Sunteți pe pagina 1din 21

SIAM J.

CONTROL AND ÜPTIMIZATION © 1995 Society for Industrial and Applied Mathematics
Vol. 33, No. 1, pp. 126-148, January 1995 006

NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS*


A. M. BLOCHt ANO P. E. CROUCHt

Abstract. This paper gives a general formulation of the theory of nonholonomic control systems on a Riemannian
manifold modeled by second-order differential equations and using the unique Riemannian connection defined by
the metric. The main concern is to introduce a reduction scheme, replacing sorne of the second-order equations by
first-order equations. The authors show how constants of motion together with the nonholonomic constraints may be
combined to yield such a reduction. The theory is applied to a particular class of nonholonomic control systems that
may be thought of as modeling a generalized rolling ball. This class reduces to the classical example of a ball rolling
without slipping on a horizontal plane.

Key words. Riemannian manifold, nonholonomic, control, optimization

AMS subject classifications. 70F25, 70005, 93A30, 93B05

l. Introduction. In this paper we consider the forrnulation of controlled classical non-


holonornic systerns on Riernannian rnanifolds. As described in Vershik [1984), Vershik and
Fadeev [1981), Vershik and Gershkovich [1988), and Amold [1988), one can divide the theory
of nonholonornic systerns into two classes: variational systerns derivable frorn aLagrangian,
and classical (rnechanical) systerns derivable frorn D'Alernbert's principle. The latter is nor-
rnally used for describing rnechanical systerns, and our goal here is to introduce sorne aspects
of control into the theory at the general level of systerns evolving on Riernannian rnanifolds.
This paper expands and extends sorne of the results announced in Bloch and Crouch [1992).
The classical nonholonornic systerns have been considered by rnany others. We rnention
sorne of the rnost irnportant here: Cartan [1952), using the general theory of equivalence;
Weber [1986), using the Hamiltonian setting; Herrnann [1982), using the Lagrangian setting
and Ehresrnann's theory of connections on jet spaces; and finally, Bates and Sniatycki [1993),
again using a Harniltonian setting but introducing a theory of reduction under syrnrnetry.
Although little work has been done on the role of control in such systerns until very
recently, for exarnple Yang [1992] and Bloch et al. [1989], [1992b], a lot of work has been
done on the kinernatic problern, where the nonintegrable velocity constraints define nonlinear
(kinernatic) control problerns through the direct control of sorne of the velocities. Exarnples of
this work include Brockett and Dai [1993), Krishnaprasad and Yang [1991), Lafferiere and
Sussrnann [1991), Montgornery [1990), and Murray and Sastry [1990); see Bloch et al. [1992b]
for further references.
Our work sets out to show that reduction of the dynarnic equations of classical nonholo-
nornic systerns with extemal forces or controls can be expressed in a general Lagrangian
setting, and that if syrnrnetries are also present they can be incorporated into the same set-
ting. The reduction is effected by introducing a bundle structure, so that the reduced systern
is defined by first-order differential equations in the fiber space and second-order differen-
tial equations in the base, defined through the Riernannian connection induced by a natural
Riernannian rnetric or kinetic energy. This approach is rnost closely aligned with the irnpor-
tant work of Koiller [1992], who considers the reduction process on a principal bundle, the
structure group of which is a syrnrnetry group for the problern. These are termed nonabelian

* Received by the editors August 3, 1992; accepted for publication (in revised form) September 28, 1993.
t Department of Mathematics, The Ohlo State University, Columbus, Ohio 432!0, and University of Michigan,
Ann Arbor, Michigan 48!09. This author's research was partially supported by National Science Foundation grant
PYI DMS-9157556, and Air Force Office of Scientific Research grant F49620-93-l-0037.
t Center for Systems Science and Engineering, Arizona State University, Tempe, Arizona 85287. This author's
research was partially supported by National Science Foundation grant DMS-9!011964.
126
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 127

Caplygin systems. Our framework incorporates this setting as a special case, even though it
is an important case.
We illustrate our general prescription by introducing an extended example. We generalize
the case of a ball rolling on a flat plane with extemal forces parallel to the plane to push the ball,
as studied by Brockett and Dai [1993]. The generalization incorporates the well-known model
of a generalized rigid body, using a compact semi-simple Lie group, together with velocity
constraints, linking the generalized rotational motion to a generalized translational motion.
We call this system the generalized rolling ball. In the case that the ball is not symmetric
(i.e., has an inertia tensor not equal to a multiple of the identity), the nonholonomic velocity
constraints are not symmetries, and our reduction procedure generalizes that of Koiller [1988].
We also consider sorne controllability and optimality properties of the generalized rolling
ball in the symmetric case. The controllability result generalizes the local argument of Bloch
et al. [1992b], using a general controllability result on principal bundles. We also formalize a
minimum force control problem as the higher-order analogue of the minimum energy control
problem. In the holonomic case the minimum force control problem for second-order New-
tonian systems was studied in Noakes et al. [1989] and Crouch and Leite [1991a], [199lb].
These works treat higher-order variational problems on Riemannian manifolds. We outline
the modifications necessary to treat the nonholonomic case.
The outline of the paper is as follows. In §2 we give a general formulation of holonomic
control systems on Riemannian manifolds, and in §3 we give the formulation of nonholonomic
systems. In §4 we introduce the concept of symmetries and describe the general reduction
procedure based on the nonholonomic constraints and the constants of motion derived from
the symmetries. In §5, we have four subsections dealing with the control of the generalized
rolling hall, in which we describe sorne preliminaries on Riemannian structures on Lie groups,
the generalized rigid body, the generalized rolling ball in both body fixed and inertial axes,
and finally controllability and optimality questions.
2. Holonomic control systems and optimal control. In this section we give a brief
formulation of mechanical systems, without any velocity constraints, under the influence of
externa} forces. We broadly follow the formulation given by Hermann [1982]. We consider
the case of systems with "nonintegrable" velocity constraints in the following section.
Mechanical systems, in which the velocity constraints are integrable, are referred to here
as holonomic mechanical systems. These integrable velocity constraints yield constraints on
the configuration variables only, and thereby determine a manifold in which the configuration
variables are constrained to evolve (see Arnold [1978]). This provides the motivation for
considering holonomic control systems in the generality discussed below.
We let M denote a smooth (infinitely differentiable), n-dimensional manifold with a
Riemannian metric denoted /C(., ·).TM will denote the tangent bundle to M. The norm of a
vector X p E TpM will be denoted by

M denotes the configuration space of a mechanical system, whereas TM denotes the phase
(or state) space. The notion of an inertia tensor will be modeled by a bundle mapping
J : TM TM ,

such that J is the identity on M. Thus for each p E M we have a linear mapping
Jp : TpM TpM .

We assume that for each p, Jp is an isomorphism satisfying for each X p, Yp E TpM


128 A. M. BLOCH AND P.E. CROUCH

(i) IC( JpX p, Yp) = K( X p, JpYp ),


(ii) K( JpX p, X p ) O (= O if and only if X p = O).
From J we may define another Riemannian metric on M by setting

(X, Y) = K( J X , Y )

for all vector fields X , Y on M. We refer to K, as the ambient metric and (·, ·) as the mechanical
metric. The norm of a vector X p E TpM with respect to the mechanical metric will be denoted
by

The mechanical metric determines aunique Riemannian connection on M,denoted V', and
thereby determines a covariant derivative, denoted D/ Ot. A mechanical system is determined
by its kinetic energy T : TM ---. R given by the expression

T _ / dq dq )
q - 2 \ dt , dt

and the potential energy U : M ---. R given by an arbitrary smooth function on M. Denote
the cotangent bundle to M by T* M. An extemal force is modeled as a covector field F on
r;
M , in general time varying. Thus, Fp E M for each p E M . We define the momentum P
as a covector field along the trajectories of the mechanical system on M by setting

We define an holonomic mechanical system on M to be that given by the basic Newtonian


system of equations in T*M

(1) := F - dU.

For each smooth vector field W on M , we have along the motion of (1)

D ( P(W ) ) - D / dq
at - at \ dt '
w) - / at w) + /\
D2q
-\ 2 '
dq DW )
dt ' at
= DP (W ) p ( DW ) = DP (W ) / dq DW ) .
at + at at + \ dt '
at

We deduce that

(2)

From the definition of kinetic energy and potential energy, equation (1) yields
d
(3)
F (
dq )
= U ( q( t) ) + d Tq( t)·
dt dt dt

We may rewrite the holonomic mechanical system ( 1) as a system of equations in TM


as follows. Let xi,
1 ::::: i ::::: N be N ::::: n independent vector fields on M, and let Ui( ·),
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 129

1 :::; i:::; N, be input or control functions (real valued functions of time). We then model the
force field F by setting
N
(
4) Fq( t) = L ui(t) (Xi(q(t)), ·).
i=l
Let J* : TM -r T*M denote the bundle isomorphism determined on fibers by

From equations (2) and (4) we may rewrite equation (1) in the form

D2q A *-1
(5) 8t2 = w UiXi( q) - Jq dUq.
i=l

Equation (5) now represents a general holonomic mechanical system with inputs. From now
on we shall ignore the potential term in this equation, but it may be added without any extra
= =
difficulty. When F O (and U O) equation (5) reduces to the geodesic equations on the
Riemannian manifold (M, (·, ·))
D2q
at2 = o.
This flow is known to be an extremal of the variational problem (Milnor [1963]),

(6) min Jo{T 11 dq 2


11 dt.
q(O)=qo, ( T )=qT Q t

We now introduce a natural optimal control problem for the system (5). First we define
a norm on fibers of T* M in the usual way:
IF(Wq)I
(7) llFqll = sup llW ll .
IW.lk10 q K

Note that we use the ambient metric in this definition. We introduce the minimum force
control problem as

(8)

subject to the dynamics f = F and the boundary conditions


(9) q(O) = qo , (O) = qo , q(T ) = qr , (T) = qr ,

or equivalently we may specify Pq(O) and Pq(T ). From (2) and (7) we obtain

llFqll = sup K ( Jq , Wq) 11 D2q 11


= Jq
at2 V .

llW.llK #O yl( K(Wq , Wq) "'


Thus the cost functional (8) may be reformulated as

(10)
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 131

lt is now natural to consider the formulation (5) of the holonomic control system. lt is
convenient to assume a little more structure for the force field F defined in (4). We modify
the definition as follows:
N

(11) Fq( t) = L Ui( t)( J;;c ) Xi( q( t)), ·),


i=l

where Xi , 1 :o::; i < n is an orthonormal base of vector fields with respect to the ambient
metric
(12) 1 :o::; i, j :o::; n.

With the force field (11) the system (5) may be rewritten as
D2 N
(13) at'! = L
J;¡ 1xi( q)ui( t).
i=l

The orthonormality assumption (12) now implies that

j D2q D2q ) N 2
\ 8t2 ' Jq at2 = ui ( t).

lt follows that for system (13) the minimum force control problem is defined by the cost
functional
{T l N
(14) mjn Jo l L u¡( t) dt
o i=l

subject to the boundary conditions (9).


In the case N = n this optimal control problem corresponds to the higher-order variational
problem posed by the functional (10) with boundary conditions (9). Similar higher-order
variational problems have been treated in various contexts, most notably as the minimum
curvature problem (Griffiths [1983), Jurdjevic [1991)), but recently they have occurred in the
context of interpolation problems in Gabriel and Kajiya [1988), Noakes et al. [1989), and
Crouch and Silva-Leite [199la], [1991b]. In the latter three works a simpler functional is
considered, namely

(15)

The "normal" extremals of such functionals satisfy an equation of the form

(16)

where R is the curvature tensor associated with the connection 'V. lt follows that for N = n,
the minimum force control problem introduced above is a natural higher-order version of the
classical variational problem (6), which is often interpreted as a rninimum energy problem.
In the control literature another class of optimal control problem has received much
attention. lt may be characterized by the cost functionals (14) subject to systems of the form

(17) ( N < n), q E M,


130 A. M. BLOCH AND P. E. CROUCH

where xi are independent vector fields on M and Vi are control functions. Although these
problems are govemed by first-order rather than second-order equations (13), the singular na-
ture of this optimal control problem, because of the fact that N < n, is of the same character as
that of the minimum force control problem when N < n. Brockett [1982] considered the con-
trol problem posed by (17) and (14) (see also Baillieul [1978]), and subsequently the analysis
has been taken up in the mathematics and control theory literature as sub-Riemannian geom-
etry. In the next section we reinterpret the class of systems (17) as kinematic nonholonomic
control systems (Remark 3).
3. Nonholonomic control systems. We now consider the formulation of controlled non-
holonomic control systems. Nonholonomic systems may be divided into two classes (see,
e.g., Vershik and Gershkovich [1988])-variational nonholonomic systems (dubbed vako-
nomic systems by Arnold [1988]) and classical (or mechanical) nonholonomic systems. In
either case the basic ingredients are "nonintegrable" constraints on phase space, defined by
m, m < n, one-forms on M , w 1 , . . • , Wm, such that their span, over the smooth functions on
M, contains no nontrivial exact forms and in particular none of the forms W¡ , . . . , Wm is exact.
These forms then define a smooth distribution H on M . For each p E M , Hv is the subspace
of TvM defined by

We also stipulate that the distribution H is nonsingular, so that the dimension of the subspace
Hv does not vary with p, although the significance of this is not fully understood.
Variational nonholonomic systems are obtained in our context as solutions of variational
problems of the form

subject to

and boundary conditions q(O) = Qo , q(T ) = Qr .


This problem is solved in the usual way by introducing Lagrange multipliers µk and
solving the new variational problem defined by

If we consider the kinematic system (17), in which the vector fields xi are orthonormal
with respect to the mechanical metric

1 ::::; i, j ::::; m,
and we define n - m independent dual one-forms wk satisfying
132 A. M. BLOCH AND P. E. CROUCH

it follows that the variational problem described above coincides with the optimal control
problem posed by the cost functional (14), and subject to the kinematics (17). In this context
the reader should be aware of work by Kirshnaprasad and Yang [1991] and many others
including Brockett [1982], Baillieul [1978], Murray and Sastry [1990], and Lafferiere and
Sussmann [1991]. See also Remark 3 below.
Classical nonholonomic systems are not obtained from a variational principie in the usual
sense (see Remark 2 below), but from D'Alembert's principie. The equations may be written
in the form

(18)

(19) 1 :=::; k :=::; m.

The vector fields Wi on M are determined through the identity

for any vector field X on M . The multipliers Ai are also uniquely determined from the
equations (18) and (19) as shown in Remark 2 below.
After our formulation of a holonomic controlled mechanical system (5), we define a
nonholonomic controlled mechanical system to have the form
D2 m N

(20) a/f = L AiWi + L uixi ,


i=l i=l

(21) 1 :=::; k :=::; m.

At this point we do not impose any relation between m and N.


We now summarize sorne remarks about the formulation (20).
Remark 1. After our general formulation of a holonomic system with external inputs as
equations in T* M, we may similarly define a nonholonomic system with externa!inputs as
follows:

(22)
1 :=::; k :=::; m.

Differentiating the constraints we obtain

(23) 0 = .:!:_W ( dq) = .:!:_ lw dq ) = / DWk dq ) + lw D 2q )


dt k dt dt \ k' dt \ at ' dt \ k' at2 ·

From equations (2) and (22) we obtain

1 :=::; k :=::; m,
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 133

and so from (23) we deduce that

(24) 1 :::; k :::; m.

Thus the force field Fin a nonholonomic system (22) is not arbitrary but satisfies the constraints
(24). ltfollows that regardless of N,or the vector fields Xi in the nonholonomic control system
(20), after solving for the multipliers >.i , the resulting system has at most n - m independent
control directions (as specified by vector fields that multiply the control functions ui).
Remark 2. We note that although the system (20) and (21) may not arise from a variational
principle in the usual Lagrangian sense, it is a solution of an instantaneous variational problem
(Vershik [1984])

i
8t2
11 - t
i=l
ui Xi 11 subject to \ Wk, ) = O, 1 :::; k :::; m.

This is made clear by differentiating the constraints, as in (23), yielding m affine constraints
in D 2 q/ 8t 2 . The multipliers >.k in (20) are the solutions of the following system of equations

(25) 1 :::; k :::; m.

Remark 3. We define the kinematic nonholonomic control system, corresponding to the


dynamic nonholonomic control system (20), (21), to be a system of the form

(26)

where Vi are also control functions and xi are vector fields spanning the distribution H. In
general xi differ from xi appearing in the dynamic equations (20), the particular choice govemed
by physical consideration of the control mechanism. However, the general principle goveming
the choice of X i and Vi is that Vi control all Of the independent COmponents of the
velocity 'f¡f . For controllability analysis of the system (26), the assumption is made that the
distribution H is completely nonholonomic (see Vershik and Gershkovich [1988]), that is, the
involutive closure HL of H satisfies

p E M.

This is also an important observation in Brockett [1982]. ltis shown in Bloch et al. [1992b] that
under certain circumstances, controllability of the associated dynamic nonholonomic control
system (20), (21) follows from the assumption above.
Remark 4. Computing the derivative of the kinetic energy along solutions of the non-
holonomic control system (20), (21) yields

Thus

d
dt Tq( t) = Fq( t ) ( dq)
dt '
134 A. M. BLOCH AND P. E. CROUCH

where F is defined by (4). Because the same equation holds for the holonomic system
(equation (3) with U = O), it is clear that the constraints do no work, as is well known (see,
e.g., Neimark and Fufaev [1972]).
Example 1. We consider a penny rolling on the x-y plane with rotation angle (} and
heading angle c/J, as in Bloch and McClamroch [1989] and Bloch et al. [1992b]. We have two
controls, one that rolls the penny about its center of mass and another that tums it about its
vertical axis. The constraints are given by

x = iJ cos c/J ,
y = iJ sin c/J.
Letting q = (x, y, (}, cjJ ) T, the dynamic equations of motion (20) may be written as

ij = u1X 1 + u2X 2 + ..\1W1 + ..\2W2,


where X ¡ (O, o, 1,o)r , X2 = (O, o, o, l)T , W¡ = (1, o, - cos c/J , o)r , and W2 =
ft ft (
(O, 1, - sin c/J , O)T . It easily follows that ..\ 1 = ( iJ cos cjJ ), ..\2 = iJ sin cjJ ), and so the
resulting system is given by
.. d (o· ,/.) .. d ( o· . ,/.)
x = dt cos 'I' , y = dt sm 'I' , 20 = U¡ ,

On the other hand, the kinematic equations are

<j = V ¡X ¡ + V2X 2 ,

where .X-1 = (cos c/J , sin c/J, 1, O)T and X-2 = (O, O, O, l )T lie in the distribution H.
Example 2. We now consider the "Heisenberg" system (so called because its vector fields
generate the Heisenberg algebra)-see Brockett [1982] and the work of Vershik et al. (see
e.g., Vershik and Gershkovich [1988]). Here we have a system on R3 in the variables (x, y , z)
and subject to the constraint
z = yx - xiJ.
We have controls in the x and y directions and wish to control the system in R3 . Letting
q = ( x , y, z )r , the natural dynamic nonholonomic control system (20) may be written as

ij = u1X 1 + u2X 2 + ..\W,


where X 1 = (1, 0, 0)r, X2 = (0, 1,0)r, and W = ( -y,x , l )r. lt is easy to see that in this
case

so that the dynamics become

c/Jx = (1 + x 2 )u1 + xyu2 ,


c/Jjj = (1 + y 2 )u2 + xyu¡,
cjJ z = yu¡ - xu2 ,
where cjJ( x , y ) = 1 + y 2 + x2.

The corresponding kinematic system, as discussed in Brockett [1982] is given by

<j = v1 X 1 + v2X 2 ,
- _ T - _ T
where X1- (1, O, y ) ,X2- (O, 1, -x) .
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 135

4. Symmetries and reduction. Symmetries in mechanics give rise to constants of the


motion (see for example Abraham and Marsden [1978]). In the Riemannian context isometries
are generated by Killing vector fields. Recall that Z is a Killing vector field with respect to
the mechanical metric if

(V'y Z, Y) = O
for ali vector fields Y. Further, a sufficient condition for (Z (q), i)
to be a constant of motion for
a geodesic ftow is that Z is a Killing vector field. For controlled nonholonomic systems we
have the following restatement of Theorem 6, p. 82 in Arnold [1988].
LEMMA l. Sufficient conditions for (Z, ¡) to be a constant of motionfor the controlled
nonholonomic system (20), (21) are
(i) Z E H,
(ii) Z E Span {X¡, ..., X N }..L ( with respect to (-,·) ),
(iii) Z is a Killing vectorfield.
Proof. As in equation (23) we have

.!!_ j z dq ) _ j D Z dq ) j z D1 q )
dt \ ' dt - \ at ' dt +\ ' at 2 •

The first term is zero by (iii), and the second is zero by the expression for D 2 q/ ot2 in (20)
and (i) and (ii). D
Note that when M = Rn and the metric ( ·, ·) is independent of the coordinate function
Xi , then a/ OX i is a Killing vector field.
Note also that if the uncontrolled nonholonomic system (18) is determined by constraints
(19) in which the vector fields Wi are indeed Killing vector fields, then equation (23) gives
(Wk , D 2 q/ 8t2 ) = O. lt then follows from (18) that >..k = O, 1 ::; k ::; m, so that the ftow
of such a nonholonomic system is a restriction of the geodesic ftow. This is illustrated in
Example 2.
lt is often convenient to introduce a bundle structure in M, 7í : M --+ B, with fiber F,
dim. B = r, and dim. F = n - r. This structure reftects the natural geometric structure
of the system induced by the constraints and must be compatible with the constraints in the
sense that

7í* HP = T7r(p )B for all p E M .


Clearly this forces dim. H = n - m 2". dim. B = r. Our aim is to reduce the controlled
nonholonomic system (20), (21) so that the evolution on the fiber is given by a first-order equa-
tion. To this end we introduce two further assumptions. Either Assumption 1 or Assumption
2 follows.
Assumption l. dim. H = dim. B, that is, n = m + r. In this case ÍI = H clearly
defines a horizontal distribution on the bundle.
Assumption 2. dim. H - dim. B = n - m - r = s > O, and there exist s linearly
independent vector fields Z 1 , • • • , Zs that satisfy conditions (i)-(iii) of Lemma l. In particular,

(27) \ Zi , ;) = ci = const.

are constants of the motion for (20), (21). We define a distribution ÍI0 on M by setting X E Ílo
if

(Wk , X ) = O 1 ::; k ::; m,


( Zk , X ) = O 1 ::; k ::; s,
136 A. M. BLOCH AND P. E. CROUCH

and assurne that Íio satisfies

(28) TpF n ÍI0 = {O} for allp E M .


Finally we define the r-dirnensional affine distribution ÍI on M by setting X E ÍI if

(Wk , X ) = 0 1 S k S m,
(29)
( Zk , X ) = Ck 1 ks s s.

Note that Assurnption 2 ensures that ÍI0 is a constant r-dirnensional distribution on M,


and (28) ensures that ÍI0 is a horizontal distribution on the bundle. If either Assurnption 1 or
2 holds, we have as a direct surn of affine spaces

Itfollows that any vector field Y on M can be decornposed uniquely into cornponents

With this structure we rnay decornpose the velocity


dq _ ·H + ·F
dt - q q '

In general q_H and q_ F are not derivatives of functions qH and qF on M, although in rnany
applications one can indeed identify such functions. Frorn equations (21) and (29) we obtain

(Wk , q_F ) = -(Wk , q_H ) , 1Sk S m,


( Zk , qF ) = -( Zk ,<j_H ) + Ck , 1 S k S s.

Condition (28) allows us to solve these equations uniquely for q_ F ·

(30) q·F = j F (q, q·H ) .

Note that f F is affine in q_H . Frorn the original controlled nonholonornic systern (20) we rnay
deduce equations of the form

(31) ;t =
D' H
Í H ( q, q, u).

Equations (30) and (31) provide areduction of the 2n first-order equations (20), with constraints
(21), to n+ r first-order equations, without constraints. Locally we can write q_H = ft qB for
sorne trajectory qB (t) E B. In sorne cases we rnay be able to rewrite equations (30) and (31)
globally in terrns of a trajectory q( t) = ( qB( t), qF ( t) ), qB E B, qF E F.
The class of Caplygin control systerns introduced in Bloch et al. [1992b] corresponds to
Assurnption 1, in the case where M is a product M = B x F, B = Rr , F = Rn-r , (·, ·) is
the Euclidean rnetric, and the whole dynarnics (30), (31) is invariant with respect to qF . A
global prescription for these systerns is given in the forrn

where f F is affine in q_ B.
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 137

Another class of systems in which a global reduction is possible is the class of controlled
nonabelian Caplygin systems. The uncontrolled systems of this type were discussed in Koiller
[1988]. We define a controlled nonabelian Caplygin system as a system (30), (31), in which M
is a principal G bundle M ( G,B) for a Lie group G, G acts by isometries on the mechanical
metric, H (in either Assumption 1 or 2) is invariant under G in the sense that

H 9 .p = g* HP , g E G, p E M ,

and finally the vector fields Xi are G invariant

g* Xi( P ) = Xi( g . p ), g E G, p E M , 1 i N.

The reduced system of equations can be written (locally) (by analogy with the work of Koiller)
in the form

g E G,
(32) Di¡_B .
at=
fB( qB , qB , u), qB E B M / G,

where f F is a G invariant vector field on G.


These reduced systems fall directly within the class of control systems on principal bundles
M ( G, B) studied by San Martin and Crouch [1984]. If D is a set of vector fields on M , we
say that D is projectable if for each X E D there exists X' on B such that n*X = X' o n. We
have the following result.
THEOREM 1 (San Martin and Crouch [1984]). Let M ( B, G) be a connected principal
fiber bundle with G a compact Lie group and D a G-invariant, projectable family of vector
fields on M defining a control system D , which is accessible. Denote by 'o the system on
B defined by D' = n(D). Then D is controllable if and only if 'o is controllable.
We have the following corollary.
COROLLARY 1. Consider a nonabelian Caplygin control system with compact structure
group. Assume that the reduced system (32) is accessible and the system on the base is
controllable. Then the reduced system on M is controllable.
5. An example: The generalized rolling hall.
5.1. Background on Lie groups. In this section we briefly review sorne of the material
relating to the dynamical equations of the generalized rigid body. The material in the next
two sections may be found in books by Amold [1978], Hermann [1977], Boothby [1975], etc.
Let G be an L-dimensional compact semi-simple Lie group with identity element e, and let g
denote its Lie algebra. There exists a positive definite inner product on g that we denote by
( ( ·, ·)), defined as a multiple of the Killing form. If X is an element of g , then we can define
left and right invariant vector fields on M by setting

x; = R .X ,
9

where L 9 and R9 are the left and right translations on G by g E G.


Let Ad denote the adjoint mapping on Q; if </>h : G -+ G is defined by

then

Ad h = </>h. l g=e·
138 A. M. BLOCH AND P. E. CROUCH

Note that with this definition Ad is an anti-homomorphism Ad gh = Ad h Ad g. Now we


have

Thus,

(33)

For compact semi-simple Lie groups, Ad g is an isometry of Q, for every g E G, with respect
to the Killing form. Thus we have

(34) ((Ad gX , Ad gY)) = ((X, Y)), X , Y E Q, g E G.

We may now define a bi-invariant Riemannian metric on G by setting

K( X , Y;) = ( (X, Y)) = K( x;, Y;).


The bi-invariance follows directly from (33) and (34).
If h(·) is a curve in G satisfying the equation

(35) h = z;;,, h(O) = e, h E G,

then
d
dt Ad h(t )Y l t=O = [Z, Y] = ad Z(Y ).

Note that the definition of Lie bracket on g coincides here, under the identification of g with
TeG, with the standard definition of the Lie bracket
[W, V] 9 (f ) = W9 (V (f ) ) - V9 (W (f ) )
for all functions f and G and vector fields W and V on G.
Applying (34) to Ad h(t ) and differentiating, yields

((ad Z(X), Y)) + ((X, ad Z(Y))) = O, X ,Y, Z E Q.

Let J be a positive definite linear mapping J : g , g


satisfying (i) ((J(X), Y) ) = ((X, J(Y))),
(ii) ((J(X), X)) 2: O (= O if and only if X = 0).
We now define a right invariant metric on G by setting

(x;,Y
;) = ( (X, J(Y))), X ,Y E Q.

We may extend J to a linear isomorphism J9 : T9 G , T9 G by setting


J9 Y; = ( JY ) , g E G, Y E Q.
It follows that

for all vector fields V and W on M . Corresponding to the right invariant metric (-, ·) there
exists a unique Riemannian connection \!.Explicit formulas for \! are given in Arnold [1978]
or Nomizu [1954]. Specifically, \! defines a bilinear form on g

(36) (X, Y) r-+ 'VxY = {[X, Y] + J- 1 [X, JY ] + J- 1 [Y, JX ]}.


NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 139

V' is now extended to right invariant vector fields on G by setting

If J is a multiple of the identity, ( ·, ·) is the same multiple of K , and V' reduces to

(37) V'xrY r = [Xr, yr ].


2

For the purposes of our later analysis we prove the following result.
LEMMA 2.

where V'' is the bilinearform on g given by


V''xY = Y'xY - [X, Y].

Proof Let X¡, . . . , X L be a basis of 9, and write Ad g- 1 y = 'L=I fk ( g )X k , where


Ík are functions on G. Thus
L L
(V'xrY 1)( g ) = L Ík ( g )(V' xr X k)( g ) + L xr (f k )( g )X k (g ),
(38) k=I k=I

X r ( fk )( g ) =:
/ k ( R9( h( t) ) ) \t=o'

where h( ·) is a solution of (35). But

dd Ad ( R9 ( h( t)))- 1 Y 1 = dd Ad h- 1 (t) Ad g- 1 y 1 = -[X, Ad g- 1Y].


t t=O t t=O
Thus from (38) we obtain

5.2. Background on the generalized rigid body. In this section we briefl.y review ma-
terial on the motion of a generalized rigid body. This is modeled by geodesic equations on
a compact semi-simple Lie group G, with the right invariant metric defined by a positive
definite mapping J on g as described in the previous section. We have two representations
of the velocity defined by a right invariant frame and left invariant frame. If X 1, . . . , X L is a
basis for g , we set
L
dg l
(39) dt = L.,, wi(t )Xi ( g ),
i=I

L
(40) := L Vi(t)X [ ( g ).
i=I

The kinetic energy expressed in terms of the representation (40) is given by

T (g ) = \dg dg ) =
2 dt ' dt 2
((
L.,,
i=I
v·X
i " L.,,
i=I
v ·Jx-)).
i i
140 A. M.
A. M. BLOCH
BLOCH ANO
AND P. E.CROUCH
P.E. CROUCH

Thus the coordinates Vi refer to the velocity with respect to a frame moving in the body,
whereas the coordinates Wi refer to the velocity with respect to a frame fixed in inertial space.
(See also Arnold [1978, p. 323], but there the roles of left and right are reversed.) We have

L L
L viX [ (g ) = L wi X J (g ).
i=I i=I
Hence, if we now assume that the basis Xi of g is orthonormal

we obtain
L
Wk = L((Xi, Ad g- 1 Xk ))vi.
i=I
We now obtain two representations for the acceleration, based on the left and right invariant
settings. First we obtain the representation based on (40). Differentiating, we obtain
L L
D2
a/.
= L Vi(t)X [( g ) + L vi( t)v j( t)('V x;X [)( g ).
i=I j,i= I

Setting vt = ¿7= 1 vi( t)Xi and 8vt / 8t = ¿7= 1 vi(t)Xi as vectors in 9, using the expression
(36) gives the usual expression (see, for example, Hermano [1977])

(41)

Now turning to the representation (39) we have the following result in which we write

L L
(42) Wt = L wi( t)Xi , ªZt = L wi( t)Xi , J( g ) = Ad g J Ad g- 1 •
i=I i=I

LEMMA 3.

(43) D2g
at2 = ( 8t
8Wt + J( g ) -1 [Wt.J( g )Wt ]) tg .

Proof Differentiating the representation (39), we obtain

D2
at;,
= LL wi( t)X J ( g ) + LL wi( t)w j( t)('V x; X J )( g ).
i=I ij= I

But 'Vx1 X f = 'V(Ad g-1x;ir Xf. so using (36) and Lemma (2)
J

1
'Vx; X J = 2{-[Ad g- 1 X j , Ad g- 1 Xi] + J-1 [Ad g- 1 X j , J Ad g- 1 Xi]
+ J-1 [Ad g- 1Xi , J Ad g- 1 Xj]} .
141 A. M. BLOCH AND P. E. CROUCH

Now using the definition of J(g) we obtain the stated result in equation (43). O

5.3. The generalized rolling hall. In this section we describe a generalization of a ball
rolling on a ftat table, in which we model the ball as a generalized rigid body described in the
preceding section. We take the configuration space to be given as M = G x RN, where G is
an L-dimensional compact semi-simple Lie group as in the previous sections with L > N.
We put two Riemannian structures on M ; the ambient structure is defined by setting

where (·, ·)E is the Euclidean structure on RN, and (X 9 , V), (Y9 , W) are vectors in T9 G x
TxRN. The mechanical metric is defined by setting

where ( ·, ·) is the right invariant mechanical metric on G,defined in terms of a positive definite
mapping J of Q. The mechanical metric on M determines a Riemannian connection on M,
but the product structure defined by the metric enables us to rely on the connections defined
by the metrics (·, ·) and (-, ·)E on G and RN, respectively.
We wish to define a nonholonomic control system on M according to the prescription
given by the equations (20) and (21). For this example we suppose that m = N, that is, the
number of independent control forces is equal to the number of kinematic constraints. The
nonholonomic control system is completely described by defining the input forces

(44)
F
= LN Ui d xi = LN Ui \ 8ax. , .)E
i=I i=I i

where xi , 1 i N are coordinates in RN, and the constraints Wk , 1 k N


E
((dg dx )) \
dg -1 l ) dx a )
(45) \

Wk dt ' dt = dt , Jg X k ( g ) - dt ' axk =o

where ( 1t , ) is the velocity of the trajectory (g , x ) in M , and X 1, . . • , X L is an orthonormal


basis for Q, with respect to ( (·, ·)). Note that in the physical (three-dimensional) case L =
3, N = 2 (see Bloch and Crouch [1992b]) the constraints are

dg l
) \ dx a
)E \ dg -1 l ) \ dx a
)E
\ -1

dt ' J9 X i (g ) - dt ' 8x2 = dt ' J9 X z( g ) + dt ' 8x1 = O.

This follows from setting the velocity of the point of contact of the ball equal to zero and using
the fact that a three-dimensional rotation has a unique axis of rotation. Because this is not true
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 141

in higher
Using thedimensions, the constraint
skew symmetry of the Lie(45) appears
bracket to be the natural generalization. The system
we get
(20) and (21) now becomes

(46)

. / dg -1 l ( ))
(47) Xk = \ dt ' Jg Xk g ' 1 :::; k :::; N.

THEOREM 2. The following are constants of motion for the controlled nonholonomic
system (46), (47).

(48) N < k ::::; L.

Proof. We show that the conditions of Lemma 1are satisfied.


(i) We must show that XL , N < k :::; L belong to the distribution H on M defined by
the constraints (45). This is equivalent to the identity
(Xk, J¡1 Xj(g)) = O, 1 :::; j :::; N, N < k :::; L.
But this follows from the definition of the mechanical metric on G and the orthonormality of
the vector fields Xj .
(ii) We must show that the vector fields XL N < k :::; L are orthogonal to the control
vector fields 8/ ax k . This follows trivially from the definition of the metric on M.
(iii) We must show that XL N < k :::; L are Killing vector fields with respect to the
mechanical metric on M, which reduces to the same problem for the mechanical metric on
G. Thus it is sufficient to show that
(Wr , Vwr Xk ) = O, for ali vectors W E g.
From Lemma 2 we obtain

(Wr , Vwr Xk ) 9 = Jc:9 (W;,J9 (V Ad g- 1 Xk ); )


= Jc: 9 (W;,( JV Ad g- 1 X k ); )
= ((W, JV Ad g- 1 xk ) )
1
= 2 ((W, -J[W, Ad g- 1 xk ] + [W, J Ad g- 1 xk ] + [Ad g- 1 xk , JW]))
= 0. D

It follows directly from this result that along the trajectories of (46), (47) we have for
suitable constants Ck ,

(49) N < k ::::; L.

We may calculate the multipliers >.k in (46) as was done in Remark 2. Differentiating (47)
and substituting for the second derivatives given in (46), we obtain
143 A. M. BLOCH AND P. E. CROUCH

Define a matrix A with components given by

The assumed positive definiteness of J ensures that A is always invertible, so that (50) defines
the multipliers Ak uniquely in equations (46).
We now describe the two reductions of equations (46), (47), described in §4. We first
consider the reduction based on Assumption l. In this case we take the base B = G and the
fiber F = RN. The N-independent constraints (47) ensure that dim. H = (N + L) - N =
L = dim. B, as required. The reduction procedure described in §4 simply rewrites the second-
order equations on the fiber by the constraint equations (47). We may employ equation (50)
to eliminate the multipliers >.k ; however, we first employ a simple feedback control, defined
by

(52) 1 :::::: k :::::: N,

which defines Üj , 1 :::::: j :::::: N uniquely. lt follows that under this reduction and feedback the
system (46), (47) becomes

X k = ( ddgt , Ju-1xkl (g ) ) , 1 :::::: k :::::: N


(53)

N
D2g
8t2
L J; 1 X k ( g )ük .
k= I

We note that along solutions of (53), we have


N
d 1 (dg dg ) '"" . -
dt 2 dt ' dt = L.,¿ X kU k
k= I
compared with a similar computation for system (46), (47), given in Remark 4,

d [21 (dg
dt ' dt
dg ) + 2(x, E]
1 . .x) = 6 N
'""X.kuk .
dt

Thus the work done by the force control ü in (53) simply changes the generalized rotational
energy and not the generalized translational energy.
Finally we may rewrite the reduced system (53) in terms of an "inertial frame" (39), as
described in the previous section. Using Lemma 3 to express the acceleration in terms of the
velocity Wt in equation (42), we obtain

1 :::::: k :::::: N,

(54)
NONHOLONOMIC CONTROL SYSTEMS ON RIEMANNIAN MANIFOLDS 143

or

We now describe the reduction based on Assumption 2 in §4. In this case we take the base
B = RN and the fiber F = G. We employ the L constraints (47) and (49) to determine the
velocity '1if . Note that we employ ali N velocity constraints and L - N constants of motion.
To calculate '1if explicitly, it is useful to employ the expression for it in terms of an "inertial
frame" as given in equations (39). (47) and (49) can then be written as

X k = Wk , 1 sk sN
L

Ck = L((Xk ,J(g)Xk ) )wi, N <k s L.


i=l

From these equations we may define functions f j , N <j s L,

by the following system of equations, using the positive definiteness of J.


N L
(55) Ck = L((J(g)Xk , Xi))±i + L ( (J(g)Xk , Xi))fi(±i, g), N + 1s k s L.
i=l i=N+I

!fjf by substituting Wk =
It follows that we may rewrite the expression (39) for the velocity
s s s s
X k , 1 k N , Wk = fk ( x , g), N + 1 k L. The remaining equations of the reduced
system are those for X k in system (46). However, we employ feedback again, expressed in
vector form as

(56) (A - I)u + e = Ail ,

where e is the N vector with components

j dg D -1 l) 1 sk sN.
\dt ' at J 9 xk '

The control uis defined uniquely by the invertibility of the matrix A defined by (51). The
resulting reduced equations have the form

1 sk sN,
N L
(57)
= L
xk X k (g ) + L fk (± , g )X k ( g ).
k= I k=N+I
Note that for this system we have

N
1( . ' )E '"' .
ddt 2 X ,X =L., X kUk ·
A

k= I
u
Thus in this formulation the force control in (57) simply changes the generalized translational
energy and not the generalized rotational energy.
145 NONHOLONOMICA.CONTROL
M. BLOCH SYSTEMS ON RIEMANNIAN MANIFOLDS
AND P. E. CROUCH 145

5.4. Controllability and optimal control. In this section we make sorne comments
about the problems associated with the controllability and optimal control of the reduced
models (54) and (57). We first comment on the controllability aspects. Clearly the system
(54) as written cannot be controllable, because of the constants of motion (49), which can be
reexpressed in the form

Reduction of the equations (54) by these constraints would be complicated to analyze, and
the reduction has already been performed in the system (57). A necessary condition for
controllability is accessibility (or weak controllability) (see Isidori [1989] or Nijmeijer and
Van der Schaft [1990]). However, as stated, the Lie algebra associated with the system (57)
is also complicated to analyze. We therefore content ourselves to the case where J = I , in
which case it is easily deduced that both systems (54) and (57) reduce to the system

Xk = Vk,
(58) vk = úk ,
d N L

d = L vk X k ( g ) + L ck X k ( g ), 1 :::; k :::; N.
k=l k=N+I

To analyze the accessibility and controllability of this system, we introduce sorne subspaces
of the Lie algebra g. Let P be the subspace spanned by the vectors X 1 , • • . , X N , let g p be the
subalgebra of g generated by P and let I p denote the ideal of g p generated by the subspace,
[P, P], of gp.
THEOREM 3. Assume that g is a simple Lie algebra. Then the reduced system (58) is
controllable and accessible if and only ij gp = g and [P, P] -=J O.
Proof We first analyze accessibility of the system. It is sufficient to analyze the Lie
algebra of the system (58) as represented by the Lie algebra L on the vector space R2N x g
with generators

a
9k = 8 ,
Vk

Bcause the elements of L depend only on v E RN, accessibility of (58) is equivalent to the
fact that Lv = R2 N x g , where Lv is the subspace of R2 N x g spanned by elements of L
evaluated at v. We construct a subalgebra L e L with the property that none of its elements
depend on v and such that L = L0• Accessibility is therefore equivalent to L = R2N x g.
(Because L does not depend on v , we may identify it with a subspace of R2N x g.) Explicitly
we set L to be the subalgebra of L generated by the vectors { gk , [gk , f ]; 1 :::; k :::; N}. All
Lie brackets of generators 9k and f not in L vanish at v = O. It is easily verified that

L, = span {Ba
vk ; Ba
xk + X k , 1 :::; k :::; N ; lp } .
Indeed, span {[[f, gj ], [f , gi]]; 1 :::; j , i :::; N} = [P, P]. Now L = R2N x g if and only
if lp = g, and so accessibility is equivalent to lp = g. Consider the situation in which
lp e:;; gp = g. Thus lp is an ideal of a simple Lie algebra g, so that lp = O or lp = g.
Because lp = O ifand only if [P, P] = 0, accessibility is equivalentto gp = g and [P, P] -=J O.
We now turn to controllability. We appeal to the result in Theorem 1. Specifically, we treat
system (58) as a system on a principal bundle G x R2 N, in which the system is G invariant and

S-ar putea să vă placă și