Muller - Hamel's Equations and Geometric Mechanics of Constrained and Floating Multibody and Space Systems
Muller - Hamel's Equations and Geometric Mechanics of Constrained and Floating Multibody and Space Systems
Geometric Mechanics of
rspa.royalsocietypublishing.org
Constrained and Floating
Multibody and Space Systems
Research
Andreas Müller 1
Cite this article: Müller A. 2023
Hamel’s equations and geometric 1 Johannes Kepler University, Linz, Austria
20220732. Proc R Soc A 479, 2023. rest on a bundle structure of the configuration
http://dx.doi.org/10.1098/rspa.2022.0732 space. The connection on this bundle allows for
an intrinsic splitting of the reduced Euler-Lagrange
equations. Hamel’s equations, on the other hand,
Article submitted to journal
provide a universal approach to non-holonomic
mechanics in local coordinates. The link between
Hamel’s formulation and geometric approaches in
Subject Areas:
local coordinates has not been discussed sufficiently.
Geometric mechanics, Lie groups,
The reduced Euler-Lagrange equations as well as
computational mechanics, robotics the curvature of the connection, are derived with
Hamel’s original formalism. Intrinsic splitting into
Keywords: Euler-Lagrange and Euler-Poincaré equations, and
Geometric mechanics, Hamel inertial decoupling is achieved by means of the locked
equations, Hamel coefficients, velocity. Various aspects of this method are discussed.
Euler-Poincaré equations, Lagrange
reduction, mechanical connection,
locked velocity, kinematic 1. Introduction
reconstruction, gauge fields, space
Many dynamical systems and controlled multibody
systems systems possess symmetry invariants, and can be
modeled on a principle bundle. The bundle formulation
Author for correspondence: (for Lagrangian systems) was developed in [1,2] as key
concept in geometric mechanics, where the configuration
Andreas Müller
space is regarded as principle bundle Q = G × Q/G with
e-mail: [email protected]
symmetry group G. Central is the notion of a connection
as it allows encoding specific symmetries of the system
[3–6]. The (natural) mechanical connection, deduced from
the system momentum, was introduced in [7–9]. Given a
G-invariant Lagrangian `, in bundle coordinates (Ω α , ṙI )
of a (left) trivialization, where Ω α is the locked velocity,
the dynamics of unconstrained floating-base systems is
governed by the Lagrange-Poincaré equations [3]
d ∂` ∂` β I
α
= β
−EIα ṙ + cβ
αλ Ωλ (1.1)
dt ∂Ω ∂Ω
d ∂` ∂` ∂` β J β
I
− I = β
−BIJ ṙ + EIα Ωα (1.2)
dt ∂ ṙ ∂r ∂Ω
with connection coefficients Aα α
I and its curvature BIJ .
© The Authors. Published by the Royal Society under the terms of the
Creative Commons Attribution License http://creativecommons.org/licenses/
by/4.0/, which permits unrestricted use, provided the original author and
source are credited.
α
The coefficients EβI = cα λ
βλ AI , and the curvature coefficients 2
α ∂Aα ∂Aα β λ
BIJ = I
− J
± cα
βλ AI AJ (1.3)
uα = ξ α + Aα I I
I (r )ṙ . (1.4)
The curvature vanishes if and only if the constraints defined by momentum conservation,
and thus the Pfaffian system (1.4), is integrable. Similarly for constrained systems, a kinematic
connection was introduced, and the scleronomic constraints are expressed as uα = 0. Mechanical
systems whose spatial motion is constrained are typical examples, where the constraints restrict
the motion of a base body in G. Now the symmetry group accounts for the invariance of the
constraints, and the curvature vanishes if and only if the constraints are integrable. The dynamics
equations for constrained systems are obviously obtained from (1.2) by setting Ω α = 0. Thus,
in both cases, a kinematic relation of the form (1.4) applies, and the curvature appears in the
equations of motion, and is a central object in motion planning and control.
Geometric mechanics provides an intrinsic and coordinate-free framework for modeling,
analysis, and control of finite-dimensional (discrete) as well as infinite-dimensional (continua)
systems. Local coordinate formulations, as the one above, are used for computations, where it
may be necessary to switch between different local coordinate charts. In this context, the fact
that (1.1,1.2) is a specific form of the Hamel equations [10,11] for finite-dimensional systems in
local coordinates deserves recognition, which is the topic of this paper. In geometric mechanics,
Hamel’s equations are now introduced in an elegant modern form, e.g. [12], on the expense
that the relation to the original Hamel equations is lost, however. Moreover, how the geometric
framework and the Hamel formalism are related, and how connection and curvature of the
configuration space (bundle) are related to the Hamel coefficients, is not discussed in the
literature. Without making explicit reference to the original formulation, (1.1,1.2) are referred to
as Hamel equations [8]. Also in robotics and multibody system dynamics, the Euler-Poincaré
equations (1.1) on SE (3) are often interchangeably referred to as Hamel equations or Euler-
Poincaré equations, e.g. [13].
It is shown in this paper that the coefficients in (1.1,1.2) are naturally derived as the
Hamel coefficients. This provides a link between the original Hamel formalism and the
bundle formulation. The explicit derivation also admits consolidating the different coordinate
expressions of local curvature found in the literature, which is crucial for applying the above
equations (see Rem. 5.3). There are also various aspects that need to be taken into account when
using the geometric formalism. One is the concept of locked velocity [3], and the related concept
of average velocity [14,15] that proved to be powerful tools for control of floating-base systems.
It is discussed that there is no frame which can be associated with this velocity whose motion
is a function of g ∈ G and r ∈ Vn , which has consequences for control of floating-base systems.
Another aspect discussed in this paper is that the mechanical connection on Q induced by
the locked velocity intrinsically splits the reduced Euler-Lagrange equations in horizontal and
vertical. As an important consequence, the equations (1.1,1.2) are inertially decoupled, which
is relevant for computational multibody dynamics. Throughout the paper, all constraints are
assumed to be linear (i.e. catastatic) and scleronomic.
1
Throughout the paper, the term floating-base systems covers a large class of mechanical (control) systems characterized by
a base body free to move in space to which further bodies (links) are geometrically connected, e.g. by joints.
The paper is organized as follows. In Sec. 2, the classical Hamel equations are summarized 3
for unconstrained and constrained systems. Hamel’s formulation is related to the formulation
on a (locally) trivial bundle in Sec. 3. The obtained relations for the Hamel coefficients in local
where Aab and Bba are smooth functions of q. With vectors of generalized coordinates q ∈ Vn and
quasi-velocities u ∈ Rn , these relations are written in matrix form as
It is assumed that q a are valid local coordinates so that A is regular, and B = A−1 . The relation of
ua to the corresponding quasi-coordinates π a , a = 1, . . . , n is described by the differential forms
dπ a = Aab dq b , and the inverse relation by dq a = Bba dπ b . The quasi-coordinates are generalized
coordinates if and only if the differential forms are exact, otherwise ua are called non-holonomic
velocities [21,22], following [11, p. 473], [23, p. 218].
Denote with L (q a , ua ) a Lagrangian in terms of quasi-velocity coordinates ua , which for
mechanical systems is defined as kinetic energy T (q a , ua ) minus potential energy U (q a ). Then
the variational form of the Hamel equations [10,11,22,24] is
∂L dδπ a − δdπ a
d ∂L ∂L a
− − Q a δπ + =0 (2.3)
dt ∂ua ∂π a ∂ua dt
where Qa are generalized forces dual to ua . The explicit form of the Boltzmann-Hamel equations 4
d ∂L ∂L ∂L b c
− b Bab + γac u = Qa (2.4)
dt ∂ua ∂ub
..........................................................
is obtained after inserting the transitory relation
dδπ a − δdπ a a b
= γcb u δπ c (2.5)
dt
a
in which γcb are the Hamel coefficients defined as
∂Acr ∂Acs
c
γab := − Bar Bbs . (2.6)
∂q s ∂q r
The Hamel coefficients vanish identically if and only if (2.1) are integrable, i.e. if ua are holonomic
velocities. The equations (2.4) along with the kinematic equations (2.1) govern the dynamics in the
non-holonomic tangent bundle (i.e. tangent space defined by non-holonomic constraints). They
are referred to as the Boltzmann-Hamel equations (e.g. in [25,26]) as they where (in a very similar
form) presented by Boltzmann in [27,28] and by Hamel in [10,11,29]. It was Hamel, however, who
generalized them to systems an a Lie group [24,29].
The motion of many mechanical systems can be described on a n-dimensional Lie-group G, so
that quasi-velocities belong the corresponding Lie algebra g. If g (t) ∈ G, such quasi-velocities can
be introduced as left- or right-invariant vector fields, û = g −1 ġ or û = ġg −1 , respectively. Let q a
be canonical coordinates on G, and let u ∈ Rn be the vector representation of û ∈ g, then (2.2)
is a map from Rn to Rn ∼ = g. When using the left-invariant definition of quasi-velocities, the
Hamel coefficients (2.6) are identical to the structure constants ccab of the Lie group. This was
already shown by Hamel [24, p. 428] using the transitory relations dδπ a − δdπ a = γcb a
dπ c δπ c .
c c
Using right-invariant quasi-velocities leads to a change of sign: γab = −cab . Typical example for
such quasi-velocities are the angular velocity (where G = SO (3)) or rigid body twists (where
G = SE (3)). Then left-invariance implies body-fixed representation of angular velocity or twists,
and right-invariance implies spatial representation [30,31]. The explicit derivation of the Hamel-
coefficients for SO (3) and SE (3) using the definition of Hamel coefficients can be found in
[32]. Clearly, π a are holonomic coordinates if and only if G is Abelian. It must be emphasized
that canonical coordinates on G are only locally valid in general. This applies in particular to
SO (3), and thus to SE (3), since it is not simply connected, which leads to the well-known
parameterization singularity of rotations.
In summary, the Hamel equations on a Lie group are the forced Euler-Poincaré equations for
Lagrangian L (q a , ua )
d ∂L ∂L ∂L
± cbac b uc = b Bab + Qa (2.7)
dt ∂ua ∂u ∂q
where the positive sign applies to left-invariant, and the negative sign to right-invariant quasi-
velocities. These are the reduced Euler-Lagrange equations for systems whose Lagrangian is (left
or right) invariant under action of a symmetry group G [9,33].
uα := Aαa q
a a
q̇ . (2.8)
It is assumed that the system of Pfaffian constraints is regular, i.e. the m̄ constraints are
independent. Then, δ̄ := n − m̄ independent quasi-velocity coordinates are introduced as
uI = AIa q a q̇ a , I = m̄ + 1, . . . , n
(2.9)
where δ̄ is the differential DOF of the system (also called instantaneous DOF) [34,35]. The overbar
of δ̄ and m̄ indicate that the constraints are generally non-holonomic. If they are integrable, there
are m = m̄ geometric constraints, and δ = n − m = δ̄ is the finite DOF. The n − m̄ independent 5
coordinates are indexed with capital lattin letters I, J, K. The Pfaffian system (2.8) and the
solution (2.9) are summarized as ua = Aab (q a ) q̇ b , as in (2.1) with index set {a} = {α, I}. The
where uα , α = 1, . . . m̄ are set to zero after taking the derivatives. The generalized forces are QI =
BIa Qa . The Hamel coefficients in (2.10) are obtained by restricting (2.6) to indices I, J as
∂Aab ∂Aac
a
γIJ = − BIb BJc . (2.11)
∂q c ∂q b
q̇ a = BIa q a uI , I = m̄ + 1, . . . , n
(2.12)
govern the system dynamics in terms of state variables (q a , uI ). Not all of the q a may be
independent if the constraints are not completely non-holonomic. Relation (2.12) is obtained from
(2.2) assuming uI are locally valid coordinates on the tangent space and constraints are regular.
Quasi-velocities uI are integrable if and only if γIJ
K
≡ 0. The constraints (2.8), and thus the
∗ ∗ ∗
co-distribution D ⊂ Tq Q with Dq := span (A (q)) defined by the constraints, are integrable
α
(in Pfaffian sense) if and only if γIJ ≡ 0. The constraint distribution D on Q, defined as Dq :=
ker A (q) ⊂ Tq Q, is thus integrable (in Cartan sense) if and only if the Hamel coefficients vanish
and the constraints are regular (rank A is constant). This may not apply to non-regular constraints.
uα := Aα β q
a β
ṡ + AαI q
a I
ṙ , α = 1, . . . , m̄ (3.1)
! !
A1 A2 ṡ
, with A1 = Aα α
u = A (q) q̇ = β , A2 = AI . (3.2)
0 I ṙ
where (3.2) resembles the matrix form (2.2). The inverse relation of (3.1) and (3.2) are, respectively, 6
α
ṡ = Bβα q a β
u + BIα q a I
ṙ (3.3)
With A in (3.2) and B in (3.4), the expression (2.11) gives rise to the Hamel coefficients
∂Aα ∂Aα
µ
α
γβλ = − ν
Bβµ Bλν
∂sν ∂sµ
∂Aα ∂Aα
α
γIJ = r
− s
BIr BJs (3.5)
∂q s ∂q r
∂Aα ∂Aα ∂Aα
! ! !
∂Aα ∂Aα β ∂Aαµ β µ ∂Aα β β β ∂Aα
= I
− J
+ − BI BJ + I
− BJ + − J
BIβ
∂rJ ∂rI ∂sµ ∂sβ ∂sβ ∂rI ∂rJ ∂sβ
of the constraints is imposed, after taking the derivatives. The dynamic equations (3.6) along with
the kinematic equations (3.7) govern the dynamics of the non-holonomically constrained system
in terms of the state (q a , q̇ a ) evolving on the non-holonomic tangent bundle.
uα := ṡα + Aα
I q
a I
ṙ , α = 1, . . . , m̄ (3.8)
−1
where (Aα α I I
I ) := A1 A2 = −B2 , with B2 = (BI ) in (3.3). Combined with u := ṙ , this is written
in matrix form as
! !
I −B2 ṡ
u = Ā (q) q̇ = . (3.9)
0 I ṙ
..........................................................
J β
α
γIJ = I
− J
+ A − I
Aβ , α = 1, . . . , m̄; I, J = m̄ + 1, . . . , n. (3.11)
∂rJ ∂rI ∂sβ I ∂sβ J
The configuration space Q = Vm̄ × Vδ̄ is regarded as a trivial bundle2 over the base manifold
V with fiber Vm̄ , and bundle coordinates (sα , rI ) ∈ Vm̄ × Vδ̄ . The horizontal space of this trivial
δ̄
bundle is the constraint distribution, i.e. the vector space of velocities satisfying the constraints.
The homogenous kinematic constraints (3.8) define a connection on this bundle. Writing the
constraints in terms of the Pfaffian forms ω α := uα dt = dsα + Aα I
I dr , a connection is introduced
α ∂
as A = ω ∂sα . This is referred to as an Ehresmann connection [3,18,33] with reference to the
original publication [36], and Aα I are the local coordinates of the connection. Since it arises
from the kinematic constraints, it is called the kinematic connection [18]. The connection relates
(independent) motions in the base manifold Vδ̄ to motions in the fiber. Whether this connection
α
(i.e. the constraints) is holonomic is revealed by its curvature, denoted BIJ . Moreover, the
curvature of the kinematic connection plays a key role in the control of constrained mechanical
systems [3,18,37] as well as in locomotion planning [38] as it encodes how motions in the base
manifold generate motions in the fiber. On the trivial vector bundle, the Lie bracket in the
∂Aα β ∂Aα β
curvature (1.3) is the Lie bracket [AJ , AI ]α = ∂s β AI − ∂sβ AJ of
J I
vector fields AI , AJ on Q,
α ∂Aα ∂AJα
so that the local curvature is BIJ
= − I
∂r J ∂r I
+ [AJ , AI ]α [18, p.
32], [3, p. 108]. The Hamel
coefficients (3.11) are thus clearly related to the coordinate form of the curvature as follows.
Proposition 3.1. The Hamel coefficients (3.11) are identical to the curvature components of the kinematic
connection A, in bundle coordinates (sI , rα ), induced by the constraints with (3.8), i.e. BIJ
α α
= γIJ .
Although local coordinates are used in this paper, it should be mentioned that the curvature
of a connection A is its covariant derivative, written coordinate-free as B (X, Y ) = dA (X, Y ) −
[A (X) , A (Y )], with horizontal vector fields X, Y , i.e. Bα (X, Y ) = BIJ
α
X I Y J = γIJ
α
XI Y J .
Remark 3.1. The constraints are holonomic if and only if the kinematic connection is flat, i.e. the
α
components BIJ of the curvature 2-form vanish identically. In this case, Vδ̄ serves as configuration space.
Constraints are said to be in Chaplygin form if A = A (r), referring to Chaplygin’s publications [39,40]. In
α ∂Aα ∂Aα
this case, the Hamel-coefficients (2.10) reduce to γIJ = ∂rJI − ∂rIJ , which implies the obvious condition
∂Aα ∂Aα
I
∂r J
≡ J
∂r I
for integrability of dsα + Aα J I
I (r )dr = 0.
ṡα = −AαI q
a I
ṙ . (3.13)
The constrained dynamics is governed by the Hamel equations (3.12) along with the kinematic
equations (3.13). The equations (3.12) are obtained as constrained Lagrange–d’Alembert
2
Q = M × F is a trivial bundle if it can be written as Cartesian product of a manifold M and F , and if there is a projection
π : Q → M [19]. F is called the standard fiber. For the considered systems, the base manifold is the coordinate subspace Vδ̄
corresponding to the independent velocities, and the fiber is the subspace Vm̄ corresponding to dependent velocities.
equations with variations satisfying the constraints 0 = δsα + Aα I I
I (r )δr [3,8,18], which is a 8
particular form of Hamel’s equations when using a connection to introduce constraints.
uα := ξ α + Aα I I
I (r )ṙ (4.1)
..........................................................
(4.2)
so that the horizontal subspace of the connection is the space of velocities satisfying the
constraints, and Akin is a g-valued one-form (which may be considered as a special type of
Ehresmann connection) called the kinematic connection as it arises from (4.1) by requiring it to be
G-equivariant [3]. The name stems from the fact that it relates base and fiber motions according to
the kinematic constraints. It is sufficient to use the local connection form Aα
I in (4.1) as it encodes
all relevant information. The symbol Akin is used to distinguish it from the coefficients of the
local form Aα I . Next, the Hamel coefficients are derived and are identified as the coefficients of
the curvature of the kinematic connection.
Lemma 4.1. The non-vanishing Hamel coefficients in (2.10) for the system subjected to left G-invariant
constraints are
α ∂Aα ∂Aα λ µ
γIJ = I
J
− J
+ cα
λµ AI AJ , α = 1, . . . , m̄. (4.3)
∂r ∂rI
Proof. In order to apply the original definition (2.6) of the Hamel coefficients, local canonical
coordinates sα , α = 1, . . . , m̄ are introduced on G. The fiber coordinates are then expressed as
ξ α = Aα α β α α α β
β (s ) ṡ , with inverse relation ṡ = Bβ (s ) ξ , and (4.1) is written as
uα = Aα
β s
α β
ṡ + Aα J I
I (r )ṙ . (4.4)
with Bβα defined in (3.4). Noting that AIa = const, the only non-zero Hamel coefficients for the
α
constrained system are γIJ . They are immediately found from (3.5), by replacing Aα J
I (r ) with
α J α α a β J
AI (r ), and BI with −Bβ (s ) AI (r ), as
The expressions (4.3) are the coefficients of the local curvature of the connection Akin on the
principal bundle possibly up to a change of sign [19,47] (for the sign convention see Rem. 5.3). In
context of geometric mechanics, this can be stated as follows [18, notice the correction on p. 44].
Proposition 4.1. The Hamel coefficients (4.3) for left-invariant kinematic constraints (4.1) are the
components of the curvature (1.3) of the kinematic connection defined in bundle coordinates (ξ α , ṙI ) on
α α
the corresponding left-trivialized principal bundle: BIJ = γIJ . Noting that the curvature coefficients are
identical to the Hamel coefficients, the connection is flat, if and only if the constraints are holonomic.
Also the curvature on the principle bundle can be defined coordinate-free as covariant 10
derivative B (X, Y ) = dA (X, Y ) − [A (X) , A (Y )], with horizontal vector fields X, Y , now with
Lie bracket on g. In local coordinates, it is Bα (X, Y ) = γIJ
α
XI Y J .
α ∂Aα ∂Aα λ µ
γIJ = I
J
− J
− cα
λµ AI AJ , α = 1, . . . , m̄. (4.7)
∂r ∂rI
Remark 4.2. The constraints (4.1) give rise to a kinematic control system on G that can be written as
ξ α = −Aα I I
I (r )ṙ . Such control systems with symmetry were addressed in various publications, e.g. [38,
44,48,49]. Controllability of this driftless control problem is encoded in the control vector fields Aα I and the
distribution on the fiber defined by them. Written as ġ = −gAdṙ shows that the connection describes how a
path in shape space is lifted to a path in the group (called the horizontal lift of the curve), which is the basis
for kinematic control. The net change in the group variable as result of the horizontal lift of a closed curve
in shape space is the holonomy (in this context called the geometric phase). The latter can be related to the
curvature of the kinematic connection. Let φ be a closed path in Vδ̄ . The geometric phase is then found as
I Z
g (φ) = − gAdr = − B + hot. (4.8)
φ Φ
using ġ = g ξ̂ = −gAṙ, where Φ is the area enclosed by the path in Vδ̄ , and B = (BIJ α
drI drJ ). If the
constraints are holonomic, i.e. the curvature vanishes, the geometric phase is zero. This is an important
relation for locomotion planning, where the closed path φ represents a gait, and the aim is to maximize the
net motion in G generated by a gait [38]. The geometric phase further discussed in Sec. 6.
where ξ = −Aṙ. The Hamel equations (4.9) along with the reconstruction equations (4.10) govern
the dynamics of the constrained system on the principle bundle Q. Chaplygin systems (Rem. 3.1)
∂Aα ∂Aα
are special cases with Abelian symmetry group G = Rm̄ , where BIJ
α
= J
∂r I
− I
∂r J
.
Remark 4.3. Equations (4.10) are the Poisson equations on G. They are known in context of rigid body 11
kinematics as the left- and right Poisson-Darboux equations, referring to [50], or as the generalized Poisson-
Darboux equations [51]. In order to solve these differential equations on G, g is expressed as the exponential
η̂˙ = dexp−1 −1
−η̂ (ξ̂) = −dexp−η̂ (Aṙ), with g = g0 exp(η̂) (left trivialization) (4.11)
η̂˙ = dexp−1 −1
η̂ (ξ̂) = −dexpη̂ (Aṙ), with g = exp(η̂)g0 (right trivialization)
with initial value g0 ∈ G, where η̂ (t) ∈ g represents a local canonical parameterization of G, and dexp is
the right-trivialized differential of the exp map on G. The latter is defined by ġg −1 = dexpη̂ (η̂) assuming
g = exp(η̂)g0 . This replacement is a key step in Lie group integration schemes [52–54]. For many Lie
groups relevant to solid mechanics, there are closed from expressions for the dexp map, in particular for
SO (3) and SE (3) [55]. The maps dexp and dexp−1 can be evaluated using truncated series expansions.
3 2
b
1
3
2 p
..........................................................
α α α I
u := ω + AI ṙ , AI := α R εα3I¯, α = 1, 2 (4.12)
0, α=3 .
α β µ
The Hamel coefficients are γIJ = −cα
βµ AI AI , I, J = 4, 5 (negative sign is due to right-
β µ
trivialization), which is only non-zero for α = 3 as β, µ = 1, 2. Evaluation yields cα
βµ AI AJ =
1 α β I ¯
1
ε ε ε
R4 αβµ βr I¯ µsJ¯
ρr ρs = 1
(δ δ − δµI¯δαr )εµsJ¯ρr ρs
R4 µr αI¯
= − R14 ρα εIs s
¯ J¯ρ = − R2 ρ AJ δβ , thus
(
α 0, α = 1, 2
γIJ = 1 I¯ (4.13)
R AJ , α=3 .
The force is obtained as cross product of this torque with ρ, noting that π1 = π3 = 0 and ω3 = 0.
Remark 4.4. The principal kinematic case can be extended by allowing the Lagrangian to be invariant
under actions of a subgroup of the symmetry group G of the constraints, i.e. infinitesimal generators of
this subgroup lie in the constraint distribution. Then the system is said to possess horizontal symmetries
(relative to the constraints) [3,4,18]. In this case, the kinematic connection along with a connection
accounting for the horizontal symmetry can be introduced. As an example for mechanical systems with
horizontal symmetries, a ball moving on a plane, i.e. a rolling ball that is free to spin about the plane
normal, was used in [4,18]. The horizontal symmetry is then due to the momentum being invariant under
the subgroup SO (2) ⊂ SO (3) of rotations about the plane normal. The m̄ rolling constraints are the 1-
and 2-component of the condition 0 = ṙ + ρωe above. The invariant momentum is the 3-component of Θω.
d ∂l ∂l
± β cβ ξλ = 0 (5.1)
dt ∂ξ α ∂ξ αλ
d ∂l ∂l
− I =0 (5.2)
dt ∂ ṙI ∂r
where in (5.1) the positive sign holds for a left-trivialization (ξˆ = g −1 ġ body velocities), and the
negative sign when right-trivialization is used (ξˆ = ġg −1 spatial velocities). The base motion is
reconstructed with the respective kinematic equation in (4.10).
The above equations obviously split into the Euler-Poincaré equations (5.1) and Euler-
Lagrange equations (5.2), where the first equation (5.1) can also be written as Lie-Poisson
∂l
equations on G when expressed with momentum Πα = ∂ξ α . While here they have been solely
derived with the Hamel formalism, they were derived as reduced Euler-Lagrange equations [9,33]
from a variational principle on Q. The involved variations are not intrinsic in the sense that they
are not split into variations in the fiber and the base manifold, respectively. Such a splitting leads
to a decoupling of the equations and to a block-diagonalization of the mass matrix defining the
kinetic energy. This is achieved by variations of ξ α with zero variations of ṙI , i.e. setting the shape
velocity to zero, which in geometric terms is equivalent to variations in the vertical space of the
principle bundle. This is formalized using a connection, as described next.
The so-defined connection, is called the mechanical connection [8,9] building upon a concept
discussed in [62]. In contrast to the kinematic connection, it is defined via the momentum. The
connection may be considered to be in Chaplygin form since it is independent of group variables.
Notice that this strictly relies on a local bundle trivialization since Q may not be a trivial principal
bundle. The locked velocity Ω is the vertical part relative to the mechanical connection.
Remark 5.1. An important aspect of the locked velocity is that it cannot be associated to a frame whose
motion is described by h (g, r) ∈ G depending on some g and r, so that Ω̂ = h−1 ḣ. This is an immediate
consequence of the fact that Ω is defined by the non-holonomic momentum Π (r, ξ, ṙ) (see Sec. 5(e)).
Remark 5.2. The principal bundle view on floating-base systems has an interesting connection to gauge
theory. In gauge theory, the symmetry is related to gauge invariance, G is called the ’gauge group’, and
the connection one-form A in (5.5) is the ’gauge potential’ [47]. This was discussed for the falling cat and
similar non-holonomic control systems in [60], and more generally, for ’deformable bodies’ (mechanical
structures that can change their shape) in [63,64], and for Maxwell or Yang-Mills fields in [1]. In gauge
theory, the equivariance condition on the connection describes a gauge transformation from a local gauge
A to a new A0 = Adg (g −1 dg + Adr) [19,65]. In case of mechanical systems, it describes a transformation
from one body-fixed frame to another in which velocities are measured. In [63,64], Aβ
I was called the master
gauge, while its curvature is considered as field strength.
Lemma 5.1. The Hamel coefficients in bundle coordinates (Ω α , ṙI ) are γab
I
≡ 0, and
α
γβδ = ±cα
βδ (5.6)
α α
γJβ = −γβJ = ±cα δ
βδ AJ (5.7)
∂Aα ∂Aα ∂Aα ∂Aα
α
γIJ = I
− J α
+ γJβ Aβ
I =
I
− J
± cα β λ
βλ AI AJ (5.8)
∂rJ ∂rI ∂rJ ∂rI
where the positive sign of ± applies to left-, and the negative sign to a right-trivialization of the G-bundle.
Proof. Canonical coordinates sα are introduced on G, which are related to the fiber coordinates 15
via ξ α = Aα α β α α α β
β (s ) ṡ and ṡ = Bβ (s ) ξ , respectively (see Sec. 4(b)). Relation (5.5) and its inverse
are then written as
With the locked velocity, the quasi-velocities are (ua ) = (uα , uI ) = (Ω α , ṙI ), thus expressed in the
form (2.1), with BJI = δJI and BβI ≡ 0, or in matrix form
! ! ! !
Aα β (AαI) ṡ ṡ Bβα − Bβα Aβ I
u= , = u. (5.11)
0 I ṙ ṙ 0 I
The relation (2.6) is separated for the coordinates Ω α and ṙI . The coefficients γβδ α
=
∂Aα α
µ ∂Aλ µ λ α
∂sλ
− ∂sµ Bβ Bδ = ±cβδ are again determined by the structure constants of G, and thus
∂Aα ∂Aα
α
γβJ = δ
− λ
Bβλ Bγδ AγJ = ±cα δ
βδ AJ .
∂sλ ∂sδ
α
The remaining coefficients γIJ are given in (4.3).
α α
The Hamel coefficients (5.8) are identical to the components of the curvature (1.3), BIJ = γIJ ,
of the mechanical connection in bundle coordinates. They are indeed formally identical to the
curvature coefficients (4.3) of the kinematic connection.
and Sec. 5(e)). Equations (5.12,5.13), admit the following geometric interpretation.
Proposition 5.1. The Hamel equations (5.12,5.13) for a left G-invariant Lagrangian `(rI , Ω α , ṙI ) are the
reduced Euler-Lagrange equations (1.1,1.2) in terms of bundle coordinates (Ω α , ṙI ), with the coefficients
α α
EβI = γIβ = cα λ α α α α
βλ AI , cβλ = γβλ , and curvature BIJ = γIJ determined by the Hamel-coefficients.
The equations (1.1,1.2) have been derived in [9], and presented in [33, p. 397], by introducing
the locked velocity (5.5) into the Lagrangian `(rI , Ω α , ṙI ) before taking the Euler-Lagrange
derivative. Their derivation as Hamel equation in terms of the locked velocity has not been
reported in the literature. The first equation (1.1), respectively (5.12), is indeed the Euler-Poincaré
equation (5.1) with ξ α replaced by the velocity Ω α of the locked system, which is why (1.1,1.2) are
β
also called Lagrange-Poincaré equations [45, p. 3395], [3, p. 146]. The terms with γαI in (5.12)
and (5.13) can be regarded as interaction (or coupling) terms. Clearly, as remarked in [4, pp.
912, 913], the equations (5.12,5.13) reduce to the Hamel equations (5.1,5.2) if the coefficients of
connection and curvature vanish, i.e. when expressed in local coordinates (ξ α , ṙI ). However, since
the Hamel formalism applies to any choice of local coordinates, as shown in this paper, it should
not be said that (5.12,5.13) reduce to the Hamel equations when using local coordinates (ξ α , ṙI ), 16
as occasionally stated, e.g. [66, p. 226].
It follows from the definition of the Hamel coefficients that the mechanical connection is flat,
α
Remark 5.3. A note on the sign convention for the curvature coefficients BIJ is in order. Given a
α α α
connection, the local curvature is usually defined as BIJ = −γIJ with γIJ in (4.3) (see [47] for right
bundles), which agrees with the definition of curvature used in gauge theory [19, p. 247]. This convention
α
is used in [9, p. 157] and [45, p. 3395], and thus BIJ appears with a positive sign in the reduced Euler-
Lagrange equations (1.2). In [4, p. 910] and [18, notice the correction on p. 44], the local curvature is
α α
introduced as BIJ = γIJ , as in this paper, along with the reduced Euler-Lagrange equations (1.1,1.2).
α α
In [3, pp. 117,146], the curvature is derived as BIJ = −γIJ , but is then used with a negative sign in the
α α
reduced Euler-Lagrange equations. Similarly in [66], the curvature is introduced as BIJ = γIJ but used
with positive sign in (1.2). These inconsistencies deserve particular attention when applying equations
(4.9) and (1.1,1.2).
and (5.13) are inertially decoupled (not coupled on accelerations level). Coupling of the equations
β
is via the velocity terms involving γαI . Inertial decoupling using the locked velocity has been
addressed for modeling of floating base robots [13,67] and space robots in [68,69].
A closely related concept for decoupling the equations is that of the centroidal momentum as
introduced in [70], which is widely used for whole-body control of humanoid robots [15,71,72]
for instance. In this context g = se (3), and V ∈ se (3) is the velocity (twist) of the base body
(using symbol V instead of ξ), and ṙ are the joint velocities. A frame FG is introduced that is
located at the total COM of the system and aligned with the inertial frame F0 . The configuration
of FG relative to the frame Fb attached at the base body is described by gbG ∈ SE (3) = G.
∗
The centroidal momentum is defined as ΠG = AdT gbG Π, with momentum co-screw Π ∈ se (3)
defined in Sec. 5(b). This is also expressed as ΠG = MG VG , where VG is referred to as the
average velocity, and MG = AdT gbG Mbb AdgbG is called the centroidal composite inertia matrix,
and AdT gbG (L, K) the centroidal momentum matrix [14,15]. Comparing this with the definition of
the locked velocity L (r) Vloc = Π shows that VG = Ad−1 gbG Vloc . Moreover, VG = (ωave , ṗcom ),
where ωave = ωloc is called the average angular velocity [70], and ṗcom is the velocity of the total
!
ΘG 0
COM of the system, both expressed in F0 . The important point is that MG = is
0 m̄I
a block diagonal matrix, where m̄ is the total mass, and ΘG is the total inertia tensor w.r.t. to the
total COM expressed in F0 . This would replace the locked inertia L in (5.14) when the EOM are
expressed with VG .
The motivation for using the centroidal momentum ΠG = (ΘG ωave , m̄ṗcom ) is that the linear
and angular momentum are decoupled (in addition to the inertial decoupling of (5.12) and
4
The vertical space is the tangent space ker Tq π to the group orbits, i.e. possible velocities of the base body for locked shape
coordinates. The horizontal space is the space of velocities not producing a locked velocity.
(5.13)), and can be controlled independently. However, there is generally no frame associated 17
to VG (Rem. 5.1) that could serve to represent the system orientation. This would imply that the
−1
motion of this frame is represented by a gbG (g, r) such that V̂G = ġbG gbG . It is clear from the
3
b
1 2
3 2
Figure 2. Schematic drawing of a satellite model. The figure shows a self-stabilizing cube reported in [79].
(i) Mixed Representation of Rigid Body Velocity –Symmetry Group G = SO (3) × R3 18
3
The configuration (pose) of the main body is represented as (R, p) ∈ SO (3) × R , where R ∈
..........................................................
Fb relative to an inertial frame (IFR) F0 . The group multiplication on the direct product group
is (R1 , p1 ) · (R2 , p2 ) = (R1 R2 , p1 + p2 ). Since rotations and translations are decoupled, this
is clearly not a frame transformation (i.e. a rigid body motion). The corresponding velocity
defined via left-trivialization is V̂b = g −1 ġ = (R−1 Ṙ, ṗ) = (ω̂, ṗ) ∈ g = so (3) × R3 , and in vector
representation Vb = (ω, ṗ) ∈ R6 , where ω ∈ R3 is the angular velocity of the main body relative
to F0 resolved in Fb . This is referred to as the mixed representation of rigid body velocities since
ω is resolved in the body frame, and ṗ in the inertia frame [30,31]. Regarding the dynamics,
decoupling of rotation and translation is valid only if the body-fixed RFR is located at the COM,
which is the main premise when using the direct product group G, since then the angular and
linear momenta are decoupled. Therefore, the velocities of main body and rotors will be measured
at the total COM of the satellite (main body including the rotors), thus p is the position vector of
the total COM resolved in the IFR, and Fb is located at the total COM. Denote with Vi = (ωi , ṗ)
the hybrid velocity of rotor i = 1, 2, 3. W.l.o.g. the RFR is aligned with the rotor axes. Then
ωi = ω + ei ϕ̇i , where ei ∈ R3 is the i-th unit vector (e.g. e1 = (1, 0, 0)), and thus Vi = Vb + Vi ,
with Vi = (ei ϕ̇i , 0).
The momentum of the main body in mixed representation is Π b = Mb Vb ∈ R6 ∼ = g∗ =
∗ 3 i i
so (3) × R , and of the i-th rotor Π = M Vi , with the inertia matrix of the main body and
the i-th rotor, respectively,
! !
b Θb 0 i Θi 0
M = , M = (5.15)
0 mb I 0 mi I
where Θ b and Θ i is the inertia tensor of the main body and the i-th rotor w.r.t. the total COM,
and mb and mi is the mass of the main body and i-th rotor.
The velocity coordinates are (ξ α , ṙI ) = (ω 1 , ω 2 , ω 3 , ṗ1 , ṗ2 , ṗ3 , ϕ̇1 , ϕ̇2 , ϕ̇3 ) = (Vbα , ϕ̇i ) = (Vb , ϕ̇),
with fiber coordinates (ξ α ) = (Vbα ) = (ω 1 , ω 2 , ω 3 , ṗ1 , ṗ2 , ṗ3 ), α = 1, . . . , 6 and (ṙI ) = (ϕ̇1 , ϕ̇2 , ϕ̇3 ),
I = 7, 8, 9. In the following, indexes i, j, k, l = 1, 2, 3 and the notation I¯ = I − 6 are used. The
kinetic energy of the satellite is
3
1 T b 1X T i
T (Vb , ϕ̇) = Vb Π + Vi Π
2 2
i=1
3
1 b α β 1 Xh i i
= Θαβ ω ω + Θjk (ω j + δij ϕ̇i )(ω k + δik ϕ̇i ) + m̄ṗi ṗi (5.16)
2 2
i=1
where m̄ := mb + 3i=1 mi is the total mass of the satellite. The kinetic energy is invariant under
P
Hamel Equations The structure coefficients on the direct product group SO (3) × R3 are cβ
αλ =
εαλβ , for α, β, λ = 1, 2, 3, and cβ
αλ = 0 otherwise. The Euler-Poincaré equations (5.1) are found as
3
d ∂T ∂T j k X j j
+ c ω = Θ̄ij ω̇ j + εikj ω k Θ̄jl ω l + (Θij ϕ̈ + εikl ω k Θlj ϕ̇j )
dt ∂ω i ∂ω j ik
j=1
d ∂T
= m̄p̈j (5.17)
dt ∂ ṗj
where Θ̄ := Θ b + 3i=1 Θ i is the composite inertia tensor of the satellite including main body
P
19
and rotors. They can also be written in vector form, with matrix adω = ω,
e
..........................................................
d ∂T ∂T X i i
− adT
ω = Θ̄ ω̇ + ω
e Θ̄ω + e i ϕ̇i
θ ϕ̈ + ωθ (5.18)
dt ∂ω ∂ω
i=1
d ∂T
= m̄p̈ (5.19)
dt ∂ ṗ
where θ i := Θ i ei is the i-th column of Θ i .
¯
The Euler-Lagrange equations (5.2) are, noting that T does not depend on ϕi (= rI ),
d ∂T i j i i
= Θij ω̇ + Θii ϕ̈ (5.20)
dt ∂ ϕ̇i
i
where the diagonal element Θii is the moment of inertia about the axis of rotator i.
The above equations are summarized to the set of EOM for the satellite
θ1 θ2 θ3
Θ̄ 0 e Θ̄ω + 3i=1 ωθe i ϕ̇i
P
ω̇ ω
0 m̄I 0 0 0
1T p̈
0
θ 1 1
0 Θ 11 0 0
ϕ̈
+
0 = 0.
(5.21)
2T
2 ϕ̈2
0
θ 0 0 Θ22 0
3T 3
θ 0 0 0 Θ 3 ϕ̈ 0
33
The mass matrix has the form (5.4) with non-zero submatrix K. It is constant due to the
assumption of symmetric rotors and axes aligned with the RFR axes.
The pose of the satellite is obtained by solving
the kinematic
reconstruction equations (4.10). To
this end, the equations Ẋ = dexp−1 −1
−X Vb = dexp−x ωb , ṗ are solved for the coordinate vector
∼ so (3) × R3 . Here dexp is the matrix form of the right-trivialized differential
X = (x, p) ∈ R6 = x
of the exp map on SO (3) [55].
Euler-Lagrange Equations on a Trivial Principle Bundle With the partitioning (5.4) of the mass
matrix, the connection coefficients Aα
I are defined by
!
−1 θ̄ 1 θ̄ 2 θ̄ 3
A=L K= . (5.22)
0 0 0
¯
The locked velocity (5.5) is Vloc = Vb + L−1 Kϕ̇, and thus with (5.22), ωloc α
= ω α − Aα I
I ϕ̇ and
α α
ṗloc = ṗ . In terms of the locked velocity, the kinetic energy is
1 b α I¯
β β J¯
T (Vloc , ϕ̇) = Θαβ ωloc − Aα¯
I ϕ̇ ω loc − A J¯
ϕ̇ (5.23)
2
3
1X i α α I¯
β J¯
m̄
+ Θαβ ωloc + δIα ¯ − AI¯ ϕ̇ ωloc + (δJβ¯ − Aβ
¯)ϕ̇
J
+ ṗα ṗα .
2 2
i=1
with T T T
1
− θ 1 θ̄ 1 −θ 1 θ̄ 2 −θ 1 θ̄ 3
Θ11
S − KT L−1 K =
T T T
−θ 2 θ̄ 1 2
− θ 2 θ̄ 2 −θ 2 θ̄ 3 (5.25)
Θ22
T T T
3 1 3 2 3 3 3
−θ θ̄ −θ θ̄ Θ33 −θ θ̄
Thus the mass matrix in the EOM in terms of the locked velocity has the block-diagonal form
(5.14). The Hamel coefficients (5.6)-(5.8) are determined by the non-zero structure coefficients
cβ
αλ = εαλβ , for α, β, λ = 1, 2, 3. Since the connection coefficients are constant, the curvature
α
coefficients (5.8) are BIJ = [AI , AJ ]α . In vector representation, AI is the I-th
¯ column in (5.22), and
20
I¯ J¯
[AI , AJ ] = (θ̄ × θ̄ , 0). It is non-zero due the non-commutativity of vector fields AI w.r.t. the
Lie bracket on g = so (3) × R3 (non-parallel rotor axes). The non-zero BIJ α
(non-flat connection)
where Θb and Θi are the inertia tensors of the main body and the i-th rotor w.r.t. the RFR, and db ,
di are the position vectors to the COM w.r.t. the RFR. The total kinetic energy of the satellite is
3 3
1 1X T i 1 1X T
T (Vb , ϕ̇) = VbT Mb Vb + Vi M Vi = VbT Mb Vb + Vb + Vi (ϕ̇) Mi Vb + Vi (ϕ̇) .
2 2 2 2
i=1 i=1
(5.27)
Hamel Equations In the following, the matrix form of the equations will be presented, for
simplicity. The structure coefficients on the semi-direct product group SE (3) give rise to the
matrix form of the adjoint operator [55,80]
!
ωeb 0
adVb = (5.28)
v
eb ω eb
so that the Lie bracket is adX Y = [X, Y]. The Euler-Poincaré equations (5.1) are
3 3
d ∂T ∂T ˙ − adT LV + X Mi V
− adT Mi V
X
Vb = LV̇b + i Vb b i (5.29)
dt ∂Vb ∂Vb
i=1 i=1
P3
with locked mass matrix L = Mb + i
i=1 M . Written explicitly yields the instructive form
3
e˙ + ωe θ i ϕ̈i + ωθ
e i − mi v
e ai ϕ̇i
X
e Θ̄ω − m̄(v
Θ̄ ω̇ + ω e v)d+ = 0
i=1
X3
m̄ v̇ + ωv e˙ + ω
e + (ω e ω)d
e − mi ai ϕ̈i + ωa
e i ϕ̇i = 0 (5.30)
i=1
θ1 θ2 θ3
Θ̄ m̄d
ω̇ ∗
e
−m̄d m̄I −m1 a1 −m2 a2 −m3 a3
∗∗
v̇
e
1T T 1
1
θ −m a Θ 0 0 ϕ̈ + 0 = 0 (5.32)
1 1 11
2 T ϕ̈2 0
−m2 aT 2
θ 2 0 Θ22 0
θ3
T
−m aT 0 0 Θ3 ϕ̈3 0
3 3 33
P3 i i P3 i
with ∗ := ω
e Θ̄ω − m̄ωe
e vd + i=1 (ωθ
e − mi v
e ai )ϕ̇ and ∗∗ := m̄ (ωv
e +ω e −
e ωd) i=1 mi ωa
e i ϕ̇ .
Eule-Lagrange Equations on a Trivial Principle Bundle The mass matrix in (5.32) is block-
partitioned, according to (5.4), with
1
!
1 2 3
! Θ11 0 0
Θ̄ m̄d
e θ θ θ 2
L= , K= , S= 0 Θ22 0 .
−m̄de m̄I −m1 a1 −m2 a2 −m3 a3 3
0 0 Θ33
Therewith, the local connection, defining the locked velocity Vloc = Vb + Aϕ̇ in (5.5), is
A = L−1 K = m̄1 m̄2 m̄3 (5.33)
!
ei
with column vectors m̄i := L−1 Mi . Explicit expressions for L, K, A are given in the
0
supplement [16]. The kinetic energy expressed with the locked velocity is
3
1 1X T
(Vloc − Aϕ̇)T Mb (Vloc − Aϕ̇) + Vloc + Vi (ϕ̇) − Aϕ̇ Mi Vloc + Vi (ϕ̇) − Aϕ̇ .
T (Vloc , ϕ̇) =
2 2
i=1
(5.34)
The partial derivatives in (5.12) and (5.13) are found (replacing Ω α with Vloc
α
and ṙI with ϕ̇i , i = I)
¯
as
3
∂T
Mi Vi = LVloc
X
= L (Vloc − Aϕ̇) + (5.35)
∂Vloc
i=1
∂T
and ∂ ϕ̇as in (5.24). Consequently, the mass matrix becomes block diagonal is in (5.14). The Hamel
coefficients (5.6-5.8) are determined by the structure constants cα βδ on SE (3). Again, the curvature
α
BIJ = [AI , AJ ]α does not vanishing because of the non-commutativity of G = SE (3), where
¯ ¯ ¯
[AI , AJ ] = [m̄I , m̄J ] = adm̄Ī m̄J is the Lie bracket on se (3) in (5.28), i.e. screw product [80]. The
Euler-Lagrange equations (5.12),(5.13) are thus determined explicitly. Finally, the inverse of the
locked mass matrix attains the closed form
!
−1 U −Ude
e −1 .
L = 1 , with U = (Θ̄ + m̄d e d) (5.36)
dU m̄ I − dUd
e e e
The satellite pose is obtained by solving the local kinematic reconstruction equations
Ẋ = dexp−1 6∼
−X Vb for the instantaneous screw coordinate vector X = (x, y) ∈ R = se (3), see
supplement [16]. The matrix form of the dexp map on SE (3) also possesses a closed form [55].
..........................................................
∂l
Πα = = Lαβ ξ β + KαJ ṙJ . (6.1)
∂ξ α
Assuming that the initial momentum is zero, the momentum conservation Πα = 0 imposes
non-holonomic dynamic constraints uα = 0, α = 1, . . . , m̄, which are expressed in terms of the
mechanical connection with
uα = ξ α + Aα I I
I (r )ṙ . (6.2)
The Hamel equations are the Lagrange–d’Alembert–Poincaré equations (4.9), now with the
curvature of the mechanical connection in (5.5). The connection encodes dynamic constraints
due to the momentum conservation. If the initial momentum is zero, then the locked velocity
is also zero. Comparing ξ α = uα − Aα I α α α I
I ṙ , obtained from (6.2), with ξ = Ω − AI ṙ , obtained
from (5.5), shows that the equations (4.9) are obtained from the equations (5.13), in terms of the
locked velocity, when Ω α is set to zero and the mechanical connection is used:
d ∂` ∂` ∂` β J
− I + B ṙ = QI (6.3)
dt ∂ ṙI ∂r ∂Ω β IJ
β β
with `(rI , Ω α , ṙI ) := l(rI , ξ α := Ω α − Aα I I
I ṙ , ṙ ), and curvature BIJ = γIJ given by the Hamel
α
coefficients in (5.8), where Ω is set to zero after taking the derivatives. The system dynamics is
thus described in terms of coordinates rI on the base manifold (shape space). The motion in G
is determined as solution of the kinematic reconstruction equations (4.10) with ξ α = −Aα I (r) ṙ
I
with ∂∂T
ϕ̇i
∂T
, and ∂V α in (5.24) if G = SO (3) × R3 , and with ∂V
∂T
α in (5.35) if G = SE (3). The
loc loc
α
components of the curvature are the Hamel coefficients γIJ in (5.8) given with the structure
constants of the respective symmetry group G.
that replace equations (4.10). Solving the reconstruction equations for a full cycle along a closed
path in Vδ̄ yields the total phase shift as in (4.8), but now with the additional term gL−1 Π 0 . The
latter delivers the dynamic phase which is intrinsically due to the (initial) momentum, and leads
to a symmetry breaking from G to the symmetry group that preserves the initial momentum. If
(6.5) is regarded as a control problem, this term is the drift vector field. As an example, consider
the satellite in Sec. 5(f) with specific parameters. The rotation of the wheels is prescribed as r (t) =
(π (cos (2πt) − 1) , π sin (2πt) , π/2 sin (4πt)), which is periodic with cycle time 1 s. The geometric
and dynamic parameters, and animations can be found in the supplementary material [16]. First
assume zero total momentum. The motion of the base (i.e. of base frame Fb located at geometric
center of the base body, as shown in Fig. 2) is found from the reconstruction equations (4.11).
Fig. 3a) shows the translation of Fb in the x-y-plane of F0 over 6 s time duration, i.e. for six
cycles of the rotor motion, starting at the origin. Indicated is the position after each cycle, which
corresponds to the translation component of the geometric phase. The translation of Fb is caused
by the rotation about the total COM, which is not the origin of Fb . Fig. 3b) shows the translation
when the initial momentum is not zero. As an example, the momentum is set to Π 0 = Kṙ (0),
which is the momentum injected by the rotors when the base is at rest. This resembles the situation
where a satellite is released with spinning fly-wheels. The base motion is caused by the turning
rotors via the non-holonomic kinematics as well as the dynamics due to the momentum, which
determine the total phase. For completeness, the translation that is generated by the conserved
momentum only when the rotors are rest, is shown as dashed line, which yields the dynamic
phase.
2
y [mm]
y [mm]
4
Dynamic
1 2 phase
Total
phase
0 0
-2
-1 Geometric
phase
-4
-2
-1 0 1 2 3 0 2 4 6 8 10 12 14
a) x [mm] b) x [mm]
Figure 3. a) Translation of base frame Fb projected onto the x-y -plane of F0 (Fig. 2), when the total momentum is zero,
and Fb and F0 initially coincide. Positions after full cycles (with 1 s) of the rotor motion (geometric phase) are indicated.
b) Base motion for non-zero initial momentum (solid line), with positions after a full cycle of rotor motions indicated. Shown
separately (dashed line) is the motion only due to the initial momentum, and the corresponding dynamic phase.
7. Remark on Classical Riemannian Geometry Formulations 24
free. While this is a classical result, there is a beautiful relation for the linearized equations, which
is less known. Denote with (xa ) ∈ Rn small perturbations superposed to the nominal trajectory
q a , so that q a (t) + xa (t) is the perturbed trajectory. The linearized equations along the nominal
trajectory q a are, in covariant form, with the Riemann-Christoffel curvature tensor Rcbda ,
D 2 xb
gab 2
+ Rcbda q̇ c q̇ d − ∇b Qa xb = Φa (7.2)
dt
the covariant derivative ∇b Qa = ∂Q a
∂q b
− Γbac
Qc , and small generalized forces Φa dual to xa . The
curvature is hence a measure of stability of the perturbed dynamics.
8. Conclusion
The classical Hamel formulation is a generally applicable approach in analytical mechanics for
describing the dynamics of finite-dimensional systems in terms of local coordinates, which can
be extended to continua [87]. Frequently, the coordinate form of equations that can be derived
coordinate-free in the framework of geometric mechanics are referred to as Hamel equations. The
link between these conceptually very different approaches has not been sufficiently addressed,
however. This link was established in this paper, where the key is to identify the Hamel
coefficients as the coefficients appearing in the coordinate form of the reduced Euler-Lagrange
equations, respectively the Lagrange-Poincaré equations. Of particular significance are the local
curvature coefficients that are central in many aspects of control and computational treatment
of mechanical systems whose configuration space is a non-linear manifold or a Lie-group.
Therewith, a clear connection between the equations governing the dynamics on a principle
bundle, defined by a connection originating from certain symmetries, and the original Hamel
formulation is established. In this context the choice of bundle coordinates is crucial. As such the
locked velocity, and the related concept of average velocity, were discussed. The locked velocity
leads to inertial decoupling, which is important for control and computational investigations.
This should motivate further research into their use for deriving formulations with improved
efficiency. A problem that is increasingly receiving attention is that the average velocity cannot be 25
used for kinematic reconstruction [67,71,73,88,89]. This could, for instance, be addressed by means
of holonomy minimizing gauge transformations, i.e. introducing a frame that is not body-fixed
A. List of Symbols
n – number of (generalized) coordinates
m̄ – i) number of Pfaffian constraints, ii) dimension of the symmetry group G
δ̄ = n − m̄ – differential (instantaneous) DOF defined by the m̄ Pfaffian constraints
a, b, c, . . . – indices running over all coordinates: a = 1, . . . , n
I, K, L, . . . – indices I = m̄ + 1, . . . , n of i) independent velocity, ii) shape coordinates
(i.e. coordinates of the base manifold of the principle bundle) of a
constrained or unconstrained system
α, β, γ, . . . – indices α = 1, . . . , m̄ of i) constraint equations, ii) dependent velocity
coordinates, iii) canonical coordinates on the symmetry group G
i, j, k, l, . . . – indices i, j, k, l = 1, 2, 3 of Cartesian vectors, e.g. x = (xi ) ∈ R3
q a , q = (q a ) – local coordinates, generalized coordinates of unconstrained system
rI , r = (rI ) – independent (local) coordinates, I = m̄ + 1, . . . , n
sα , s = (sα ) – dependent (local) coordinates, α = 1, . . . , m̄
a α α
γbc , γIJ , γβJ – Hamel coefficients
α
u – i) quasi-velocities, ii) bundle coordinates, iii) Pfaffian constraints
Ωα – local coordinates of the locked velocity
Aα α
I , BIJ – coefficients of the local connection and of the local curvature
εijk – Levi-Civita symbol
δij – Kronecker delta symbol
x
e – skew symmetric matrix x e = (εikj xk ) associated to vector x = (xk ) ∈ R3
x×y – cross product of x, y ∈ R3 , can be written as x ey
η̂ ∈ g – element of Lie algebra g, corresponding to vector η ∈ Rn ∼ =g
cα
βλ – structure coefficients of the Lie algebra g
[X, Y ] – Lie bracket [X, Y ] = cα β λ
βλ X Y of X, Y ∈ g
F0 , Fb – Inertial frame (IFR) F0 , body-fixed frame Fb
SE (3) – special Euclidean group (rigid body motion group) SE (3) = SO (3) n R3
SO (3) – special orthogonal group (rotation group)
Ricci’s summation convention: e.g. BIa uI = I BIa uI , cα β λ
P P P α β λ
βλ X Y = β λ cβλ X Y
References
1. R. Montgomery, J. E. Marsden, and T. S. Ratiu, “Gauged Lie-Poisson structures,” Contemp.
Math, vol. 28, pp. 101–114, 1984.
2. R. W. Montgomery, The bundle picture in mechanics. University of California, Berkeley, 1986.
3. A. M. Bloch, Nonholonomic mechanics and control. Springer, 2003.
4. W.-S. Koon and J. E. Marsden, “Optimal control for holonomic and nonholonomic mechanical
systems with symmetry and Lagrangian reduction,” SIAM Journal on Control and Optimization,
vol. 35, no. 3, pp. 901–929, 1997.
5. N. Sreenath, “Nonlinear control of planar multibody systems in shape space,” Mathematics of 26
Control, Signals and Systems, vol. 5, no. 4, pp. 343–363, 1992.
6. E. A. Shammas, H. Choset, and A. A. Rizzi, “Towards a unified approach to motion