Conference PaperPDF Available

Relative equilibria and conserved quantities in symmetric Hamiltonian systems

Authors:

Figures

Content may be subject to copyright.
RELATIVE EQUILIBRIA AND CONSERVED QUANTITIES
IN SYMMETRIC HAMILTONIAN SYSTEMS
JAMES MONTALDI
Institut Non Lin´eaire de Nice
Published in:
Peyresq Lectures in Nonlinear Phenomena
R. Kaiser, J. Montaldi (eds)
World Scientific, 2000
1 Introduction
In this introduction, we first recall the basic phase space structures involved
in Hamiltonian systems, the symplectic form, the Poisson brackets and the
Hamiltonian function and vector fields, and the relationship between them.
Afterwards we describe a few examples of Hamiltonian systems, both of
the classical ‘kinetic+potential’ type as well as others using the symplec-
tic/Poisson structure more explicitly.
There are many applications of the ideas in these notes that have been
investigated by different people, but which I shall not cover. The major ex-
ample, the one for which classical mechanics was invented, is the gravitational
N-body problem. But there are many others too, such as rigid bodies, cou-
pled rigid bodies, coupled rods, underwater sea vehicles, . . . not to mention
the infinite dimensional systems such as water waves, fluid flow, plasmas and
elasticity. The interested reader should consult the books in the list of refer-
ences at the end of these notes. Note that many of the items in the list of
references are not in fact referred to in the text!
1.1 Hamilton’s equations
The archetypal Hamiltonian system describes the motion of a particle in a
potential well. If the particle has mass m, and V(x) is the potential energy
at the point x(in whatever Euclidean space), then Newton’s laws state that
m¨x=−∇V(x).
In the 18th century, Lagrange introduced the phase space by defining y= ˙x
and passing to a first order differential equation, and Hamilton carried this
further by introducing his now-famous equations
˙q=H/∂p,
˙p=∂H/∂q,
where qreplaces x, and p=m˙xis the momentum, and
H(q, p) = 1
2m|p|2+V(q)
J. Montaldi, Relative equilibria and conserved quantities . . . . 239
is the Hamiltonian, or total energy (kinetic + potential). This first order
system is equivalent to Newton’s law above, as is easily checked.
The principal advantage of Lagrange/Hamilton’s approach is that it is
more readily generalized to systems where the configuration space is not a
Euclidean space, but is a manifold. Such systems usually come about because
of constraints imposed (eg in the rigid body the constraints are that the dis-
tances between any two particles is fixed, and the configuration space is then
the set of rotations and translations in Euclidean 3-space).
The other advantage of the Hamilton’s approach is that it lends itself to
generalizations to systems that are not of the kinetic + potential type, such as
the model of a system of Npoint vortices in the plane or on a sphere, which
we will see below, or Euler’s equations modelling the “reduced” motion of a
rigid body.
These two generalizations lead to defining the dynamics in terms of a
Hamiltonian on a phase space, where the phase space has the additional
structure of being a Poisson or a symplectic manifold. The ‘canonical’ Poisson
structure is given by
{f, g}=
n
X
j=1
∂f
∂pj
∂g
∂qjg
∂pj
∂f
∂qj
,
where fand gare any two smooth functions on the phase space. The canonical
symplectic structure is given by,
ω=
n
X
j=1
dpjdqj=dα,
where α=Pn
j=1 pjdqjis the canonical Liouville 1-form.a
We shall see other examples of Poisson and symplectic structures in the
course of these lectures.
The Hamiltonian vector field XHis determined by the Hamiltonian H, a
smooth function on the phase space, in either of the following ways:
˙
F={H, F }
ω(v, XH) = dH(v),(1.1)
where ˙
F=XH(F) is the time-rate of change of the function Falong the
trajectories of the dynamical system. Combining these two expressions gives
aNot all authors agree on the choices of signs in the definition of the symplectic form or
the Poisson brackets, so when using formulae involving either from a text or paper, it is
necessary first to check the definitions. Our choice ensures that X{f,g}= [Xf, Xg].
J. Montaldi, Relative equilibria and conserved quantities . . . . 240
the useful formula relating the symplectic and Poisson structures,
ω(Xf, Xg) = {f, g}(1.2)
for any smooth functions f, g.
The first property of such Hamiltonian systems is their conservative na-
ture: the Hamiltonian function His conserved under the dynamics and so
too is the natural volume in phase space (Liouville’s theorem). This has an
important effect, not only on the type of dynamics encountered in such sys-
tems, but also on the types of generic bifurcations that can occur. Indeed, the
first consequence of these conservation laws (energy and volume) is that one
cannot have attractors in Hamiltonian systems, and in particular the notion
of asymptotic stability is not available.
A further feature of Hamiltonian systems is that symmetries lead to con-
served quantities. The two best-known examples of this are rotational sym-
metry leading to conservation of angular momentum, and translational sym-
metry to conservation of ordinary linear momentum. These further conserved
quantities could in principle complicate the types of dynamics and bifurca-
tions one sees. However a well defined process called reduction (or symplectic
reduction) can be used to replace the symmetry and conservation laws by a
family of Hamiltonian systems, parametrized by these conserved quantities,
on which there are general Hamiltonian systems whose bifurcations are those
expected in generic systems.
That said, there is a further complication which is that the phase space(s)
for these reduced systems may be singular, or change or degenerate in a family,
and we are only just beginning to understand the effects of these degenerations
on the dynamics and bifurcations.
1.2 Examples
Spherical pendulum A spherical pendulum is a particle constrained to move
on the surface of a sphere under the influence of gravity. As coordinates,
one can take spherical polar coordinates θ, φ (θmeasuring the angle with the
downward vertical, and φthe angle with a fixed horizontal axis). Of course,
this system has the defect of being singular at θ= 0, π. The kinetic energy is
T(q,˙
q) = 1
2mℓ2˙
θ2+ sin2θ˙
φ2,
while the potential energy is V(q) = mg(1 cos θ). The momenta conjugate
to the spherical polars are
pθ=L/∂ ˙
θ=mℓ2˙
θ, pφ=L/∂ ˙
φ=mℓ2sin2θ˙
φ,
J. Montaldi, Relative equilibria and conserved quantities . . . . 241
where L=TVis the Lagrangian, and the Hamiltonian is then
H(q,p) = 1
2mℓ2p2
θ+1
sin2θp2
φ+V(q).
The equations of motion are given as usual by Hamilton’s equations. In
particular, one can see from these equations that the angular momentum pφ
about the vertical axis is conserved since His independent of φ(one has
˙pφ=∂H/∂φ = 0).
Point vortices Since the work of Helmholtz, Kirchhoff and Poincar´e systems
of point vortices on the plane have been widely studied as finite dimensional
approximations to vorticity evolution in fluid dynamics. Small numbers of
point vortices model the dynamics of concentrated regions of vorticity while
large numbers can be used to approximate less concentrated regions. The
equations of motion can be derived by substituting delta functions into Euler’s
equation for a two dimensional ideal fluid. For general surveys of planar point
vortex systems see for example [17,3,15].
For this system of point vortices, each vortex has a vorticity—a real non-
zero number λ—and it is convenient to use complex numbers to describe
the positions of the vortices. The evolution is described by the differential
equation
˙
¯zj=1
2πi X
k6=j
λk
zjzk
,(1.3)
where zjis a complex number representing the position of the j-th vortex.
This system is Hamiltonian, with the Hamiltonian given by a pairwise inter-
action depending on the mutual distances:
H(z) = 1
4πX
j<k
λjλklog |zjzk|2
This is clearly not of the form ‘kinetic+potential’. The Poisson structure is
{f, g}=X
j
λ1
j(fzg¯zgzf¯z)
and the symplectic form is ω(u,v) = Pjλj(uj¯vj¯ujvj). Being of dimension
3, the Euclidean symmetry has 3 conserved quantities associated to it—see
equation (6.2).
A similar model can be obtained for point vortices on the sphere, providing
a simple model for cyclones and hurricanes in planetary atmospheres. See
Section 6.
J. Montaldi, Relative equilibria and conserved quantities . . . . 242
1.3 Symmetry
A transformation of the phase space T:P P is a symmetry of the Hamil-
tonian system, if
(i) H(T x) = H(x) for all x P,
(ii) Tpreserves the symplectic structure: Tω=ω, or
(ii’) Tpreserves the Poisson structure: {fT , g T}(x) = {f, g}(T(x)).
There are three basic ways that symmetries affect Hamiltonian systems:
(a) The image by Tof a solution is also a solution;
(b) A solution with initial point fixed by Tlies entirely within the set
Fix(T, P), where
Fix(T, P) = {x P | T x =x}.
(c) If Tis part of a continuous group, then the group gives rise to conserved
quantities (Noether’s theorem).
The first of these is clear, and in fact is also true of more generalized
symmetries for which HTHis constant, and Tω= (ca constant).
This occurs for example for homotheties of the plane in the planar point vortex
model described above. The second (b) is less obvious, but very well-known;
it follows from a very simple calculation as follows. If T x =xthen
σt(x) = σt(T x) = T σt(x),
where σtis the time tflow associated to the Hamiltonian system, and so σt(x)
is fixed by T. If Tis part of a compact group, then not only is Fix(T , P) invari-
ant, but it is a symplectic submanifold, and the restrictions of the Hamiltonian
and the symplectic form (or Poisson structure) to Fix(T , P) is a Hamiltonian
system which coincides with the restriction of the given Hamiltonian system—
an exercise for the reader. This technique of restricting to fixed point spaces
is sometimes called discrete reduction.
In these notes we will be concentrating on the effect of (c). The central
force problem provides the basic motivating example of this.
1.4 Central force problem
Consider a particle of mass mmoving in the plane under a conservative
force, whose potential depends only on the distance to the origin (a sim-
ilar analysis is possible for the spherical pendulum). It is then natural
J. Montaldi, Relative equilibria and conserved quantities . . . . 243
to use polar coordinates (r, φ) which are adapted to the rotational sym-
metry of the problem, so that V=V(r). The velocity of the particle is
˙
x= ( ˙rcos φ+ (rsin φ)˙
φ, ˙rsin φ(rcos φ)˙
φ), so that the kinetic energy is
T=m
2|˙
x|2=m
2˙r2+r2˙
φ2.
The Lagrangian is given by L=TV, and the Hamiltonian is H=T+V
with associated momentum variables given by pr=L/∂r =m˙rand pφ=
L/∂ ˙
φ=mr2˙
φ. Substituting for the velocities ˙rand ˙
φin terms of the
momenta determines the Hamiltonian to be:
H(r, φ, pr, pφ) = 1
2mp2
r+1
r2p2
φ+V(r).
Then Hamilton’s equations with respect to these variables are
(˙r=1
mpr,˙pr=1
mr3p2
φ+V(r)
˙
φ=1
mr2pφ˙pφ= 0.(1.4)
The last equation says that pφis preserved under the dynamics. In fact, pφ
is the angular momentum about the origin r= 0.
Since pφis preserved, let us consider a motion with initial condition for
which pφ=µ. Then (r, pr) evolve as
(˙r=1
mpr,
˙pr=1
mr3µ2+V(r)(1.5)
This is in fact a Hamiltonian system, with Hamilton Hµ(r, pr) obtained by
substituting µfor pφ. So that
Hµ(r, pr) = 1
2mp2
r+µ2
mr2+V(r).
This is a 1-degree of freedom problem, called the reduced system, with
“effective” potential energy
Vµ(r) = µ2
mr2+V(r),
and for a given potential energy function V(r), one can study how the be-
haviour of the system depends on µ.
For example, with the gravitational potential V(r) = 1/r, one obtains
an effective potential of the form in Figure 1 below, where the fist graph shows
the potential V(r) as a function of r, while the second and third show Vµ(r)
for increasing values of µ.
J. Montaldi, Relative equilibria and conserved quantities . . . . 244
It is clear that for µ= 0 there is no equilibrium for the reduced system,
while for µ > 0 there is an equilibrium, at rµsatisfying V
µ(rµ) = 0—here
rµ= 3µ2/m. Indeed, in this example it is a stable equilibrium for the effective
potential has a minimum at the rµ.
Figure 1. Effective potential for increasing values of µ, for V(r) = 1/r
Since r=rµand pr= 0 is an equilibrium of the reduced system, it is
natural to substitute these values into the original equation (1.4). The two
remaining equations are then
(˙
φ=µ
mr2
µ
˙pφ= 0.
This describes a simple periodic orbit in the original phase space:
(r, φ, pr, pφ)(t) = (rµ,µ
mr2
µt, 0, µ),
and there is one such periodic orbit for each value of µ6= 0 or more in
the general case if V
µ(r) = 0 has several solutions. In the case of a stable
equilibrium in the reduced space, the nearby solutions in the phase space will
be linear flows on invariant tori.
There are three things one should learn from this example:
the reduced dynamics can vary with the conserved quantity, giving rise
to bifurcation problems where the bifurcation parameter is an “internal”
variable;
the reduction is possible due to the symmetry of the original problem,
the relationship between the equilibrium in the reduced system and the
dynamic of the corresponding trajectory (relative equilibrium) in the full
phase space is given by the group action.
J. Montaldi, Relative equilibria and conserved quantities . . . . 245
1.5 Lie group actions
These notes assume the reader has a basic knowledge of actions of Lie groups
on manifolds. Here I recall a few basic formulae and properties that are used.
A useful reference is the new book by Chossat and Lauterbach [5].
Let Gbe a Lie group acting smoothly on a manifold P, and let gbe its
Lie algebra. We denote this action by (g, x)7→ g·x. The orbit through xis
G·x={g·x|gG}.
To each element ξgthere is associated a vector field on Pwhich we
denote ξP. It is defined as follows
ξP(x) = d
dt t= 0 (exp()·x).
The tangent space to the group orbit at xis then g·x={ξP(x)|ξg}.
A simple calculation relates the vector field at xwith its image at g·x:
dgxξP(x) = (Adgξ)P(g·x),(1.6)
where Adgξis the adjoint action of gon ξ, which in the case of matrix groups
is just
Adgξ=g1.
The adjoint representation of gon gis the infinitesimal version obtained by
differentiating the adjoint action of G:
adξη=d
dt t= 0 Adexp()η= [ξ, η ].
Dual to the adjoint action on gis the coadjoint action on g:
hCoadgµ, ηi:= µ, Adg1η,(1.7)
and similarly there is the infinitesimal version,
hcoadξµ, ηi:= hµ, adξηi=hµ, [η, ξ]i.(1.8)
Examples 2.5 describe the coadjoint actions for the groups SO(3),SE(2) and
SL(2).
Given x P, the isotropy subgroup of xis
Gx={gG|g·x=x}.
The Lie algebra gxof Gxconsists of those ξgfor which ξP(x) = 0, and the
fixed point set of K
Fix(K, P) = {x P | K·x=x},
J. Montaldi, Relative equilibria and conserved quantities . . . . 246
consists of those points whose isotropy subgroup contains K. It is not hard
to show that it’s a submanifold of P. Moreover, those points with isotropy
precisely Kform an open (possibly empty) subset of Fix(K, P).
Stratification by orbit type If Gacts on a manifold P, then the orbit space
P/G is smooth at points where Gpis trivial, and more generally where the
orbit type in a neighbourhood of pis constant.
More generally, for each subgroup H < G one defines the orbit type stra-
tum P(H)to be the set of points pfor which Gpis conjugate to H. This is a
union of G-orbits, and its image in P/G is also called the orbit type stratum
(now in the orbit space). These orbit type strata are submanifolds of Pand
P/G, and they fit together to form a locally trivial stratification (i.e. locally
it has a product structure).
For dynamical systems, the importance of this partition in to orbit type
strata, is that for an equivariant vector field, the strata are preserved by the
dynamics.
Slice to a group action Aslice to a group action at x P is a submanifold of
Pwhich is transverse to the orbit through xand of complementary dimension.
If possible, the slice is chosen to be invariant under the isotropy subgroup Gx
(this is always possible if Gxis compact). A basic result of the theory of Lie
group actions is that under the orbit map P P/G the slice pro jects to a
neighbourhood of the image of G·xin the orbit space.
Principle of symmetric criticality This principle is the variational version of
discrete reduction, and provides a useful method for finding critical points of
invariant functions. It states that, if Gacts on a manifold P, and if f:P R
is a smooth invariant function, then xFix(G, P) is a critical point of fif
and only if it is a critical point of the restriction f|Fix(G,P)of fto Fix(G, P).
One proof is to use an invariant Riemannian metric to define an equivariant
vector field f, which being equivariant, is tangent to Fix(G, P). For a full
proof, valid also in infinite dimensions, see [58].
2 Noether’s Theorem and the Momentum Map
The purpose of this section is to bring together facts about symmetry and
conserved quantities that are useful for studying bifurcations. They are all
found in various places, more or less explicitly, but not together in a single
source. Furthermore, there appear to be misconceptions about whether “non-
equivariant” momentum maps cause extra problems. Essentially, anything
true for the equivariant ones remains true for non-equivariant ones (which are
J. Montaldi, Relative equilibria and conserved quantities . . . . 247
in fact equivariant, but for a modified action, as we shall see below). Many
of the details of this chapter can be found in the books [16,8,10,13].
Examples 2.1 We will be using 3 examples of symplectic group actions in
this section to illustrate various points. These arise in the models of systems
of point vortices.
(A) G=SO(3) acts by rotations on the sphere P=S2, with symplectic
form given by the usual area form with total area 4π(for example in spherical
polars, ω= sin(θ) ). The Lie algebra so(3) can be represented by skew-
symmetric matrices, and the vector field corresponding to the skew-symmetric
matrix ξis simply x7→ ξx.
(B) G=SE(2) acts on the plane P=R2, with its usual symplectic form ω=
dx dy. This group acts by translation and rotations; indeed, SE(2) R2
SO(2) (semidirect product), where R2is the normal subgroup of translations
of the plane, and SO(2) is the group of rotations about some point, e.g.
the origin. The Lie algebra se(2) is represented by constant vector fields
(corresponding to the translation subgroup) and by infinitesimal rotations.
(C) G=SL(2) = SL(2,R) acts by isometries on the hyperbolic plane P=H.
There are several ways to realize this action, of which perhaps the best-known
is to use obius transformations on the upper-half plane. However, we will
use one that is more in keeping with the others, which is to represent the
hyperbolic plane as one sheet of a 2-sheeted hyperboloid in R3:
H={(x, y, z)R3|z2x2y2= 1, z > 0}.(2.1)
(The hyperbolic metric on His induced from the Minkowski metric (dx2+
dy2dz2) on R3.) The symplectic form on His given by
ωx(u,v) = 1
2kxk2R(x)·uv,(2.2)
where R(x, y, z) = (x, y, z) and u,vTxH. One way to realize the action
of SL(2) on His to embed Hinto the set of 2 ×2-trace zero matrices sl2(R):
x=
x
y
z
7→ ˆ
x=x y +z
yzx.(2.3)
Then A·ˆ
x=Aˆ
xA1, for ASL(2). Note that the image of the embedding
consists of those matrices Xof trace zero, unit determinant and such that
X12 > X21. Under this identification, the symplectic form (2.2) becomes the
Kostant-Kirillov-Souriau symplectic form on the coadjoint orbit (see Example
2.5(C)). 2
J. Montaldi, Relative equilibria and conserved quantities . . . . 248
2.1 Noether’s theorem
With such a set-up, the famous theorem of Emmy Noether states that any
1-parameter group of symmetries is associated to a conserved quantity for
the dynamics. In fact one needs some hypothesis such as the phase space
being simply connected, or the group being semisimple (see [8] for details).
For example, the circle group acting on the torus does not produce a globally
well-defined conserved quantity.
How do these conserved quantities come about? We already have a pro-
cedure for passing from Hamiltonian function to Hamiltonian vector field, and
here we apply the reverse procedure. For each ξglet φξ:P Rbe a
function that satisfies Hamilton’s equation
ξ=ω(ξP,),(2.4)
if such a function exists. Of course each of these φξis only defined up to a
constant, since only ξis determined.
Such functions are known as momentum functions, and a symplectic ac-
tion for which such momentum functions exist is said to be a Hamiltonian
action. See [8] for conditions under which symplectic actions are Hamilto-
nian.
Theorem 2.2 (Noether) Consider a Hamiltonian action of the Lie group G
on the symplectic (or Poisson) manifold P, and let Hbe an invariant Hamil-
tonian. Then the flow of the Hamiltonian vector field leaves the momentum
functions φξinvariant.
Proof: A simple algebraic computation:
XH(φξ) = {H, φξ}=−{φξ, H }=ξP(H) = 0.
This last equation holds because His G-invariant.
Momentum map We leave questions of dynamics now, and consider the
structure of the set of momentum functions. The first observation is that the
momentum functions φξcan be chosen to depend linearly on ξ. For example,
let {ξ1,...,ξd}be a basis for g, and let φξ1,...,φξdbe Hamiltonian functions
for the associated vector fields. Then for ξ=a1ξ1+···+adξd, one can put,
φξ=a1φξ1+···+adφξd.
It is easy to check that such φξsatisfy the necessary equation (2.4).
Thus, for each point p P we have a linear functional ξ7→ φξ(p), which
we call Φ(p), so that
Φ(p) = (φξ1(p),...,φξd(p))
J. Montaldi, Relative equilibria and conserved quantities . . . . 249
is a map Φ : P g, where gis the vector space dual of the Lie algebra
g. Such a map is called a momentum map. The defining equation for the
momentum map is,
hdΦp(v), ξi=ωp(ξP(p),v).(2.5)
for all p P all vTpPand all ξg. An immediate and important
consequence of (2.5) is:
ker(dΦp) = (g·p)ω(2.6)
im(dΦp) = g
pg(2.7)
Here, Uωis the linear space that is orthogonal to Uwith respect to the
symplectic form. In particular, the momentum map is a submersion in a
neighbourhood of any point where the action is free (or locally free: gp= 0).
The fact that Φ is only determined by the differential condition (2.5)
means that it is only defined up to a constant. It follows that if Φ is a
momentum map, then Φ1is a momentum map if and only if there is some
Cgfor which, for all p P,
Φ1(p) = Φ(p) + C.
We return to the possibility of choosing a different Φ below.
Examples 2.3 Here we see momentum maps for three symplectic actions
that arise for the point-vortex models, firstly for point vortices on the sphere,
secondly for those in the plane and thirdly on the hyperbolic plane. We also
give the general formula for momentum maps for coadjoint actions.
(A) Let G=SO(3) act diagonally on the product P=S2×... ×S2(N
copies). On the ith factor we put the symplectic form ωi=λiω0, where ω0is
the canonical area form on the unit sphere (RS2ω0= 4π). Then a momentum
map is given by
Φ(x1,...,xN) = X
j
λjxj,(2.8)
where the Lie algebra so(3) (consisting of skew symmetric 3 ×3 matrices)
is identified with R3in the “usual way”: Bso(3) corresponds to bR3
satisfying Bu =b×ufor all vectors uR3. It is clear that this momentum
map is equivariant with respect to the coadjoint action, which under this
identification becomes the usual action of SO(3) on R3.
(B) An analogous example is G=SE(2) acting on a product of Nplanes,
with symplectic form ω=jλjωj, where ωjis the standard symplectic form
J. Montaldi, Relative equilibria and conserved quantities . . . . 250
on the jth plane. If we write SE(2) as R2SO(2) (semidirect product), then
the natural momentum map is
Φ(x1,...,xN) =
X
j
λjJxj,1
2Pjλj|xj|2
.(2.9)
where J=01
1 0 is the matrix for rotation through π/2.
(C) A further analogous example is G=SL(2) acting on a product of N
copies of the hyperbolic plane, P=HN, with symplectic form ω=jλjωj,
where ωjis the standard symplectic form on the jth plane, see (2.2). The
natural momentum map is given by,
Φ(x1,...,xN) = X
j
λjxj.(2.10)
2
Remark 2.4 Consider any group acting on a manifold X, the configuration
space. Classical mechanics of the “kinetic + potential” type takes place on the
cotangent bundle of a configuration space, and in this setting the given action
on Xinduces a symplectic action on the cotangent bundle, by the formula
g·(x, p) = (g·x, (dgx)Tp),
where ATis the inverse transpose of the operator A. Such actions are called
cotangent actions or cotangent lifts and they always preserve the canonical
symplectic form ωon the cotangent bundle.
The momentum map for cotangent actions always exists, and is given by
hΦ(x, p), ξi=hp, ξP(x)i,
where the pairing on the right is between T
xXand TxX. For example, if
P=TR3is the phase space for a central force problem, which has SO(3)
symmetry, then after identifying so(3)with R3as above, the momentum
map Φ : P so(3)is just the angular momentum.
2.2 Equivariance of the momentum map
A natural question arises: since a momentum map Φ : P gis defined on a
space Pwith an action of the group G, is there an action of Gon gfor which
Φ commutes with (intertwines) the two actions? The answer was given in the
affirmative by Souriau [16]. Usually, but not always, this turns out to be the
coadjoint action of Gon g, and before stating Souriau’s result we give some
examples of coadjoint action.
J. Montaldi, Relative equilibria and conserved quantities . . . . 251
Examples 2.5 The coadjoint actions for the groups described in Examples
2.1 are as follows.
(A) For G=SO(3), the Lie algebra g=so(3) consists of the 3 ×3 skew-
symmetric matrices, and via the inner product hA, Bi= tr(ABT) we can
identify the dual gwith g. An easy computation shows that the coadjoint
action is given by
CoadAµ=AµAT.
With the usual identification of so(3) with R3(see Example 2.3) the coadjoint
action is just the usual action of SO(3) by rotations, so that the orbits for the
coadjoint actions are spheres centered at the origin, and the origin itself. For
each of the 2 types of orbit, the isotropy subgroup is either all of SO(3), or
it is a circle subgroup SO(2) of SO(3). This variation of the symmetry type
of the orbits has interesting repercussions for the dynamics, and in particular
for the families of relative equilibria.
(B) Consider now the 3-dimensional non-compact group G=SE(2) of
Euclidean motions of the plane. As a group this is a semidirect product
R2SO(2), where R2acts by translations and SO(2) by rotations about
some given point (the “origin”). Elements (u, R)R2SO(2) can be iden-
tified with elements
Ru
0 1 GL(R2×R),
by introducing homogeneous coordinates. A calculation shows that the adjoint
action is
Ad(u,R)(v, B) = (RvBu, B),
since Band Rcommute. The coadjoint action is then
Coad(u,R)(ν, ψ) = (Rν, ψ +uT).(2.11)
Note that since elements of so(2) are skew-symmetric, it follows that only
the skew-symmetric part of ψ+uTis relevant. One can therefore replace
ψ+uTby ψ+1
2(uTuνTRT). A nice representation of this using
complex numbers is given in Section 6.
The coadjoint orbits are again of two types: first the cylinders with axis
along the 1-dimensional subspace of gwhich annihilates the translation sub-
group (or its subalgebra), and secondly the individual points on that axis;
that is, the points of the form (0, ψ). In this case the two types of isotropy
subgroup for the coadjoint action are firstly the translations in a given direc-
tion (orthogonal to ν), so isomorphic to R, or in the case of a single point
J. Montaldi, Relative equilibria and conserved quantities . . . . 252
ψ
Figure 2. Coadjoint orbits for SO(3) and for SE(2)
on the axis, it is the whole group. In both cases the isotropy subgroup is
non-compact, a fact to be contrasted with the modified coadjoint action to be
defined below.
(C) For G=SL(2). Let ASL(2) and µsl(2), then
CoadAµ=ATµAT.
Notice that the determinant of the matrix in sl(2)is constant on each coad-
joint orbit. In fact the coadjoint orbits are of 4 types: (i) the 1-sheeted
hyperboloids (where GµR), (ii) one sheet of the 2-sheeted hyperboloids
(where GµSO(2)), (iii) each sheet of the cone with the origin removed
(where GµR), and (iv) the origin itself (where Gµ=SL(2)). 2
It is easy to see that the momentum maps given in Examples 2.3(A,C)
are equivariant with respect to the coadjoint actions described above, which
is not surprising in the light of the theorem below. However this is not always
true in the case of G=SE(2).
To describe the action that makes Φ equivariant we follow Souriau and
define the cocycle
θ:Gg
g7→ Φ(g·x)CoadgΦ(x),(2.12)
It is of course necessary to show that this expression is independent of x, which
it is provided Pis connected. We leave the details to the reader: it suffices to
differentiate with respect to xand use the invariance of the symplectic form.
The map θdefined above allows one to define a modified coadjoint action,
by
Coadθ
gµ:= Coadgµ+θ(g).(2.13)
A short calculation shows that this is indeed an action. Moreover, this action
is by affine transformations whose underlying linear transformations are the
coadjoint action.
J. Montaldi, Relative equilibria and conserved quantities . . . . 253
Theorem 2.6 (Souriau) Let the Lie group Gact on the connected symplec-
tic manifold Pin such a way that there is a momentum map Φ : P g. Let
θ:Ggbe defined by (2.12). Then Φis equivariant with respect to the
modified action on g:
Φ(g·x) = Coadθ
gΦ(x).
Furthermore, if Gis either semisimple or compact then the momentum map
can be chosen so that θ= 0.
For proofs see [16] or [8]; the first proof of equivariance in the compact
case appears to be in [48].
Examples 2.7 In Examples 2.3 we gave the momentum maps for the three
point-vortex models, and pointed out that for SO(3) and SL(2) the mo-
mentum map is equivariant with respect to the usual coadjoint action (not
surprisingly in view of the theorem above since SO(3) is compact and SL(2)
is semisimple). However, this is not always true in the planar case:
(B) Consider the momentum map given in Example 2.3(B) for the point
vortex model in the plane:
Φ(x1,...,xN) =
X
j
λjJxj,1
2Pjλj|xj|2
.
To find the action that makes this momentum map equivariant, we compute
Φ(Ax1+u,...,AxN+u) = AX
j
λjJxj,1
2Pjλj|xj|2+PjλjAxj.u+
+ ΛJu,1
2|u|2
= Coad(u,A)Φ(x1,...,xN) + Λ(Ju,1
2|u|2),(2.14)
where Λ = PjλjR, and Coad is given in (2.11). The cocycle associated to
this momentum map is thus given by θ(u, A) = Λ(Ju,1
2|u|2). If Λ = 0 then
Φ is equivariant with respect to the usual coadjoint action, while if Λ 6= 0
it is equivariant with respect to a modified coadjoint action. Furthermore,
one can show that in this latter case there is no constant vector Cgfor
which Φ + Cis coadjoint-equivariant. Indeed, it is enough to see that the
orbits for this modified coadjoint action are in fact paraboloids, with axis the
annihilator in gof R2gand these are not translations of the coadjoint
orbits, which are either cylinders or points. See Figure 3. Furthermore, a
short calculation shows that the isotropy subgroups for this action are all
compact: GµSO(2), for all µse(2), which is quite different from the
coadjoint action. 2
J. Montaldi, Relative equilibria and conserved quantities . . . . 254
ψ
Figure 3. Modified coadjoint orbits for SE(2)
2.3 Reduction
Since by Noether’s theorem the dynamics preserve the level sets of the mo-
mentum map Φ, it makes sense to treat the dynamical problem one level set at
a time. However, these level sets are not in general symplectic submanifolds,
and the system on them is therefore not a Hamiltonian system. However, it
turns out that if one passes to the orbit space of one of these level sets of
Φ, then the resulting reduced space is symplectic, and the induced dynamics
are Hamiltonian. Historically, this process of reduction was first used by Ja-
cobi, in what is called “elimination of the nodes”. However, its systematic
treatment is much more recent, and is due to Meyer [47] and independently
to Marsden and Weinstein [43].
As usual suppose the Lie group Gacts in a Hamiltonian fashion on the
symplectic manifold P, and let Φ : P gbe a momentum map which is
equivariant with respect to a (possibly modified) coadjoint action as discussed
above. Consider a value µgof the momentum map. Then since Φ is
equivariant, the isotropy subgroup Gµof this modified coadjoint action acts
on the level set Φ1(µ). Define the reduced space to be
Pµ:= Φ1(µ)/Gµ.(2.15)
That is, two points of Φ1(µ) are identified if and only if they lie in the same
group orbit. This defines Pµas a set, but to do dynamics one needs to give
it more structure.
If the group Gacts freely near p(i.e. Gpis trivial), then Pµis a smooth
manifold near the image of pin Pµ. This is for two reasons: firstly Φ is a
submersion near pby (2.7), and secondly the orbit space by Gµwill have
no singularities. This is the case of regular reduction. We discuss singular
reduction briefly below.
It is important to know whether the induced dynamics on the reduced
J. Montaldi, Relative equilibria and conserved quantities . . . . 255
spaces are also Hamiltonian. The answer of course is “yes”. To see this it
is necessary to define the symplectic form ωµon Pµ. Let ¯u, ¯vT¯pPµ, be
projections of u, v TpP, and define
ωµu, ¯v) := ω(u, v).
Of course, one must show this to be well defined, non-degenerate and closed:
exercises left to the reader!
The Hamiltonian is an invariant function on Pso its restriction to Φ1(µ)
is Gµ-invariant and so induces a well-defined function on Pµ, denoted Hµ. The
dynamics induced on the reduced space is then determined by a vector field
Xµwhich satisfies Hamilton’s equation: dHµ=ωµ(Xµ,).
The orbit momentum map It is clear that Pµ= Φ1(µ)/G can also be
defined by Pµ= Φ1(Oµ)/G, where Oµis the coadjoint orbit through µ.
Since Φ1(Oµ)/G P/G it is natural to use the orbit momentum map:
ϕ:P/G g/G defined by
PΦ
g
P/G ϕ
g/G
(2.16)
where the vertical arrows are the quotient maps. Then Pµ=ϕ1(Oµ). This is
very useful for studying bifurcations as the momentum value varies. However,
it may not be so useful if Gµis not compact, for there the orbit space g/G
is not in general a reasonable space (it is not even Hausdorff for example for
the coadjoint action of SE(2) near the ψ-axis).
2.4 Singular reduction
If the action of Gon Pis not free, then the reduced space is no longer a
manifold. However, it was shown by Sjamaar and Lerman [67] that if Gis
compact then the reduced space can be stratified—that is, decomposed into
finitely many submanifolds which fit together in a nice way—and each stratum
has a symplectic form which determines the dynamics on that stratum. In
fact these strata are simply the sets of constant orbit type in Φ1(µ), or
rather their images in Pµ. For the case of a proper action of a non-compact
group see [20]. The theorem follows from the local normal form of Marle and
Guillemin-Sternberg [8,42].
J. Montaldi, Relative equilibria and conserved quantities . . . . 256
2.5 Symplectic slice and the reduced space
Recall that a slice to a group action at a point p P is a submanifold S
through psatisfying TpSg·p=TpP. If Gpis compact, it can be chosen
to be invariant under Gp. The slice, or more precisely S/Gp, provides a local
model for the orbit space P/G.
In the symplectic world, one needs to take into account the symplectic
structure, and one wants the “symplectic slice” to provide a local model for
the reduced space Pµ. In the case of free (or locally free) actions this is fairly
straightforward, but in the general case, this is more delicate because the
momentum map is singular. For this reason, the symplectic slice is usually
taken to be a subspace of TpPrather than a submanifold of P.
Definition 2.8 Suppose Gpis compact. Define NTpPto be a Gpinvariant
subspace satisfying TpP=Ng·p. The symplectic slice is then
N1:= Nker(dΦp).
It follows from the implicit function theorem that local coordinates can
be chosen that identify a transversal to Gµ·pwithin the possibly singular set
Φ1(µ) with a subset of the symplectic slice N1.
3 Relative Equilibria
An equilibrium point is a point in the phase space that is invariant under the
dynamics: p P for which XH(p) = 0, or equivalently dHp= 0, and one way
to define a relative equilibrium is as a group orbit that is invariant under the
dynamics. Although geometrically appealing, this is not the most physically
transparent definition.
Definition 3.1 Arelative equilibrium is a trajectory γ(t) in Psuch that for
each tRthere is a symmetry transformation gtGfor which γ(t) =
gt·γ(0).
In other words, the trajectory is contained in a single group orbit. It
is clear that if a group orbit is invariant under the dynamics, then all the
trajectories in it are relative equilibria; and conversely, if γ(t) is the trajectory
through p, then g·γ(t) is the trajectory through g·pand consequently the
entire group orbit is invariant as claimed above.
For N-body problems in space, relative equilibria are simply motions
where the shape of the body does not change, and such motions are always
rigid rotations about some axis.
Proposition 3.2 Let Φbe a momentum map for the G-action on Pand let
Hbe a G-invariant Hamiltonian on P. Let p P and let µ= Φ(p). Then
J. Montaldi, Relative equilibria and conserved quantities . . . . 257
the following are equivalent:
1The trajectory γ(t)through pis a relative equilibrium,
2The group orbit G·pis invariant under the dynamics,
3ξgsuch that γ(t) = exp()·p, tR,
4ξgsuch that pis a critical point of Hξ=Hφξ,
5pis a critical point of the restriction of Hto the level set Φ1(µ).
Remarks 3.3 (i) The vector ξappearing in (3) is the angular velocity of the
relative equilibrium. It is the same as the vector ξappearing in (4). The
angular velocity is only unique if the action is locally free at p; in general it
is well-defined modulo gp.
(ii) If Φ1(µ) is singular then it has a natural stratification (see §2.4), and
condition (4) of the proposition should be interpreted as being a stratified
critical point; that is all derivatives of Halong the stratum containing p
vanish at p.
(iii) Notice that (3) implies that relative equilibria cannot meandre around
a group orbit, but must move in a rather rigid fashion. It follows from this
equation that the trajectory is in fact a dense linear winding on a torus, at
least if Gis compact. The dimension of the torus is at most equal to the rank
of the group G. For G=SO(3), the rank is 1, and this means that any re
that is not an equilibrium is in fact a periodic orbit.
Proof: The equivalence of (1) and (2) is outlined above.
Equivalence of (1) and (3): (3) (1) is clear. For the converse, since
γ(t) lies in G·pso its derivative XH(p) = ˙γ(0) lies in the tangent space
g·p=Tp(G·p). Let ξgbe such that XH(p) = ξP(p). Then by equivariance
XH(g·p) = (Adgξ)P(g·p), see (1.6). Let gt= exp(). Then since Adgtξ=ξ,
d
dt(gt·p) = ξP(gt·p) = XH(gt·p).
That is, t7→ gt·pis the unique solution through p.
Equivalence of (3) and (4): (3) is equivalent to XH(p) = ξP(p). Using
the symplectic form, this is in turn equivalent to dH(p) = ξ(p).
Equivalence of (4) and (5): If Φ is submersive at pthen Φ1(µ) is a
submanifold of Pand this is just Lagrange multipliers, since ξ=hdΦ(p), ξi.
In the case that Φ is singular, the result follows from the principle of symmetric
criticality (see the end of the Introduction for a statement). If Φ is singular at
p, then by (2.7) gp6= 0. Consider the restriction of Hto Fix(Gp,P). By the
J. Montaldi, Relative equilibria and conserved quantities . . . . 258
theorem of Sjamaar and Lerman, the set Φ1(µ) is stratified by the subsets
of constant orbit type—see §2.4. Consider then the set PGpof points with
isotropy precisely Gp(this an open subset of Fix(Gp,P) containing p), and
restrict both Φ and Hto this submanifold. Now Φ restricted to PGpis of
constant rank, so that Φ1(µ)PGpis a submanifold. The result then follows
as before, since by the principle of symmetric criticality Hrestricted to PGp
has a critical point at pif and only if Hhas a critical point at p.
One of the earliest systematic investigations of relative equilibria was in
a paper of Riemann where he classified all possible relative equilibria in a
model of affine fluid flow that is now called the Riemann ellipsoid problem
(or affine rigid body or pseudo-rigid body)—see [25] and [66] for details. The
configuration space is the set of all 3×3 invertible matrices, and the symmetry
group is SO(3)×SO(3). In that paper he identified the 6 conserved quantities
and found geometric restrictions on the possible forms of relative equilibria
using the fact that the momentum is conserved, and that the solutions are
of the form given in (3) of the proposition above. In terms of general group
actions, the geometric condition Riemann used is the following, which follows
immediately from Proposition 3.2 above together with the conservation of
momentum.
Corollary 3.4 Let p P be a point of a relative equilibrium, of angular
velocity ξ, and let θbe the cocycle associated to the momentum map, then
coadθ
ξΦ(p) = 0.
If Gis compact, so the adjoint and coadjoint actions can be identified and
θ= 0 for a suitable choice of momentum map, this means that the angular
velocity and momentum of a relative equilibrium commute. For example, for
any system with symmetry SO(3) this means that at any relative equilibrium,
the angular velocity and the value of the momentum are parallel.
Definition 3.5 A relative equilibrium through pis said to be non-degenerate
if the restriction of the Hessian d2Hξ(p) to the symplectic slice N1is a non-
degenerate quadratic form.
Definition 3.6 A point µgis a regular point of the (modified) coadjoint
action if in a neighbourhood of µall the isotropy subgroups are conjugate.
Examples of regular points are: all points except the origin for the coad-
joint action of SO(3), all points except the special axis for the coadjoint action
of SE(2), and all points for the modified coadjoint action of SE(2) described
in Example 2.7(B).
The following result, the first on the structure of the families of relative
equilibria, was observed by V.I. Arnold in [2]. The proof is an application of
J. Montaldi, Relative equilibria and conserved quantities . . . . 259
the implicit function theorem.
Theorem 3.7 (Arnold) Suppose that plies on a non-degenerate relative
equilibrium, with Gp= 0 and µ= Φ(p)a regular point of the (modified)
coadjoint action. Then in a neighbourhood of pthere exists a smooth family
of relative equilibria parametrized by µg.
Lyapounov stability A compact invariant subset Sof phase space is said to be
Lyapounov stable if “any motion that starts nearby remains nearby”, or more
precisely, for every neighbourhood Vof Sthere is another neighbourhood
VVsuch that every trajectory intersecting Vis entirely contained in V.
A compact subset is said to be Lyapounov stable relative to or modulo G
if the Vand Vabove are only required to be G-invariant subsets.
The principal tool for showing an equilibrium to be Lyapounov stable is
Dirichlet’s criterion, which is that if the equilibrium point is a non-degenerate
local minimum of the Hamiltonian, then it is Lyapounov stable. The proof
consists of noting that in this case the level sets of the Hamiltonian are topolog-
ically spheres surrounding the equilibrium, and so by conservation of energy,
if a trajectory lies within one of these spheres, it remains within it.
It is reasonably clear firstly that it is sufficient if any conserved quantity
has a local minimum at the equilibrium point, not necessarily the Hamiltonian
itself, and secondly that the local minimum may in fact be degenerate. These
observations lead to the notion of . . .
Extremal relative equilibria A special role is played by extremal relative
equilibria. This is partly due to Dirichlet’s criterion for Lyapounov stability,
and partly because of their robustness.
A relative equilibrium is said to be extremal if the reduced Hamiltonian
Hµon Pµhas a local extremum (max or min) at that point. This is usually
established by showing the restriction to the symplectic slice of the Hessian
of Hξ=Hφξto be positive (or negative) definite, for some ξfor which Hξ
has a critical point at p, see Proposition 3.2.
It is of course conceivable that a relative equilibrium is extremal while the
Hessian matrix is degenerate. The simplest case of this is for H(x, y) = x2+y4
in the plane. In fact this arises in the case of 7 identical point vortices in the
plane. The configuration where they lie at the vertices of a regular heptagon
is a relative equilibrium, and the Hessian of Hon the symplectic slice is only
positive semi-definite. However, a lengthy calculation shows that the relevant
fourth order terms do not vanish, and the reduced Hamiltonian does indeed
have a local minimum there.
For the following statement, recall that any momentum map is equivariant
with respect to an appropriate action of Gon g(Section 2), and for µg
J. Montaldi, Relative equilibria and conserved quantities . . . . 260
we write Gµfor the isotropy subgroup of µfor that action.
Theorem 3.8 ([48,34]) Let Gact properly on Pwith momentum map Φ,
and suppose pΦ1(µ)is an extremal relative equilibrium, with Gµcompact.
Then,
(i) The relative equilibrium is Lyapounov stable, relative to G;
(ii) There is a G-invariant neighbourhood Uof psuch that, for all µΦ(U)
there is a relative equilibrium in UΦ1(µ).
The proof is mostly point-set topology on the orbit space and using the
orbit momentum map (2.16), though part (ii) uses the deeper property of
(local) openness of the momentum map. In fact the proof of (i) also holds if µis
a regular point for the (modified) coadjoint action, even if Gµis not compact.
By Proposition 2.4 of [35] this result can be refined to conclude that the
relative equilibrium is stable relative to Gµ, as described by Patrick [59]. Note
that the compactness of Gµensures that gµhas a Gµ-invariant complement in
g, as required in [35]. A consequence of using point-set topology is that there
is very little information on the structure of the family of relative equilibria;
for such information see Section 5.
The crucial remaining point is how to determine whether a given relative
equilibrium is extremal. In the case of free actions this was done by the
so-called energy-momentum and/or energy-Casimir methods of Arnold and
Marsden and others. Recently this was extended to the general case of proper
actions:
Proposition 3.9 (Lerman, Singer [35]) Let pΦ1(µ)be a relative equi-
librium satisfying the same hypotheses as the theorem above, and let ξgbe
any angular velocity of the re as in Proposition 3.2. If the restriction to the
symplectic slice of the quadratic form d2Hξis definite, then the re is extremal,
and so Lyapounov stable relative to Gµ.
4 Bifurcations of (relative) equilibria
In this section we take an extremely brief look at the typical bifurcations of
equilibria in families of Hamiltonian systems as a single parameter is varied.
These results also apply to relative equilibria, provided the reduced space is
smooth in a neighbourhood of the relative equilibrium, and failing that, it
applies to the stratum containing the relative equilibrium. Bifurcations of
relative equilibria near singular points of the reduced space have not been
investigated systematically.
J. Montaldi, Relative equilibria and conserved quantities . . . . 261
4.1 One degree of freedom
Saddle-centre bifurcation
This is the generic bifurcation of equilibria in 1 d.o.f. See Fig. 4. For
λ < 0 there are two equilibrium points: at (x, y) = (±λ, 0), one local
extremum and one saddle. As λ0 these coalesce and then disappear
for λ > 0. One notices also a homoclinic orbit for λ < 0, connecting the
saddle point with itself, which can also be seen as a limit of the family of
periodic orbits surrounding the stable equilibrium. In terms of eigenvalues of
the associated linear system, this bifurcation can be seen as a pair of simple
imaginary eigenvalues (a centre) decreases along the imaginary axis, collide at
0 and “then” emerge along the real axis (a saddle). This description is slightly
misleading as the centre and saddle coexist. At the point of bifurcation, the
linear system is 0 1
0 0 ; that is, it has non-zero nilpotent part.
Note that this bifurcation is compatible with an antisymplectic (i.e. time-
reversing) symmetry (x, y)(x, y).
λ < 0λ= 0 λ > 0
H(x, y) = 1
2y2+1
3x3+λx + h.o.t.
Figure 4. Saddle-centre bifurcation
Symmetric pitchfork This is usually caused by symmetry, and is 2 saddle
centre bifurcations occurring simultaneously. On one side of the critical pa-
rameter value, there are 3 coexisting equilibria, while on the other side there
is only one. In fact there are two types of pitchfork: a supercritical pitchfork
involves two centres collapsing into a central saddle, leaving a single centre,
and a subcritical pitchfork involves two saddles collapsing into a central centre,
leaving a single saddle. See Figures 5 and 6.
These results and normal forms are derived from Singularity/Catastrophe
Theory.
J. Montaldi, Relative equilibria and conserved quantities . . . . 262
λ < 0λ= 0 λ > 0
H(x, y) = 1
2y2+x4+λx2+ h.o.t.
Figure 5. Supercritical Pitchfork
λ < 0λ= 0 λ > 0
H(x, y) = 1
2y2x4λx2+ h.o.t.
Figure 6. Subcritical Pitchfork
4.2 Higher degrees of freedom
From the point of view of bifurcations of equilibria, or of critical points of
the Hamiltonian, the results for 1 degree of freedom carry over to higher
dimensions, by adding a sum of quadratic terms in the other variables. So for
example the saddle-centre bifurcations has as “normal form”,
Hλ(x,y) = 1
2y2
1+1
3x3
1+λx1+1
2Pj>1(±x2
j±y2
j) + h.o.t.
Whether any of the equilibria are stable depends of course on the signs of the
quadratic terms.
On the other hand, it is a much more subtle question as to whether any
of the associated dynamics, such as the heteroclinic connections, survive this
J. Montaldi, Relative equilibria and conserved quantities . . . . 263
passage into higher dimensions. For the saddle-centre bifurcation, see [22] and
for the time-reversible case, see the lectures of Eric Lombardi in this volume,
and for more detail [39].
5 Geometric Bifurcations
In this section we discuss bifurcations of the families of relative equilibria
(res) due to degenerations in the geometry of the momentum map. One
understands fairly well now the geometry of the family of relative equilibria
in the neighbourhood of a point where the reduced phase spaces change in
dimension, which occurs at special values of the momentum map, provided
however that the group action on the phase space is (locally) free, so that
the momentum map is submersive. However, the general structure of relative
equilibria near points with continuous isotropy is not so well understood,
although some recent progress has been made.
Notation Throughout the theorems below, we will suppose that p P is a
point on a non-degenerate re, that the angular velocity of this re is ξand the
momentum value is µ. Recall that an re is said to be non-degenerate if the
Hessian d2
Hξrestricted to the symplectic slice is a non-degenerate quadratic
form.
In an important paper [60], George Patrick investigates the structure
of the set of relative equilibria as quoted in the following theorem. He also
studied the nearby dynamics and introduced the notion of drift around relative
equilibria in terms of the linearized vector field there, but we will not be
describing that aspect here.
Theorem 5.1 (Patrick [60]) Assume Gis compact, Gpis finite and Gξ
Gµis a maximal torus. Then in a neighbourhood of pthe set of relative
equilibria forms a smooth symplectic submanifold of Pof dimension dim(G)+
rank(G).
Note that, as matrices, since ξand µcommute they are simultaneously
diagonalizable. Consequently, they are both contained in a common maximal
torus, so that GξGµalways contains a maximal torus. The condition of the
theorem is therefore a generic condition. For example, for SO(3) the condition
is satisfied if and only if ξand µare not both zero. This theorem has been
refined by Patrick and Mark Roberts [62], where they show that assuming
a generic transversality hypothesis and that as before Gpis finite, the set of
relative equilibria near pis a stratified set, with the strata corresponding to
the conjugacy class of the group GξGµ.
J. Montaldi, Relative equilibria and conserved quantities . . . . 264
The following result is more of a bifurcation theorem, as it is aimed at
counting the number of relative equilibria on each reduced phase space, near a
given non-degenerate re. Recall that, if µis a regular point for the coadjoint
action and Gpis trivial, then by Arnold’s theorem (Theorem 3.7) there is a
unique re on each nearby reduced space.
Theorem 5.2 (Montaldi [48]) Assume Gµis compact and Gptrivial.
Then for regular µnear µthere are at least w(Gµ)res on the reduced space
Pµ, where w(Gµ)is the order of the Weyl group of Gµ. Furthermore, if ξis
regular then there are precisely this number of relative equilibria on Pµ. (For
non-regular µsee [48].)
Note that if Gµis a torus, then w(Gµ) = 1, and this result reduces
to Arnold’s in the case that Gis compact. The lower bound of w(G) for
general ξfollows from the Morse inequalities on the coadjoint orbits, and
so presupposes that the res on Pµare all non-degenerate. Without that
assumption one can use the Lyusternik-Schnirelman category giving a lower
bound of 1
2dim(Oµ) + 1.
The proof of this result relies on the local normal form of Marle [42] and
Guillemin-Sternberg [28]. Here I will outline the idea of the proof in the case
that µ= 0. The idea is to use the reduced Hamiltonian rather than the
augmented Hamiltonian Hξ. So, the relative equilibria in question are critical
points of the Hamiltonian restricted to Pµ, and near pone has
Pµ P0× Oµ P0×g.
The reduced space P0can be identified with the symplectic slice N1(see §2.5),
so that by hypothesis p P0=P0× {0}is a non-degenerate critical point of
the restriction of Hto P0. Write coordinates (y, ν ) P × g. Then for each
ν, the function H(·, ν) has an isolated non-degenerate critical point y=y(ν).
Define h:gRby
h(ν) = H(y(ν), ν).
Then one can show that the restriction hµof hto Oµhas a critical point
at νiff H|Pµhas a critical point at (y(ν), ν). In this manner, the problem
is reduced to finding critical points of the restrictions of a smooth function
hto coadjoint orbits Oµ, and then one can use Morse theory or Lyusternik-
Schnirelman techniques.
The above proof assumes that Gpis trivial. However, if Gpis finite,
then the proof can be modified to show that the nearby relative equilibria
correspond to critical points of a smooth Gp-invariant function h, constructed
in the same manner as above, but on P0×grather than on the full orbit
space P0×Gpg. This time though, different critical points correspond to
J. Montaldi, Relative equilibria and conserved quantities . . . . 265
the same relative equilibrium if they lie in the same Gp-orbit on Oµ. This
argument gives the following ‘equivariant’ version of Theorem 5.2.
Theorem 5.3 (Montaldi, Roberts [50]) If Gpis finite, then it acts on
nearby coadjoint orbits Oµ. If a subgroup Σof Gpis such that Fix(Σ,Oµ)
consists of isolated points, then they are all relative equilibria.
This result uses the fact that His Gp-invariant, and not that Gpacts
symplectically. In [50] this bifurcation result is applied to finding relative
equilibria of molecules, and in [37] it is applied to finding relative equilibria
of systems of point vortices on the sphere, and in both we use antisymplectic
symmetries of Has well as symplectic ones. An action of Gwhere the elements
act either symplectically or antisymplectically, i.e. gω=±ω, is said to be
semisymplectic [51].
In [50], the stability of the bifurcating relative equilibria is also calculated
using these methods. The reader should also see [65,36,62,57] for further
developments.
6 Examples
Let us look at three examples of symmetric Hamiltonian systems.
Point vortices on the sphere.
Point vortices in the plane.
Molecules (as classical mechanical systems).
The motivation for choosing these models is that the first is relatively
simple: it has no points where the action fails to be free (unless there are only
2 point vortices), and the group is compact. The bifurcations that arise are
therefore of the types described in Section 5.
The second is similar, except that the group of symmetries is no longer
compact.
The study of the classical mechanics of molecules is also of interest in
molecular spectroscopy, where it is common practice to label particular quan-
tum states in terms of the corresponding classical behaviour. We treat this
example extremely briefly!
6.1 Point vortices on the sphere
The model is a finite set of point vortices on the unit sphere in 3-space. A
point vortex is an infinitesimal region of vorticity in a 2-dimensional fluid
flow, though we ignore the fluid, and just concentrate on the vortices. The
equations of motion for this system were obtained by V.A. Bogomolov [21].
A study of the dynamics of 3 point vortices has been carried out by Kidambi
J. Montaldi, Relative equilibria and conserved quantities . . . . 266
and Newton [30] and by Pekarsky and Marsden [63]. The case of Nidentical
point vortices has been treated in [37].
If x1,...,xNare the distinct locations of these point-vortices (unit vectors
in R3), then the differential equation describing their motion is
˙xj=X
k6=j
λk
xk×xj
1xk·xj
.
Here λ1,...,λNare the strengths of the vortices: each λjis a non-zero real
number. It turns out that this vector field is Hamiltonian, with
H(x) = 1
4πX
j<k
λjλklog (1 xj·xk),
and Poisson structure
{f, g }(x) = X
j
λ1
jdjf×djg·xj,
where djf=djf(x)R3is the differential of fwith respect to the point
xjS2, and x= (x1,...,xN).
The phase space is P=S2×S2×... ×S2\∆, where is the ‘big
diagonal’ where at least one pair of points coincides, which is removed to
avoid collisions. The symplectic form on Pis given by
ω=λ1ω1 · ·· λNωN,
where ωjis the standard area form on the jth copy of S2.
This system has full rotational symmetry G=SO(3), and hence a 3-
component conserved quantity. After identifying so(3) with R3as usual, this
momentum map is the so-called centre of vorticity:
Φ(x) = X
j
λjxj.
There are a number of immediate general consequences for this system
that can be drawn from the Hamiltonian structure. For example, if all the
vorticities are of the same sign, then as x in P, so H(x)+and
it follows that Hattains its minimum at some point; this point is necessar-
ily an equilibrium point, and the set on which this minimum is attained is
Lyapounov stable (one would expect this set to be a finite union of SO(3)-
orbits). Moreover, for any µso(3)R3the same argument can be applied
on Φ1(µ), and the minimum on that set is necessarily a relative equilibrium,
by Proposition 3.2.
J. Montaldi, Relative equilibria and conserved quantities . . . . 267
However, if the vorticities are of mixed signs, then there is no general
statement about the existence of equilibria or relative equilibria, except for
N= 3 where all res are known (see below).
If there are some identical vortices, then there is an extra finite symmetry
group—a subgroup of the permutation group SN. This extra finite symmetry
group is used to considerable effect in [37], from which Figures 7 and 8 are
taken. Moreover, the time-reversing symmetries obtained from the reflexions
in O(3) are also used together with the principle of symmetric criticality to
prove the existence of many types of relative equilibria.
For example, let Chdenote the group of order 2 generated by reflexion in
the horizontal plane. Then xFix(Ch) if all the vortices are on the equator.
An application of the arguments above shows that if all the vorticities are of
the same sign then there is a point on Fix(Ch,P) where the restriction of the
Hamiltonian attains its minimum, and by the principle of symmetric critical-
ity (see §1.5) it follows that this is an equilibrium point for the full system
(though no longer a local minimum in general). And the same extension to this
argument as before provides relative equilibria in Fix(Ch,Φ1(µ)). In general
there will be several relative equilibria in each component of Fix(Ch,Φ1(µ)).
However, it is shown in [37] that if all the vorticities are of the same sign then
in each component of Fix(Ch,P) there is a unique equilibrium point.
2 vortices If there are only 2 point vortices, then every solution is a relative
equilibrium. Indeed, if Φ(x) = µ6= 0 then the two points rotate at the same
angular velocity about the axis containing µ. The only possible case where
µ= 0 is if λ1=λ2and x1=x2which is an equilibrium point.
3 vortices There have been two recent studies of the system of 3 point
vortices, by Kidambi and Newton [30] (who also treat the 2 vortex case in
an appendix) and by Pekarsky and Marsden [63]. The former describes not
only the relative equilibria, but also self-similar collapse, where triple collision
occurs in finite time with the three vortices retaining their same shape up to
similarity. The latter describes the res and their stability using the techniques
described in these lectures (many of which were in fact developed by Marsden
and co-workers).
One of the results of Kidambi and Newton is that there are two classes
of re: those lying on a great-circle and those lying at the vertices of an
equilateral triangle, and any re belongs to one of these classes. Moreover,
every equilateral triangle on the sphere is an re, as we shall prove below.
The following results are mostly taken from [30] and [63], and some are
discussed below. Denote λ1λ2+λ2λ3+λ3λ1by σ2(λ).
Proposition 6.1 For the system of N= 3 point vortices on the sphere,
J. Montaldi, Relative equilibria and conserved quantities . . . . 268
C (R)
3v C (R,p)
2v
Figure 7. Relative equilibria for 3 identical vortices on the sphere
(i) There exist equilibria iff σ2(λ)>0, and they always lie on a great circle.
(ii) All equilateral configurations are res, and they are Lyapounov stable
modulo SO(2) if σ2(λ)>0, and are unstable if σ2(λ)<0.
(iii) Self-similar collapse occurs iff σ2(λ) = 0.
(iv) The configurations where the vortices lie on a great circle, at the vertices
of a right-angled isosceles triangle is always an re. Furthermore, the triangle
with x1at the right-angle is Lyapounov stable provided
λ2
2+λ2
3>2σ2(λ).
The phase space is of dimension 6 in this case, and so the orbit space
P/SO(3) is of dimension 3, and points in the orbit space correspond to the
shapes of the triangle formed by the 3 vortices. The obvious set of coordinates
consisting of the three pairwise distances has a problem for great circle con-
figurations since nearby such a configuration these distances do not determine
the configuration uniquely. Indeed, these three distances are O(3) invariants,
as they do not distinguish the orientation, and configurations on a great circle
have non-trivial isotropy for the O(3)-action so it is not surprising that these
coordinates have a problem there. For a good set of SO(3)-invariants, one
must use the oriented volume as well, which is what is done in the two works
cited above.
However, the three distances do form a good set of coordinates away from
the great circle configurations, and we will restrict our attention to those. So,
let r1be the chord distance kx2x3ketc. Then the Hamiltonian and the
orbit momentum map (2.16) are given by
H(r1, r2, r3) = λ1λ2
2πlog(r3)λ2λ3
2πlog(r1)λ3λ1
2πlog(r2)
ϕ(r1, r2, r3) = |Φ|2= (λ1+λ2+λ3)2λ1λ2r2
3λ2λ3r2
1λ3λ1r2
2
(6.1)
J. Montaldi, Relative equilibria and conserved quantities . . . . 269
The reduced spaces are then Pµ:= ϕ1(µ2). The relative equilibria
are determined by the critical points of the restriction of Hto the reduced
spaces, and are therefore critical points of Hηϕ, for some η(the Lagrange
multiplier). A short calculation shows that the only solutions to this equation
are
r1=r2=r3=r, say, and η= 1/(4πr2).
Thus all relative equilibria away from great circle configurations are equilateral
triangles. Note that the relation between the angular velocity ξ(Proposition
3.2) and ηis simply ξ= 2ηΦ(x), since
0 = d(Hηϕ) = d(Hη|Φ|2) = dH 2ηΦ.dΦ.
Of course, Φ 6= 0 since the vortices do not all lie on a great circle.
For the stability of these equilateral res, we first calculate the Hessian of
Hηϕ (this is in fact Arnold’s “energy-Casimir” method):
d2(Hηϕ)(r) = 4
r2
λ2λ30 0
0λ3λ10
0 0 λ1λ2
.
Now, the tangent space to ϕ1(µ2) is spanned by (λ1,λ2,0),(λ1,0,λ3),
and a computation then shows that the Hessian of the reduced Hamiltonian
is definite if and only if σ2(λ)>0.
The computations for the configurations of vortices lying on great circles
are longer, and I will not go into them further here; details can be found in
the original papers [30,63].
Remarks 6.2 (i) The case σ2(λ) = 0 remains to be understood.
(ii) It is not known whether any other form of collapse (i.e. not self-similar)
can occur.
(iii) A further bifurcation occurs at equilateral res which lie on a great
circle which to my knowledge has not been investigated.
4 vortices Much less is known in general about the case of 4 vortices. It is
shown in [63] that the configuration where the four vortices lie at the vertices
of a regular tetrahedron is always a relative equilibrium. It is also easy to show
(from the equations of motion) that a square lying in a great circle is always
an re, independently of the values of the vorticities. However, in contrast to
the 3-vortex case, squares not lying in great circles are not res unless all the
vortices are identical.
In the case that all the vorticities coincide, a classification of symmetric
res is given in [37], from which Figure 8 is taken. Moreover, using the implicit
J. Montaldi, Relative equilibria and conserved quantities . . . . 270
C (R,R’)
2v C (2R)
2v
C (R,p)
3v
C (R,2p)
2v
C (R)
4v
Figure 8. Relative equilibria for 4 identical vortices on the sphere
function theorem one can show that many of the res shown to exist in [37]
persist under small perturbations of the Hamiltonian, and so in particular
under small changes of the values of the vorticities. There is one interesting
occurrence of a symmetric pitchfork bifurcation: consider the family of square
configurations, on the co-latitude θ—type C4v(R) in the figure. When the
ring is close to the North pole (say), the re is stable. As θincreases, one
pair of eigenvalues approaches 0, and at θ= arccos(1/3) there is a pitchfork
bifurcation (of type I in the terminology of §4.1). The bifurcating pair of
relative equilibria are of type C2v(R, R).
Stability of a ring of vortices It was shown by Dritschel and Polvani [64]
that the stability of a single ring of identical vortices depends on the latitude.
They show that if θis the angle subtended by any of the vortices with the axis
of symmetry of the ring (the colatitude), then the configuration of a regular
J. Montaldi, Relative equilibria and conserved quantities . . . . 271
ring of Nidentical vortices is linearly stable as follows
Nrange of stability Nrange of stability
3 all θ4 cos2θ > 1/3
5 cos2θ > 1/2 6 cos2θ > 4/5
while for N > 6 the ring is never stable. See also [38], where it is shown that
the rings are not only linearly stable but Lyapounov stable.
Bifurcations The changes in stability that occur in the table above involve
bifurcations of the relative equilibria, and indeed the supercritical pitch-
fork bifurcation described in Section 4, since they all involve a loss of C2
symmetry. Consider for example the case of N= 4 identical vortices. A
single ring (square) near the pole is a Lyapounov stable relative equilib-
rium. As θcos1(1/3), so one of the eigenvalues tends to 0, and for
θ > cos1(1/3), there appears a new family of relative equilibria consisting
of vortices alternately above and below the vortices in the now-unstable square
configuration. These bifurcating res—denoted C2v(R, R) in [37]—are then
Lyapounov stable. Other stability transitions have been observed in [37,38],
but the corresponding bifurcations have not been studied.
The other type of bifurcation that occurs in this problem is the geometric
bifurcation due to the different geometry of the reduced spaces for µ= 0 and
µ6= 0—see Section 5. Consider a relative equilibrium peon µ= 0, for example
the ring of Nidentical vortices on the equator. This corresponds to a point
with symmetry DN h in the phase space (the dihedral group in the equatorial
plane together with inversion in that plane). Nearby reduced spaces are then
locally of the form Pµ P0× Oµ, where Oµis the coadjoint orbit through
µ, which here is a sphere, as described very briefly in Section 5. The relative
equilibria on Pµnear peare the critical points of some function h:OµR,
and moreover this function is invariant under some action of Dnh DN×C2
on Oµ. An analysis of this action shows that there must be critical points with
symmetry of types CN v and C2v; see Figure 9. The corresponding relative
equilibria have configurations of types CN v (R) (a regular ring) and if N= 2m
then C2v(mR) and C2v((m1)R, 2p), while if N= 2m+ 1 then C2v(mR, p).
Here the configuration C2v(mR, ℓp) consists of mpairs and poles all lying
on a common great circle, while the great circle is rotating rigidly about an
axis containing the poles. See Figure 8, and see [37] for details.
6.2 Point vortices in the plane
This system has a much older history than the model of vortices on the sphere,
going back to Helmoltz and Kirchhoff, and has been studied by many people
J. Montaldi, Relative equilibria and conserved quantities . . . . 272
Figure 9. Level sets of a typical function on a sphere with symmetry D3h, showing half of
the 8 critical points
since; for reviews see [17,3,15]. Consider an ideal fluid in the plane whose
vorticity is concentrated in Npoint vortices, of strengths λ1,...,λN. These
points move according to the differential system
˙
zj=1
2πi X
k6=j
λk
zjzk
,
where zjis a complex number representing the position of the j-th vortex,
after identifying the plane with C, see (1.3). The Hamiltonian for this system
is
H(z) = 1
4πX
j<k
λjλklog |zjzk|2.
The symmetry here is the group of 2-D Euclidean motions SE(2)—which is
not compact. The corresponding conserved quantity (momentum map) is
Φ(z1,...,zN) = iX
j
λjzj,1
2Pλj|zj|2.(6.2)
From the geometric point of view, this is interesting because if the total
vorticity Λ Pjλjis non-zero, the coadjoint action on se(2)must be mod-
ified in order that the momentum map be equivariant (see Section 2). If we
identify SE(2) with CU(1) and u(1) with R, then the modified coadjoint
action (2.7) becomes
CoadΛ
(u,θ)(ν, ψ) = e ν, ψ +(eν¯u)+ Λ iu, 1
2|u|2(6.3)
where (z) is the imaginary part of z.
J. Montaldi, Relative equilibria and conserved quantities . . . . 273
If Λ = 0 then the orbits are points on the ψ-axis, and cylinders around
that axis, while if Λ 6= 0 the orbits are all paraboloids—see Examples 2.5 and
2.7 respectively and Figures 2 and 3. Indeed, one can show that the modified
coadjoint orbits are given by the level sets of f:C×RRdefined by
f(ν, ψ) = |ν|2ψ, (6.4)
at least if Λ 6= 0. If Λ = 0 then the non-zero level sets are the cylindrical
orbits, but the zero level set is the whole ψ-axis.
2 vortices in the plane As in the case of 2 vortices on the sphere, here they
are also always relative equilibria. It is simple to prove from the differential
equation that if Λ λ1+λ26= 0 then the two point vortices rotate about the
fixed point (λ1z1+λ2z2)/Λ. On the other hand, if Λ = 0 then they translate
together towards infinity, in the direction orthogonal to the segment joining
them. These two types of motion are both relative equilibria.
From a geometrical point of view, the reduced spaces are all just single
points, so the corresponding motions are indeed all relative equilibria, and
furthermore the relative equilibria are trivially extremal, and so provided Gµ
is compact (ie. Λ 6= 0, so GµSO(2)) they are stable modulo Gµ.
3 vortices in the plane The classical work on three planar point vortices is
a beautiful paper by J.L. Synge [69]. We approach this problem as we did
for three point vortices on the sphere. That is, points in the quotient of the
phase space (R2)3C3by SE(2) correspond to shapes of oriented triangles.
Again we ignore the orientation, which only causes problems near collinear
configurations. Then a point in the quotient space is determined by the three
lengths (r1, r2, r3), and on that space
4πH(r1, r2, r3) = λ1λ2log(r3)λ2λ3log(r1)λ3λ1log(r2)
ϕ(r1, r2, r3) = λ1λ2r2
3λ2λ3r2
1λ3λ1r2
2,(6.5)
the second is just fΦ. It is remarkable that apart from constant term in ϕ,
these are identical to equations (6.1) for 3 point vortices on the sphere.
The non-collinear relative equilibria are given by the critical points of H
restricted to the level-sets of ϕ, and the computation has already been done.
Thus the relative equilibria are again equilateral triangles, of side rsay, with
Lagrange multiplier η= 1/(4πr2) again. Note however, that this time the
relation between ηand the “angular velocity” ξis not so simple:
0 = d(Hηϕ)(p) = d(Hηf Φ)(p) = dH (p)η df(µ)dΦ(p),
so that ξ=η df (µ). Thus if Φ(p) = (ν, ψ) then
ξ= 2η(ν, Λ).
J. Montaldi, Relative equilibria and conserved quantities . . . . 274
One consequence of this expression for ξis that if Λ = 0 then the “angu-
lar velocity” is in fact rectilinear motion, with constant velocity ξ= 2ην =
(1/2πr2)ν, where ν=iPjλjzj, which in modulus is independent of the rep-
resentative triangle. Note that if the 3 vortices are not collinear, then ν6= 0
so that the function fseparates the relevant coadjoint orbits even in the case
Λ = 0. On the other hand, if Λ 6= 0, and η6= 0, then the relative equilibrium
is a periodic orbit.
Furthermore, the Lyapounov stabilities modulo Gof these equilateral
relative equilibria are the same as for the spherical case.
The symplectic reduction for a particular class of planar 4-vortex prob-
lems has recently been considered by Patrick [61]. In particular he treats the
case Λ = 0 so that the momentum map is coadjoint-equivariant; it would be
interesting to see how the results are affected by changing to Λ 6= 0.
6.3 Molecules
Consider a molecule consisting of Natoms. The Born-Oppenheimer approxi-
mation consists of ignoring the movement of the electrons (which is reasonable
as they are so light). The system then has 3Ndegrees of freedom, the 3 di-
rections of motion for each nucleus, or 3N3 after fixing the centre of mass.
The rotational symmetry of the system gives rise to the conservation of
angular momentum J. If we put J= 0, then there are 3N6 degrees of
freedom which describes the shape of the molecule, and correspond to the
vibrational motions. In other words, the J= 0 reduced space is of dimension
6N12. The simplest motions beyond the equilibria are the periodic orbits,
which near the stable equilibrium are given by Lyapounov’s theorem and
its generalizations. For J6= 0, the motion has a rotational aspect, and the
simplest type of motion is the relative equilibrium. Beyond that are motions
that are a combination of rotations and vibrations, so-called rovibrational
states. The reduced spaces for J6= 0 are of dimension 6N10.
Consider now the simplest interesting case of a triatomic molecule. The
reduced spaces are of dimension 6 (for J= 0) or 8 (for J6= 0). There is
one complication that we will not discuss here, namely the reduced space for
J= 0 is singular at points corresponding to collinear equilibria. Consider then
an equilibrium of a triatomic molecule which is not collinear. The geometric
bifurcation methods discussed in Section 5, and at the end of §6.1, show that
there are relative equilibria which bifurcate from the equilibria, and in fact
there at at least 6 such families of re parametrized by kJkand corresponding
to critical points of functions on a sphere. The stabilities of these bifurcating
families are discussed in [50]. For a complete investigation into the relative
J. Montaldi, Relative equilibria and conserved quantities . . . . 275
equilibria of a specific molecule with 3 identical atoms—namely H+
3—see [33].
Another interesting example is the tetra-atomic molecule ammonia NH3.
This has an equilibrium where the three hydrogen atoms form an equilateral
triangle, and the nitrogen atom is slighly above (or below) the centre of the
triangle—so that the molecule is nearly planar. The analysis is similar to the
triatomic case, with 6 families of res that bifurcate from each equilibrium.
The stability analyses should also be similar to the triatomic case. However,
the presence of two very close stable equilibria (with the nitrogen atom on
one side and on the other of the hydrogen-plane) and an unstable planar
equilibrium betewen them suggests that there will be further bifurcations. An
interesting “semilocal” analysis could be obtained by adding a new parameter
λso that the equilibria undergo a pitchfork bifurcation (of type I) at λ= 0,
with the genuine system corresponding to say λ=1 and an artificial one
for λ > 0 with the symmetric planar equilibrium being stable. One needs
to investigate how the pitchfork bifurcation within the reduced space P0,
obtained by varying λ, couples with the geometric bifurcation, obtained by
varying kJk.
References
Books
1. R. Abraham and J. Marsden. Foundations of Mechanics. Benjamin-
Cummings, 1978.
2. V.I. Arnold. Mathematical Methods of Classical Mechanics. Springer-
Verlag, 1978.
3. V.I. Arnold and B.A. Khesin, Topological Methods in Hydrodynamics,
Springer-Verlag, New York, 1998.
4. V.I. Arnold, V.V. Kozlov and A.I. Neishtadt. Mathematical Aspects of
Classical and Celestial Mechanics, 2nd edition. Springer-Verlag, 1997.
5. P. Chossat and R. Lauterbach, Methods in Equivariant Bifurcation The-
ory and Dynamical Systems. World Scientific, to appear 2000.
6. R.H. Cushman and L.M. Bates, Global Aspects of Classical Integrable
Systems. Birkhuser Verlag. 1997.
7. M. Golubitsky, I. Stewart and D. Schaeffer. Singularities and Groups in
Bifurcation Theory, Vol. II. Springer-Verlag, New York. 1988.
8. V. Guillemin and S. Sternberg Symplectic Techniques in Physics. Cam-
bridge University Press. 1984.
9. V. Guillemin, E. Lerman and S. Sternberg Symplectic Fibrations and
Multiplicity Diagrams. Cambridge University Press. 1996.
J. Montaldi, Relative equilibria and conserved quantities . . . . 276
10. P. Libermann and C.-M. Marle. Symplectic Geometry and Analytical
Mechanics. Reidel, 1987.
11. R. MacKay and J. Meiss. Hamiltonian Dynamical Systems, a reprint
collection. Adam Hilger, Bristol. 1988.
12. J. Marsden. Lectures on Mechanics. L.M.S. Lecture Note Series 174,
Cambridge University Press, 1992.
13. J. Marsden and T. Ratiu. Introduction to Mechanics and Symmetry.
Springer-Verlag, New-York, 1994. [Second ed. 1999]
14. K. Meyer and G. Hall. Hamiltonian Systems and the N-Body Problem.
Springer-Verlag, New-York, 1992.
15. P.G. Saffman, Vortex Dynamics. Cambridge University Press, Cam-
bridge, 1992.
16. J.-M. Souriau Structure des Syst`emes Dynamiques. Dunod, Paris, 1970.
[English translation: Structure of Dynamical Systems: A Symplectic View
of Physics, Birkhauser, Boston, 1997.]
Research papers
17. H. Aref, Integrable, chaotic and turbulent vortex motion in two-
dimensional flows, Ann. Rev. Fluid Mech. 15 (1983), 345–389.
18. J.M. Arms, A. Fischer and J.E. Marsden, Une aproche symplectique pour
des th´eor`emes de ecomposition en eom´etrie ou relativit´e en´erale. C.
R. Acad. Sci. Paris 281 (1975), 517 520.
19. J.M. Arms, J.E. Marsden and V. Moncrief, Bifurcations of momentum
mappings. Comm. Math. Phys. 78 (1981), 455–478.
20. L.M. Bates and E. Lerman, Proper group actions and symplectic strati-
fied spaces. Pacific J. Math. 181 (1997), 201–229.
21. V.A. Bogomolov, Dynamics of vorticity at a sphere, Fluid Dynamics 6
(1977), 863–870.
22. H. Broer, S.-N. Chow, Y. Kim and G. Vegter, A normally elliptic Hamil-
tonian bifurcation. Z. Angew. Math. Phys. 44 (1993), 389–432.
23. M. Dellnitz, I. Melbourne and J. Marsden, Generic bifurcation of Hamil-
tonian vector fields with symmetry. Nonlinearity 5(1992), 979–996.
24. J.J. Duistermaat, Bifurcations of perodic solutions near equilibrium
points of Hamiltonian systems. In Bifurcation Theory and Applications,
Montecatini, 1983 (ed. L. Salvadori), LNM 1057, Springer, 1984.
25. F. Fass`o and D. Lewis, Stability properties of the Riemann ellipsoids.
Preprint, 2000.
26. I.M. Gelfand and L.D. Lidskii, On the structure of stability of linear
Hamiltonian systems of differential equations with periodic coefficients.
J. Montaldi, Relative equilibria and conserved quantities . . . . 277
Usp. Math. Nauk. 10 (1955), 3–40. (English translation: Amer. Math.
Soc. Translations (2) 8(1958), 143–181.)
27. M. Golubitsky and I. Stewart, Generic bifurcations of Hamiltonian sys-
tems with symmetry. Physica D 24 (1987), 391–405.
28. V. Guillemin and S. Sternberg, A normal form for the moment map. In
Differential Geometric Methods in Mathematical Physics (S. Sternberg
ed.) Mathematical Physics Studies, 6. D. Reidel Publishing Company
(1984).
29. G. Iooss and M.-C. erou`eme, Perturbed homoclinic solutions in re-
versible 1:1 resonance vector fields. J. Differential Equations 102 (1993),
62–88.
30. R. Kidambi and P. Newton, Motion of three point vortices on a sphere,
Physica D 116 (1998), 143–175.
31. Y. Kimura, Vortex motion on surfaces with constant curvature. Proc. R.
Soc. Lond. A 455 (1999), 245–259.
32. F.C. Kirwan, The topology of reduced phase spaces of the motion of
vortices on a sphere, Physica D 30 (1988), 99–123.
33. I. Kozin, R.M. Roberts and J. Tennyson, Symmetry and structure of
rotating H+
3.Preprint, University of Warwick, 1999.
34. E. Lerman, J. Montaldi and T. Tokieda, Persistence of extremal relative
equilibria. In preparation
35. E. Lerman and S.F. Singer, Relative equilibria at singular points of the
momentum map. Nonlinearity 11 (1998), 1637-1649
36. E. Lerman and T.F. Tokieda, On relative normal modes. C. R. Acad.
Sci. Paris Sr. I 328 (1999), 413-418.
37. C.C. Lim, J. Montaldi and R.M. Roberts, Systems of point vortices on
the sphere. Preprint, INLN, 2000.
38. C.C. Lim, J. Montaldi and R.M. Roberts, Stability of relative equilibria
for point vortices on the sphere. In preparation.
39. E. Lombardi, Oscillatory integrals and phenomena beyond any alge-
braic order; with applications to homoclinic orbits in reversible systems.
Springer Verlag Lecture Notes in Mathematics. To appear.
40. A.M. Lyapounov, Probl`eme en´erale de la stabilit´e du mouvement. Ann.
Fac. Sci. Toulouse 9, (1907). (Russian original: 1895.)
41. R. MacKay, Stability of equilibria of Hamitonian systems. Nonlinear
Phenomena and Chaos (1986), 254–70. Also reprinted in [11].
42. C.-M. Marle, Mod`ele d’action hamiltonienne d’un groupe the Lie sur une
vari´et´e symplectique. Rend. Sem. Mat. Univers. Politecn. Torino 43
(1985), 227–251.
43. J. Marsden and A. Weinstein, Reduction of symplectic manifolds with
J. Montaldi, Relative equilibria and conserved quantities . . . . 278
symmetry. Pep. Math. Phys 5(1974), 121–130.
44. J.-C. van der Meer, The Hamiltonian Hopf bifurcation, Lecture Notes in
Math. 1160, Springer, 1985.
45. I. Melbourne, Versal unfoldings of equivariant linear Hamiltonian vector
fields. Math. Proc. Camb. Phil. Soc. 114 (1993), 559–573.
46. I. Melbourne and M. Dellnitz, Normal forms for linear Hamiltonian vector
fields commuting with the action of a compact Lie group. Math. Proc.
Camb. Phil. Soc. 114 (1993), 235–268.
47. K. Meyer, Symmetries and integrals in mechanics. Dynamical Systems
(M. Peixoto, ed.), 259–273. Academic Press, New York, 1973.
48. J. Montaldi, Persistence and stability of relative equilibria. Nonlinearity
10 (1997), 449–466.
49. J. Montaldi, Perturbing a symmetric resonance: the magnetic spheri-
cal pendulum. In SPT98 Symmetry and Perturbation Theory II, (A.
Degasperis and G. Gaeta eds.), World Scientific 1999.
50. J. Montaldi and R.M. Roberts, Relative equilibria of molecules. J. Non-
lin. Sci. 9(1999), 53–88.
51. J. Montaldi and R.M. Roberts, A note on semisymplectic actions of Lie
groups. In preparation.
52. J. Montaldi, R.M. Roberts and I. Stewart, Periodic Solutions near Equi-
libria of Symmetric Hamiltonian Systems. Proc. Roy. Soc. London 325
(1988), 237–293.
53. J. Montaldi, R.M. Roberts and I. Stewart, Existence of nonlinear normal
modes of symmetric Hamiltonian systems. Nonlinearity 3(1990), 695–
730.
54. J. Moser, Lectures on Hamiltonian Systems. Memoirs of the A.M.S. 81
(1981). (Also reprinted in [11]).
55. J. Moser, Periodic orbits near equilibrium and a theorem by Alan Wein-
stein. Communs. Pure Appl. Math. 29 (1976), 727–747.
56. J.P. Ortega, Symmetry, Reduction ans Stability in Hamilonian Systems.
Thesis, University of California, Santa Cruz, 1988.
57. J.P. Ortega and T.S. Ratiu, Stability of Hamiltonian relative equilibria.
Nonlinearity 12 (1999), 693–720.
58. R. Palais. The principle of symmetric criticality, Commun. Math. Phys.
69 (1979), 19–30.
59. G.W. Patrick, Relative equilibria in Hamiltonian systems: The dynamic
interpretation of nonlinear stability on the reduced phase space. J. Geom.
Phys. 9(1992), 111-119.
60. G.W. Patrick, Relative equilibria of Hamiltonian systems with symmetry:
linearization, smoothness and drift. J. Nonlin. Sci. 5(1995), 373–418.
J. Montaldi, Relative equilibria and conserved quantities . . . . 279
61. G.W. Patrick, Reduction of the planar 4-vortex system at zero momen-
tum. Preprint math-ph/9910012
62. G.W. Patrick and R.M. Roberts, The transversal relative equilibria of
Hamiltonian systems with symmetry. Preprint, University of Warwick
(1999).
63. S. Pekarsky and J.E. Marsden, Point vortices on a sphere: Stability of
relative equilibria, J. Mathematical Physics 39 (1998), 5894–5906.
64. L.M. Polvani and D.G. Dritschel, Wave and vortex dynamics on the sur-
face of a sphere, J. Fluid Mech. 255 (1993), 35–64.
65. R.M. Roberts and M.E.R. Sousa Dias, Bifurcations from relative equilib-
ria of Hamiltonian systems. Nonlinearity 10 (1997), 1719–1738.
66. R.M. Roberts and M.E.R. Sousa Dias, Symmetries of Riemann ellipsoids.
To appear in Resenhas do IME - USP, 2000.
67. R. Sjamaar and E. Lerman, Stratified symplectic spaces and reduction.
Ann. Math 134 (1991), 375–422.
68. S. Smale, Topology and mechanics I, II. Invent. Math. 10 (1970), 305–
331, and 11 (1970), 45–64.
69. J.L. Synge, On the motion of three vortices. Can. J. Math. 1(1949),
257–270.
70. A. Vanderbauwhede and J.C. van der Meer, A general reduction method
for periodic solutions near equilibria in Hamiltonian systems. Fields In-
stitute Comm. 4(1995), 273–294.
71. A. Weinstein, Normal modes for nonlinear Hamiltonian systems. Invent.
Math. 20 (1973), 47–57.
72. A. Weinstein, Bifurcations and Hamilton’s principle. Math. Z. 159
(1978), 235–248.
J. Montaldi, Relative equilibria and conserved quantities . . . . 280
... This approach highlights the application of Noether's theorem, which states that the components of the momentum map are preserved by the dynamics. If S is the plane, the sphere or the hyperbolic plane then the respective actions of the groups SE(2), SO (3) and SL(2) are indeed Hamiltonian (see above references and [19]). However, if the point vortices lie on a torus or cylinder then the actions of T 2 and R × S 1 respectively, while being symplectic, fail to be Hamiltonian [22]. ...
... Relative equilibria and allowed velocity vectors. We consider for the moment the general setting of a G-invariant Hamiltonian system on a symplectic manifold P. See for example [16] or [19] for definitions. For ξ ∈ g, the associated vector field on P is denoted by ξ P and, given a Hamiltonian function H : P → R, the associated vector field is denoted by X H . ...
... The symmetry groups we use in this paper are SO(3), the group of isometries of the sphere, and SO(2, 1) (or SL(2, R)), the isometries of the Lobachevsky plane (also called the hyperbolic plane), and we describe their one-parameter subgroups in the relevant section. See also (Montaldi 2000;Montaldi and Nava-Gaxiola 2014) for details about relative equilibria for these groups. ...
Article
Full-text available
We perform the reduction of the two-body problem in the two dimensional spaces of constant non-zero curvature and we use the reduced equations of motion to classify all relative equilibria (RE) of the problem and to study their stability. In the negative curvature case, the nonlinear stability of the stable RE is established by using the reduced Hamiltonian as a Lyapunov function. Surprisingly, in the positive curvature case this approach is not sufficient and the proof of nonlinear stability requires the calculation of Birkhoff normal forms and the application of stability methods coming from KAM theory. In both cases we summarize our results in the Energy-Momentum bifurcation diagram of the system.
... These equilibrium points represent the periodic solutions for the original Hamiltonian, which are called the relative equilibria. It should be noted that the equilibrium point in the reduced phase space is related to the dynamic of the corresponding relative equilibria in the full phase space by the group action [24]. ...
Article
Full-text available
A fourth-order Hamiltonian describing the planar full two body problem is obtained, allowing for a mapping out of the geography of spin–spin–orbit resonances. The expansion of the mutual potential function up to the fourth-order results in the angles to come through one single harmonic and consequently the rotation of both bodies and mutual orbit are coupled. Having derived relative equilibria, stability analysis showed that the stability conditions are independent of physical and orbital characteristics. Simultaneously chaotic motion of bodies is investigated through the Chirikov diffusion utilizing geographic information of the complete resonances. The results show that simultaneous chaos among the binary asteroids is not expected to be prevalent due to the mass distribution of primary in compare with secondary. If mass distribution of bodies is of the same order, simultaneous chaos and global instability are achievable.
... A thorough historical summary of research of these studies can be found in [1, 3, 9, 21] for the plane, and [11] for the sphere. On the other hand, the case of point vortices on the hyperbolic plane has only been treated briefly in [3, 5, 10, 17, 18] and in some greater detail in [7, 8], although none take advantage of the geometry of the conserved quantities. As on the plane and sphere, the governing equations of the system of point vortices on the hyperbolic plane are Hamiltonian. ...
Article
Full-text available
We investigate the dynamical system of point vortices on the hyperboloid. This system has non-compact symmetry $SL(2, R)$ and a coadjoint equivariant momentum map. The relative equilibrium conditions are found and the trajectories of relative equilibria with non-zero momentum value are described. We also provide the classification of relative equilibria and the stability criteria for a number of cases, focusing on 2 and 3 vortices. Unlike the system on the sphere, this system has relative equilibria with non-compact momentum isotropy subgroup, and these are used to illustrate the different stability types of relative equilibria.
... The reason for this behaviour is that the corresponding relative equilibrium cannot be generated from a single element of the natural subgroup SO(2) × SO(2) of the Lie group SO(4) that arises from the wx and yz coordinate pairs of the chosen reference frame. Nevertheless, a classical result, which claims that in a semisimple compact Lie group every element is contained in a maximal torus, [20], shows that a suitable change of coordinates leads to a reference frame in which all functions r i are constant. In particular, SO(4) is a semisimple compact Lie group, so it satisfies the above result. ...
Article
We consider the N-body problem in spaces of constant curvature and study its rotopulsators, i.e.\ solutions for which the configuration of the bodies rotates and changes size during the motion. Rotopulsators fall naturally into five groups: positive elliptic, positive elliptic-elliptic, negative elliptic, negative hyperbolic, and negative elliptic-hyperbolic, depending on the nature and number of their rotations and on whether they occur in spaces of positive or negative curvature. After obtaining existence criteria for each type of rotopulsator, we derive their conservation laws. We further deal with the existence and uniqueness of some classes of rotopulsators in the 2- and 3-body case and prove two general results about the qualitative behaviour of rotopulsators. More precisely, for positive curvature we show that there is no foliation of the 3-sphere with Clifford tori such that the motion of each body is confined to some Clifford torus. For negative curvature, a similar result is proved relative to foliations of the hyperbolic 3-sphere with hyperbolic cylinders.
Article
Vortex crystals is one name in use for the subject of vortex patterns that move without change of shape or size. Most of what is known pertains to the case of arrays of parallel line vortices moving so as to produce an essentially two-dimensional (2D) flow. The possible patterns of points indicating the intersections of these vortices with a plane perpendicular to them have been studied for almost 150 years. Analog experiments have been devised, and experiments with vortices in a variety of fluids have been performed. Some of the states observed are understood analytically. Others have been found computationally to high precision. Our degree of understanding of these patterns varies considerably. Surprising connections to the zeros of ‘special functions’ arising in classical mathematical physics have been revealed. Vortex motion on 2D manifolds such as the sphere, the cylinder (periodic strip) and the torus (periodic parallelogram) has also been studied because of the potential applications, and some results are available regarding the problem of vortex crystals in such geometries. Although a large amount of material is available for review, some results are reported here for the first time. The subject seems pregnant with possibilities for further development.
Article
Multiplicity diagrams can be viewed as schemes for describing the phenomenon of "symmetry breaking" in quantum physics. The subject of this book is the multiplicity diagrams associated with the classical groups U(n), O(n), etc. It presents such topics as asymptotic distributions of multiplicities, hierarchical patterns in multiplicity diagrams, lacunae, and the multiplicity diagrams of the rank 2 and rank 3 groups. The authors take a novel approach, using the techniques of symplectic geometry. The book develops in detail some themes which were touched on in the highly successful Symplectic Techniques in Physics by V. Guillemin and S. Sternberg (CUP, 1984) , including the geometry of the moment map, the Duistermaat-Heckman theorem, the interplay between coadjoint orbits and representation theory, and quantization. Students and researchers in geometry and mathematical physics will find this book fascinating.