34
Vol. 42 (1998) REPORTSONMATHEMATICALPHYSICS No. 112 POISSON REDUCTION FOR NONHOLONOMIC MECHANICAL, SYSTEMS WITH SYMMETRY WANG SANG KOON and JERROLD E. MARSDEN Control and Dynamical Systems, California Institute of Technology 107-81 Pasadena, CA 91125, USA (e-mails: [email protected], [email protected]) (Received January 6, 1998) This paper continues the work of Koon and Marsden [lo] that began the comparison of the Hamiltonian and Lagrangian formulations of nonholonomic systems. Because of the necessary replacement of conservation laws with the momentum equation, it is natural to let the value of momentum be a variable and for this reason it is natural to take a Poisson viewpoint. Some of this theory has been started in van der Schaft and Maschke [24]. We build on their work, further develop the theory of nonholonomic Poisson reduction, and tie this theory to other work in the area. We use this reduction procedure to organize nonholonomic dynamics into a reconstruction equation, a nonholonomic momentum equation and the reduced Lagrange-d’Alembert equations in Hamiltonian form. We also show that these equations are equivalent to those given by the Lagrangian reduction methods of Bloch, Krishnaprasad, Marsden and Murray [4]. Because of the results of Koon and Marsden [lo], this is also equivalent to the results of Bates and Sniatycki [2], obtained by nonholonomic symplectic reduction. Xvo interesting complications make this effort especially interesting. First of all, as we have mentioned, symmetry need not lead to conservation laws but rather to a momentum equation. Second, the natural Poisson bracket fails to satisfy the Jacobi identity. In fact, the so-called Jacobiizer (the cyclic sum that vanishes when the Jacobi identity holds), or equivalently, the Schouten bracket, is an interesting expression involving the curvature of the underlying distribution describing the nonholonomic constraints. The Poisson reduction results in this paper are important for the future development of the stability theory for nonholonomic mechanical systems with symmetry, as begun by Zenkov, Bloch and Marsden [2.5]. In particular, they should be useful for the development of the powerful block diagonalization properties of the energy-momentum method developed by Simo, Lewis and Marsden [23]. 1. Introduction The general setting. Many important problems in robotics, the dynamics of wheeled vehicles and motion generation, involve nonholonomic mechanics, which typically means mechanical systems with rolling constraints. Some of the important issues are trajectory tracking, dynamic stability and feedback stabilization (including nonmini- mum phase systems), bifurcation and control. Many of these systems have symmetry, WI

Poisson reduction for nonholonomic mechanical systems with symmetry

Embed Size (px)

Citation preview

Page 1: Poisson reduction for nonholonomic mechanical systems with symmetry

Vol. 42 (1998) REPORTSONMATHEMATICALPHYSICS No. 112

POISSON REDUCTION FOR NONHOLONOMIC MECHANICAL, SYSTEMS WITH SYMMETRY

WANG SANG KOON and JERROLD E. MARSDEN

Control and Dynamical Systems, California Institute of Technology 107-81

Pasadena, CA 91125, USA (e-mails: [email protected], [email protected])

(Received January 6, 1998)

This paper continues the work of Koon and Marsden [lo] that began the comparison of the Hamiltonian and Lagrangian formulations of nonholonomic systems. Because of the necessary replacement of conservation laws with the momentum equation, it is natural to let the value of momentum be a variable and for this reason it is natural to take a Poisson viewpoint. Some of this theory has been started in van der Schaft and Maschke [24]. We build on their work, further develop the theory of nonholonomic Poisson reduction, and tie this theory to other work in the area. We use this reduction procedure to organize nonholonomic dynamics into a reconstruction equation, a nonholonomic momentum equation and the reduced Lagrange-d’Alembert equations in Hamiltonian form. We also show that these equations are equivalent to those given by the Lagrangian reduction methods of Bloch, Krishnaprasad, Marsden and Murray [4]. Because of the results of Koon and Marsden [lo], this is also equivalent to the results of Bates and Sniatycki [2], obtained by nonholonomic symplectic reduction.

Xvo interesting complications make this effort especially interesting. First of all, as we have mentioned, symmetry need not lead to conservation laws but rather to a momentum equation. Second, the natural Poisson bracket fails to satisfy the Jacobi identity. In fact, the so-called Jacobiizer (the cyclic sum that vanishes when the Jacobi identity holds), or equivalently, the Schouten bracket, is an interesting expression involving the curvature of the underlying distribution describing the nonholonomic constraints.

The Poisson reduction results in this paper are important for the future development of the stability theory for nonholonomic mechanical systems with symmetry, as begun by Zenkov, Bloch and Marsden [2.5]. In particular, they should be useful for the development of the powerful block diagonalization properties of the energy-momentum method developed by Simo, Lewis and Marsden [23].

1. Introduction

The general setting. Many important problems in robotics, the dynamics of wheeled vehicles and motion generation, involve nonholonomic mechanics, which typically means mechanical systems with rolling constraints. Some of the important issues are trajectory tracking, dynamic stability and feedback stabilization (including nonmini- mum phase systems), bifurcation and control. Many of these systems have symmetry,

WI

Page 2: Poisson reduction for nonholonomic mechanical systems with symmetry

102 WANG SANG KOON and J. E. MARSDEN

such as the group of Euclidean motions in the plane or in space and this symmetry plays an important role in the theory.

Bloch, Krishnaprasad, Marsden and Murray [4], hereafter denoted [BKMM], ap- plied the methods of geometric mechanics to the Lagrange-d’Alembert formulation and extended the use of connections and momentum maps associated with a given symmetry group to this case. The resulting framework, including the nonholonomic momentum equation and nonholonomic mechanical connection, provides a setting for studying nonholonomic mechanical control systems that may have a nontrivial evolu- tion of their nonholonomic momentum.

The setting is a configuration space Q with a (nonintegrable) distribution 2) c TQ describing the constraints. For simplicity, we consider only homogeneous velocity constraints. We are given a Lagrangian L on TQ and a Lie group G acting on the configuration space that leaves the constraints and the Lagrangian invariant. In many examples, the group encodes position and orientation information. For example, for the snakeboard, the group is SE(2) of rotations and translations in the plane. The quotient space Q/G is called shape space.

The dynamics of such a system is described by a set of equations of the following form:

g& = -dnh(+ + I-i(r)p, (1.1)

p = i%(r)+ + ?%(T)p + pTD@)p, (1.2)

M(T)j: = S(T, i, p) + 7. (1.3)

The first equation is a reconstruction equation for a group element g, the sec- ond is an equation for the nonholonomic momentum p (strictly speaking p is the body representation of the nonholonomic momentum map, which is not conserved in general), and the third are equations of motion for the reduced variables T which describe the “shape” of the system. The momentum equation is bilinear in (i,p). The variable 7 represents the external forces applied to the system, and is assumed to affect only the shape variables, i.e., the external forces are G-invariant. Note that the evolution of the momentum p and the shape T decouple from the group variables.

This framework has been very useful for studying controllability, gait selection and locomotion for systems such as the snakeboard. It has also helped in the study of optimality of certain gaits, by using optimal control ideas in the context of non- holonomic mechanics [9, 221. Hence, it is natural to explore ways to develop similar framework on the Hamiltonian side.

Bates and Sniatycki [2], developed the symplectic geometry on the Hamiltonian side of nonholonomic systems, while [BKMM] explored the Lagrangian side. It was not obvious how these two approaches were equivalent, especially how the momentum equation, the reduced Lagrange-d’Alembert equations and the reconstruction equation correspond to the developments in [2].

Koon and Marsden [lo], established the specific links between these two sides and used the ideas and results of each to shed light on the other, deepening our understanding of both points of view. For example, in proving the equivalence of

Page 3: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 103

Lagrangian reduction and symplectic reduction, we have shown where the momentum equation lies on the Hamiltonian side and how this is related to the organization of the dynamics of nonholonomic systems with symmetry into the three parts displayed above: a reconstruction equation for the group element g, an equation for the non- holonomic momentum p and the reduced Hamilton equations for the shape variables r,p,.. Koon and Marsden [lo] illustrate the basic theory with the snakeboard, as well as a simplified model of the bicycle (see [7]). The latter is an important prototype control system because it is an underactuated balance system.

However, on the Hamiltonian side, besides the symplectic approach, one can also develop the Poisson point of view. Because of the momentum equation, it is natural to let the value of momentum be a variable and for this a Poisson rather than a symplectic viewpoint is more natural. Some of this theory has been started in van der Schaft and Maschke [24], hereafter denoted [VM]. We build on their work and de- velop the theory of Poisson reduction for the nonholonomic systems with symmetry. We use this Poisson reduction procedure to obtain specific formulae for the non- holonomic Hamiltonian dynamics. We also show that the equations given by Poisson reduction are equivalent to those given by the Lagrangian reduction via a reduced constrained Legendre transform.

Bvo special features of nonholonomic systems make this study interesting. First, as we have mentioned, symmetry need not lead to conservation laws but rather to a momentum equation. Second, the natural Poisson bracket fails to satisfy the Jacobi identity. In fact, the Jacobiizer (the cyclic sum that vanishes when the Jacobi identity holds) is an interesting expression involving the curvature of the underlying distri- bution describing the nonholonomic constraints. As is well known (see, for example, Marsden and Ratiu [17, 310.81, the failure of Jacobi’s identity may be equivalently measured by the Schouten bracket of the Poisson tensor with itself.

These results are important for the future development of the stability theory for nonholonomic mechanical systems with symmetry. Important progress towards a stability theory for these systems has been made in [25]. However, additional insight will be required for the development of the powerful block diagonalization properties of the energy-momentum method developed by Simo, Lewis and Marsden [23]. This technique is very important for the applicability of the energy-momentum method to complex systems.

Outline of the paper. In Section 2, we first consider general nonholonomic systems without symmetry assumptions. In this section,

1. We recall the basic ideas and results of [BKMM] on general nonholonomic sys- tems: in particular, how to describe constraints using Ehresmann connection and how to write the Lagrange-d’Alembert equations of motion using the curvature of this connection.

2. We review the Poisson formulation of nonholonomic systems in [VM]. 3. With the help of the Ehresmann connection, we use the Poisson procedure to

write a compact formula for the equations of the nonholonomic dynamics.

Page 4: Poisson reduction for nonholonomic mechanical systems with symmetry

104 WANG SANG KOON and J. E. MARSDEN

4. We prove the equivalence of the Poisson and Lagrange-d’Alembert formulations for the nonholonomic mechanics.

5. We develop a formula for the Jacobiizer that involves the curvature of the Ehres- mann connection.

In Section 3, we add the hypothesis of symmetry to the preceding development. In this section:

1. We recall the basic ideas and results of [BKMM] on simple nonholonomic me- chanical systems, especially on how it extended the Lagrangian reduction theory of Marsden and Scheurle [18, 191 to the context of nonholonomic systems. We shall describe briefly how [BKMM] modifies the Ehresmann connection associated with the constraints to a new connection, called the nonholonomic connection, that also takes into account the symmetries, and how the reduced equations, relative to this new connection, break up into two sets: a set of reduced Lagrange-d’Alembert equations, and a momentum equation. When the reconstruction equation is added, one recovers the full set of equations of motion for the system.

2. We build on the work of [VM] and develop the Poisson reduction, using the tools like the nonholonomic connection and nonholonomic momentum. We write the equations of motion for the reduced constrained Hamiltonian dynamics using a reduced Poisson bracket. This Poisson reduction procedure breaks the Hamiltonian nonholonomic dynamics into a reconstruction equation, a momentum equation and a set of reduced Hamilton equations.

3. We prove that the set of equations given by the Poisson reduction is equivalent to those given by the Lagrangian reduction via a reduced Legendre transform.

4. We apply the Poisson reduction procedure to the example of the snakeboard.

In the conclusions, we give a few remarks on the future research directions.

2. General nonholonomic mechanical systems

Following the approaches of [BKMM], we first consider mechanics in the presence of homogeneous linear nonholonomic velocity constraints. For now, no symmetry as- sumptions are made; we add such assumptions in the following section.

2.1. Review of the Lagrangian approach

We start with a configuration space Q with local coordinates denoted $, i = 1,. . . , n and a distribution D on Q that describes the kinematic nonholonomic constraints. The distribution is given by the specification of a linear subspace V, c T,Q of the tangent space to Q at each point q E Q. Consistent with the fact that each D, is a linear subspace, we consider only homogeneous velocity constraints. The extension to affine constraints is straightforward, as in [BKMM].

The dynamics of a nonholonomically constrained mechanical system is governed by the Lagrange-d’Alembert principle. The principle states that the equations of motion of a curve q(t) in configuration space are obtained by setting to zero the variations in

Page 5: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 105

the integral of the Lagrangian subject to variations lying in the constraint distribution and that the velocity of the curve q(t) itself satisfies the constraints; that is, G(t) E 2),tt). Standard arguments in the calculus of variations show that this “constrained variational principle” is equivalent to the equations

-6L:= ( ;g-$ sqi=o, > (2.1)

for all variations Sq such that Sq E D, at each point of the underlying curve q(t). These equations are often equivalently written as

ddL dL=x. ---- dt a$ aqi %’

where Xi is a set of Lagrange multipliers (i = 1,. . . ,n), representing the force of constraint. Intrinsically, this multiplier X is a section of the cotangent bundle over q(t) that annihilates the constraint distribution. The Lagrange multipliers are often determined by using the condition that Q(t) lies in the distribution.

In Bloch and Crouch [3] and Lewis [13], the Lagranged’Alembert equations are shown to have the form of a generalized acceleration condition

VgQ = 0

for a suitable affine connection on Q and the force of constraint X is interpreted as a generalized second fundamental form (as is well known for systems with holo- nomic constraints; see Abraham and Marsden [l], for example). In this form of the equations, one can add external forces directly to the right-hand sides so that the equations take the form of a generalized Newton law. This form is convenient for control purposes.

To explore the structure of the Lagranged’Alembert equations in more detail, let {w”}, a = l,..., lc, be a set of k independent one-forms whose vanishing describes the constraints; i.e., the distribution V. One can introduce local coordinates qi = (P, 8) where Q = l,..., n - k, in which wa has the form

wa(q) = dsa + A”,(r, S)dP,

where the summation convention is in force. In other words, we are locally writing the distribution as

The equations of motion (2.1) may be rewritten by noting that the allowed vari- ations 6qi = (&“, 6s”) satisfy 6s” + A:&” = 0. Substitution into (2.1) gives

d aL aL ---- dt a+ aTa

Equation (2.2) combined with the constraint equations

(2.2)

S .a = -A:F (2.3)

Page 6: Poisson reduction for nonholonomic mechanical systems with symmetry

106 WANG SANG KOON and J. E. MARSDEN

gives a complete description of the equations of motion of the system; this proce- dure may be viewed as one way of eliminating the Lagrange multipliers. Using this notation, one finds that X = Xawa, where X, = -$ $$ - $$.

Equations (2.2) can be written in the following way:

d 8L, dL, -- -- dt di* W

+A& = _dLBb +P Cl ds” fjib 00 ’ (2.4)

where LC(rcI, P, i*) = L(P, 8, F, -AE(r, s)P)

is the coordinate expression of the constrained Lagrangian defined by L, = LID and where

(2.5)

Letting dwb be the exterior derivative of wb, a computation shows that

dwb(i, .) = B$iadrP

and hence the equations of motion have the form

This form of the equations isolates the effects of the constraints, and shows, in particular, that in the case where the constraints are integrable (i.e., dw = 0), the equations of motion are obtained by substituting the constraints into the Lagrangian and then setting the variation of L, to zero. However in the non-integrable case the constraints generate extra (curvature) terms, which must be taken into account.

The above coordinate results can be put into an interesting and useful intrinsic geometric framework. The intrinsically given information is the distribution and the Lagrangian. Assume that there is a bundle structure ~Q,R : Q -+ R for our space Q, where R is the base manifold and ?TQ,$ is a submersion and the kernel of TQTQ,$ at any point q E Q is called the vertical space V,. One can always do this locally. An Ehresmann connection A is a vertical vector-valued one-form on Q such that

1. A, : T,Q + V, is a linear map and 2. A is a projection: Am) = zlq for all vq E V,. Hence, TPQ = V, @ I&, where H4 = kerA, is the horizontal space at q, sometimes

denoted hor,. Thus, an Ehresmann connection gives us a way to split the tangent space to Q at each point into a horizontal and vertical part.

If the Ehresmann connection is chosen in such a way that the given constraint distribution 29 is the horizontal space of the connection, that is, HP = VD,, then in the bundle coordinates qi = (P , sa), the map TQ,R is just projection onto the factor r and the connection A can be represented locally by a vector-valued differential form wa:

d Acwa- dsa ’

wa(q) = dsa + Az(r, s)dP,

Page 7: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 107

and the horizontal projection is the map

(P) P) H (i”, -A;(r, s)P).

The curvature of an Ehresmann connection A is the vertical-valued two-form de- fined by its action on two vector fields X and Y on Q as

B(X, Y) = -A([hor X, hor Y]),

where the bracket on the right-hand side is the Jacobi-Lie bracket of vector fields obtained by extending the stated vectors to vector fields. This definition shows the sense in which the curvature measures the failure of the constraint distribution to be integrable.

In coordinates, one can evaluate the curvature B of the connection A by the following formula:

B(X, Y) = dw”(hor X, hor Y)&,

so that the local expression for curvature is given by

B(X,Y)” = B$$XaYP,

where the coefficients BE0 are given by (2.5). The Lagrange-d’Alembert equations may be written intrinsically as

in which 6q is a horizontal variation (i.e., it takes values in the horizontal space) and B is the curvature regarded as a vertical-valued two-form, in addition to the constraint equations

A(q) . Q = 0.

Here ( , ) denotes the pairing between a vector and a dual vector and

As shown in [BKMM], when there is a symmetry group G present, there is a natural bundle one can work with and put a connection on, namely the bundle Q + Q/G. In the generality of the preceding discussion, one can get away with just the distribution itself and can introduce the corresponding Ehresmann connection locally. In fact, the bundle structure Q + R is really a “red herring”. The notion of curvature as a T,Q/D, valued form makes good sense and is given locally by the same expressions as above. However, keeping in mind that we eventually want to deal with symmetries and in that case there is a natural bundle, the Ehresmann assumption is nevertheless a reasonable bridge to the more interesting case with symmetries.

Page 8: Poisson reduction for nonholonomic mechanical systems with symmetry

108 WANG SANG KOON and J. E. MARSDEN

2.2. Review of the Poisson formulation

The approach of [VIM] starts on the Lagrangian side with a configuration space Q and a Lagrangian L (possibly of the form kinetic energy minus potential energy, i.e.,

L(% 4 = ; ((44)) - V(Q),

where (( , )) is a metric on Q defining the kinetic energy and V is a potential energy function.)

As above, our nonholonomic constraints are given by a distribution 23 c TQ. We also let 2)” c T*Q denote the annihilator of this ‘distribution. Using a basis ua of the annihilator Do, we can write the constraints as

w”(Q) = 0,

where a = l,...,k. As above, the

are written as basic equations are given by the Lagrange-d’Alembert principle and

where A, is a set The Legendre

d 823 aL =A (Jl ---- dt a$ dqi a 2,

of Lagrange multipliers. transformation IFL : TQ --t T*Q, assuming that it is a diffeomorphism,

is used to define the Hamiltonian H : T*Q -+ R in the standard fashion (ignoring the constraints for the moment):

H = (p, Q) - L = pig - L.

Here, the momentum is p = lFL(v,) = aL/a& Under this change of variables, the equations of motion are written in the Hamiltonian form as

aH !i”= dpi,

where i = l,... , n, together with the constraint equations

The preceding constrained Hamiltonian equations can be rewritten as

(2.6)

(2.7)

Page 9: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 109

Recall that the cotangent bundle T*Q is equipped with a canonical Poisson bracket and is expressed in the canonical coordinates (q,p) as

{J’,G)(q,p) = $$g - g$ = 2 2

(z,z)J(;).

Here J is the canonical Poisson tensor

J=

which is intrinsically determined by the Poisson bracket &I as

(2.9)

On Lagrangian side, we saw that one can get rid of the Lagrangian multipliers. On the Hamiltonian side, it is also desirable to model the Hamiltonian equations without the Lagrange multipliers by a vector field on a submanifold of T*Q. In [VM], it is done through a clever change of coordinates. We now recall how they do this.

First, a constraint phase space M = lFL(D) c T*Q is defined in [2] so that the constraints on the Hamiltonian side are given coordinates,

Let {Xa} be a local basis for the constraint distribution 23 and let {w”} be a local basis for the annihilator 2)O. Let {wa} span the complementary subspace to V such that (u’“,wb) = ~5; where 6; is the usual Kronecker delta. Here a = 1,. . . , k and cr=l ,*a*, n - k. Define a coordinate transformation (q,p) -+ (q,&,&) by

in the same way as by p E M. In local

_ PC2 = x~Pi, Ija = W~pi. (2.10)

[VM] shows that in the new (generally not canonical) coordinates (q,& ,&), the Poisson tensor becomes

and the constrained Hamiltonian equations (2.8) transform into

(2.11)

(2.12)

Page 10: Poisson reduction for nonholonomic mechanical systems with symmetry

110 WANG SANG KOON and J. E. MARSDEN

where fi(q,fi) is the Hamiltonian H(q,p) expressed in the new coordinates (q,lj).

Let (j&,&) satisfy the constraint equations &q, @) = 0. Since

[VM] uses (q,&) as induced local coordinates for M. It is easy to show that

where HM is the constrained Hamiltonian on M expressed in the induced coordinates.

Now we are ready to eliminate the Lagrange multipliers. Notice that $$(q,$) = 0 on M, and by restricting the dynamics on M, we can disregard the last equations involving X in Eqs. (2.12). In fact, we can also truncate the Poisson tensor j in (2.11) by leaving out its last Ic columns and last k rows and then describe the constrained dynamics on M expressed in the induced coordinates (qi, Ija) as follows

Here JM is the (2n-Ic) x (2n- Ic) truncated matrix of 9 restricted to M and is expressed in the induced coordinates.

The matrix JM defines a bracket {, }M on the constraint submanifold M as follows

for any two smooth functions FM, GM on the constraint submanifold M. Clearly this bracket satisfies the first two defining properties of a Poisson bracket, namely, skew symmetry and Leibniz rule, and it is shown in [WI] that it satisfies the Jacobi identity if and only if the constraints are holonomic. Furthermore, the constrained Hamiltonian HM is an integral of motion for the constrained dynamics on M due to the skew symmetry of the bracket.

In Section 2.5, we will develop a general formula for the Jacobiizer (the cyclic sum that vanishes when the Jacobi identity holds) which is an interesting expression involving the curvature of the underlying distribution that describes the nonholonomic constraints. From this formula, one can see clearly that the Poisson bracket defined here satisfies the Jacobi identity if and only if the constraints are holonomic.

REMARKS. In [14] it has been shown that the bracket {, }M obtained in [VM] can be given an intrinsic interpretation as follows:

Page 11: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 111

1. Given a constrained Hamiltonian system {T* Q, {, }, H, M, V} where the first four objects are defined as above and the last object V is a vector subbundle of TJM(T*Q) defined by

Here, Tm(T* Q) is the restriction of the tangent bundle of T(T*Q) to the constraint submanifold M, and vert,(n) E T,(T*Q) is the vertical lift of 77 E T,*Q where p E T*Q; in coordinates, vert(,,,)(q,q) = (q,p,O,v). Marle [14] uses this subbundle V to encode the fact that the constraint forces obey the Lagrange- d’Alembert principle. It can be shown that the sum of the vector subbundles TM and V of TM(T*Q) is a direct sum, and

TM $ V = TM(T*Q).

Moreover, the restriction XH]~ of the Hamiltonian vector field to the constraint submanifold M splits into a sum

XHIM = XM +xV,

where XM is the constrained Hamiltonian vector field which is tangent to M and Xv is a smooth section of the subbundle V (whose opposite is the constraint force field).

2. The canonical Poisson tensor A (associated with the Poisson bracket {, }) of T*Q can be projected on M, and its projection AM is a contravariant skew symmetric 2-tensor on M. More precisely, let p E M, Q and /3 E TPM. There

exists a unique pair (15, p) of elements of T,*(T* Q) which vanish on the vector subspace V, of T,(T*Q), and whose restrictions to the vector subspace T,M are CY and p, respectively. We can therefore define AM@) by setting

AM(P)(‘U, P) = A(P)(& b>.

Then the vector field XM, whose integral curves are the motions of the system, is

X,,,, = AL@&).

Using the 2-tensor AM, Marle [14] defines an intrinsic bracket {, }M on the space of smooth functions on M by setting

{f, S’)M = A,(#, ‘-67).

In the local coordinates (q,$), the truncated matrix JM(q,&) obtained in [VM] is exactly the matrix associated with the 2-tensor AM and is nothing but the projection of the Poisson matrix j(q,$) onto the constraint submanifold M.

The above considerations show that the nonholonomic brackets constructed in the present paper by a quotient operation, are also intrinsic.

Page 12: Poisson reduction for nonholonomic mechanical systems with symmetry

112 WANG SANG KOON and J. E. MARSDEN

2.3. The constrained Hamilton equations

In holonomic mechanics, it is well known that the Poisson and the Lagrangian formulations are equivalent via a Legendre transform. And it is natural to ask whether the same relation holds for the nonholonomic mechanics as developed in [VM] and [BKMM]. But before we answer this question in the next section, we would like first to use the general procedures of [VM] to write down a compact formula for the nonholonomic equations of motion.

THEOREM 2.1. Assume that we have same setup as in the preceding section. Let q’ = (T”, sa) be th e 1 ocal coordinates in which wa has the form

wa(q) = dsa + Az(r, s)drQ, (2.14)

where A;(r, s) is the coordinate expression of the Ehresmann connection described in Section 2.1. Then the nonholonomic constrained Hamilton equations of motion on M can be written as

OHM b OHM b aHM &=--+AaT- - PbBcQ 85, T

(2.15)

(2.16)

(2.17)

where B& are the coeficients of the curvature of the Ehresmann connection given in Eq. (2.5). Here, Pb should be understood as Pb restricted to M and more prectiely should be denoted as (Pb),,,&

Proof: As mentioned in Section 2.1, no additional assumption is needed since one can always choose local coordinates in which

w”(q) = dsa + A;(r, s)dr*.

In this local coordinate system,

D = span{&.- - A”,&}.

Then the new coordinates (ra, sa, $,, J?J~) of [VM] are defined

@a=pa-A~pa, Ija=pa+A~pa,

and we can use (ra, sa,&) as the induced coordinates on M.

by

(2.18)

(2.19)

Moreover, we can find the constrained Poisson structure matrix J,$,j(T”, sa, &) by computing {qi, qj}, {qi,fio}, {&,@p} and then restrict them to M. Recall that JM is constructed out of the Poisson tensor 3 in Eq. (2.11) by leaving out its last k columns and last k rows and restricting its remaining elements to M.

Page 13: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON IEDUCTI-ION FOR NONHOI_.ONOMIC SYSTEMS

Clearly

In addition, we have

{+%x) = {@A - A:p,} = {&p~} - {TO, A&z} = 6,p,

(~b,2j,} = {s~,P, - A&z} = (s*,p,) - {sb, A;p,} = -A;,

where 6: is the usual Kronecker delta. It is also straightforward to find

113

{Pa - &pa, pp - A$&} --{pm A$pb) - {A:P~,P~) + (&pa, A$%}

After restricting the above results to M, all other terms remain the same but the last line should be understood as -@&(&)M. But for notational simplicity, we keep

writing it as ~~~~pb. Putting the above computations together, we can write the nonholonomic equations of motion as follows

which is the desired result. Notice that the order of the variables r” and S= have been switched to make the block diagonalization of the constrained Poisson tensor more apparent. Also, it may be important to point out here that @b)M can be expressed in terms of the induced coordinates (Q,&) on the constraint submanifold M by the following formula

where gba and i& are the components of the kinetic energy metric and Kf is defined by the last equality.

Page 14: Poisson reduction for nonholonomic mechanical systems with symmetry

114 WANG SANG KOON and J.E. MARSDEN

2.4. The equivalence of Poisson and Lagrange-d’Alembert formulations

Now we are ready to state and prove the equivalence of the Poisson and Lagrange- d’Alembert formulations.

THEOREM 2.2. The Lagrange-d’Alembeti equations

$= = -AZ=, (2.21)

d dL, dL, dL, dL ---- dt W W’

+ A+ = -=B$@ (2.22)

are equivalent to the constrained Hamilton equations

.a - s a OHM

- -A, &j, ’

via the constrained Legendre transfomt given by

(2.23)

(2.24)

(2.25)

(2.26)

Proof: Recall that

~={(~,s,~,~)ETQ(~+A~~~=O}.

We can use (T, s, i) as the induced coordinates for the submanifold 2). Since the con- strained Lagrangian is given by

-Ura, 8, ta) = L(r=, sa, F, -A:(r, s)F),

we have dL, dL -= -- a+” a+ EA; = p, -p,A; = &. (2.27)

Hence, $$$ = Ija does define the right constrained Legendre transform between the submanifolds 2, and M with the corresponding induced coordinates (TO, 8, F) and

(rQ, sa, Ax). Now notice that if E = $$ .’ qz - L, then restricting it to D we will get

aLc.a_L, =-_T

a+ c

Page 15: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 115

Hence, the constrained Hamiltonian is given by

HM = jaia - L,.

And it is straightforward to show that

(2.28)

which gives the Eq. (2.24). Clearly, da = -A;@ together with Eq. (2.24) gives Eq. (2.23).

Furthermore, we have

and

OHM ata dL, aL, di” aL, -=?Ljijp-pj-$qp= -- ara aI@ ’

(2.29)

(2.30)

Substituting the results of (2.29) and (2.30) into Eq. (2.22), we get the remaining Eq. (2.25). 0

2.5. A formula for the Jacobiizer

Recall that in the proof of Theorem 2.1, we have obtained

Clearly

~{qi~d)M7qk)M + qclic=o, {{$,q’}M,fia},4A + cyclic = 0.

It is also straightforward to obtain

{-iTY,fh)M,ljp)~ + cyclic = KiY@p,

{{~~,&}~,ljp}M + cyclic = -BEa - A~K~B$.

As for {{fia,~~}M}M + cyclic, it takes slightly more work to find

(2.31)

(2.32)

(2.33)

(2.34)

t3B$ aB$ drr - +jp- + cyclic.

Page 16: Poisson reduction for nonholonomic mechanical systems with symmetry

116 WANG SANG KOON and J. E. MARSDEN

Notice that the right-hand side of the last equation involves the derivatives of the curvature. However, it can be shown that the following (Bianchi) identity

a%3 dry

AC aB:,, aA; -- - 7 asc + BED-

asa + cyclic = 0

holds, and we can apply it to rewrite the last equation using only the curvature but not its derivatives

~~~~~I$}M,&}M + cyclic =&K,‘B&K,6B$ -g z - A;%) Bip

+&K,’ - aA’ B$ + cyclic. ad

(2.35)

Eqs. (2.31X2.35) give the Jacobiizer of the Poisson bracket on M. Of course, one can also use the formalism of the Schouten bracket to do the computations and obtain the same results.

Notice that from the formulae for the Jacobiizer, one can see clearly that if the constraints are holonomic and hence the Ehresmann connection has zero curvature, then the Jacobiizer is zero and the Jacobi identity holds. Conversely, if the Jacobi identity holds, then we have

0 = KYBb b ap,

0 = -B” - A”KTBb 4 Y b a@’

Therefore, BE, = 0 and the constraints are holonomic.

3, Nonholonomic mechanical systems with symmetry

Now we add the hypothesis of symmetry to the preceding development. Assume that we have a configuration manifold Q, a Lagrangian of the form kinetic minus potential, and a distribution D that describes the kinematic nonholonomic constraints. We also assume there is a symmetry group G (a Lie group) that leaves the Lagrangian invariant, and that acts on Q (by isometries) and also leaves the distribution invariant, i.e., the tangent of the group action maps VD, to ID,, (for more details, see [BKMM].) Later, we shall refer this as a simple nonholonomic mechanical system.

3.1. Review of Lagrangian reduction

We first recall how [BKMM] explains in general terms how one constructs reduced systems by eliminating the group variables.

PROPOSITION 3.1. Under the assumptions that both the Lagrangian L and the distri- bution D are G-invariant, we can form the reduced velocity phase space TQ/G and the constrained reduced velocity phase space D/G. The Lagrangian L induces well defined functions, the reduced Lagrangian

1: TQ/G + IR

Page 17: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 117

satisjj&g L = 1 o TTQ where TTQ : TQ + TQ/G is the projection, and the constrained reduced Lagrangian

1, : D/G + IR,

which satisfies L(V = 1, o XV where RD : V ---) D/G is the projection. Also, the Lagrange-d’Alembert equations induce well defined reduced Lagrange-dAlembert equa- tions on D/G. That is, the vector field on the manifold D determined by the Lagrange-d’Alembert equations (including the constraints) is G-invariant, and so de- fines a reduced vector field on the quotient manifold D/G.

This proposition follows from general symmetry considerations, but to compute the associated reduced equations explicitly and to reconstruct the group variables, one defines the nonholonomic momentum map Jnh, and extends the Noether theorem to nonholonomic system and synthesizes, out of the mechanical connection and the Ehresmann connection, a nonholonomic connection dnh which is a connection on the principal bundle Q + Q/G.

The nonholonomic momentum map. Let the intersection orbit and the distribution at a point q E Q be denoted

S, = D, n T,JOrb(q)).

Define, for each q E Q, the vector subspace gQ to be the in g whose infinitesimal generators evaluated at q lie in

gq = (< E g 1 <Q(q) E sq).

of the tangent to the group

set of Lie algebra elements s,:

We let gv denote the corresponding bundle over Q whose fiber at the point q is given by gQ. The nonholonomic momentum map Jnh is the bundle map taking TQ to the bundle (gn)’ (whose fiber over the point q is the dual of the vector space gq) that is defined by

(Jnh(uq), t.) = g(tQ)“> (3-l)

where c E g’J. Notice that the nonholonomic momentum map may be viewed as encoding some of the components of the ordinary momentum map, namely the pro- jection along those symmetry directions that are consistent with the constraints.

[BKMM] extends the Noether theorem to nonholonomic systems by deriving the equation for the momentum map that replace the usual conservation law. It is proven that if the Lagrangian L is invariant under the group action and <q is a section of the bundle gn, then any solution q(t) of the Lagranged’Alembert equations must satisfy, in addition to the given kinematic constraints, the momentum equation:

(3.2)

When the momentum map is paired with a section in this way, we will just refer to it as the momentum. Examples show that the nonholonomic momentum map may or may not be conserved.

Page 18: Poisson reduction for nonholonomic mechanical systems with symmetry

118 WANG SANG KOON and J. E. MARSDEN

The momentum equation in body representation. Let a local trivialization (T, g) be chosen on the principal bundle T : Q -+ Q/G. Let n E gq and c = g-i& Since L is G-invariant, we can define a new function 1 by writing L(r,g,b, 4) = l(r,i,<). Define J p,“, : TQ/G + @D>* by

(JP,“,<vX),rl) = (6,~).

As with connections, Jnh and its version in a local trivialization are related by the Ad map; i.e.,

Jnh(y g, +, 4) = Ad;-1 J;:(r, i, 0.

Choose a q-dependent basis e,(q) for the Lie algebra such that the first m elements span the subspace g 9. In a local trivialization, one chooses, for each r, such a basis at the identity element, say

cl(T), e2(T), . . . , e,(T), h+l(T), . . . , a(T).

Define the body fixed bash by

ea(T, g) = Ad, . e&T);

thus, by G invariance, the first m elements span the subspace gq. In this basis, we have

( Jnh(r, g, f, i), eb(T, 9)) = ($e,(,,) := Pa, (3.3)

which defines Pb, a function of T, 1: and c. Note that in this body representation, the functions pb are invariant rather than equivariant, as is usually the case with the mo- mentum map. It is shown in [BKMM] that in this body representation, the momentum equation is given by

d -gi = (g, [5, ei] + $+O) ,

where the range of i is 1 to m. Moreover, the momentum equation in this representation is independent of, that is, decouples from, the group variables g.

The nonholonomic connection. Recall that in the case of simple holonomic mechanical system, the mechanical connection A is defined by d(v,) = II(q)-lJ(vq) where J is the associated momentum map and II(q) is the locked inertia tensor of the system. Equivalently the mechanical connection can also be defined by the fact that its horizontal space at q is orthogonal to the group orbit at q with respect to the kinetic energy metric. For more information, see, for example, [15, 171.

As [BKMM] points out, in the principal case where the constraints and the orbit directions span the entire tangent space to the configuration space, that is,

v, + T,(OrW) = TqQ7 (3.5)

Page 19: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 119

the definition of the momentum map can be used to augment the constraints and provide a connection on Q + Q/G. Let Jnh be the nonholonomic momentum map and define similarly as above a map A;Y” : TPQ + S, given by

Asym(wq) = (Ph(q)-lJ”h(wq))Q

(this defines the momentum “constraints”), where Inh : gD + (g”)* is the locked inertia tensor defined in a similar way as in holonomic systems.

Choose a complementary space to S, by writing T,(Orb(q)) = S, $ U,. Let Atin : T,Q + U, be a U, valued form that projects U, onto itself and maps V, to zero. Then the kinematic constraints are defined by the equation

Akin(q)4 = 0.

This kinematic constraints equation plus the momentum “constraints” equation can be used to synthesize a nonholonomic connection dnh which is a principal connection on the bundle Q + Q/G and whose horizontal space at the point q E Q is given by the orthogonal complement to the space S, within the space VD,. Moreover,

dnh(w,) = IInh(q)-lJnh(v,).

In a body fixed basis, (3.6) can be written as

(3.6)

Ad,(g-‘8 + df,h,(r)i) = Ad,(I;i#-)-lp).

Hence, the constraints can be represented in a nice way by

g-lb = < = -d(r)+ + In, (3.7)

where d(T) is the abbreviation for d::(r) and I’(P) = II&t(r)-‘. Moreover, with the help of nonholonomic mechanical connection, the Lagrange-

d’Alembert principle may be broken up into two principles by breaking the variations 6q into two parts, namely parts that are horizontal with respect to the nonholonomic connection and parts that are vertical (but still in D), and the reduced equations break up into two sets: a set of reduced Lagrange-d’Alembert equations (which have curvature terms appearing as ‘forcing’), and a momentum equation, which have a form generalizing the components of the Euler-Poincare equations along the symmetry directions consistent with the constraints. When one supplements these equations with the reconstruction equations, one recovers the full set of equations of motion for the system.

3.2. Poisson reduction

Now let G be the symmetry group of the system and assume that the quotient space fi = M/G of the G-orbit in M is a quotient manifold with projection map p : M + J% Since G is a symmetry group, all intrinsically defined vector fields push down to M. In this section, we will write the equations of motion for the reduced

Page 20: Poisson reduction for nonholonomic mechanical systems with symmetry

120 WANG SANG KOON and J. E. MARSDEN

constrained Hamiltonian dynamics using a reduced “Poisson” bracket on the reduced constraint phrase space fi. Moreover, an explicit expression for this bracket will be provided.

The crucial step here is how to represent the constraint distribution 2) in a way that is both intrinsic and ready for reduction. The work in both [BKMM] and [lo] suggests that we should use the tools like nonholonomic momentum p and the non- holonomic connection A in [BKMM] to describe 27

Recall that in [BKMM], a body fixed basis

et&, T) = Ad, . a(r)

has been constructed such that the infinitesimal generators (ei(g,r))Q of its first m elements at a point q span S, = DD, f~ T,(Orb(q)). Assume that G is a matrix group and et is the component of ei(r) with respect to a fixed basis {b,} of the Lie algebra g where (b,)~ = dso, then

(ci(% r))Q = g$ef&.

Since DD, is the direct sum of S, and the horizontal space of the nonholonomic connec- tion A, it can be represented by

(3-g)

Before we state the theorem and do some computations, we want to make sure that the readers understand the index convention used in this section:

1. The first batch of indices is denoted a b c , , 7-e. and range from 1 to Ic corresponding to the symmetry direction (Ic = dim g).

2. The second batch of indices will be denoted i, j, Ic, . . . and range from 1 to m corresponding to the symmetry direction along constraint space (m is the number of momentum functions).

3. The indices a, /3,. . . on the shape variables T range from 1 to n - k (n - k = dim (Q/G), i.e., the dimension of the shape space).

Then the induced coordinates (ga, ra ,$i, fia) for the constraint submanifold M are de- fined by

Iji = gietpp, = met, (3.9) _

~a = P, - ga”&pa = P, - CL&-$ (3.10)

Here p is an element of the dual of the Lie algebra g* and pa are its coordinates with respect to a fixed dual basis. Notice that ji are nothing but the corresponding momentum functions on the Hamiltonian side.

We can find the constrained Poisson structure matrix JM (ga, ~~~ &, &) by computing {ga, g*}, etc. and then restrict them to M. Recall that JM is constructed out of the Poisson tensor f in (2.11) by leaving out its last k columns and last k rows and restricting its remaining elements to M.

Page 21: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON RJ2DUCTION FOR NONHOLONOMIC SYSTEMS 121

Clearly

{ga, g*} = 0, {g”, P} = 0, {P, P} = 0.

And we also have

It is also straightforward to find

where 15’:~ are the structure constants of the Lie algebra 0. Similarly, we have

and

aA; dAb - - = pb ara pb 87-P 2 - &,Aa,A;

where B$ are the coefficients of the curvature of the nonholonomic connection and are given by

Bb _ aA: aA; 4---- 87-0 87-a

+ C;,AzA;.

Page 22: Poisson reduction for nonholonomic mechanical systems with symmetry

122 WANG SANG KOON and J.E.MAFtSDEN

Therefore the constrained Hamilton equations can be written as follows

where

0 573; 0 -@A;

-(d4T -p,Cbqzefejd 0 hF,ap (3.11)

0 0 0 9

(SX )T -(cLJ$)~ -St -Q$

F$ is defined by

F” - de,a ; c” e6Ad 20 - &4 bdz 0’ (3.12)

Notice that the order of the variables TV and ji have been switched to make the diagonalization of the constrained Poisson tensor more apparent.

Since G is the symmetry group of the system and the Hamiltonian H is G-invariant, we have HM = ha. Hence

aH,=O 8gb ’

OHM dhM -=- ?@j afij ’

OHM _ ah

&-a &-P ’ OHM 8hM

XT=-. %9

After the reduction by the symmetry group G, we have

0 e j” 0 -A;

-(ez)T -paCtdepej” 0 i-kF,ap

0 0 0 &P P

GwT -~#“a)~ -6: -PUB&

(3.13)

where Eb = (g-‘)i$. The above computations prove the following theorem

THEOREM 3.2. The momentum equation and the reduced reduced constraint submanifold &I can be written as follows

.Q ah&f r =-

afi, ’

Hamilton equations on the

(3.14)

(3.15)

(3.16)

Page 23: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 123

Adding in the following reconstruction equation

i” = -APdlig bah&f +e?% 3 a& ’

(3.17)

we recover the full dynamics of the system.

Notice that Eq. (3.14) can be considered as the momentum equation on the Hamiltonian side which corresponds to the momentum equation developed in [BKMM]. It generalizes the Lie-Poisson equation to the nonholonomic case.

Moreover, if we now truncate the reduced Poisson matrix j in Eq. (3.13) by leaving out its first column and first row, the new matrix JM given by

defines a bracket {, }M on the reduced constraint submanifold ti as follows

for any two smooth functions FM, GM on the reduced constraint submanifold fi whose induced coordinates are (iii, ra, &). Clearly this bracket satisfies the first two defining properties of a Poisson bracket, namely, skew-symmetry and Leibniz rule. Moreover, the reduced constrained Hamiltonian hA is an integral of motion for the reduced Hamiltonian dynamics on fi due to the skew symmetry of the reduced bracket.

3.3. The equivalence of Poisson and Lagrangian reduction

THEOREM 3.3. The Eqs. (3.14)-(3.17) gz ken by the Poisson reduction are equivalent to the equations given by the Lagrangian reduction

5’ = -Ai@ + rbipi = -A$$ + e!@, f (3.19)

(3.20)

(3.21)

via a reduced Legendre transform

Page 24: Poisson reduction for nonholonomic mechanical systems with symmetry

124 WANG SANG KOON and J. E. MARSDEN

Proof: Define the reduced constrained Lagrangian

ZC(~, i, Cl) = Z(r, i, -A+ + fle),

where fi is the body angular velocity and e(r) is the body fked basis at the identity defined earlier. Notice first that

a1 dL -= - =p,. a+* aia

Since

we have

Hence,

al - = pa. ae

a1 a1 Aa =--- df” aca a

= Pa - PaA: _ = PCY,

and

That is, & = s and fii = $$ do define the right reduced constrained Legendre trans- form between the reduced constraint submanifolds 3 and fi with the corresponding reduced coordinates (P , ia, ni) and (P, &, &).

To find the reduced constrained Hamiltonian hm, notice first that since E is G- invariant, we have

After restricting it to the submanifold 23, we have

Page 25: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 125

Therefore, we have hM = $-li + ljafa - l,, (3.22)

via the Legendre transform (P, ia, 0”) - (P,&,&). Differentiate hM with respect to Ija and jj and use the Legendre transform, we have

which is Eq. (3.15). Also, we have

ai, a+* ai, aiY --- -- ai, aj+ aa @j

= fij,

which, together with Eq. (3.19), gives Eq. (3.17). Moreover, since $ = &$i = pb

and fii = pi, we have

which is Eq. (3.14). Finally, differentiate hM with respect to P, we have

which together with Eq. (3.21) gives

ah&f ah&f jj, = -- -p,Fjaol- - a5j

which is Eq. (3.16).

Remark: Notice that Eqs. (3.21) are the same as the reduced Lagrange-d’Alem- bert equations in [BKMM]. The only difference is that in this paper, the reduced constrained Lagrangian 1, is a function of r, 7;,Q where in [BKMM] and [lo] it is considered as a function of r, i,p. Since it is more natural to use the body angular velocity as a variable on the Lagrangian side, the formulation here looks better.

Page 26: Poisson reduction for nonholonomic mechanical systems with symmetry

126 WANG SANG KOON and J. E. MARSDEN

3.4. Example: the snakeboard

The snakeboard is a modified version of a skateboard in which the front and back pairs of wheels are independently actuated. The extra degree of freedom enables the riders to generate forward motion by twisting their body back and forth, while simultaneously moving the wheels with the proper phase relationship. For details, see [BKMM] and the references listed there.

The snakeboard is modeled as a rigid body (the board) with two sets of independently actuated wheels, one on each end of the board. The human rider is modeled as a momentum wheel which sits in the middle of the board and is allowed to spin about the vertical axis. Spinning the momentum wheel causes a counter-torque to be exerted on the board. The configuration of the board is given by the position and orientation of the board in the plane, the angle of the momentum wheel, and the angles of the back and front wheels. Let (x, y, 0) represent the position and orientation of the center of the board, $-the angle of the momentum wheel relative to the board, and & and $2 the angles of the back and front wheels, also relative to the board. Take the distance between the center of the board and the wheels to be r. See Fig. 3.1.

Fig. 3.1. The geometry of the snakeboard.

In [BKMM], a simplification is made which we shall also assume in this paper, namely 41 = ~$2, Ji = Ja. The parameters are also chosen such that J + JO + Ji + Jz = mr2, where m is the total mass of the board, J is the inertia of the board, JO is the inertia of the rotor and JI, Ja are the inertia of the wheels. This simplification eliminates some terms in the derivation but does not affect the essential geometry of the problem. Setting 4 = 41 = -$a, then the configuration space becomes Q = SE(2) x S1 x S1 and the Lagrangian L : TQ -+ R is the total kinetic energy of the system and is given by

L(q, 4) = frm(ci2 + jl”) + irnr2b2 + f 504~ + J&b + Jld”.

Clearly .Z is independent of the configuration of the board and hence it is invariant to all possible group actions by the Euclidean group SE(2).

The constraint submanifold. The rolling of the front and rear wheels of the snakeboard is modeled using nonholonomic constraints which allow the wheels to spin about the vertical axis and roll in the direction that they are pointing. The wheels are not allowed

Page 27: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 127

to slide in the sideways direction. This gives the following constraint one-forms

wl(q) = - sin(8 + ~$)da: + cos(0 + ~$)dy - T cos $de,

wz(q) = - sin(8 - 4)dz + cos(0 - 4)dy + r cos +dO,

which are also invariant under the SE(2) action. matic distribution VD,:

D, = span{&, 84, a& +

where a = -2r cos2 4 cos 8, b = -2r cos2 4 sin 0, c orbits of the SE(2) action is given by

The constraints determine the kine-

b&I + c&?),

= sin24. The tangent space to the

T,@rb(q)) = span{&, a,, 80)

The intersection between the tangent space to the group orbits and the constraint distribution is thus given by

S, = Dq n Z’,(Orb(q)) = span{a& + MY + CC?,}.

The momentum can be constructed by choosing a section of S = D r-l T(Orb) regarded as a bundle over Q. Since DD, n T,(Orb(q)) is one-dimensional, the section can be chosen to be

(6 = ad, + bL$, + c&,

which is invariant under the action of SE(2) on Q. The nonholonomic momentum is thus given by

= mai + mbyj + mr2c6 + Jo&.

The kinematic constraints plus the momentum are given by

0 = - sin(8 + 4): + c0s(e + c#I)$ - r cos $4,

0 = - sin(0 - ~~5)i + c0qe - $)jl + r cos ~$4,

p = -2mr cos2 4 cos ek - 2mr cos2 C$ sin e?j

+ mr2 sin 246 + Jo sin 244.

Adding, subtracting, and scaling these equations, we can write (away from the point b = n/2),

Page 28: Poisson reduction for nonholonomic mechanical systems with symmetry

128 WANG SANG KOON and J. E. MARSDEN

These equations have the form

g-‘jr + A(T)+ = r(r)p,

where

A(r)=-& sin 24 e, dlC, + -$$ sin’ 4 ee d$,

r(r) = $ez + & tan$ee.

These are precisely the terms which appear in the nonholonomic connection relative to the (global) trivialization (T, g).

Since rp = ne, we can rewrite the constraints using the angular momentum fi as follows

The reduced constrained Hamiltonian. From the Lagrangian L, we find the reduced Lagrangian

Z(r,i,<) = fm((~l)2 + (C2)“) + fmr2(<3)2 + 3Ja?j” + +Jc?j(<s) + J&

where t = 9-l). After plugging in the constraints (3.24), we have the reduced con- strained Lagrangian

I,(r,i,il) = - J,2 2mr2 sin2 44” + 2mr2 cos2$fi2 + i 504~ + JI~!J”. (3.25)

Then the reduced constrained Legendre transform is given by

And its inverse is

fi= p 4mr2 cos2 C$ ’

+ mr2&

Jo(mr2 - JO sin2 4) ’

Page 29: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS

Therefore, the reduced constrained Hamiltonian hM is

h,=pil+$&+&&E,

129

se3 4 2 mr2 1

= ?GiFp + 2Jo(mr2 - JO sin2 #J) P$ + 4J,P”T

The reduced Poisson matrix. Recall that in computing the reduced structural matrix, we only need to calculate {&,$p}, etc. and then restrict them to &f. Since

p= -2rcos2+cost9p, -2rcos2f$sinOp, +sin2$pe,

Jo p$ = 2mr2 sin24cosBp, + & sin2$sinBp, - $ sin2 4p0 + P$,

64 = P$,

we have

{p,&} = {-2rcos2$~1,p4} + {sin2+ps,p$} = 2rsin2$pl + 21x1~24~3.

Similarly, we find

{P&l = 0.

As for ~1, ~2, p3 (restricted to &f ), recall that

p1 = costlp, + sinfIp,

= cos e(mk) + sin O(mG)

= $cos$sin#$ - $p

= mrsin$cos$ _

2 P*- mr2 - JO sin 4

iP.

We can also find ~2, ,LL~ in a similar way. Therefore

- mrsinfjcos4

II _ lj$

--1 - Pl (mr2 - JO sin2 4) 2rp l-42 = 0 + 0

CL3 mr2cos2 4 _ tan 4

(mr2 - JO sin2 4)” A _TP_

So after substituting the constraints (3.29) into Eqs. (3.26)-(3.28), we have

(3.26)

(3.27)

(3.28)

(3.29)

(3.30)

(3.31)

(3.32)

Page 30: Poisson reduction for nonholonomic mechanical systems with symmetry

130 WANG SANG KOON and J. E. MAFCSDEN

Therefore the reduced Poisson matrix is given by

/ -2rcos2$ ’ 0 0 0 0 0 & sin 24 0

0 0 0 0 0 0 0 0

0 0 0 sin 24 0 0 -&sin24 0

2rcos2qi 0 - sin 2f$ 0 0 0 0 {PJ%LGl

0 0 0 0 0 0 1 0

0 0 0 0 0 0 0 1

-&sin24 0 &sin24 0 -1 0 0 ~~*&~&f

\ 0 0 0 -~P,Ij$Jhl 0 -1 4c&hl 0 ,

where {p,&,}~ and {&,&}M are given as above by (3.30) and (3.31).

The reduced constrained Hamilton equations. It is straightforward to find that

dhfi _ sec2 C#I --wP’ ap

\

,

I

ah, _ sec2 C#J tan4 2 mr2 sin 24 - - a$ 4mr2 ’ + 2(mr2 - Ja sin2 $)2”’

ah&l _ mr2 -- 85, J0(mr2 - Jo sin2 4)

13+

Then by using the formula in (3.13) and after some computations, we obtain the mo- mentum equation and the reduced constrained Hamilton equations as follows

li= (

- tan4p + 2mr2 cos2 C$

2 ” mr2 - Jo sin 4 > i,,, 2Jr

+ mr2

Jo(mr2 - JO sin2 C#J) filil,

$ = &$, 1

6, = - Jo sin 24

2(mr 2 - JO sin2 4) ‘* ZJ, > i,$,

j34 =o.

(3.33)

(3.34)

(3.35)

(3.36)

(3.37)

Page 31: Poisson reduction for nonholonomic mechanical systems with symmetry

POISSON REDUCTION FOR NONHOLONOMIC SYSTEMS 131

Also, we can obtain the following reconstruction equations on the Hamiltonian side

k=[1cos0-[2sin8 = ( -&+ T sin 24

2(mr2 - JO sin2 4) j@ case,

> (3.38)

$=<1sin8-~2cos0= (

-&+ r sin 24

2(mr2 - Jo sin2 4) $li, sint9,

> (3.39)

e=~3=gE&_ sin2 4

mr2 - Jo sin2 4 ?% (3.40)

Together, these two sets of equations give us the dynamics of the full constrained systems but in a form that is suitable for control theoretical purposes.

4. Conclusions and future work

This paper has continued the work of Koon and Marsden [lo] in comparing the Hamiltonian and Lagrangian approaches to nonholonomic systems., This paper, together with [lo], builds on the recent advances made by Bates and Sniatycki [2], van der Schaft and Maschke [24], Bloch, Krishnaprasad, Marsden and Murray [15] and others in the study of nonholonomic systems. It helps to lay a firm foundation for a gauge viewpoint of such systems; that is, the viewpoint of principal bundles (described in, for example, [15]) that has been so useful in stability theory, geometric phases and related matters.

[BKMM] has started this work on the Lagrangian side and generalized the use of connections and momentum maps associated with a given symmetry group to non- holonomic systems. It has shown how Ehresmann connections can be introduced to write the kinematic constraints as the condition of horizontality with respect to this connection and it has shown how the equations of motion can be written in terms of base variables and that these equations involve the curvature of the connection. It has also shown that the presence of symmetries in the nonholonomic case may or may not lead to conservation laws and has developed the momentum equation, which plays an important role in control problems for such systems. The process of reduction and reconstruction for these systems is worked out by making use of a nonholonomic connection which is obtained by synthesizing the mechanical connec- tion and constraint connection. Moreover, using the tools of Lagrangian reduction, it developed the reduced Lagrange-d’Alembert equations.

This paper together with Koon and Marsden [lo] extend this gauge viewpoint of such systems to the Hamiltonian side, building on the works of Bates and Sniatycki [2] and Van der Schaft and Maschke [24]. With the help of nonholonomic connections and momentum maps, the present paper develops the Poisson reduction of nonholo- nomic systems with symmetry. It shows that Lagrangian reduction for nonholonomic mechanics is equivalent to both the symplectic reduction and the Poisson reduction via a reduced constrained Legendre transform. But most importantly, it shows where the momentum equation lies on the Hamiltonian side and how this is related to the

Page 32: Poisson reduction for nonholonomic mechanical systems with symmetry

132 WANG SANG KOON and J. E. MARSDEN

organization of the Hamiltonian dynamics of such systems into a reconstruction equa- tion, a momentum equation and the reduced Hamilton equations. We also explore the failure of the Jacobi identity when the constraints are nonholonomic and show that the so-called Jacobiizer (the cyclic sum that vanishes when the Jacobi identity holds) is an interesting expression involving the curvature of the underlying distri- bution describing the nonholonomic constraints. Using this formula, one sees clearly that the Poisson bracket satisfies the Jacobi identity if and only if the constraints are holonomic.

Some interesting topics for future work include:

A deeper understanding of the geometry behind this gauge viewpoint of nonholonomic systems. In studying the Lagrangian reduction by stages of holonomic systems, Cendra, Marsden and Ratiu [6] has developed an intrinsic approach to the Lagrangian reduction theory of Marsden and Scheurle [18, 191. Their work is based on a deeper understanding of the geometry of the bundle (TQ)/G over the shape space Q/G as a Whitney sum T(Q/G)$Ad(Q) where Ad(Q) is the adjoint bundle (i.e., the associated bundle with fiber g where the action on g is the adjoint action). Extending this theory to the nonholonomic case will clarify further the geometry of Lagrangian reduction in [BKMM]. It will also shed light on the geometry of the gauge Hamiltonian reductions for nonholonomic dynamics.

Stability, stabilization and bifurcation theories for nonholonomic systems. Because of the momentum equation, it is natural to let the value of the momentum be a variable and for this a Poisson rather than a symplectic viewpoint is more natural. This approach may also allow one to extend to the nonholonomic systems the block diagonalization procedure in the energy-momentum method developed by Simo, Lewis and Marsden [23]. With the development of the Poisson geometry in this paper, we hope that these results will lead to further progress on the stability issues started by Zenkov, Bloch and Marsden [25]. In this light, the stability and stabilization of a simplified model of the bicycle (see [lo]) is an important problem to tackle.

Optimal control and numerical methods. Closely related to the work mentioned above is the optimal control of the bicycle, which is an underactuated balance system. Koon and Marsden [9] initiated the investigation of optimal control for nonholonomic systems like the snakeboard, using the Lagrangian framework developed in [BKMM] and coupling it with the method of Lagrange multipliers and Lagrangian reduction. Interestingly, Gregory and Lin [8] has used the same method of Lagrange multipliers to devise a general, accurate and efficient numerical method to solve the constrained optimal control problem. Ostrowski, Desai and Kumar [22] have built on these advances to study the optimal gait selection for nonholonomic locomotion systems. This kind of finite element method applied to the variational problem in integral form developed in [8] fits well with the Lagrangian framework and gives good and interesting results in the case of a relatively complicated problem, namely the optimal control of a snakeboard. We would like to use this Lagrangian approach to study the optimal control of a simplified model of the bicycles,

Page 33: Poisson reduction for nonholonomic mechanical systems with symmetry

133

Geometric phases for nonholonomic systems. The geometric effect of holonomy plays an important role in the understanding of phase drifts and is a crucial ingredient in problems of stabilization and tracking. The basic idea of holonomy is that if the system undergoes cyclic motion in the shape space (this is sometimes the control space), then it need not undergo cyclic motion in the configuration space. The difference between the beginning and the end of the motion is given by a drift in the group variables and this is the geometric phase. But the basic theory for the holonomy is not as well developed in the case of nonholonomic systems as for holonomic ones.

The geometric tools to further develop the theory for systems with nonholonomic constraints are laid in [16, 41. We aim to develop the theory by combining the ap- proaches in these two papers and also by making the calculations more concrete and accessible. In particular, in [4] the notion of the nonholonomic connection is defined and this is what replaces the mechanical connection in the case of holonomic con- straints. What makes this theory more interesting is the presence of the constraint distribution as well as the fact that the momentum need not be conserved. A start on this problem is made in [ll].

REFERENCES

PI PI 131

141

R. Abraham and J. E. Marsden: Foundations of Mechanics, 2nd Edition, Addison-Wesley, 1978. L. Bates and J. Sniatycki: Rep. Math. Phys. 32 (1993), 99-115. A. M. Bloch and P. Crouch: On the dynamics and control of nonholonomic systems on Riemannian manifolds. Proceedings of NOLCOS ‘92, Bordeaux, 1992, p. 368-372. A. M. Bloch, P. S. Krishnaprasad, J. E. Marsden and R. Murray: Arch. Rut. Mech. AnaL, 136 (1996), 21-99.

PI

161 [71

PI

191 WI VI

WI

1131 u41

A. M. Bloch, N. Leonard and J. E. Marsden: Stabilization of Mechanical Systems Using Controlled Lagrangians, Proc. CDC 36 (1998), 2356-2361. H. Cendra, J. E. Marsden and T. S. Ratiu: Lugmngian Reduction by Stages, preprint, 1997. N. H. Getz and J. E. Marsden: Control for an autonomous bicycle, International Conference on Robotics and Automation, IEEE, Nagoya, Japan, May 1995. J. Gregory and C. Lin: Constrained Optimization in the Calculus of Variations and Optimal Control Theory, Van Nostrand Reinhold, NY, 1992. W. S. Koon and J. E. Marsden: SZAM J. Control Optim. 35 (1997) 901-929. W. S. Koon and J. E. Marsden: Rep. Math. Phys. 40, 21 (1997). W. S. Koon and J. E. Marsden: The Geomehic Structure of Nonholonomic Mechanics, Proc. CDC 36 (1998), 485W862. N. E. Leonard and J. E. Marsden: Stability and Drift of Underwater Vehicle Dynamics: Mechanical Systems with Rigid Motion Symmetry, Physicu D, to appear. A. Lewis: Afine Connections and Distributions, preprint, 1996. C.-M. Made: Various Approaches to Conservative and Nonconsewative Nonholonomic Systems, preprint, 1997.

WI

WI

J. E. Marsden: Lectures on Mechanics, London Mathematical Society Lecture Note Series. 174, Cam- bridge University Press, 1992. J. E. Marsden, R. Montgomery and T. S. Ratiu: Reduction, Symmetry, and Phases in Mechanics, Memoirs AMS 436, 1990.

1171 J. E. Marsden and T. S. Ratiu: Symmetry and Mechanics, Texts in Appl. Math., 17, Springer, 1994.

WI J. E. Marsden and J. Scheurle: ZAMP 44 (1993), 1743.

1191 J. E. Marsden and J. Scheurle: Fields Institute Commun. 1 (1993), 139-164.

Page 34: Poisson reduction for nonholonomic mechanical systems with symmetry

134 WANG SANG KOON and J. E. MARSDEN

[243] J. E. Marsden and J. Ostrowskiz Symmetries in Motion: Geometric Foundations of Motion Control, to appear.

[21] J. Ostrowski: The Mechanics and Control of Undulatory Robotic Locomotion, PhD thesis, California Institute of Txhnology, 1996.

[22] J. Ostrowski, J. P. Desai and V. Kumar: Optimal gait selection for nonholonomic locomotion systems, Proc. CDC, to appear.

[23] J. C. Simo, D. R. Lewis and J. E. Marsden: Arch. Rat. Mech. Anal. 115 (1991), 15-59. [24] A. J. van der Schaft and B. M. Maschke: Rep. Math. Phys. 34 (1994), 225-233. [25] D. V. Zenkov, A. M. Bloch and J. E. Marsden: The energy momentum method for the stability of

nonholonomic systems, Dynamics and Stability of Systems (to appear).