞ X

h[m] (ξ)

m=0

we denote its Taylor series expansion with respect to (ξr+1 , . . . , ξn )T at 0 ∈ Rr ×Rn−r , where h[m] (ξ) stands for a homogeneous polynomial of degree m of the variables ξr+1 , . . . , ξn whose coefficients are in C0∞ (Rr ). Similarly, throughout this section, for a map φ of an open subset of Rr × Rn−r to Rr × Rn−r (resp. for a vector field f on an open subset of Rr ×Rn−r ) we will denote by φ[m] (resp. by f [m] ) the term of degree m of its Taylor series expansion with respect to (ξr+1 , . . . , ξn )T at 0 ∈ Rr × Rn−r , [m] [m] that is, each component φj of φ[m] (resp. fj of f [m] ) is a homogeneous polynomial of degree m of the variables ξr+1 , . . . , ξn whose coefficients are in C0∞ (Rr ). Consider the Taylor series expansion of the system Σ given by Σ∞ : ξ˙ = F ξ + Gu + f [0] (ξ) +

∞ X

(f [m] (ξ) + g [m−1] u).

(7.1)

m=1

Notice that, although we assume f (0) = 0, the term f [0] (ξ) is present because the degree is computed with respect to the variables ξr+1 , . . . , ξn only and thus f [0] is, in general, a function of ξ1 , . . . , ξr . Consider also the Taylor series expansion Γ∞ of the feedback transformation Γ given by x = Tξ + Γ∞ :

∞ P

φ[m] (ξ)

m=0

u = Kξ + Lv + α[0] (ξ) +

∞ P

(7.2) (α[m] (ξ) + β [m−1] (ξ)v),

m=1

where T is an invertible matrix and L 6= 0. The method proposed by Kang and Krener is to study the action of Γ∞ on the system Σ∞ step by step, that is, to analyze successively the action of the homogeneous parts of Γ∞ on the homogeneous parts, of the same degree, of Σ∞ . Notice, however, that in their approach, Kang and Krener consider Taylor series expansions with respect to all state variables ξ1 , . . . , ξn of the system and, as a consequence, homogeneity is considered with respect to all variables ξ1 , . . . , ξn . Following our approach [79], [83] we propose a slight modification of that homogeneity. In view of a different nature of the controllable and uncontrollable parts of the linear approximation, we consider Taylor series expansions with respect to the linearly controllable variables ξr+1 , . . . , ξn only. This leads to considering as homogeneous parts of the system and of the feedback transformations, according to our definition, terms that are polynomial with respect to ξr+1 , . . . , ξn with smooth coefficients depending on ξ1 , . . . , ξr . When analyzing the action of a homogeneous transformation Γm (understood as homogeneity with respect to the controllable variables) on the system Σ∞ , we can notice three undesirable phenomena (see Section 7.5) that are not present in the action of homogeneous transformations in the controllable case (where homogeneity is 35

considered with respect to all variables). In order to deal with those problems caused by the presence of the uncontrollable linear part, we will introduce, in Section 7.5, different weights for the components corresponding to the controllable and uncontrollable parts.

7.3

Linear part and resonances

Let λ = (λ1 , . . . , λr ) ∈ Cr be the set of eigenvalues associated to the uncontrollable part of the linear system ξ˙ = F ξ + Gu. By a linear feedback transformation Γ1 :

x = Tξ u = Kξ + Lv

it is always possible to bring the linear system into the following Jordan-Brunovsk´ y canonical form µ ¶ µ ¶ J 0 0 A= and B = , 0 A2 n×n B2 n×1 where J is the Jordan canonical form of dimension r and (A2 , B2 ) the Brunovsk´ y canonical form of dimension n − r. In the case when all eigenvalues λi are real, we have   λ1 σ 2 · · · 0 . .. ..  . . ..   0  J = . . σi ∈ {0, 1}.  , . .. . . σr   .. 0 · · · 0 λr r×r In the case of complex eigenvalues, we replace in J the eigenvalue λj by the matrix µ ¶ αj βj Λj = , −βj αj 2×2 where λj = αj + iβj , and we replace the integer σj ∈ {0, 1} either by the matrix µ ¶ µ ¶ 0 0 1 0 Ξj = or Ξj = . 0 0 2×2 0 1 2×2 Recall from Section 2 the notion of a resonant eigenvalue and the resonant set associated with it. Definition 7.1 An eigenvalue λj is called resonant if there exists a r-tuple α = (α1 , . . . , αr ) ∈ Nr of nonnegative integers, satisfying |α| = α1 + · · · + αr ≥ 2, such that λj = λ1 α1 + · · · + λr αr . For each eigenvalue λj , where 1 ≤ j ≤ r, we define Rj = { α = (α1 , . . . , αr ) ∈ Nr : |α| ≥ 2 and λj = λ1 α1 + · · · + λr αr } , which is called the resonant set associated to λj . 36

7.4

Notations and definitions

The method described in Section 3 (proposed by Kang and Krener [52], and then followed by Kang [48], [49] and by the authors [77], [81]) is to analyze step by step the action of the transformation Γ∞ on the system Σ∞ . In the controllable case, it consists of bringing the linear part of the system into the Brunovsk´ y canonical form and then applying, step by step, homogeneous feedback transformations of the form Γm :

x = ξ + φ[m] (ξ) u = v + α[m] (ξ) + β [m−1] (ξ)v

in order to normalize the homogeneous part of degree m of the system. The advantage of this method, in the controllable case, follows from the fact that homogeneous transformations Γm leave invariant all terms of degree smaller than m of the system. The situation gets very different in the uncontrollable case with the modified notion of homogeneity. In order to see the difference, we analyze the action of Γm on the system Σ∞ , given by (7.1). Let ˜ ∞ : x˙ = F x + Gv + f˜[0] (x) + Σ

∞ X

(f˜[m] (x) + g˜[m−1] (x)v)

m=1 ∞

m

be the system Σ transformed by Γ . Recall that for both systems, the first r components of the state correspond to the uncontrollable part and that the degree of homogeneity (for all terms of the systems and for the transformation Γm ) is computed with respect to the controllable variables, that is, the last n − r variables only. We can observe the three following undesirable phenomena. Firstly, notice that the homogeneous transformation Γm does not preserve homogeneous (with respect to linearly controllable variables) terms of degree smaller than m. It only preserves terms of degree smaller than m − 1, that is, f˜[k] = f [k] and g˜[k−1] = g [k−1] , for any 0 ≤ k ≤ m − 2, while transforms those of degree m − 1 as follows [m−2]



=g

[m−2]

and

f˜[m−1] = f [m−1] +

n X ∂φ[m] [0] f . ∂ξi i i=r+1

Notice that, when comparing the homogeneous parts of the same degree k of two systems, we have to compare the homogeneous parts of degree k of the drifts and the homogeneous parts of degree k − 1 of control vector fields. Indeed, since the homogeneity is considered with respect to the state and the control, the homogeneous part of degree k of the system is represented by the homogeneous part f [k] , of degree k, of the drift and the homogeneous part g [k−1] , of degree k − 1, of the control vector field multiplied by the control u. Secondly, Γm transforms the homogeneous part (f [m] , g [m−1] ), of degree m, not according to a homological equation, but in such a way that ¶ n µ [1] [m] X ∂f ∂φ [1] [m] [m] [m] [m] [m] f − φ f˜ = f + [Aξ, φ ] + Bα + ∂ξi i ∂ξi i i=r+1 ¶ r µ X ∂φ[m] [0] ∂f [0] [m] + fi − φi + g [0] α[m] ∂ξ ∂ξ i i i=1 g˜[m−1] = g [m−1] + adB φ[m] + Bβ [m−1] + g [0] β [m−1] +

n P i=r+1

37

∂φ[m] [0] gi . ∂ξi

Thirdly, the Lie bracket of two homogeneous vectors fields f [m] and g [k] is given by ¶ n µ X ∂g [k] [m] ∂f [m] [k] [m] [k] [f , g ](ξ) = fi (ξ) − gi (ξ) ∂ξ ∂ξ i i i=1 and thus fails, in general, to be homogeneous of degree m + k − 1 because the terms

∂g [k] [m] f (ξ) ∂ξi i

∂f [m] [k] g (ξ), for 1 ≤ i ≤ r, are, in general, homogeneous of degree m + k. ∂ξi i Those three inconveniences are caused only by the fact that differentiating with respect to the variables ξ1 , . . . , ξr does not decrease the degree (in particular, by the presence of terms of degree 0 with respect to the variables ξr+1 , . . . , ξn ). To overcome this, we define, for any m ≥ 0, and

[m−1]

f hmi = (f1

[m−1]

g hmi = (g1

[−1]

[m−1]

, . . . , gr

[m−1]

φhmi = (φ1

[m−1]

, . . . , fr

[−1]

[m]

[m]

[m]

, gr+1 , . . . , gn )T

[m−1]

, . . . , φr

[m]

, fr+1 , . . . , fn )T [m]

[m]

, φr+1 , . . . , φn )T ,

[−1]

where, for any 1 ≤ i ≤ r, we set fi = gi = φi = 0. Control systems, vector fields, feedback transformations, etc., that are homogeneous with respect to the above defined weights, will be called weighted homogeneous. Notice that just defined weighted homogeneity is related with the decomposition of the state-space Rn into uncontrollable and controllable parts and therefore it does not apply to applications with values in R. In particular, for real-valued homogeneous polynomials h[m] we will write hhmi = h[m] . We will denote by P [m] (ξ) the space of homogeneous polynomials of degree m of the variables ξr+1 , . . . , ξn (with coefficients depending on ξ1 , . . . , ξr ) and by P ≥m (ξ) the space of formal power series of the variables ξr+1 , . . . , ξn (with coefficients depending on ξ1 , . . . , ξr ) starting from terms of degree m. We will denote by V hmi (ξ) the space of weighted hmi-homogeneous vector fields, that is, the space of vector fields whose first r components are in P [m−1] (ξ) and the last n − r components are in P [m] (ξ). Moreover, V h≥mi (ξ) will denote the space of vector fields formal power series whose first r components are in P ≥m−1 (ξ) and the last n − r components are in P ≥m (ξ).

7.5

Weighted homogeneous systems

Applying a linear feedback transformation, we can bring the linear approximation (F, G) of the system into Jordan-Brunovsk´ y canonical form (A, B), that is, the uncontrollable part, of dimension r, is in the Jordan form and the controllable part, of dimension n−r, in the Brunovsk´ y form. Notice, however, that contrary to the controllable case (where there are no zero degree terms while the terms of degree one are just linear terms that we bring to the Brunovsk´ y canonical form), after having normalized linear terms of an uncontrollable system, we are still left with weighted homogeneous terms of degree zero and one. We can normalize them as follows. Proposition 7.2 Consider the system Σh≤1i : ξ˙ = Aξ + Bu + f h0i (ξ) + f h1i (ξ) + g h0i (ξ)u, where (A, B) is in the Jordan-Brunovsk´y canonical form. 38

(i) There exists a smooth feedback transformation of the form Γh≤1i :

x = ξ + φh0i (ξ) + φh1i (ξ) u = v + αh0i (ξ) + αh1i (ξ) + β h0i (ξ)v,

(7.3)

which takes the system Σh≤1i into the system ˜ h≤1i : x˙ = Ax + Bv + f˜h1i (x), Σ

(7.4)

h1i modulo terms in V h≥2i , where the vector field f˜h1i satisfies f˜j = 0 for r + 1 ≤ j ≤ n. (ii) Assume that all eigenvalues λ1 , . . . , λr of the Jordan-Brunovsk´y canonical form (A, B) are real and distinct (in particular all σj = 0). Then a formal feedback transformation Γh1i of ˜ h≤1i defined by (7.4), can be the form (7.3), which takes the system Σh≤1i into the system Σ h1i chosen in such a way that the vector field f˜ is replaced by f¯h1i (x) whose components are given by  P γjα xα1 1 · · · xαr r if 1 ≤ j ≤ r  h1i α∈Rj f¯j (x) = (7.5)  0 if r + 1 ≤ j ≤ n. h1i The existence of a transformation x = ξ + φh1i (ξ) yielding the form of f¯j (x) in (ii) is an immediate consequence of Theorem 2.3. Now we will study the action of the weighted homogeneous feedback

Γhmi :

x = ξ + φhmi (ξ) u = v + αhmi (ξ) + β hm−1i (ξ)v

(7.6)

on the following weighted homogeneous system Σhmi : ξ˙ = Aξ + Bu + f¯h1i (ξ) + f hmi (ξ) + g hm−1i (ξ)u,

(7.7)

where, because of Proposition 7.2, we assume that f¯j is of the form (7.5). Notice that a weighted homogeneous feedback is a smooth feedback, actually, it depends polynomially on the variables ξr+1 , . . . , ξn and smoothly on the variables ξ1 , . . . , ξr . ˜ hmi given by Consider another weighted homogeneous system Σ h1i

˜ hmi : x˙ = Ax + Bv + f˜h1i (x) + f˜hmi (x) + g˜hm−1i (x)v, Σ

(7.8)

˜ hmi , and we will denote it by where f˜h1i = f¯h1i . We will say that Γhmi transforms Σhmi into Σ hmi ˜ hmi , if Γhmi transforms Σhmi into Γ∗ (Σhmi ) = Σ x˙ = Ax + Bv + f˜h1i (x) + f˜hmi (x) + g˜hm−1i (x)v + Rh≥m+1i (x, v), where Rh≥m+1i (x, v) ∈ V h≥m+1i (x, v). Recall that λj , for 1 ≤ j ≤ r, are the eigenvalues of the uncontrollable part of the linear approximation and that σj for 2 ≤ j ≤ r, define the corresponding Jordan form (see Section 7.3). We define additionally σr+1 = 0 and for any r + 1 ≤ j ≤ n − 1, we put λj = 0 and σj+1 = 1. Analysis of weighted homogeneous systems is based on the following result which generalizes, to the uncontrollable case, that proved by Kang [48] (and recalled in Proposition 3.1): 39

Proposition 7.3 For any m ≥ 2, the feedback transformation Γhmi , of the form (7.6), brings ˜ hmi , given by (7.8), if and only if the following relations the system Σhmi , given by (7.7), into Σ  hmi hmi hmi hmi ˜hmi   LAξ+f h1i φj − λj φj − σj+1 φj+1 = fj − fj       hmi hm−1i hm−1i  LB φj = g˜j − gj  (7.9)  hmi hmi hmi  hmi  LAξ+f h1i φn + α = f˜n − fn        hmi hm−1i hm−1i LB φn + β hm−1i = g˜n − gn hold for any 1 ≤ j ≤ n − 1. This proposition can be viewed as the weighted control homological equation for systems with uncontrollable linearization (its proof follows the same line as that of Kang [48] for the standard control homological equation (CHE)). Once again, the solving of a system of 1-st order partial differential equations may be avoided if we perform the analysis step by step, and thus reduce ˜ hmi (with uncontrollable linearization) to the feedback equivalence of two systems Σhmi and Σ solving the algebraic system (7.9). The following result gives our normal form for weighted homogeneous systems with uncontrollable linearization. Recall the notation πi (x) = (x1 , . . . , xi ). Theorem 7.4 For any m ≥ 2, there exists a weighted feedback transformation Γhmi , that transforms the weighted homogeneous system Σhmi , given by (7.7), into its weighted homogeneous normal form hmi ΣN F : x˙ = Ax + Bv + f¯h1i (x) + f¯hmi (x), (7.10) where f¯h1i (x) is given by Proposition 7.3 (ii) and the vector field f¯hmi (x) satisfies  n P hm−3i m−1  x2i Qj,i (πi (x)) if 1 ≤ j ≤ r, x S (π (x)) +  j,m r r+1   i=r+2   hmi n P 2 hm−2i (7.11) f¯j (x) = xi Pj,i (πi (x)) if r + 1 ≤ j ≤ n − 2,    i=j+2    0 if n − 1 ≤ j ≤ n, hm−2i

where Sj,m (πr (x)) ∈ C0∞ (Rr are C ∞ -functions of the variables x1 , . . . , xr and the functions Pj,i hm−3i and Qj,i are homogeneous polynomials, respectively of degree m − 2 and m − 3, of the variables xr+1 , . . . , xi , with coefficients in C0∞ (Rr ). The proof of Theorem 7.4 is based on Theorem 7.7, stated in Section 7.7, which explicitly gives hmi transformations bringing Σhmi into its normal form ΣN F . The above normal form generalizes [m] to the uncontrollable case the normal form ΣN F of Kang [48] for systems with controllable linearization, which we stated in Theorem 3.2. It can also be viewed as a generalization of the normal form obtained by Kang [49] in the uncontrollable case for second order terms. Other normal forms, for systems with uncontrollable linearization, have been obtained in [76], and in [56], [57] and [58] (for third order terms). Notice, however, that the latters coincide neither with h2i h3i our normal form ΣN F nor ΣN F because of different weights we use, as we explain in the following example. 40

Example 7.5 Consider the system ξ˙ = f (ξ) + g(ξ)u on R3 and assume that the dimension of the linearly controllable subsystem is 2 and that the linear part is in the Jordan-Brunovsk´ y canonical form. The homogenous system Σ[2] (homogeneity being calculated with respect to all variables ξ1 , ξ2 , ξ3 ) is [2] [1] ξ˙1 = λξ1 + f1 (ξ) + g1 (ξ)u [2] [1] ξ˙2 = ξ3 + f2 (ξ) + g2 (ξ)u [2] [1] ξ˙3 = u + f (ξ) + g (ξ)u. 3

3

The linearly uncontrollable subsystem is of dimension one (with ξ1 being the linearly uncontrollable variable and (ξ2 , ξ2 )T being the linearly controllable variables) and the resonant set associated with the eigenvalue λ is empty if and only if λ 6= 0. Kang [49] proved that Σ[2] is [2] equivalent via a homogeneous feedback Γ2 to the following normal form ΣN F : x˙ 1 = λx1 + γ2 x21 + x2 s1,2 (x1 ) + x23 q1,3 x˙ 2 = x3 x˙ 3 = u where γ2 = 0 if λ 6= 0 (no resonances) and γ2 ∈ R if λ = 0. Moreover, s1,2 is a linear function of x1 and q1,3 is a constant. h2i h3i Now we will compare this normal form with the normal forms ΣN F and ΣN F . To this end, we start with h1i ξ˙1 = λξ1 + f1 (ξ) Σh1i : ξ˙2 = ξ3 + f h1i (ξ) ξ˙3 = u + h1i

2 h1i f3 (ξ),

h1i

where f2 (ξ) and f3 (ξ) are linear functions with respect to ξ2 and ξ3 with coefficients that are h1i arbitrary functions of ξ1 while f1 (ξ) is an arbitrary function of ξ1 whose development at zero h1i starts with quadratic terms. Now Proposition 7.2 (i) implies that we can annihilate f2 and h1i h1i f3 . If λ 6= 0, then there are no resonances so we can also annihilate f1 and without loss of h1i h1i generality we can assume that f1 (ξ) is in the normal form f¯1 (ξ) = 0 assured by Proposition 7.2. h1i h1i Otherwise, that is, if λ = 0 then all terms of f1 are resonant so we can assume that f1 is in P P h1i ∞ ∂ i i ∂ the normal form f¯1 (ξ) = ( ∞ i=2 γi ξ1 ) ∂ξ1 , where γi ∈ R. Actually, the vector field ( i=2 γi ξ1 ) ∂ξ1 can be normalized by a local diffeomorphism around ξ1 = 0 into the form (γp xp1 + γ2p−1 x2p−1 ) ∂x∂ 1 , 1 but we will not use it. Consider the system Σh2i : ξ˙ = Aξ + Bv + f¯h1i (ξ) + f h2i (ξ) + g h1i (ξ)v, where (A, B) is in the Jordan-Brunovsk´ y canonical form and f¯h1i is in the normal form described h1i h1i above, that is Aξ + f¯h1i (ξ) + Bu = (λξ1 + f¯1 (ξ)) ∂ξ∂1 + ξ3 ∂ξ∂2 + u ∂ξ∂3 , where f1 (ξ) equals 0 P∞ h2i h2i i or i=2 γi ξ1 (depending on λ). Moreover, the components f2 (ξ) and f3 (ξ) are quadratic h2i h1i h1i functions of ξ2 , ξ3 , the components f1 (ξ), g2 (ξ) and g3 (ξ) are linear functions of ξ2 , ξ3 (all h1i coefficients depending on ξ1 ) and g1 (ξ) depends on ξ1 only. By a weighted homogenous (of 41

degree 2 with respect to ξ2 , ξ3 ) feedback transformation Γh2i we can bring Σh2i into the normal form h1i x˙ 1 = λx1 + f¯1 (x) + x2 S1,2 (x1 ) h2i ΣN F : x˙ 2 = x3 x˙ 3 = u, where S1,2 is an arbitrary function of x1 . Now we will consider Σh3i : ξ˙ = Aξ + Bu + f¯h1i (ξ) + f h3i (ξ) + g h2i (ξ)u, where (A, B) is in the Jordan-Brunovsk´ y canonical form and f¯h1i is in the normal form described h3i h3i above. Moreover, the components f2 (ξ) and f3 (ξ) are cubic functions of ξ2 , ξ3 , the components h3i h2i h2i h2i f1 (ξ), g2 (ξ) and g3 (ξ) are quadratic functions of ξ2 , ξ3 , and g1 (ξ) is a linear function of ξ2 , ξ3 (all coefficients depending on ξ1 ). By a weighted homogenous (of degree 3 with respect to ξ2 , ξ3 ) feedback transformation Γh3i we can bring Σh3i into the normal form h3i

ΣN F

h1i x˙ 1 = λx1 + f¯1 (x) + x22 S1,3 (x1 ) + x23 Q1,3 (x1 ) : x˙ 2 = x3 x˙ 3 = v,

where S1,3 and Q1,3 are arbitrary functions of x1 . [2] Observe that the term x2 s1,2 (x1 ) of the Kang normal form ΣN F , where s1,2 (x1 ) is linear, shows h2i up as the first term in the development of x2 S1,2 (x1 ) of ΣN F while the term x23 q1,3 (x1 ), where q1,3 is constant, shows up as the first term in the development of x23 Q1,3 (x1 ) but both S1,2 (x1 ) and Q1,3 (x1 ) contain, in general, terms of arbitrary degrees (except for constant terms in S1,2 ). [2] h2i This example illustrate mutual differences between the form ΣN F of Kang and of ours ΣN F and h3i ΣN F . ¤

7.6

Weighted homogeneous invariants

In this section, we will define weighted homogeneous invariants ahmij,i+2 of weighted homogeneous systems Σhmi under weighted homogeneous feedback transformations Γhmi and we will state for them a result of the authors [79], [83] generalizing, to the uncontrollable case, a result established by Kang [48] in the controllable case. Consider the weighted homogeneous system (7.7). For any i ≥ 0, let us define the vector field Xim−1 (ξ) = (−1)i adiAξ+f h1i (ξ)+f hmi (ξ) (B + g hm−1i (ξ)) hm−1i

and let Xi be the homogeneous part of degree hm − 1i of Xim−1 . It means that the first r components are homogeneous of degree m − 2 and the last n − r components homogeneous of degree m − 1 with respect to the variables ξr+1 , . . . , ξn . One can easily check that à ! i X hm−1i i−k hmi (ξ) = (−1)i adiAξ+f h1i g hm−1i + (−1)k adAξ+f . Xi h1i adAk−1 B f k=1

42

Define the set of indices ∆r = ∆1r ∪ ∆2r ⊂ N × N by taking ∆1r = { (j, i) ∈ N × N :

1 ≤ j ≤ r and 0 ≤ i ≤ n − r − 1 } ,

∆2r = { (j, i) ∈ N × N :

r + 1 ≤ j ≤ n − 2 and 0 ≤ i ≤ n − j − 2 } .

For any r + 1 ≤ k ≤ n define the following subspaces © ª Wk = (ξ1 , . . . , ξr , ξr+1 , . . . , ξn )T ∈ Rr × Rn−r : ξk+1 = · · · = ξn = 0 , and let πk (ξ) denote the projection on Wk , that is, πk (ξ) = (ξ1 , . . . , ξr , ξr+1 , . . . , ξk , 0, . . . , 0)T . For any 1 ≤ j ≤ n, we denote by Cj = (0, . . . , 0, 1, 0, . . . , 0) the row vector in Rn whose all components are zero except the j th component which equals 1. For any (j, i) ∈ ∆1r (resp. (j, i) ∈ ∆2r ), we define ahmij,i+2 (ξ) as the homogeneous part of degree hm − 3i (resp. of degree hm − 2i ) of the function £ ¤ m−1 Cj Xim−1 , Xi+1 (πn−i (ξ)). One can easily establish that ³ ´ hm−1i hm−1i ahmij,i+2 (ξ) = Cj adAi B Xi+1 − adAi+1 B Xi (πn−i (ξ)). The functions ahmij,i+2 thus defined will be called weighted homogeneous hmi-invariants of the weighted homogeneous system Σhmi . ˜ hmi given by (7.8) and denote by Consider, along with Σhmi defined by (7.7), the system Σ hmij,i+2 a ˜ the weighted homogeneous hmi-invariants of the latter system. The following result, which generalizes that obtained by Kang [48] for systems with controllable linearization, asserts that the weighted homogeneous hmi-invariants ahmij,i+2 are complete invariants of weighted homogeneous feedback and, moreover, illustrates their meaning for the hmi normal form ΣN F . Theorem 7.6 For any m ≥ 2 we have the following properties. ˜ hmi , given by (7.8), are (i) Two weighted homogeneous systems Σhmi , given by (7.7), and Σ equivalent via a weighted homogeneous feedback Γhmi if and only if, for any (j, i) ∈ ∆r , we have ahmij,i+2 = a ˜hmij,i+2 . hmi

¯hmij,i+2 of the weighted homogeneous normal form ΣN F , defined by (ii) The hmi-invariants a (7.10)-(7.11), are given by hmi ∂ 2 f¯j hmij,i+2 (πi (x)). a ¯ (ξ) = ∂x2i

43

7.7

Explicit normalizing transformations

We will give in this section explicit weighted homogeneous transformations bringing the syshmi tem Σhmi into its normal form ΣN F . Their interest is two-fold. Firstly, they are easily computable (via differentiation and integration of polynomials). Secondly, the proof of Theorem 7.4 giving hmi the normal form ΣN F is based on them. hm−1i

by

For any r + 1 ≤ i ≤ n and any 1 ≤ j < i ≤ n, the homogeneous polynomial ψj,i hm−1i

ψj,i

hm−1i

= Cj Xn−i . hm−1i

For any r + 1 ≤ i ≤ j ≤ n, we define recursively the polynomials ψj,i hm−1i

ψj,r

is defined

by setting

hm−1i

= ψr+1,r+1 = 0

and by taking hmi

hm−1i ψj,i

∂fj−1 hm−1i hm−1i = + LAξ+f h1i ψj−1,i + ψj−1,i−1 + ∂ξi

Z

hm−1i

Notice that the degree of the homogeneous polynomial ψj,i m − 1 if r + 1 ≤ j ≤ n.

ξi 0

hm−1i

∂ψj−1,i dξi . ∂ξi−1

is either m − 2 if 1 ≤ j ≤ r or

Consider the weighted homogeneous feedback transformation Γhmi :

x = ξ + φhmi (ξ) u = v + αhmi (ξ) + β hm−1i (ξ)v

defined, for any 1 ≤ j ≤ n, by hmi φj (ξ)

=

n Z X

ξi 0

i=r+1

hm−1i

ψj,i

(ξ¯i )dξi ,

hmi

hmi

αhmi (ξ) = −fn (ξ) − LAξ+f h1i φn (ξ), hm−1i

β hm−1i (ξ) = −gn

(7.12)

hmi

(ξ) − LB φn (ξ).

We have the following result: Theorem 7.7 For any m ≥ 2, the weighted homogeneous feedback transformation Γhmi , defined by (7.12), brings the weighted homogeneous system Σhmi , given by (7.7), into its weighted hmi homogeneous normal form ΣN F , defined by (7.10).

7.8

Weighted normal form for single-input systems with uncontrollable linearization

We will present in this section our main result giving a normal form under a formal feedback transformation Γ∞ (see Section 3 for some comments on formal feedback) of any single-input control system (with controllable or uncontrollable linearization). For any 1 ≤ i ≤ n we denote πi (x) = (x1 , . . . , xi )T . 44

Theorem 7.8 Consider the system Σ∞ , given by (7.1) and assume that all eigenvalues of the uncontrollable part of the linear approximation are real. There exists a formal feedback transformation Γ∞ of the form (7.2), which brings the system Σ∞ , given by (7.1), into its normal form Σ∞ ˙ = Ax + Bv + f¯(x), NF : x given by

  λ x + σj xj+1 + f¯j (x) if 1 ≤ j ≤ r,   j j xj+1 + f¯j (x) if r + 1 ≤ j ≤ n − 1, x˙ j =    v if j = n,

(7.13)

where

 n P α α1 P  αr  γ x · · · x + x S (π(x) ) + x2i Qj,i (πi (x)), if 1 ≤ j ≤ r,  r+1 j r+1 j 1 r   i=r+2   α∈Rj n P 2 f¯j (x) = xi Pj,i (πi (x)), if r + 1 ≤ j ≤ n − 2,    i=j+2     0, if n − 1 ≤ j ≤ n, (7.14) α where Pj,i , Qj,i , and Sj are formal power series of the indicated variables and γj ∈ R. Remark 7.9 In the general case of complex eigenvalues of the uncontrollable part, for each complex eigenvalue λj = αj + iβj , we replace the expression for x˙ j by µ ¶ µ ¶µ ¶ µ ¶ x˙ j,1 αj βj xj,1 xj,1 = + Ξj + f¯j , x˙ j,2 −βj αj xj,2 xj,2 α α T where f¯j = (f¯j,1 , f¯j,2 )T , for 1 ≤ j ≤ r, being defined by formula (7.14), with γjα = (γj,1 , γj,2 ) ∈ 2 2 R , and Sj and Qj,i being R -valued formal power series of the indicated variables. Of course, ¯j , the resonant set of a complex eigenvalue λj is the same as that corresponding to its conjugate λ which explains why we have the same Rj for both components xj,1 and xj,2 . ¤

Notice that the above normal form is a natural combination of the two extreme cases: that of dynamical systems and that of systems with controllable linearization. Indeed, if r = n, that is we deal with a dynamical system, then the coordinates (xr+1 , . . . , xn ) are not present and the normal form Σ∞ ˙ = Jx + f¯(x) containing resonant terms N F reduces to a dynamical system x P α only, that is f¯j (x) = γjα x1 1 · · · xαr r , for 1 ≤ j ≤ n. This is, of course, Poincar´e normal α∈Rj

form of a vector field under a formal diffeomorphism (see, e.g., [1] and Theorem 2.4). On the other hand, if r = 0, that is the linear approximation is controllable, the coordinates (x1 , . . . , xr ) [m] are not present and our normal form reduces to ΣN F of Kang [48] (see Section 3), for which n P 2 f¯j (x) = x Pj,i (πi (x)), for 1 ≤ j ≤ n − 2 and f¯j (x) = 0 otherwise. i=j+2

i

Another normal form for nonlinear single-input systems with uncontrollable linearization was obtained by the authors in [76] (see also [75]) and by Krener et al in [56], [57] and [58]. Notice, however, that those normal forms differ substantially from the one proposed in this paper and in [79]. Indeed, in the presented approach the homogeneity is calculated with respect to the 45

linearly controllable variables while it is calculated with respect to all variables in [76], [56], [57] and [58].

7.9

Example

Example 7.10 (Kapitsa Pendulum) We consider in this example the Kapitsa Pendulum whose equations (see [5] et [17]) are given by α˙ = p + wl sin α 2 p˙ = (gl − wl2 cos α) sin α − wl p cos α z˙ = w,

(7.15)

where α denotes the angle of the pendulum with the vertical z-axis, w the velocity of the suspension point z, and p is proportional to the generalized impulsion; g is the gravity constant and l the length of the pendulum. We assume to control the acceleration a = w. ˙ Introduce the coordinate system (ξ1 , ξ2 , ξ3 , ξ4 ) = (α, p, z/l, w/l) and take u = a/l as the control. The system (7.15) considered around an equilibrium point (α0 , p0 , z0 , u0 ) = (kπ, 0, 0, 0), where k ∈ Z, rewrites as ξ˙1 ξ˙2 ξ˙3 ξ˙4

= = = =

ξ2 + ξ1 ξ4 + ξ4 T1 (ξ1 ) ²g0 ξ1 − ξ2 ξ4 + ξ2 ξ4 T2 (ξ1 ) + ξ42 Q2 (ξ1 ) + R2 (ξ1 ) ξ4 u,

(7.16)

where g0 = g /l , ² = ±1 and T1 , T2 , R2 are analytic functions whose 1-jets at (kπ, 0, 0, 0) vanish and Q2 is an analytic function vanishing at (kπ, 0, 0, 0). Above, the case ² = 1 corresponds to α0 = 2nπ and the case ² = −1 to α0 = (2n + 1)π. One can easily check that the quadratic feedback transformation y1 = ξ1 − ξ1 ξ3 y = ξ2 + ξ2 ξ3 Γ2 : 2 y3 = ξ3 y4 = ξ4 brings the system (7.16) into the system y˙ 1 y˙ 2 y˙ 3 y˙ 4

= = = =

y2 + y2 y3 S˜1 (y3 ) + y4 T˜1 (y1 , y3 ) ˜ 2 (y1 , y3 ) + R ˜ 2 (y1 ) ²g0 y1 + y1 y3 S˜2 (y1 , y3 ) + y4 T˜2 (y1 , y2 , y3 ) + y42 Q y4 u,

˜ 1, Q ˜ 2, R ˜ 2 , S˜2 , T˜1 , and T˜2 are analytic functions. where Q Since the vector field defined in R3 by f = T˜1 (y1 , y3 )∂/∂y1 + T˜2 (y1 , y2 , y3 )∂/∂y2 + ∂/∂y3

46

(7.17)

does not vanish at (0, 0, 0) ∈ R3 (resp. at (π, 0, 0) ∈ R3 ), there exists, in a neighborhood of (0, 0, 0) ∈ R3 (resp. of (π, 0, 0) ∈ R3 ), an analytic transformation x = φ(y) of the form x1 = φ1 (y1 , y3 ) x2 = φ2 (y1 , y2 , y3 ) x3 = y 3 such that (φ∗ f )(x) = ∂/∂x3 . This latter transformation, completed with x4 = y4 and u = v, brings the system (7.17) into the normal form (compare with Theorem 7.8) ¯ 1 (x1 , x2 ) + x3 S¯1 (π3 (x)) x˙ 1 = x2 + R ¯ 2 (x1 , x2 ) + x3 S¯2 (π3 (x)) + x2 Q ¯ x˙ 2 = ²g0 x1 + R 4 2 (π3 (x)) (7.18) x˙ 3 = x4 x˙ 4 = v, where π3 (x) = (x1 , x2 , x3 ). Clearly, the dimension of the linearly controllable part of (7.16), that is that of (7.18), equals 2, which means that r = 2. √ In the case ² = 1, the eigenvalues of the uncontrollable linear part are λ1 = −λ2 = g0 while √ in the case ² = −1, they are given by λ1 = −λ2 = i g0 . In both cases, those eigenvalues are resonant and satisfy, for any m ≥ 2, the relations λ1 = mλ1 + (m − 1)λ2

and λ2 = (m − 1)λ1 + mλ2 .

Applying Poincar´e’s method (see [1]), we get rid, by a formal diffeomorphism in the space (x1 , x2 ), of all nonresonnant terms of the dynamical system ¯ 1 (x1 , x2 ) x˙ 1 = x2 + R ¯ 2 (x1 , x2 ) x˙ 2 = ²g0 x1 + R and thus we transform the system, for ² = 1 and ² = −1, into one of the following normal forms. √ Set λ = g0 . For ² = 1, which is the case of real eigenvalues, the normal form is given by (compare with Theorem 7.8) ∞ P ¯ 1 (π3 (x)) x˙ 1 = λx1 + am x1 (x1 x2 )m−1 + x3 S¯1 (π3 (x)) + x24 Q m=2 ∞ P

x˙ 2 = −λx2 + x˙ 3 = x4 x˙ 4 = v.

¯ 2 (π3 (x)) bm x2 (x1 x2 )m−1 + x3 S¯2 (π3 (x)) + x24 Q

m=2

(7.19)

For ² = −1, which corresponds to the case of complex eigenvalues (see Remark following Theorem 7.8, the normal form is given by ∞ P ˜ 1 (π3 (x)) x˙ 1 = λx2 + (cm x1 + dm x2 )(x21 + x22 )m−1 + x3 S˜1 (π3 (x)) + x24 Q m=2 ∞ P

x˙ 2 = −λx1 + x˙ 3 = x4 x˙ 4 = v.

˜ 2 (π3 (x)) (−dm x1 + cm x2 )(x21 + x22 )m−1 + x3 S˜2 (π3 (x)) + x24 Q

m=2

47

(7.20)

¯ 1 , S¯2 , and Q ¯ 2 on the one hand side, and the functions S˜1 , Q ˜ 1 , S˜2 , Above, the functions S¯1 , Q ˜ and Q2 on the other hand side, are formal power series which, in general, are different from the objects denoted earlier by the same symbols. Notice that we transform the original system into its normal form (7.18) using analytic feedback transformations and that the only passage defined by a formal feedback transformation is that transforming the system (7.18) into (7.19) (resp. (7.20)). ¤

8

Normal forms for multi-input nonlinear control systems

In this section we give a generalization of normal forms obtained in Section 3 for multi-input nonlinear control systems with controllable linearization (see [87] for more details). Normal forms for two-input nonlinear control systems has been obtained previously by the authors [85], and will be derived here as a particular case. The general case of multi-input systems with uncontrollable linearization will appear in [88]. We will illustrate normal forms in this section by considering three physical examples: A model of a crane in Example 8.5, a model of a planar vertical takeoff and landing aircraft as Example 8.6, and finally the Example 8.7 deals with a prototype of a wireless multi-vehicle testbed [13] and [15]. Consider control systems of the form Π : ξ˙ = F (ξ, u),

ξ(·) ∈ Rn , u(·) = (u1 (·), . . . , up (·))T ∈ Rp

around the equilibrium point (0, 0) ∈ Rn × Rp , that is, f (0, 0) = 0, and denote by Π[1] : ξ˙ = Fξ + Gu = Fξ + G1 u1 + · · · + Gp up , its linearization at this point, where F=

∂F ∂F ∂F (0, 0), G1 = (0, 0), . . . , Gp = (0, 0) . ∂ξ ∂u1 ∂up

We will assume for simplicity (see [87], [88] for general case) that G1 ∧ · · · ∧ Gp 6= 0, and the linearization is controllable, that is © ª span Fi Gk : 0 ≤ i ≤ n − 1, 1 ≤ k ≤ p = Rn . Let (r1 , . . . , rp ), 1 ≤ r1 ≤ · · · ≤ rp = r, be the largest, in the lexicographic ordering, p-tuple of positive integers, with r1 + · · · + rp = n, such that © ª span Fi Gk : 0 ≤ i ≤ rk − 1, 1 ≤ k ≤ p = Rn . (8.1) Without loss of generality we can assume that the linearization is in Brunovsk´ y canonical form [1] ΠCF : ξ˙ = Aξ + Bu = Aξ + B1 u1 + · · · + Bp up ,

where A = diag(A1 , . . . , Ap ), B = (B1 , . . . , Bp ) = diag(b1 , . . . , bp ), that is,     b1 · · · 0 A1 · · · 0     , B =  ... . . . ...  A =  ... . . . ...  0 · · · bp n×p 0 · · · Ap n×n 48

(8.2)

with (Ak , bk ) in Brunovsk´ y single-input canonical forms of dimensions rk , for any 1 ≤ k ≤ p. With the p-tuple (r1 , . . . , rp ) we associate the p-tuple (d1 , . . . , dp ) of nonnegative integers, 0 = dp ≤ · · · ≤ d1 ≤ r − 1, such that r1 + d1 = · · · = rp + dp = r. Our aim is to give a normal form of feedback classification of such systems under invertible feedback transformations of the form x = φ(ξ) u = γ(ξ, v) ,

Υ :

where φ(0) = 0 and γ(0, 0) = 0. We study, step by step, the action of the Taylor series expansion Υ∞ of the feedback transformation Υ, given by x = φ(ξ) = ξ + Υ∞ :

∞ P

φ[m] (ξ)

m=2 ∞ P

u = γ(ξ, v) = v +

γ [m] (ξ, v) .

(8.3)

m=2

on the Taylor series expansion Π Π





of the system Π, given by : ξ˙ = Aξ + Bu +

∞ X

f [m] (ξ, u).

(8.4)

m=2

Throughout this section, in particular in formulas (8.3) and (8.4), the homogeneity of f [m] and γ [m] will be taken with respect to the variables ξ, v and ξ, u respectively.

8.1

Non-affine normal forms

Let 1 ≤ s ≤ t ≤ p. We denote by xs = (xs,ds +1 , . . . , xs,r )T ,

xs,r+1 = vs

and we set x¯s,i = (xs,ds +1 , . . . , xs,i )T for any ds + 1 ≤ i ≤ r + 1. We also define the projections s πt,i (x) = (¯ xT1,i , . . . , x¯Ts,i , x¯Ts+1,i−1 , . . . , x¯Tt−1,i−1 , x¯Tt,i , x¯Tt+1,i−1 , . . . , x¯Tp,i−1 )T ,

where x¯s,i is empty whenever 0 ≤ i ≤ ds . Our main result for multi-input nonlinear control systems with controllable linearization is as following. Theorem 8.1 The control system Π∞ , defined by (8.4), is feedback equivalent, by a formal feedback transformation Υ∞ of the form (8.3), to the normal form Π∞ ˙ = Ax + Bv + NF : x

∞ X

f¯[m] (x, v) ,

m=2

where for any m ≥ 2, we have f¯[m] (x, v) =

p r−1 X X k=1 j=dk

∂ k[m] , f¯j (x, v) ∂xk,j +1

49

(8.5)

with, k[m] f¯j (x, v) =

r+1 X

X

1≤s≤t≤p i=j+2

r+1 X

X

k[m−2]

s xs,i xt,i Pj,i,s,t (πt,i (x)) +

k[m−2]

s xs,i xt,i−1 Qj,i,s,t (πt,i−1 (x)) (8.6)

1≤s