Marsden, J.E. Ratiu, T.S. - Supplement
Marsden, J.E. Ratiu, T.S. - Supplement
Marsden, J.E. Ratiu, T.S. - Supplement
Second Edition
Jerrold E. Marsden
CDS 107-81
Caltech
Pasadena, CA 91125
USA
[email protected]
Tudor S. Ratiu
Departement de Mathematiques
Ecole Polytechnique Federale de Lausanne
CH - 1015 Lausanne
Switzerland
[email protected]fl.ch
Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
N6 Cotangent Bundles . . . . . . . . . . . . . . . . . . . . . . 3
N6.A Linearization of Hamiltonian Systems . . . . . . . 3
N7 Lagrangian Mechanics . . . . . . . . . . . . . . . . . . . . 9
N7.A The Classical Limit and the Maslov Index . . . . 9
N9 Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . 23
N9.A Automatic Smoothness . . . . . . . . . . . . . . . 23
N9.B Abelian Lie Groups . . . . . . . . . . . . . . . . . 25
N9.C Lie Subgroups . . . . . . . . . . . . . . . . . . . . 26
N9.D Lie’s Third Fundamental Theorem . . . . . . . . . 29
N9.E The Symplectic, Orthogonal, and Unitary Groups 31
N9.F Generic Coadjoint Isotropy Subalgebras are Abelian 40
N9.G Some Infinite Dimensional Lie Groups . . . . . . . 44
N9.G.1 Basic Facts about Sobolev Spaces and
Manifolds. . . . . . . . . . . . . . . . . . 52
N10 Poisson Manifolds . . . . . . . . . . . . . . . . . . . . . . 67
N10.A Proof of the Symplectic Stratification Theorem . . 67
N11 Momentum Maps . . . . . . . . . . . . . . . . . . . . . . 71
N11.A Another Example of a Momentum Map . . . . . . 71
N13 Lie-Poisson Reduction . . . . . . . . . . . . . . . . . . . . 73
N13.A Proof of the Lie–Poisson Reduction Theorem for
Diff vol (M ) . . . . . . . . . . . . . . . . . . . . . . 73
N13.B Proof of the Lie–Poisson Reduction Theorem for
Diff can (P ) . . . . . . . . . . . . . . . . . . . . . . 75
N13.C The Linearized Lie–Poisson Equations . . . . . . . 78
iv Contents
Preface
This supplement contains a number of topics that are somewhat periph-
eral to the main flow of the text itself, so that to keep the book within a
reasonable size, we have placed them here. This does not mean that they
are any less important, but as usual, one has to make choices, sometimes
difficult ones. We have organized the material by Chapter to match that of
the text as far as possible.
This supplement is being continually updated and we appreciate com-
ments and suggestions from readers. Please also note that you can get the
current errata for the main text from the site
http://www.cds.caltech.edu/~marsden
Jerry Marsden
Pasadena, California
Tudor Ratiu
Lausanne, Switzerland
December, 1998
2
Page 3
N6
Cotangent Bundles
∂H ∂H
q̇ i = , ṗi = − . (N6.A.3)
∂pi ∂q i
Linearizing along a solution curve (q i (t), pi (t)) and calling the new vari-
ables (δq i , δpi ) we get the equations
∂2H ∂2H
(δq i )· = δq j
+ δpj ,
∂q j ∂pi ∂pj ∂pi
∂2H ∂2H
(δpi )· = − δq j
− δpj . (N6.A.4)
∂q j ∂q i ∂q i ∂pj
The same argument and formulas hold for infinite-dimensional weak sym-
plectic vector spaces E × E , where E and E are (weakly) paired. One of
the goals of Marsden, Ratiu, and Raugel [1991] is to generalize this simple
procedure to arbitrary symplectic manifolds. Formula (N6.A.5) cannot be
correct, in general, since the second variation of a function does not make
intrinsic sense, except at critical points. Additional structure is needed to
N6.A Linearization of Hamiltonian Systems 5
correct the second variation by the addition of terms making the resulting
formula invariant.1
Examples
1 Another motivation for working in this general context is to deal with Hamiltonian
systems in Lie–Poisson spaces, which, as we explore in detail in Chapters 13 and 14,
is equivalent to G-invariant Hamiltonian systems on T ∗ G, where G is a Lie group. At
critical points of H + C, where C is a Casimir on g∗ (g∗ is the dual of the Lie algebra g
of G), such a linearization has been carried out in Holm, Marsden, Ratiu, and Weinstein
[1985] and Abarbanel, Holm, Marsden, and Ratiu [1986]; as expected, the Hamiltonian
function of the linearized equations is the second variation of H + C, but the Poisson
structure instead of being Lie–Poisson is a “frozen coefficient” Poisson bracket.
6 N6. Cotangent Bundles
∂H ∂H
H(q i , pi , v i , wi ) = v i + wj . (N6.A.7)
∂q i ∂pj
∂H0 ∂H0
H1 (q i , pi , v i , wi ) = v i + wi + H 1 (q i , pi )
∂q i ∂pi
= H0 (q i , pi , v i , wi ) + H 1 (q i , pi ), (N6.A.10)
where H0 is given in terms of H0 by (N6.A.7). Hamilton’s equations for H1
on T P relative to the symplectic form ΩT are
dq i ∂H0 dpi ∂H0
= , =− i ,
dt ∂p
dt ∂q
dv i
i
∂ ∂ ∂H ∂H 1
j 0
= v + w j + , (N6.A.11)
dt ∂q
j ∂pj ∂pi ∂pi
dwi ∂ ∂ ∂H0 ∂H 1
= − v j j + wj i
− .
dt ∂q ∂pj ∂q ∂q i
One calls this the first variation equation relative to a parameter .
If we set H 1 = 0 we recover the first variation equation (N6.A.4) for XH0
discussed earlier, with H0 = H, v i = δq i , and wi = δpi .
Further details on the linearization of Hamiltonian systems and the use
of symplectic connections to accomplish this may be found in Marsden,
Ratiu, and Raugel [1991].
8 N6. Cotangent Bundles
Page 9
N7
Lagrangian Mechanics
Lψ = Eψ, (N7.A.1)
where
2
Lψ = − ψ +Vψ (N7.A.2)
2m
10 N7. Lagrangian Mechanics
1 Of course Planck’s constant is a constant and cannot literally tend to zero, any
more than the velocity of light can tend to infinity. However, when is small, compared
to quantities of interest in classical mechanics, this is expressed by mathematically by
taking the limit → 0 or by letting related parameters tend to zero (see Littlejohn
[1988] and de de Gosson [1997]).
2 van Hove’s theorem in R n is proved in Abraham and Marsden [1978], §5.4. van Hove
also found some positive results that were extended by Segal, Souriau and Kostant in a
procedure now called prequantization. Recent references in this direction may be found
in Gotay, Grundling, and Tuynman [1996].
N7.A The Classical Limit and the Maslov Index 11
ψ = a exp(iS/) (N7.A.7)
2a S + aS = 0, (N7.A.9)
2 a
Eψ = Lψ + ψ (N7.A.10)
2m a
which differs from (N7.A.1) by a term of order 2 . The idea is now to
continue this process by writing
N
ψ= ak (i)k exp(iS/) (N7.A.11)
k=0
iN exp(iS/) N +2
Eψ = Lψ + aN . (N7.A.14)
2m
Thus, we have “solved” (N7.A.1) up to an error of order N +2 . Therefore,
if we let N → ∞ we have found an asymptotic solution
∞
ψ∼ ak k exp(iS/) (N7.A.15)
k=0
p 2
p
__ +V=E
2m
+
q
q1 q2
–
3. Use a modified WKB method near the turning point and an asymp-
totic expansion (Maslov). We shall describe this method shortly.
There are other approaches too. For instance, Miller and Good [1953]
effectively used area–preserving maps to deform Figure N7.A.1 into that
for a harmonic oscillator. The same idea was used by Maslov [1965] for
higher superpositions of such expressions.
To study the behavior near q1 and q2 , we replace ψ = a exp(iS/) by a
superposition of such expressions, that is, by
∞
ψ(q) = a(q, p) exp(iϕ(q, p)/) dp (N7.A.18)
−∞
where x(z0 ) = x0 .
To show this, we can clearly assume that
Write first
1 1
d
ϕ(x) = ϕ(tx) dt = x ϕ (tx) dt = xα(x),
0 dt 0
where 1
α(x) = ϕ (tx) dt
0
is again a C ∞ function. Since
and ϕ (0) = α(0) = 0, the same argument shows that α(x) = xβ(x) for
some C ∞ function β(x). Therefore,
1
2
ϕ(x) = x β(x) and β(x) = α (tx) dt,
0
whence
β(0) = α (0) = 12 ϕ (0).
√
Define z(x) = 2|β(x)| 2 x which is C ∞ in a neighborhood of 0, since β(0)
=
1
0, and satisfies √
z (0) = 2|β(0)| 2
= 0.
1
ϕ(x) = x2 β(x) = (sgn β(x))x2 |β(x)| = 12 (sgn β(0))z 2 = 12 (sgn ϕ (0))z 2 .
where + or − is taken in accordance with sgn ϕ (x0 ) and γ(z) is C ∞ with
(z − z0 )γ(z) bounded together with all its derivatives. (The bound for each
derivative may be different.)
Indeed,
∞
∞
iϕ(x0 ) iϕ(x0 ) i(z − z0 )2 dx
a(x) exp dx = a(x(z)) exp ± dz dz
−∞ −∞ 2
dz 2
for some C ∞ function γ(z) (z0 denotes the point given by x(z0 ) = x0 ), that
is,
dx 1
a(x(z)) = a(x0 ) 1 + (z − z0 )γ(z).
dz
|ϕ (x0 )| 2
Since
1
d dx
(z − z0 )γ(z) = (z − z0 ) a(x(zt ) (zt ) dt,
0 dt dz
ables z → −z, the same result holds for the integral from −∞ to 0 and
N7.A The Classical Limit and the Maslov Index 17
A
B
1 −λz2 /2
= −λ−1 (e ) h(z) dz
A z
B B
−1 −λz 2 /2 −1 h(z)
−λz 2 /2
= −(λz) e h(z) + λ e dz
A A z
B B
−1 −λz 2 /2 −2 −λz 2 /2 1 h(z)
= −(λz) e h(z) − λ (e ) dz
A A z z
B
−1 −λz 2 /2 −1 h(z)
= −(λz) e h(z) + λ
z
A
B
1 h(z)
+ λ−2 e−λz /2
2
dz.
A z z
is uniformly convergent.
Arguing in the same manner for the λ–derivative, we conclude that the
integral is holomorphic for Re λ > 0. Similarly one shows continuity for
Re λ ≥ 0.
Step 4.
∞
±i(z − z0 )2 √
exp dz = 2πe±πi/4 .
−∞ 2
From the previous step it follows that this integral exists, by taking
λ = ∓i and h(z) = 1. Moreover, the classical formula
∞ √
e−u
2
/2
du = 2π
−∞
−∞
18 N7. Lagrangian Mechanics
By analytically continuing both sides for Re λ > 0, the same formula holds
for complex λ in the right half–plane. Now let λ → ∓i/ to obtain
∞ π
±i(z − z0 )2 √ √
exp dz = 2π exp ± i
−∞ 2 2
√ πi
= 2π exp ± .
4
The previous proof shows that the same formula holds if all functions
depend smoothly on additional parameters. In particular, we shall use the
following expression in analyzing the right–hand side of (N7.A.19):
∞
f (q, p)
c(q, p) exp i dp
−∞
√ iπ c(q, p) exp(if /)
= 2π exp sgnfpp + O(3/2 ), (N7.A.22)
f =0
4 |f pp |1/2
p
N7.A The Classical Limit and the Maslov Index 19
where the sum is over all p such that fp = ∂f /∂p vanishes; these critical
points are assumed to be finite in number and nondegenerate, that is,
fpp = ∂ 2 f /∂p2
= 0.
Applying (N7.A.22) to (N7.A.20) gives
Lψ − Eψ
2
√ exp[−iπ sgn T (p)/4] p
= 2π a(q, p) + V (q) − E
|T (p)|1/2 2m
q=T (p)
pq − T (p)
× exp i + O(3/2 ), (N7.A.23)
p2
+ V (T (p)) = E. (N7.A.24)
2m
Thus the graph of q = T (p) (as a function of p) is contained in the en-
ergy surface. Equation (N7.A.24) is the Hamilton–Jacobi equation in the
variable p, which is approximated near the turning points q1 and q2 .
Applying (N7.A.22) to formula (N7.A.19) gives
√ 1
ψ(q) = 2π exp[−iπ sgn T (p)/4]
|T (p)|1/2
q=T (p)
pq − T (p)
× exp i a(q, p) + O(3/2 )
√ 1
= 2π exp[−iπ sgn T (p)/4]
|T (p)|1/2
pT (p) − T (p)
× exp i a(T (p), p) + O(3/2 )
= O(1/2 ). (N7.A.25)
20 N7. Lagrangian Mechanics
d dp dp
(pq − T (p)) = p + q − T (p) = p, (N7.A.26)
dq dq dq
the term
e−iπ sgn T (p)/4
(N7.A.27)
p
2
p
__ +V=E
2m
+
S
D
T1 + T2
2 1 q
C
q1 A 3 4
q2
−
B
S–
1
p dq, (N7.A.28)
since both S and pq − T (p) are given by integrating p and the line integral
is over the energy curve. On the other hand, the phase change due to the
term (N7.A.27) is
π π
−2 × − − = −π, (N7.A.29)
4 4
1 p dq
p dq − π = 2πn, i. e. , = n + 12 . (N7.A.30)
2π
The 12 is the correction to the Bohr–Sommerfeld rules which one sees, for
example, in the harmonic oscillator solution. Equation (N7.A.30) is the
quantization condition. Its generalization to arbitrary manifolds reads
1
pi dq i − 14 Iγ = integer, (N7.A.31)
2π γ
N9
Lie Groups
Lie groups is a large subject and Chapter 9 of the text as well as this
supplement cover only a part of the subject.
Proof. It suffices to prove smoothness of γ for |t| < ! for some small
! > 0. Indeed, γ(t + s) = γ(t)γ(s) shows that if |s| < ! then γ is smooth in
an !–neighborhood of each t; thus γ(t) is smooth in a 2!–neighborhood of
zero. Repeating, we see γ is smooth everywhere.
To show that γ is smooth for |t| small, the strategy is to show that
it coincides with exp(tζ) for some ζ ∈ g and for small t. The strategy
of the proof is to show this equality for small rational numbers t using
algebraic properties of γ and exp and then to invoke continuity for a limiting
argument.
To carry this strategy out, fix some n ∈ N and let BR be the open ball
of radius R about the origin in g on which exp is a diffeomorphism.
By
continuity of γ, there is some ! > 0 such that γ(t) ∈ exp BR/2 for all
24 N9. Lie Groups
|t| < !. Fix s > 0, s < ! and define η ∈ BR/2 by exp η = γ(s). Similarly,
since s/n < !, define ξn ∈ BR/2 by γ(s/n) = exp ξn and note that
n n
exp (nξn ) = (exp ξn ) = γ (s/n) = γ(s) = exp η
γ(qs) = exp(qη).
has derivative at the origin equal to the identity map (if we identify g
with Rn via the chosen basis). Therefore, one can find open neighborhoods
V of e in G and U of the origin in Rn such that ψ|U : U → V is a
diffeomorphism. Let ϕ : V → U be given by ϕ = (ψ|V )−1 . Then (V, ϕ)
N9.B Abelian Lie Groups 25
(i) ϕ : U × U → V is a diffeomorphism,
(ii) V ∩ H = exp(U ).
exp : B × B → exp (B × B)
are diffeomorphisms. Let Bn , Bn denote the balls of radius r/n centered at
the origin in h⊥ and h respectively. We claim that for some n large enough,
The definition of h immediately implies that exp (Bn ) ⊂ H and hence that
To show the converse, assume the contrary, namely that for any n ∈ N
there exists a ξn ∈ Bn such that exp ξn ∈ H but ξn
= 0. Clearly, ξn → 0
as n → ∞ and by compactness of the unit sphere, ξn /ξn has a con-
vergent subsequence ξnk /ξnk → ξ ∈ h⊥ , ξ = 1. Step 1 then implies
that exp tξ ∈ H for all t ∈ R, that is, ξ ∈ h, by definition of h. Thus
ξ ∈ h⊥ ∩ h = {0} which contradicts ξ = 1.
Therefore, if n is large enough,
The last term equals [ξ, η]L (g) ∈ h̃g . The first three terms all have the
following structure: U and V are sections of h̃ and V (g) = 0. If we can
prove that [U, V ] (g) ∈ h̃g , this will show that each of the first three terms
lies in h̃g and we can then conclude that h̃ is an involutive distribution.
The following Lemma solves this problem.
Lemma N9.D.1. Let M be a manifold and let E be a subbundle of T M .
If Y is a section of E such that Y (m0 ) = 0 for a given point m0 ∈ M ,
then [X, Y ] (m0 ) ∈ Em0 for any X ∈ X(M ).
Proof. Let E be the Banach space modeling M . Since the problem is
local, we can replace M by an open neighborhood U of 0 ∈ E, T M by
U × E, and m0 by 0. Because E is a subbundle of T M , there is a splitting
E = E1 × E2 such that, locally, E can be replaced by U × E1 . A section Y
of E is of the form x ∈ U → (x, (f (x), 0)), where f : U → E1 is a smooth
function. The condition X (m0 ) = 0 is equivalent to f (0) = 0.
Let X ∈ X(M ) be arbitrary. Represent it locally in the chart with domain
U by X(x) = (x, g(x)), where g : U → E is a smooth function. Then,
locally, in U ,
since f (0) = 0 and since Df (0) ∈ L (E, E1 ). Therefore [X, Y ] (m0 ) ∈ Em0 .
Returning to the proof of the theorem and applying the theorem of Frobe-
nius to the involutive subbundle h̃ ∈ T G, it follows that h̃ is integrable. Let
H be the maximal integral submanifold of h̃ through the identity, that is,
e ∈ H, Th H = h̃h for any h ∈ H, and H is the maximal (relative to the
inclusion) immersed submanifold of G having these properties.
We shall prove that H is a subgroup of G. If g ∈ G, then gH is the
maximal integral manifold containing g. Indeed, g ∈ gH since e ∈ H and
if h ∈ H, then
(h, k) ∈ H × H D→ G × G → hk ∈ H D→ G
h ∈ H1 , Th H1 = Te Lh (Te H1 ) = Te Lh (h) = hh ,
then
and
The relation
where N is the nilpotent matrix with 1’s occupying some places on the first
upper diagonal given by the complex Jordan canonical form
diag (λ1 , . . . , λn ) + N.
Interchanging columns and rows (for each column interchange do the same
for the rows) one can transform this matrix to a block upper triangular
matrix, each block on the diagonal being of the form
Re λ1 − Im λ1
,
Im λ1 Re λ1
the upper 2 × 2 blocks being either the zero matrix or the matrix
0 −1
,
1 0
N9.E The Symplectic, Orthogonal, and Unitary Groups 33
and all other entries being zero. This matrix has the same determinant as
the original one (because an even number of columns and row changes have
been performed) and, since it is block upper triangular, this determinant
equals the product of the determinants of the diagonal blocks, that is,
(Re λk )2 + (Im λk )2 = |λk |2 , which proves the statement.
Using the embedding C defined above, it follows that, U(n) is embedded
in GL(2n, R) as the set of matrices of the form (N9.E.3) with a certain
additional property to be determined below. If A + iB ∈ U(n) then
(A + iB)† (A + iB) = I.
(A + iB)† = AT − iB T
which is equivalent to
As we shall see, this is the first in a series of three parallel results of this
sort.
Proof. We have already seen that A + iB ∈ U(n) iff (N9.E.4) holds.
Now let us characterize all matrices of the form
A B
∈ Sp(2n, R) ∩ O(2n, R).
C D
Recall from the main text that a block matrix like this is symplectic iff
which is equivalent to
Now, multiply on the right by D the first identity in (N9.E.6), to get from
(N9.E.5)
D = AAT D + BB T D
= A(I + C T B) + BB T D
= A + AC T B + BDT B
= A + (AC T + BDT )B = A
B = CC T B + DDT B
= C(AT D − I) + DDT B
= CAT D − C + DB T D
= −C + (CAT + DB T )D = −C
Notice that it follows from this and the fact that elements of Sp(2n, R)
have determinant 1, that
Hn = {a = (a1 , . . . , an ) | ai ∈ H}.
N9.E The Symplectic, Orthogonal, and Unitary Groups 35
This satisfies all axioms of an n-dimensional vector space over H with the
sole exception that H is not a field, being non-commutative.The group
GL(n, H) is defined to be the set of all invertible H–linear maps T : Hn →
Hn defined by left multiplication by a n × n matrix [tpr ], with tpr ∈ H, that
is,
n
(T a)r = tpr ap ,
p=1
x, y ∈ R. We have
χ((u, v)α) = χ(u, v)α
for all α ∈ C. So, again, we get only right linearity. The key property of
χ is that it turns a left quaternionic matrix multiplication operator into
a usual complex linear operator on C2n . Indeed, if [tpr ] is a quaternionic
n × n matrix, then χ−1 T χ : C2n → C2n is complex linear . To verify this,
let α ∈ C, u, v ∈ Cn and note that
−1
χ T χ (α(u, v)) = χ−1 T χ ((u, v)α) = χ−1 T ((χ(u, v))α)
= χ−1 ((T χ(u, v))α) = χ−1 T χ(u, v) α
= α χ−1 T χ(u, v) .
Let us determine, for example, the 2n × 2n complex matrix J that cor-
responds to the right linear quaternionic map given by left multiplication
with the diagonal map jI. We have
J(u, v) = (χ−1 jIχ)(u, v)
= (χ−1 jI)(u + jv) = χ−1 (ju − v)
= (−v, u),
36 N9. Lie Groups
that is,
0 I
J=
−I 0
T = A + jB
Thus
A −B
H : A + jB ∈ gl(n, H) → ∈ gl(2n, C)
B A
satisfies
†
H (A + jB)† = [H(A + jB)] ,
T T
where (A + jB)† = A − B j and
Define
sl(n, H) = {T ∈ gl(n, H) | Re trace T = 0} .
N9.E The Symplectic, Orthogonal, and Unitary Groups 37
Note that the trace of any element in sp(2n) is necessarily zero and hence
that any element in Sp(2n) has determinant equal to 1. Thus, unlike the
N9.E The Symplectic, Orthogonal, and Unitary Groups 39
case of real or complex matrices, where the isometry condition did not
imply that the determinant is zero (and hence we distinguished between
O(n) and SO(n), U(n) and SU(n)), in the case of quaternionic matrices
there is only one group of isometries, namely Sp(2n) and the determinant
equal to 1 condition is automatically satisfied.
Proposition N9.E.3. The unitary symplectic group Sp(2n) is the group
of isometries of Hn . It is a compact connected subgroup SL(n, H) ∼
= SU∗ (2n)
2
of complex dimension 2n + n whose Lie algebra is sp(2n).
Compactness is proved exactly as in the real or complex
√ case by showing
that the norm of an element in Sp(2n) is equal to 2n and the proof of
connectedness is, as usual, deferred to §9.3. From our previous consideration
it immediately follows that:
Proposition N9.E.4.
Sp(2n) = SU∗ (2n) ∩ U(2n).
The Complex Symplectic Group Sp(2n, C) is defined exactly as in
the real case by the condition
! "
Sp(2n, C) = T ∈ GL(2n, C) | T T JT = J .
It is a noncompact connected closed Lie subgroup of GL(2n, C) of complex
dimension 2n2 + n and whose Lie algebra is
! "
sp(2n, C) = T ∈ gl(2n, C) | T T J + JT = 0 .
Proposition N9.E.5.
Sp(2n) = Sp(2n, C) ∩ U(2n).
Proof. Recall that
A C
T = ∈ Sp(2n, C)
B D
if and only if AT B and C T D are symmetric and AT D − B T C = I (see
9.2.12). Also, T ∈ U(2n), if and only if
A† A + B † B = I, C † C + D† D = I and A† C + B † D = 0.
From the characterization (N9.E.7) it follows that all these conditions hold.
Conversely, if these conditions hold, then
−1
J = T T JT = T T T T JT
† −1 −1
=T T T JT = (T † T ) T JT
−1
= T JT
since T † T = I because T ∈ U(2n). Therefore T J = JT which forces C =
−B, D = A. But then, these conditions imply those in (N9.E.7).
40 N9. Lie Groups
(U1 × U2 ) ∩ U = ∅ and U1 ⊂ R, U2 ⊂ V2
p1 , . . . , pN ∈ R[X1 , . . . , Xn ], n = dim V,
g∗reg := {µ ∈ g∗ | dim gµ = r}
is Zariski open and thus open and dense in the usual topology of g∗ . If
dim gµ = r, then gµ is abelian.
Proof (Due to J. Carmona, as presented in Rais [1972]). Define the map
ϕµ : G → g∗ by g → Ad∗g−1 µ. This is a smooth map whose range is the
coadjoint orbit Oµ through µ and whose tangent map at the identity is
Te ϕµ (ξ) = − ad∗ξ µ. Note that ker Te ϕµ = gµ and
range Te ϕµ = Tµ Oµ .
rank Te ϕµ = n − dim gµ ≤ n − r
U = {µ ∈ g∗ | dim gµ = r} = {µ ∈ g∗ | rank(Te ϕµ ) = n − r}
Te ϕµ : g → g∗ , µ ∈ g∗ .
L(g, g∗ ).
Let
Si = {µ ∈ g∗ | rank Te ϕµ = n − r − i}, 1 ≤ i ≤ n − r.
Then Si is the zero set of the polynomials in µ obtained by taking all
determinants of the (n − r − i + 1)-minors of the matrix $ representation
n−r
of Te ϕµ in these bases. Thus, Si is an algebraic set. Since i=1 Si is the
∗
complement of U , if follows that U is a Zariski open set in g , and hence
open and dense in the usual topology of g∗ .
Now let µ ∈ g∗ be such that dim gµ = r and let V be a complement to
gµ in g, that is,
g = V ⊕ gµ .
Then Te ϕµ |V is injective. Fix ν ∈ g∗ and define
S = {t ∈ R | Te ϕµ+tν |V is injective.}
42 N9. Lie Groups
Note that 0 ∈ S and that S is open in R because the set of injective linear
maps is open in L(g, g∗ ) and µ → Te ϕµ is continuous. Thus, S contains
an open neighborhood of 0 in R. Since the rank of a linear map can only
increase by slight perturbations, we have rank
Te ϕµ+tν |V ≥ rank Te ϕµ |V = n − r,
rank Te ϕµ+tν = n − r
This formula shows that for |t| small, t → ξ(t) is a smooth curve in V and
ξ(0) = 0. However, since
ξ(t) + ξ ∈ gµ+tν .
[ξ, η] = 0.
Since the linear map ξˆ : g∗ → R factors through g∗ /[µ, g], it follows that
there exists a smooth map ϕ : U/G → R for which
(dϕ)π(µ) · (Tµ π · ν) = ν (ξ)
44 N9. Lie Groups
(ξ · g) · µ = ξ · (g · µ) = (g · µ) · ξ = g · (µ · ξ) = g · (ξ · µ) = (g · ξ) · µ.
ξ·g =g·ξ
Exercise
# N9.F-1. Prove the following generalization of the Duflo–Vergne Theorem
due to Guillemin and Sternberg [1984]. Let S be an infinitesimally invariant
submanifold of g∗ , that is, ad∗ξ µ ∈ S, whenever µ ∈ S and ξ ∈ g. Let
r = min{dim gµ |µ ∈ S}. Then dim gµ = r implies
Examples
A. Consider a compact manifold M (possibly with boundary) and the
infinite dimensional vector space C ∞ (M ) of all smooth real value functions
on M . Evidently, C ∞ (M ) is an abelian group with pointwise defined group
operations; that is,
µ : C ∞ (M ) × C ∞ (M ) → C ∞ (M ), (f, g) → f + g
I : C ∞ (M ) → C ∞ (M ), f → −f
the bracket on the right-hand side being taken in g. For example, if s > 3/2,
then H s (R3 , S 1 ) is an abelian Banach Lie group under pointwise multipli-
cation, with Lie algebra H s (R3 ), [· , ·] = 0. If G is a compact Lie group with
Lie algebra g and s > n/2, then H s (Rn , G) has Lie algebra H s (Rn , g) and
bracket [f, g](x) = [f (x), g(x)], the latter bracket being in g.
Diffeomorphism Groups Among the most important “classical” ex-
amples of infinite dimensional groups are the diffeomorphism groups of
manifolds. Let M be a compact boundaryless manifold and denote by
H s Diff (M ) the set of all H s diffeomorphisms of M to M , s > n/2 + 1.
We will now outline the sense in which H s Diff (M ) is a Lie group with
Lie algebra H s -X(M ), the H s vector fields on M . Similar results will be
valid for C k Diff (M ) [resp. W s,p Diff (M )], the group of C k -diffeomorphisms
of M [resp. W s,p -diffeomorphisms, of M ].
The set H s Diff (M ) is a smooth Banach manifold and is a group under
composition; explicitly, the group operations are
and
I : H s - Diff (M ) → H s - D Diff(M ); I(f ) = f −1 .
N9.G Some Infinite Dimensional Lie Groups 47
The unit element e is the identity map. For s > (n/2) + 1, H s Diff(M )
is a Banach manifold and in fact is an open submanifold of the Banach-
manifold H s (M, M ). The condition s > (n/2) + 1 guarantees that elements
of H s Diff(M ) are C 1 and a map C 1 close to a diffeomorphism is a diffeo-
morphism by the inverse function theorem (plus an additional argument to
guarantee it is globally one to one and onto; see Marsden, Ebin, and Fis-
cher [1972]). The manifold H s Diff(M ) is not, however, a Banach Lie group,
since group multiplication is differentiable only in the following restricted
sense. Right multiplication
Rg : H s Diff(M ) → H s Diff(M ); Rg (f ) = f ◦ g
Lg : H s Diff(M ) → H s Diff(M ), Lg (f ) = g ◦ f
is a C k -map. Therefore,
The idea behind the above assertions is as follows. An element of the tan-
gent space at the point f , namely, Tf (H s Diff(M )) is the tangent vector to a
curve f (t) ∈ H s Diff(M ) at t = 0 where f (0) = f . But each f (t) maps M to
M , so for x ∈ M , f (t)(x) is a curve in M . Thus (d/dt)f (t)(x)|t=0 ∈ Tf (x) M
and so we get a map Xf of M to T M taking x to an element of Tf (x) M . In
particular, the tangent space at the identity of e is the space of H s vector
fields on M . The tangent manifold T (H s Diff(M )) can be identified with
the set of all mappings from M to T M that cover diffeomorphisms and it
48 N9. Lie Groups
X : H s Diff(M ) → T (H s Diff(M ))
such that
T Lg : T (H s Diff(M ) → (H s Diff(M ))
T Rg (Xf ) = Xf ◦ g ∈ Tf ◦g (H s Diff(M ))
and
The diagrams in Figure N9.G.1 may help to clarify the situation: in them,
T Rg (Xf ) = Xf ◦ g is a vector field along f ◦ g and T Lg (Xf ) = T g ◦ Xf is
a vector field along g ◦ f .
Tg ✲T M
TM TM
✼ ✻❙
✻ ✒
T Rg (Xf ) ❙ τM
❙ Xf τM
Xf ❙
❙❙
✇
T Lg (Xf )
g f ❄
M ✲ M ✲ M M ✲M ✲M
f g
Tf ✲
M TM TM
✒ ✻ ✒
Yξ (f ) Xξ (f )
ξ ξ τ
f ❄ f ❄
M ✲M M ✲M
For the corresponding left invariant vector fields Xξ1 , Xξ2 on H s Diff(M ),
there is a sign change:
Note that the bracket [ξ1 , ξ2 ] is the ordinary Lie bracket of the vector fields
ξ1 , ξ2 on the manifold M . Due to this fact, we define the right “Lie” algebra
of the “Lie” group H s Diff(M ) to be the space of right invariant vector fields
on H s Diff(M ).
d d
c(t)(x) = ϕt (x) = ξ(ϕt (x)) = (ξ ◦ c(t))(x) = Yξ (c(t))(x).
dt dt
In particular, note that c (0) = ξ. Thus the exponential map
is given by exp(ξ) = ϕ1 . The map exp is continuous, but unlike the case
for Banach Lie groups, it is not C 1 ; in fact, there is no neighborhood of the
identity onto which it maps surjectively (Kopell [1970]). As a result in this
direction, let us prove that if a diffeomorphism of S 1 has no fixed points
and is in the image of the exponential map, then it must be conjugate to a
rotation (Hamilton [1982]). Let η ∈ Diff(S 1 ) have no fixed points. If there
is ξ ∈ X(S 1 ) such that exp(ξ) = η, then ξ is nowhere vanishing. Write
ξ = f (t)(d/dt), where t ∈ R is a parameterization of S 1 modulo 2π. Now
reparametrize the circle by
dt 1
θ=c where c = 2π
f (t) dt
0 f (t)
f ∈ H s Diff 0 (M ) → f˜ ∈ H s Diff(M̃ )
for any function f for which the right side makes sense, · p being the
Lp (Ω)-norm. It is clear that · k,p defines a norm on any vector space
N9.G.1 Basic Facts about Sobolev Spaces and Manifolds. 53
of functions on which the right side takes finite values, provided that two
functions are identified in the space if they are equal almost everywhere in
Ω. Let H k,p (Ω) be the completion of
! "
f ∈ C k (Ω) | f k,p < ∞
with respect to the norm · k,p . Then H k,p (Ω) is a Banach space called a
Sobolev space. For p = 2 we denote H k,2 (Ω) by H k (Ω).
For the reader familiar with distributional derivatives, another definition
may be helpful. Let
Let W s,p (Ω) denote the completion of {f ∈ C ∞ (Ω) | f s,p < ∞} and set
H s (Ω) ≡ W s,2 (Ω). One treats W s,p (Ω, Rm ) in a similar way. Also, by
completing the space of functions of Ω that extend to C ∞ functions in
an open neighborhood of the closure Ω̄ one similarly defines W s,p (Ω̄) and
W s,p (Ω̄, Rm ).
Proof. This follows from the SNG inequality (N9.G.2) and Young’s in-
equality:
xa y 1−a ≤ ax + (1 − a)y,
where Kε = 1/εa/(1−a) .
Let us illustrate how Fourier transform techniques can be used to directly
prove the special case of the preceding Corollary in which n = 3, j = 0,
p = ∞, m = 2, r = 2, and q = 2.
Proposition N9.G.4. There is a constant c > 0 such that for any ε > 0
and function f : R3 → R smooth with compact support, we have
f ∞ ≤ c ε3/2 f L2 + ε−1/2 ∆f L2 .
denote the Fourier transform. Recall that (∆fˆ)(k) = −k2 fˆ(k). From
Schwarz’ inequality, we have
2 2 2
ˆ dk 2 2ˆ
f (k) dk ≤ 2 ε + k f (k) dk
(ε2 + k2 )
c1
= ε2 − ∆ f 2L2 ,
ε
where
dξ
c1 = 2 < ∞.
R3 (1 + ξ2 )
Here we have used the fact that h → ĥ is an isometry the L2 -norm
inik·x
(Plancherel’s theorem). Thus, from f (x) = 1/(2π)3/2
R3
e fˆ(k)dk, we
get
c2
(2π)3/2 f ∞ ≤ fˆL1 ≤ √ ε2 − ∆ f L2
ε
≤ c2 ε f L2 + ε−1/2 ∆f L2 .
3/2
56 N9. Lie Groups
Thus we have shown that H 2 (R3 ) ⊂ C 0 (R3 ) and that the inclusion
is continuous. More generally, one can show by similar arguments that
H 2 (Ω) ⊂ C k (Ω) provided s > (n/2) + k and
one gets
∞ 2 ∂u
sup u3 (x, y, z) ≤ 3 u
∂x dx.
x −∞
N9.G.1 Basic Facts about Sobolev Spaces and Manifolds. 57
Set I = R3
u6 dx and write
3 3
∞
I=
u u dydz dx
−∞
∞ ∞ ∞
3 3
≤ sup
u dz
sup u dy dx
−∞ y −∞ −∞ x
∞
2 ∂u 2 ∂u
≤9
u ∂y dydz u ∂z dydz dx.
−∞
gives
3/2
√ 1
I ≤ 36 I grad u2 ;
R3 33/2
58 N9. Lie Groups
i.e.,
3
(36)2
I≤ grad u2 .
33 R3
Dj u · Dk v2L2 ≤ const. Dj u2L2s/j Dk v2L2s/k ≤ const. u2H s v2H s .
not generally true for W+s,p (M ) in W s,p (M ). However, if s > n/p then
W s,p (M ) is continuously embedded in C 0 (M ). Then
W+s,p (M ) = W s,p (M ) ∩ C+
0
(M ),
F (η, f ) = f ◦ η.
Then F is C ∞ .
Sobolev inverse functions. If η ∈ H s (M, M ), s > (n/p) + 1 and η has
a C 1 inverse, then the inverse is W s,p so η ∈ W s,p Diff(M ).
Now consider two maps related to the composition map F : let
Fη : W s,p (M ) → W s,p (M ),
be defined by
Fη (f ) := f ◦ η
for fixed η ∈ (W s ,p -Diff (M )), with s ≥ s + 1) and let
60 N9. Lie Groups
Of : (W s ,p -Diff (M )) → W s,p (M ),
be defined by
Of (η) := f ◦ η
for fixed f ∈ W s ,p (M ) with s ≥ s and s ≥ s).
The first of these, Fη , is the pullback map. It is linear since
Fη (f + g) = (f + g) ◦ η = f ◦ η + g ◦ η = Fη (f ) + Fη (g).
and
Lη : W s,p -Diff (M ) → W s,p -Diff (M ); Lη (µ) = η ◦ µ.
To ensure that W s,p Diff(M ) is a group, one requires that s > (n/p) + 1
so that an inverse exists. One can use the results on composition above
to show that while right multiplication is smooth, left multiplication and
inversion are only C 0 . Hence W s,p Diff(M ) is a topological group and not
a Banach Lie group.
Nested Groups. In view of the important technical properties of group
multiplication in H s Diff(M ), we introduce an abstract context for this
phenomenon developed by Omori [1974] and Adams, Ratiu, and Schmid
[1986].
Definition N9.G.7. A collection of groups {G∞ , Gs | s ≥ s0 } is called
a nest if
Examples
A. The classical examples of Lie group nests are the diffeomorphism
groups
{Diff(M ), H s -Diff (M ) | s > (dim M )/2}
with Lie algebra nests
tomorphisms of T ∗ M \{0}. But right away we are faced with the problem
of non-compactness of T ∗ M \{0}. We sketch below, following Ratiu and
Schmid [1981] how
{Diff θ (T ∗ M \{0}), H s Diff θ (T ∗ M \{0}) | s ≥ dim M + 1/2}
is a Lie group nest with Lie algebra nest
! ∞ ∗ "
S (T M \{0}), S s+1 (T ∗ M \{0}) | s > dim M + 1/2 ,
where
S s (T ∗ M \{0}) = {H : T ∗ M \{0} → R | H is of
class H s and homogeneous of degree one}
with the Poisson bracket as Lie algebra bracket. Note the gain in one deriva-
tive at the Lie algebra level. The basic idea of the ensuing discussion is
that H s Diff θ (T ∗ M \{0}) is algebraically isomorphic to the group of all H s
contact transformations of the cosphere bundle of M , which is a compact
manifold if M is. We start by recalling the relevant facts.
The multiplicative group of strictly positive reals R+ acts smoothly on
T ∗ M \{0} by αx → τ αx , τ > 0, αx ∈ Tx∗ M , αx
= 0. This action is free
and proper and therefore π : T ∗ M \{0} → Q ≡ T ∗ M \{0}/R+ is a smooth
principal fiber bundle over Q, the cosphere bundle of M . Note that Q
is compact (supposing M is) and odd-dimensional. Q carries no canonical
contact one-form but for each global section σ : Q → T ∗ M \{0} we can de-
fine an exact contact one-form θσ on Q by θσ = σ ∗ θ. Such global sections
exist in abundance; for example, any Riemannian metric on M identifies
T ∗ M with T M and Q with the unit sphere bundle. Then the usual in-
clusion of the sphere bundle into T M gives a section σ. The section σ is
uniquely determined by a smooth function fσ : T ∗ M \{0} → R+ defined
by σ(π(αx )) = fσ (αx )αx . In other words, fσ measures how far from the
section σ an element αx ∈ T ∗ M \{0} lies. The function fσ is homogeneous
of degree −1 and π ∗ θσ = fσ θ.
An H s+1 contact transformation on Q is a diffeomorphism ϕ ∈
H ∗ Diff θ (Q) such that for any two sections σ, ζ : Q → T ∗ M \{0}, there
N9.G.1 Basic Facts about Sobolev Spaces and Manifolds. 63
n−1
θσ = y i dxi + dt,
i=1
Cons+1 (Q)
!σ "
= (ϕ, h) ∈ H s+1 -Diff (Q) H s+1 (Q, R\{0}) | ϕ∗ θσ = hθσ
The Lie algebra of this semidrect product group is the semidirect product
Lie algebra
σ (Q) = {(Y, g) ∈ H
cons+1 - X(Q) H s+1 (Q, R) | LY θσ = gθσ }.
s+1
64 N9. Lie Groups
In Ratiu and Schmid [1981] (Theorem 4.1) it is shown that the group
Since Cons+1
σ (Q) and Conζ
s+1
(Q) are isomorphic as nested Lie groups, for
any two global sections σ and ζ, the isomorphism Φ determines a nested
Lie group structure on H s+1 Diff θ (T ∗ M \{0}) which is independent of σ
(or independent of the Riemannian metric if σ is induced from such). Fur-
thermore, the Lie algebra
f ◦ σ −1 : σ(Q) ⊂ (T ∗ M \{0}) → R,
N9.G.1 Basic Facts about Sobolev Spaces and Manifolds. 65
for f ∈ H s+2 (Q, R). The composition of these two isomorphisms with
Te Φ−1 gives an isomorphism
T Φ−1
σ (Q) −→ H
F : cons+1 e s+1
-Xθ (T ∗ M \{0}) → S s+2 (T ∗ M \{0})
j −1
−→ H s+2 (Q, R),
it follows that
dH : (T ∗ M \{0}) → T ∗ (T ∗ M \{0})
s+2
is C 1 -close to zero in the strong C 1 -Whitney topology. Let SW (T ∗ M \{0})
∗
be the set S (T M \{0}) equipped with this new topology. One checks
s+2
that
j : H s+2 (Q, R) → SW
s+2
(T ∗ M \{0}),
Remarks.
1. The gain of one derivative at the Lie algebra level has a corresponding
statement in H s+1 Diff θ (T ∗ M \{0}): for η ∈ H s+1 Diff θ (T ∗ M \{0}),
τ ∗ ◦ η : (T ∗ M \{0}) → M
N10
Poisson Manifolds
is an embedding.
:
(ii) The ranges of (T Ψzf1 ,... ,fk )(t) and BΨ z are equal for t ∈ Uδ .
f1 ,... ,fk (t)
(iv) If
Ψyg1 ,... ,gk : Uη → P
is another map constructed as above and y ∈ Ψzf1 ,... ,fk (Uδ ), then there
is an open subset, U
⊂ Uη , such that Ψyg1 ,... ,gk is a diffeomorphism
from U
to an open subset in Ψzf1 ,... ,fk (Uδ ).
Proof. (i) The smoothness of Ψzf1 ,... ,fk follows from the smoothness of
Φj,t in both the flow parameter and manifold variables. Then
which shows that T0 Ψzf1 ,... ,fk is injective. It follows that Ψzf1 ,... ,fk is an
embedding on a sufficiently small neighborhood of 0, say Uδ . Notice also
that the ranges of T0 Ψzf1 ,... ,fk and of Bz: coincide.
(ii) Recall from the main text that for any invertible Poisson map Φ on
P , we have
T Φ · Xf = Xf ◦Φ−1 ◦ Φ
and also recall that that Hamiltonian flows are Poisson maps. Therefore, if
t = (t1 , . . . , tk ),
where
−1
hj = fj ◦ Φ1,t1 ◦ . . . ◦ Φj−1,tj−1 .
This last equality, the previous inclusion, and the last remark in the proof
of (i) above conclude (ii).
(iii) This is obvious since Ψzf1 ,... ,fk is built from piecewise Hamiltonian
curves starting from z.
(iv) Note that Xg (z) ∈ range Bz: for any z ∈ P and any smooth func-
tion g. Using (ii), we see that Xg is tangent to the image of Ψzf1 ,... ,fk .
N10.A Proof of the Symplectic Stratification Theorem 69
Therefore, the integral curves of Xg remain tangent to Ψzf1 ,... ,fk (Uδ ) if
they start from that set. To get Ψyg1 ,... ,gk we just have to find Hamiltonian
curves which start from y. Therefore, we can restrict ourselves to the sub-
manifold Ψzf1 ,... ,fk (Uδ ) when computing the flows along the Hamiltonian
vector fields Xgj ; therefore we can consider that the image of Ψyg1 ,... ,gk is
in Ψzf1 ,... ,fk (Uδ ). The derivative at 0 ∈ Rk of Ψyg1 ,... ,gk is an isomorphism
to the tangent space of Ψzf1 ,... ,fk (Uδ ) at y (that is, range By: ), using (ii)
above. Thus, the existence of the neighborhood U
follows from the inverse
function theorem.
Proposition N10.A.2. Let P be a Poisson manifold and B its Poisson
tensor. Then for each symplectic leaf Σ ⊂ P , the family of charts satisfying
(i) in the previous proposition, namely,
! z "
Ψf1 ,... ,fk | z ∈ Σ, {Bz: dfj }1≤j≤k a basis for range Bz: ,
h1 ,... ,hk (Uγ ) ⊂ Ψf1 ,... ,fk (Uδ ) ∩ Ψg1 ,... ,gk (U
)
Ψw z y
is a diffeomorphic embedding in both Ψzf1 ,... ,fk (Uδ ) and Ψyg1 ,... ,gk (U
). This
shows that the transition maps for the given charts are diffeomorphisms
and so define the structure of a differentiable manifold on Σ. The fact that
the inclusion is an immersion follows from (i) of the above proposition. We
get the tangent space of Σ using (i), (ii) of the previous proposition; then
the equality of dimensions follows.
It follows from the definition of an immersed Poisson submanifold that
Σ is such a submanifold of P . Thus, if i : Σ → P is the inclusion,
{f ◦ i, g ◦ i}Σ = {f, g} ◦ i.
Hence if {f ◦ i, g ◦ i}Σ (z) = 0 for all functions g then {f, g}(z) = 0 for
all g, that is, Xg [f ](z) = 0 for all g. This implies that df |Tz Σ = 0 since
the vectors Xg (z) span Tz Σ. Therefore, i∗ df = d(f ◦ i) = 0, which shows
that the Poisson tensor on Σ is nondegenerate and thus Σ is a symplectic
manifold. This proves the proposition and also completes the proof of the
symplectic stratification theorem.
70 N10. Poisson Manifolds
Proof. The “only if” part follows easily. For the “if” part we remark
that the proof of the theorem above can be reproduced here replacing the
range of Bz: by Dx and the Hamiltonian vector fields with vector fields
in D. The crucial property needed to prove (ii) in the above proposition
(i.e. Hamiltonian fields remain Hamiltonian under Hamiltonian flows) is
replaced by the invariance of D given in the hypothesis.
Remarks.
N11
Momentum Maps
[z0 : z1 : · · · : zn ] = [(z0 , · · · , zn )]
72 N11. Momentum Maps
N13
Lie-Poisson Reduction
that is,
Rϕ η = η ◦ ϕ. (N13.A.2)
T Rϕ · V = V ◦ ϕ. (N13.A.3)
T ∗ Rϕ · π, V = π, T Rϕ · V
= π · (V ◦ ϕ) dx dy dz = (π ◦ ϕ−1 ) · V dx dy dz,
Ω Ω
so
T ∗ Rϕ · π = π ◦ ϕ−1 . (N13.A.5)
FR (η, π) = F (π ◦ η −1 ). (N13.A.6)
M ◦ η = π.
{FR , HR }(Id, M)
δH δF δF δH
=− ·∇ M· − ·∇ M· dx dy dz.
Ω δM δM δM δM
(N13.A.9)
{FR , HR } (Id, M)
δH δF δF δH
= M· ·∇ − ·∇ dx dy dz
δM δM δM δM
δF δH
= M· , dx dy dz, (N13.A.10)
δM δM LA
which is the “+” Lie–Poisson bracket. In doing this step note div(δH/δM) =
0 and since δH/δM and δF/δM are parallel to the boundary, no boundary
term appears. When doing free boundary problems, these boundary terms
are essential to retain (see Lewis, Marsden, Montgomery, and Ratiu [1986]).
For other diffeomorphism groups, it may be convenient to treat M as a
one-form density rather than a vector field.
Hamiltonian vector field can be identified with its generator (modulo ad-
ditive constants being understood). From the formula
we see that g may be identified with F(P ) with the Lie bracket given by
the Poisson bracket. One could then identify g∗ with functions f on P via
the pairing
f, h = f h dµ, (N13.B.2)
P
π = πi dq i + π i dpi
∂π i ∂πi
− = 0, (N13.B.5)
∂q i ∂pi
so by (N13.B.10)
0 1
δF
δπ, = δπ · XδF/δf dµ, (N13.B.11)
δπ P
i.e.,
δF
= XδF/δf . (N13.B.12)
δπ
Thus (N13.B.9) becomes, with the aid of (N13.B.1) and (N13.B.6),
. /
{FR , HR } (Id, π) = π · XδF/δf , XδH/δf LA dµ
P
= π · X{δF/δf,δH/δf } dµ
P
δF δH
= f , dµ (N13.B.13)
P δf δf
1 δ(δ 2 H)
.
2 δ(δµ)
δH δH 1 δ(δ 2 H)
= + ε + O(ε2 ) (N13.C.6)
δµ δµe 2 δ(δµ)
80 N13. Lie-Poisson Reduction
dµe d(δµ)
+ε = − ad∗δH/δµe µe
dt dt
1
− ε ad∗δ(δ2 H)/δ(δµ) µe − ad∗δH/δµe δµ + O(ε2 ).
2
Thus, the linearized equations are
d(δµ) 1
= − ad∗δ(δ2 H)/δ(δµ) µe − ad∗δH/δµe δµ. (N13.C.7)
dt 2
If H is replaced by HC := H + C, with the Casimir function C chosen to
satisfy δHC /δµe = 0, we get ad∗δHC /δµe δµ = 0, and so
d(δµ) 1
= − ad∗δ(δ2 HC )/δ(δµ) µe . (N13.C.8)
dt 2
Equation (N13.C.8) is Hamiltonian with respect to the linearized Poisson
bracket (see Example (f) of §10.1):
0 1
δF δG
{F, G} (µ) = µe , , . (N13.C.9)
δµ δµ
Ratiu [1982] interprets this bracket in terms of a Lie–Poisson structure
of a loop extension of g. The Poisson bracket (N13.C.9) differs from the
Lie–Poisson bracket (N13.C.1) in that it is constant in µ. With respect to
the Poisson bracket (N13.C.9), Hamilton’s equations given by δ 2 HC are
(N13.C.8), as an easy verification shows. Note that the critical points of
δ 2 HC are stationary solutions of the linearized equation (N13.C.8), that is,
they are neutral modes for (N13.C.8).
If δ 2 HC is definite, then either δ 2 HC or −δ 2 HC is positive-definite and
hence defines a norm on the space of perturbations δµ (which is g∗ ). Being
twice the Hamiltonian function for (N13.C.8), δ 2 HC is conserved. So, any
solution of (N13.C.8) starting on an energy surface of δ 2 HC (i.e., on a
sphere in this norm) stays on it and hence the zero solution of (N13.C.8) is
(Liapunov) stable. Thus, formal stability, i.e., definiteness of δ 2 HC , implies
linearized stability. It should be noted, however, that the conditions for
definiteness of δ 2 HC are entirely different from the conditions for “normal
mode stability,” that is, that the operator acting on δµ given by (N13.C.8)
have a purely imaginary spectrum. In particular, having a purely imaginary
spectrum for the linearized equation does not produce Liapunov stability
of the linearized equations.
The difference between δ 2 HC and the operator in (N13.C.8) can be made
explicit, as follows. Assume that there is a weak Ad-invariant metric ,
on g and a linear operator L : g → g such that
L is symmetric with respect to the metric , , that is, ξ, Lη = Lξ, η
for all ξ, η ∈ g. Then the linear operator in (N13.C.8) becomes
δµ → [Lδµ, µe ] (N13.C.11)
which, of course, differs from L, in general. However, note that the kernel
of L is included in the kernel of the linear operator (N13.C.11), that is, the
zero eigenvalues of L give rise to “neutral modes” in the spectral analysis
of (N13.C.11). There is a remarkable coincidence of the zero-eigenvalue
equations for these operators in fluid mechanics: for the Rayleigh equation
describing plane-parallel shear flow in an inviscid homogeneous fluid, taking
normal modes makes the zero-eigenvalue equations corresponding to L and
to (N13.C.11) coincide (see Abarbanel, Holm, Marsden, and Ratiu [1986]).
For additional applications of the stability method, see Holm, Marsden,
Ratiu, and Weinstein [1985], Abarbanel and Holm [1987], Simo, Posbergh,
and Marsden [1990, 1991], and Simo, Lewis, and Marsden [1991]. For a
more general treatment of the linearization process, see Marsden, Ratiu,
and Raugel [1991].
Exercises
# N13.C-1. Write out the linearized rigid body equations about an equi-
librium explicitly.
# N13.C-2. Let g be finite dimensional. Let e1 , . . . , en be a basis for g
and e1 , . . . , en a dual basis for g∗ . Let µ = µa ea ∈ g∗ and H(µ) =
H(µ1 , . . . , µn ) : g∗ → R. Let [µa , µb ] = Cab
d
µd . Derive a coordinate ex-
pression for the linearized equations (N13.C.7):
d(δµ) 1
= − ad∗δ(δ2 H)/δµ µe − ad∗δH/δµe δµ.
dt 2
82 N13. Lie-Poisson Reduction
Page 83
N14
Coadjoint Orbits
where
1
u = − Im(ax + αby).
1 + α2
Thus, if at least one of a, b is not zero, then
0 0 x
gµ = 0 0 y Im(ax + αby) = 0 ,
0 0 0
84 N14. Coadjoint Orbits
The same condition could have been obtained by lengthier direct cal-
culations involving the Lie–Poisson bracket. Here are the highlights. The
commutator bracket on g is given by
is 0 x iu 0 z 0 0 i(sz − ux)
0 iαs y , 0 iαu w = 0 0 iα(sw − uy) ,
0 0 0 0 0 0 0 0 0
Therefore,
References
References
Abarbanel, H. D. I. and D. D. Holm [1987] Nonlinear stability analysis of inviscid
flows in three dimensions: incompressible fluids and barotropic fluids, Phys.
Fluids 30, 3369–3382.
Abarbanel, H. D. I., D. D. Holm, J. E. Marsden, and T. S. Ratiu [1986] Nonlinear
stability analysis of stratified fluid equilibria. Phil. Trans. Roy. Soc. London
A 318, 349–409; also Richardson number criterion for the nonlinear stability
of three-dimensional stratified flow. Phys. Rev. Lett. 52 [1984], 2552–2555.
Abraham, R. and J. E. Marsden [1978] Foundations of Mechanics. Second Edi-
tion, Addison-Wesley.
Abraham, R., J. E. Marsden, and T. S. Ratiu [1988] Manifolds, Tensor Analy-
sis, and Applications. Second Edition, Applied Mathematical Sciences 75,
Springer-Verlag.
Abraham, R. and Robbin, J. [1967] Transversal mappings and flows. Benjamin-
Cummings, Reading, Mass.
Adams, J. F. [1969] Lectures on Lie groups. Benjamin-Cummings, Reading,
Mass.
Adams, R.A. [1975] Sobolev Spaces. Academic Press.
Adams, M. R., T. S. Ratiu, and R. Schmid [1986] A Lie group structure for pseu-
dodifferential operators. Math. Ann. 273, 529–551 and A Lie group structure
for Fourier integral operators. Math. Ann. 276, 19–41.
Arms, J.M., J.E. Marsden, and V. Moncrief [1981] Symmetry and bifurcations of
momentum mappings, Comm. Math. Phys. 78, 455–478.
88 References
Arms, J. M., J. E. Marsden, and V. Moncrief [1982] The structure of the space
solutions of Einstein’s equations: II Several Killings fields and the Einstein-
Yang-Mills equations. Ann. of Phys. 144, 81–106.
Arnold, V. I. [1967] Characteristic class entering in conditions of quantization.
Funct. Anal. Appl. 1, 1–13.
Atiyah, M. [1982] Convexity and commuting Hamiltonians. Bull. London Math.
Soc. 14, 1–15.
Aubin, T. [1976] Espaces de Sobolev sur les variètès riemanniennes. Bull. Sci.
Math. 100, 149–173.
Bates, S. and A. Weinstein [1997] Lectures on the Geometry of Quantization,
CPAM/UCB, Am. Math. Soc.
Bialynicki-Birula, I., J. C. Hubbard, and L. A. Turski [1984] Gauge-independent
canonical formulation of relativistic plasma theory. Physica A 128, 509–519.
Cantor, M. [1975] Perfect fluid flows over Rn with asymptotic conditions. J. Func.
Anal. 18, 73–84.
Chernoff, P. R. and J. E. Marsden [1974] Properties of Infinite Dimensional
Hamiltonian systems. Springer Lect. Notes in Math. 425.
de Gosson, M. [1997] Maslov classes, metaplectic representation and Lagrangian
quantization. Mathematical Research, 95. Akademie-Verlag, Berlin.
Duflo, M. and M. Vergne [1969] Une proprieté de la représentation coadjointe
d’une algébre de Lie. C.R. Acad. Sci. Paris 268, 583–585.
Ebin, D. G. [1970] On the space of Riemannian metrics. Symp. Pure Math., Am.
Math. Soc. 15, 11–40.
Ebin, D. G. and J. E. Marsden [1970] Groups of diffeomorphisms and the motion
of an incompressible fluid. Ann. Math. 92, 102–163.
Eckmann J.-P. and R. Seneor [1976] The Maslov-WKB method for the anhar-
monic oscillator, Arch. Rat. Mech. Anal. 61 153–173.
Fischer, A. E., J. E. Marsden, and V. Moncrief [1980] The structure of the space of
solutions of Einstein’s equations, I: One Killing field. Ann. Inst. H. Poincaré
33, 147–194.
Friedman, A. [1969] Partial differential equations. Holt, Rinehart and Winston.
Gotay, M. J., H. B. Grundling, G. M. Tuynman [1996] Obstruction results in
quantization theory. J. Nonlinear Sci. 6, 469–498.
Guillemin, V. and S. Sternberg [1977] Geometric Asymptotics. Amer. Math. Soc.
Surveys, 14. (Revised edition, 1990.)
Guillemin, V. and S. Sternberg [1982] Convexity properties of the moment map.
Inv. Math. 67, 491-513, 77 (1984) 533–546.
Guillemin, V. and S. Sternberg [1984] Symplectic Techniques in Physics. Cam-
bridge University Press.
Hamilton, R. [1982] The Inverse function theorem of Nash and Moser. Bull. Am.
Math. Soc. 7, 65–222.
Hermann, R. [1962] The differential geometry of foliations, J. Math. Mech. 11,
303–315.
References 89
Morrey, Jr. C.B. [1966] Multiple Integrals in the Calculus of Variations. Springer–
Verlag.
Nirenberg, L. [1959] On elliptic partial differential equations. Ann. Scuola. Norm.
Sup. Pisa 13(3), 115–162.
Omori, H. [1974] Infinite dimensional Lie transformation groups. Lecture Notes
in Mathematics, 427. Springer-Verlag.
Palais, R. S. [1965] Seminar on the the Atiyah - Singer Index Theorem. Princeton
University Press, Princeton, N.J.
Palais, R. S. [1968] Foundations of Global Non-Linear Analysis. Benjamin.
Rais, M. [1972] Orbites de la représentation coadjointe d’un groupe de Lie,
Représentations des Groupes de Lie Résolubles, P. Bernat, N. Conze, Mj.
Duflo, M. Lévy-Nahas, M. Aais, P. Renoreard, Mj. Vergne, eds. Monogra-
phies de la Société Mathématique de France, Dunod, Paris 4, 15–27.
Ratiu, T. S. [1982] Euler-Poisson equations on Lie algebras and the N -
dimensional heavy rigid body, Am. J. Math. 104, 409–448, 1337.
Ratiu, T. S. and R. Schmid [1981] The differentiable structure of three remarkable
diffeomorphism groups, Math. Zeitschrift 177, 81–100.
Sánchez de Alvarez, G. [1986] Thesis. University of California at Berkeley.
Sánchez de Alvarez, G. [1989] Controllability of Poisson control systems with
symmetry. Cont. Math. AMS 97, 399–412.
Simo, J. C., D. R. Lewis, and J. E. Marsden [1991] Stability of relative equilibria
I: The reduced energy momentum method, Arch. Rat. Mech. Anal. 115,
15-59.
Simo, J. C., T. A. Posbergh, and J. E. Marsden [1990] Stability of coupled rigid
body and geometrically exact rods: block diagonalization and the energy-
momentum method, Physics Reports 193, 280–360.
Simo, J. C., T. A. Posbergh, and J. E. Marsden [1991] Stability of relative equi-
libria II: Three dimensional elasticity, Arch. Rat. Mech. Anal. 115, 61–100.
Slawianowski, J. J. [1971] Quantum relations remaining valid on the classical
level. Rep. Math. Phys. 2, 11–34.
Stefan, P. [1974] Accessible sets, orbits and foliations with singularities, Proc.
Lond. Math. Soc. 29, 699–713.
Stein, E. M. [1970] Singular integrals and differentiability properties of functions.
Princeton Mathematical Series,30 Princeton University Press.
Sussman, H. [1973] Orbits of families of vector fields and integrability of distri-
butions. Trans. Am. Math. Soc. 180, 171–188.
Tulczyjew, W. M. [1977] The Legendre transformation. Ann. Inst. Poincaré 27,
101–114.
Varadarajan, V. S. [1974] Lie Groups, Lie Algebras and Their Representations.
Prentice Hall. (Reprinted in Graduate Texts in Mathematics, Springer-
Verlag.)
Weinstein, A. [1983] The local structure of Poisson manifolds. J. Diff. Geom. 18,
523–557
H. Whitney [1943] Differentiable even functions. Duke Math. J. 10, 159–160.
References 91