A Globally Exponentially Convergent Immersion
and Invariance Speed Observer for n Degrees of
Freedom Mechanical Systems with constraints
Alessandro Astolfi
1
, Romeo Ortega
2
and Aneesh Venkatraman
3
1
Dept. of Electrical and Electronic Engineering, Imperial College London
London, SW7 2AZ, U.K.
and DISP, University of Roma ”Tor Vergata”, Via del Politecnico 1, 00133 Rome, Italy
2
Laboratoire des Signaux et Syst`emes, CNRS–SUPELEC, 91192, Gif–sur–Yvette, France
3
Institute of Mathematics and Computing Science, University of Groningen
PO Box 407, 9700 AK, Groningen, The Netherlands
Abstract. The problem of velocity estimation for mechanical systems is of great
practical interest. Although many partial solutions have been reported in the liter-
ature the basic question of whether it is possible to design a globally convergent
speed observer for general n degrees of freedom mechanical systems remains
open. In this paper an affirmative answer to the question is given by proving the
existence of a 3n + 1–dimensional globally exponentially convergent speed ob-
server. Instrumental for the construction of the speed observer is the use of the
Immersion and Invariance technique, in which the observer design problem is re-
cast as a problem of rendering attractive and invariant a manifold defined in the
extended state–space of the plant and the observer.
Notation. For general mappings S : R
n
× R
p
R
q
, (x, ζ) 7→ S we define
x
S(x, ζ) :=
S(x,ζ)
x
and
ζ
S(x, ζ) :=
S(x,ζ)
ζ
. For brevity, when clear from
the context, the subindex of and, in general, the arguments of all the functions
are omitted.
1 Problem Formulation
We consider general n degree of freedom mechanical systems with nonholonomic con-
straints described in Lagrangian form by [11], [13],
M(q)¨q + C(q, ˙q) ˙q + U (q) = G(q)u + A(q)λ, (1)
A
(q) ˙q = 0, (2)
where q(t), ˙q(t) R
n
are the generalized positions and velocities, respectively, u(t)
R
m
is the control input, A(q)λ are the constraint forces with A : R
n
R
n×k
, λ R
k
,
G : R
n
R
n×m
is the input matrix, M : R
n
R
n×n
is the mass matrix with
Astilfi A., Ortega R. and Venkatraman A. (2009).
A Globally Exponentially Convergent Immersion and Invariance Speed Observer for n–Degrees of Freedom Mechanical Systems.
In Proceedings of the International Workshop on Networked embedded and control system technologies: European and Russian R&D cooperation,
pages 134-145
Copyright
c
SciTePress
M = M
> 0 and U : R
n
R is the potential energy function. C(q, ˙q) ˙q is the
vector of Coriolis and centrifugal forces, with the (ik)–th element of the matrix C :
R
n
× R
n
R
n×n
defined by
C
ik
(q, ˙q) =
n
X
j=1
C
ijk
(q) ˙q
j
,
where C
ijk
: R
n
R are the Christoffel symbols of the first kind defined as
C
ijk
(q) :=
1
2
M
ik
q
j
+
M
jk
q
i
M
ij
q
k
. (3)
We consider q(t) to be measurable and assume that the input u(t) is such that q(t), ˙q(t)
exist for all time, that is, the system is forward complete. Our objective is to design a
globally asymptotically convergent observer for ˙q(t).
Speed observation is a longstanding problem whose complete theoretical solution
has proven highly elusive. The first results were reported in 1990 in the fundamental
paper [14], and many interesting partial solutions have been reported afterwards. Par-
ticular attention has been given to the case in which the system (1) can be rendered
linear in the unmeasurable velocities via partial changes of coordinates, see, e.g., [6,
16]. An intrinsic local observer, exploiting the Riemannian structure of the system, has
been recently proposed in [1] (see also [2] for a Lyapunov analysis and [7] for a gener-
alization). A solution for a class of two degrees of freedom systems has been recently
reported in [8]. The reader is referred to the recent books [5,?,?] for an exhaustive list
of references.
A complete solution to the problem is given by the proposition below. As will be-
come clear in the proof, the construction of the observer relies on the use of the Im-
mersion and Invariance (I&I ) technique—first reported in [4] and further developed in
[3,10]. In I&I the observer design is recast as a problem of rendering attractive a suit-
ably selected invariant manifold defined in the extended state–space of the plant and
the observer. It should be mentioned that the observers in [8, 16] are also based on the
I&I approach.
2 Main Result
Proposition 1. Consider the system (1), and assume u is such that trajectories exist for
all t 0 . There exist smooth mappings A : R
3n2k+1
× R
n
× R
m
R
3n2k+1
and
B : R
n
R
(nk)×(3n2k+1)
such that the dynamical system
˙χ = A(χ, q, u) (4)
with state χ(t) R
3n2k+1
, inputs q(t) and u(t), and output
η = B(q)χ, (5)
has the following property.
All trajectories of the interconnected system (1), (2), (4) are such that
lim
t→∞
e
αt
[N (q) ˙q(t) η(t)] = 0, (6)
for some α > 0 and for all initial conditions (q(0), ˙q(0), χ(0)) R
n
×R
n
×R
3n2k+1
,
where N (q) : R
n
R
nk
× R
n
is a left invertible matrix. That is, (4), (5) is a globally
exponentially convergent speed observer for the mechanical system (1)-(2).
Remark 1. For the special case of a mechanical system with no nonholonomic con-
straints, it is clear that k = 0 and subsequently the matrix N (q) becomes an invertible
square matrix.
3 A Preliminary Lemma
Before giving the proof of the main result, we recall that the system (1)-(2) can be
written in the port-Hamiltonian form [13] as
˙q
˙p
=
0 I
I 0
q
H(q, p)
p
H(q, p)
+
0
G(q)
u +
0
A(q)
λ, (7)
A
(q)λ = 0, (8)
where p = M (q) ˙q are the generalized momenta and
H(q, p) =
1
2
p
M
1
(q)p + U (q)
represents the total energy stored in the system. Further, as per [13], the system (7)-(8)
when restricted to the constrained space
X
c
= {(q, ˙q)|A
(q) ˙q = 0},
takes the form
˙q
˙
˜p
=
0
˜
S(q)
˜
S
(q) J(q, ˜p)
q
H(q, ˜p)
˜p
H(q, ˜p)
+
0
B
c
(q)
u +
0
A(q)
λ, (9)
H(q, ˜p) = V (q) +
1
2
˜p
˜
M
1
(˜q)˜p, (10)
with ˜p R
nk
being given as ˜p =
˜
S
(q)p where
˜
S(q) R
n×nk
is the full rank
annihilator of the matrix A(q) satisfying the condition A
(q)
˜
S(q) = 0. The matrix
J(q, ˜p) is skew-symmetric and is given by
J
ij
(q, ˜p) = p
[
˜
S
i
,
˜
S
j
], (11)
where [
˜
S
i
,
˜
S
j
] denotes the standard Lie bracket of the column vectors S
i
, S
j
and the
matrix
˜
M(q) R
nk×nk
is symmetric positive-definite.
In order to streamline the presentation in this section, we introduce a factorization
of the mass matrix
˜
M
1
(q) = T
(q)T (q), (12)
where T : R
n
R
nk×nk
is a full rank matrix
4
and define the mappings L : R
n
R
n×nk
and F : R
n
× R
m
R
nk
as
L(q) =
˜
S(q)T
(q), (13)
F (q, u) = L
(q)[B
c
(q)u U(q)]. (14)
Notice that, since q and u are measurable, these mappings are known. We next state the
following proposition.
Lemma 1. The system dynamics (9)-(10) when expressed in the newcoordinates(y, x) =
(q, T (q)˜p), admits a state space representation of the form
˙y = L(y)x, (15)
˙x = S(y, x)x + F (y, u), (16)
where
S = T JT
+
n
X
i=1
({
T
y
i
T
1
x}{L
e
i
}
{L
e
i
}{
T
y
i
T
1
x}
), (17)
and e
i
is the i
th
basis vector of R
nk
.
Proof. We directly obtain (15) by differentiating y and by using (9), (10), (13). We next
compute the following,
˙x =
˙
T ˜p + T
˙
˜p, (18)
=
˙
T ˜p T
˜
S
y
(
1
2
˜p
˜
M
1
(˜q)˜p) T
˜
S
U + T
˜
S
B
c
u + T JT
x, (19)
=
˙
T ˜p L
y
(
1
2
˜p
˜
M
1
(˜q)˜p) + F + T JT
x, (20)
where we have made use of (10), (13) and (14). We now compute that,
˙
T ˜p =
n
X
i=1
(
T
y
i
˜p)(e
i
˜
S
˜
M
1
˜p),
=
n
X
i=1
(
T
y
i
˜p)(e
i
L)x, (21)
and further obtain
y
{
1
2
˜p
˜
M
1
˜p} =
y
{
1
2
˜p
T
T p}
=
n
X
i=1
e
i
{
T
y
i
˜p}
x. (22)
4
Since M is positive definite this factorization always exists. It may be taken to be the (univo-
cally defined) Cholesky factorization, as proposed in [9].
Substituting (21) and (22) in (20) we obtain the dynamics of x as,
˙x =
n
X
i=1
({
T
y
i
T
1
x}{L
e
i
}
{L
e
i
}{
T
y
i
T
1
x}
)x + T JT
x + F,
= Sx + F, (23)
where we have used (17) to obtain the equation (23). This concludes the proof.
Remark 2. It can be verified easily that matrix S(y, x) defined in (17) satisfies the fol-
lowing properties:
(i) S is skew–symmetric, that is,
S + S
= 0.
(ii) S is linear in the second argument, that is,
S(y, a
1
x + a
2
¯x) = a
1
S(y, x) + a
2
S(y, ¯x),
for all y, x, ¯x R
n
, and a
1
, a
2
R.
(iii) There exists a mapping
¯
S : R
n
× R
n
R
n×n
such that
S(y, x)¯x =
¯
S(y, ¯x)x,
for all y, x, ¯x R
n
.
Remark 3. Lemma 1 implies that the speed observer problem for system (1)-(2) can be
recast as an observer problem for system (15)-(16) with output y.
Remark 4. For the special case of no nonholonomic constraints, we have k = 0 and
J(q, ˜p) = 0,
˜
S(q) = I,
˜
M(q) = M(q), L(q) = T
(q).
This subsequently simplifies the expression for S(y, x) as
S(y, x) =
n
X
i=1
({
T
y
i
T
1
x}{T e
i
}
{T e
i
}{
T
y
i
T
1
x}
). (24)
It can be shown (refer to [16]) that the (jk)–th element of S is S
jk
= −{T
1
x}
[T
j
, T
k
].
4 Proof of the Main Result
The observer is constructed in four steps.
(S1) Following the I&I procedure [3], we define a manifold (in the extended state-space
of the plant and the observer) that should be rendered attractive and invariant
5
. As
is well–known, to achieve the latter objective a partial differential equation (PDE)
should, in principle, be solved.
5
The manifold should be such that the unmeasurable part of the state can be reconstructed from
the function that defines the manifold.
(S2) To avoid the need to solve the PDE the “approximation” technique proposed in
[10] is adopted. Using this approximation induces some errors in the observer error
dynamics.
(S3) Always borrowing from [10], we introduce a dynamic scaling that dominates—
in a Lyapunov–like analysis—the effect of the aforementioned disturbance terms,
proving that the scaled observer error converges to zero.
(S4) To prove that the dynamic scaling factor is bounded and, consequently, that the ac-
tual observer error converges,exponentially, to zero, high gain terms are introduced
in the observer dynamics to, again, dominate sign–indefinite terms in a Lyapunov–
like analysis.
Step 1. (Definition of the manifold) For the system (15)-(16), we propose the manifold
M := {(y, x, ξ, ˆy, ˆx) : ξ x + β(y, ˆy, ˆx) = 0} R
5n3k
, (25)
where ξ R
nk
, ˆy R
nk
, ˆx R
n
are (part of) the observer state, the dynamics of
which are defined below, and the mapping β : R
3n2k
R
nk
is also to be defined.
To prove that the manifold M is attractive and invariant it is shown that the off–
the–manifold coordinate
z = ξ x + β(y, ˆy, ˆx), (26)
the norm of which determines the distance of the state to the manifold M, is such that:
(C1) z(0) = 0 z(t) = 0, for all t 0 (invariance);
(C2) z(t) asymptotically (exponentially) converges to zero (attractivity).
Notice that, if z(t) 0, an asymptotic estimate of x is given by ξ + β.
To obtain the dynamics of z differentiate (26), yielding
˙z =
˙
ξ ˙x +
˙
β
=
˙
ξ S(y, x)x F +
y
β ˙y +
ˆy
β
˙
ˆy +
ˆx
β
˙
ˆx.
Let
˙
ξ = F
ˆy
β
˙
ˆy
ˆx
β
˙
ˆx + S(y, ξ + β)(ξ + β)
y
βL(y)(ξ + β), (27)
where
˙
ˆy and
˙
ˆx are to be defined. Replacing (27) in the equation of ˙z above, and invoking
properties (ii) and (iii) of Lemma 1, yields
˙z = S(y, ξ + β z)(ξ + β z) +
+S(y, ξ + β)(ξ + β)
y
βL(y)z
= S(y, x)z + S(y, z)(ξ + β)
y
βL(y)z
= S(y, x)z +
¯
S(y, ξ + β)z
y
βL(y)z. (28)
From (28) it is clear that condition (C1) above is satisfied. On the other hand, condition
(C2) would be satisfied if we could find a function β that solves the PDE
y
β = [k
1
I +
¯
S(y, ξ + β)]L
1
(y), (29)
with k
1
> 0, where L
1
(y) : R
n
R
nk×n
is the full rank left inverse of the
matrix L(y). Indeed, in this case, the z–dynamics reduce to ˙z = (S k
1
)z, achieving
the desired exponential convergence property. Unfortunately, solving the PDE (29) is a
daunting task, and we don’t even know if a solution exists. Therefore, in the next step
of the design we proceed to “approximate” its solution.
Step 2. (“Approximate solution” of the PDE) Define the “ideal
y
β as
H(y, ξ + β) := [k
1
I +
¯
S(y, ξ + β)]L
1
(y), (30)
and denote the columns of this n k × n matrix by H
i
: R
n
× R
nk
R
nk
for
i = 1, . . . , n, that is,
H(y, ξ + β) =
H
1
(y, ξ + β) | · · · | H
n
(y, ξ + β)
.
Now, mimicking [10], define
6
β(y, ˆy, ˆx) :=
Z
y
1
0
H
1
([s, ˆy
2
, . . . , ˆy
n
], ˆx)ds + · · · +
+
Z
y
n
0
H
n
([ˆy
1
, . . . , ˆy
n1
, s], ˆx)ds. (31)
From the definition of the mapping β, and adding and substracting H(y, ξ + β), we
have that
y
β can be written as
y
β(y, ˆy, ˆx) = H(y, ξ + β)
n
H(y, ξ + β)
H
1
(y
1
, ˆy
2
, . . . , ˆy
n
, ˆx) . . . H
n
(ˆy
1
, . . . , ˆy
n1
, y
n
, ˆx)
o
.
Since the term in brackets is equal to zero if ˆy = y and ˆx = ξ + β, and all functions are
smooth, there exist mappings
y
: R
n
× R
nk
× R
n
R
nk×n
,
x
: R
n
× R
nk
×
R
nk
R
nk×n
such that
y
β(y, ˆy, ˆx) = H(y, ξ + β)
y
(y, ˆx, e
y
)
x
(y, ˆx, e
x
), (32)
with
e
y
:= ˆy y, e
x
:= ˆx (ξ + β), (33)
and such that
y
(y, ˆx, 0) = 0,
x
(y, ˆx, 0) = 0, (34)
for all y, ˆy, R
n
and x, ˆx R
nk
.
Replacing (30) and (32) in (28) yields
˙z = (S k
1
)z + (
y
+
x
)L(y)z. (35)
6
We attract the readers attention to the particular selection of the arguments used in the inte-
grands. Namely that, with some abuse of notation, the vector ˆy has been spelled out into its
components.
Recalling that S is skew–symmetric and k
1
> 0, it is clear that the mappings
y
and
x
play the role of disturbances that we will try to dominate with a dynamic scaling in
the next step of the design.
Step 3. (Dynamic scaling) Define the scaled off–the–manifold coordinate
η =
1
r
z, (36)
with r a scaling dynamic factor to be defined below. Differentiating (36), and using
(35), yields
˙η =
1
r
˙z
˙r
r
η
= (S k
1
)η + (
y
+
x
)L(y)η
˙r
r
η.
Consider the function
V
1
(η) =
1
2
|η|
2
,
and note that its time derivative is such that
˙
V
1
= (k
1
+
˙r
r
)|η|
2
η
(
y
+
x
)L(y)η
k
1
2
+
˙r
r
1
2k
1
k[
y
+
x
]Lk
2
|η|
2
k
1
2
+
˙r
r
1
k
1
k
y
Lk
2
+ k
x
Lk
2
|η|
2
,
(37)
where k·k is the matrix induced 2—norm and we have applied Young’s inequality (with
the factor k
1
) to get the second bound. Let
˙r =
k
1
4
(r 1 ) +
r
k
1
k
y
Lk
2
+ k
x
Lk
2
, r(0) 1. (38)
Notice that the set {r R | r 1} is invariant for the dynamics (38). Replacing (38)
in (37) yields the bounds
˙
V
1
k
1
2
k
1
4
r 1
r
|η|
2
k
1
4
|η|
2
, (39)
where the property
r1
r
1 has been used to get the second bound. From (39) we
conclude that η(t) converges to zero exponentially.
Step 4. (High–gain injection) From (36) and the previous analysis it is clear that z(t)
0 if we can provethat r L
, which is the property established in this step. To enhance
readability the procedure is divided into two parts. First, we make the function
V
2
(η, e
y
, e
x
) = V
1
(η) +
1
2
(|e
y
|
2
+ |e
x
|
2
),
a strict Lyapunov function. Then, the derivative of the function
V
3
(η, e
y
, e
x
, r) = V
2
(η, e
y
, e
x
) +
1
2
r
2
, (40)
is shown to be non–positive—establishing the desired boundedness of r. In both steps
the objectives are achieved adding, via a suitable selection of the observer dynamics,
negative quadratic terms in η, e
y
, e
x
in the Lyapunov function derivative. We recall that
e
y
and e
x
are measurable quantities, defined in (33).
Towards this end, define
˙
ˆy = L(y)(ξ + β) ψ
1
(y, r)e
y
, (41)
with ψ
1
: R
n
× R
+
R
+
a gain function to be defined. The error dynamics, obtained
combining (15) and (41), are
˙e
y
= Lz ψ
1
e
y
. (42)
Now, select
˙
ˆx = F + S(y, ξ + β)(ξ + β) ψ
2
(y, r)e
x
, (43)
with ψ
2
: R
n
× R
+
R
+
a gain function to be defined. Recalling (27) the error
dynamics for e
x
become
˙e
x
=
y
βLz ψ
2
e
x
. (44)
Using (39), (42) and (44) and doing some basic bounding, yields
˙
V
2
k
1
4
|η|
2
+ re
y
ψ
1
|e
y
|
2
+
+re
x
y
βLη ψ
2
|e
x
|
2
(45)
(
k
1
4
1)|η|
2
ψ
1
r
2
2
kLk
2
|e
y
|
2
ψ
2
r
2
2
k∇
y
βk
2
kLk
2
|e
x
|
2
. (46)
Selecting
ψ
1
= k
2
+ ψ
3
+
r
2
2
kLk
2
,
ψ
2
= k
3
+ ψ
4
+
r
2
2
k∇
y
βk
2
kLk
2
, (47)
with k
2
, k
3
> 0 and ψ
3
, ψ
4
: R
n
× R
+
R
+
to be defined, we conclude that
˙
V
2
1
2
(k
1
2)|η|
2
k
2
|e
y
|
2
k
3
|e
x
|
2
,
which, selecting k
1
> 4, establishes that η, e
y
, e
x
L
2
L
and the origin of the
(non-autonomous) subsystem with state η, e
y
, e
x
is uniformly globally exponentially
stable.
We are now ready for the coup de gr
ˆ
ace, namely the selection of ψ
3
and ψ
4
to
guarantee that r L
. For, recall (34), which ensures the existence of mappings
¯
y
:
R
n
× R
nk
× R
n
R
nk×n
,
¯
x
: R
n
× R
nk
× R
nk
R
nk×n
such that
k
y
(y, ˆx, e
y
)k k
¯
y
(y, ˆx, e
y
)k |e
y
|
k
x
(y, ˆx, e
x
)k k
¯
x
(y, ˆx, e
x
)k |e
x
|. (48)
Now, evaluate the time derivative of V
3
, defined in (40), replace (47) in (46), and use
the bounds (48) to get
˙
V
3
(
k
1
4
1)|η|
2
ψ
3
r
2
k
1
k
¯
y
k
2
kLk
2
|e
y
|
2
ψ
4
r
2
k
1
k
¯
x
k
2
kLk
2
|e
x
|
2
.
Fixing
ψ
3
=
r
2
k
1
k
¯
y
k
2
kLk
2
ψ
4
=
r
2
k
1
k
¯
x
k
2
kLk
2
ensures
˙
V
3
0, which ensures r L
.
To prove condition (6) note that equation (39) implies
|η(t)| |η(0)|e
k
1
8
t
,
hence
|z(t)|
r(t)
r(0)
|z(0)|e
k
1
8
t
sup
t0
{r(t)}|z(0)|e
k
1
8
t
,
which yields the claim, by boundedness of r(t).
The proof is completed defining the state vector of the observer as χ = (ˆx, ˆy, ξ, r),
obtaining A(χ, q, u) from (43), (41), (27), and (38), and defining
B(y) :=
T
1
(y) 0 0 0
.
Remark 5. The four components ˆx, ˆy, ξ and r of the state vector of the observer can be
given the following interpretation. The component ˆx is the estimate of x and a filtered
version of ξ + β. The component ˆy is a filtered version of the measured variable y.
The ξ-dynamics render the set z = 0 invariant, regardeless of the selection of the other
dynamics, and ξ can be regarded as the state of a reduced order observer
7
. Finally, the
r-dynamics are used to trade stability of the nominal design for robustness against the
disturbances
y
and
x
.
7
To clarify this point note that, ideally, the PDE (29)should have a solution β which is a function
of y alone. In this case the variable ξ would play the role of the state of the (reduced) order
observer(see the examples in [8]).
Remark 6. Although the analysis of the performance of the proposed observer in the
presence of noise is not within the scope of the paper, it is worth noting the following.
The Lyapunov argument establishing uniform asymptotic stability of the zero equilib-
rium of the (η, e
y
, e
x
)-subsystem yields robustness againsts small additiveperturbations
on the measured variables u and y. In the presence of such perturbations the variables
e
y
and e
x
do not converge to zero. Nevertheless, as long as they are sufficiently small,
equation (38) can be regarded as describing a linear (non-autonomous) scalar differen-
tial equation in which, by equations (34), the coefficient of the linear term is uniformly
negative. This ensures boundedness of r(t) for all t.
5 Conclusions
A definite affirmative answer has been given to the question of existence of a globally
convergent speed observer for general mechanical systems of the form (1). No assump-
tion is made on the existence of an upperbound for the inertia matrix, hence the result
is applicable for robots with prismatic joints. Also, no conditions are imposed on the
potential energy function. The only requirement is that the system is forward complete,
i.e., that trajectories of the system exist for all times t 0—which is a rather weak
condition.
In some sense, our contribution should be interpreted more as an existence result
than an actual, practically implementable, algorithm. Leaving aside the high complex-
ity of the observer dynamics, that can be easily retraced from the proof of Section 4, the
difficulty stems from the fact that the key function β is defined via the integrals (31),
whose explicit analytic solution cannot be guaranteed a priori. Of course, the (scalar)
integrations can always be numerically performed leading to a numerical implementa-
tion of the observer. Given the recent spectacular advances in computational technology
this does not seem to constitute an unsurmountable difficulty.
References
1. N. Aghannan and P. Rouchon, An intrinsic observer for a class of Lagrangian systems”,
IEEE Trans. Automat. Contr., Vol. 48, No. 6, pp. 936–945, 2003.
2. D.A. Anisi, J. Hamberg, “Riemannian observers for Euler-Lagrange systems”, 16th IFAC
World Congress, Prague, Czech Republic, 2005.
3. A. Astolfi, D. Karagiannis, R. Ortega, Nonlinear and Adaptive Control with Applications,
Springer-Verlag, Berlin, Communications and Control Engineering, 2007.
4. A. Astolfi, R. Ortega, “Immersion and Invariance: A new tool for stabilization and adaptive
control of nonlinear systems”, IEEE Trans. Automat. Contr., Vol. 48, No. 4, pp. 590-606,
April 2003.
5. G. Besanc¸on (Ed.), Nonlinear Observers and Applications, Lecture Notes in Control and
Information Science 363, Springer Verlag, 2007
6. G. Besanc¸on, “Global output feedback tracking control for a class of Lagrangian systems”,
Automatica, Vol. 36, pp. 1915-1921, April 2000.
7. S. Bonnabel, P. Martin, P. Rouchon, “Symmetry-Preserving Observers”, IEEE Trans. Au-
tomat. Contr., Vol. 53, No. 11, pp. 2514–2526, 2008.
8. D. Karagiannis, D. Carnevale and A. Astolfi, “Invariant manifold based reduced-order ob-
server design for nonlinear systems”, IEEE Trans. Automat. Contr., Vol. 53, No. 11, pp.
2602–2614, 2008.
9. A. Jain and G. Rodriguez, “Diagonalized Lagrangian robot dynamics,Proc. IEEE Int. Conf.
Robot. & Autom., Vol. 11, no. 4, pp. 571-583, 1995.
10. D. Karagiannis and A. Astolfi, “Observer design for a class of nonlinear systems using dy-
namic scaling with application to adaptive control”, 47th IEEE Conference on Decision and
Control, Cancun, Mexico, December 9–11, 2008.
11. R. Kelly, V. Santibanez, A. Loria, Control of Robot Manipulators in Joint Space, Springer,
London, 2005.
12. H. K. Khalil, Nonlinear Systems, Printice-Hall, Pearson Education International, 1996.
13. Arjan van der Schaft, Nonlinear Systems, Springer-Verlag, 1996.
14. S. Nicosia and P. Tomei, Robot control by using only joint position measurements, IEEE
Trans. Automat. Contr., Vol. 35, No. 9, pp. 1058–1061, 1990.
15. R. Ortega, A. Loria, P. J. Nicklasson and H. Sira–Ramirez, Passivity–Based Control of
Euler–Lagrange Systems, Springer-Verlag, Berlin, Communications and Control Engineer-
ing, 1998.
16. A. Venkatraman, R. Ortega, I. Sarras and A. J. van der Schaft, ”Speed observation and posi-
tion feedback stabilization of partially linearizable mechanical systems”, 47th IEEE Confer-
ence on Decision and Control, Cancun, Mexico, December 9–11, 2008.