symplectic-geometry-lecture-notes
symplectic-geometry-lecture-notes
Gilles Castel
2 Symplectic manifolds 9
2.1 Vector bundles, vector fields, differential forms . . . . . . . . . . 9
2.2 Symplectic manifolds . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Hamiltonian vector fields . . . . . . . . . . . . . . . . . . . . . . 15
2.5 Hamiltonian Mechanics . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Almost Kähler Manifolds . . . . . . . . . . . . . . . . . . . . . . 18
5 Lie groups 44
5.1 Lie group actions . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.2 Moment maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Examples of moment maps . . . . . . . . . . . . . . . . . . . . . 48
5.4 Existence and uniqueness of moment maps . . . . . . . . . . . . . 53
6 Symplectic reduction 58
6.1 Symplectic reduction theorem . . . . . . . . . . . . . . . . . . . . 58
6.2 Examples of symplectic reduction . . . . . . . . . . . . . . . . . . 62
6.3 Reduction in stages . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.4 The convexity theorem . . . . . . . . . . . . . . . . . . . . . . . . 66
1
Lecture 1: Introduction
Lecture 1: Introduction
wo 12 feb 10:38
• Charlotte Kirchhofflukat
CONTENTS 2
Chapter 1
Example (Standard example of symplectic vector space). Consider R2n with stan-
dard basis coordinates (e1 , . . . , en , f1 , . . . , fn ), with Ωcan the unique antisym-
metric bilinear form s.t.
3
Lecture 1: Introduction
W Ω := {v ∈ V | Ω(v, w) = 0 ∀w ∈ W }.
Exercise. What is the symplectic orthogonal to the following spaces in (R2n , Ωcan )
• (R2n )Ω = {0}, because Ω is non-degenerate.
• {0}Ω = R2n , because Ω is bilinear.
• span(e1 , . . . , en )
• span(e1 , f1 )
• span(e1 , . . . , en , f1 )
Exercise. (W Ω )Ω = W for all subspaces W ⊂ V .
Exercise. Show that dim W + dim W Ω = dim V.
Exercise. W Ω ∩ W = 0
iff (W, Ω|W ×W ) is a symplectic vector space
iff (W Ω , Ω|W Ω ×W Ω ) is a symplectic vector space.
for some n.
Ω(v1 , w1 ) = Ω(v1 − v2 + v2 , w1 − w2 + w2 )
= Ω(v1 − v2 , w1 − w2 ) + Ω(v2 , w1 − w2 ) + Ω(v1 − v2 , w2 ) + Ω(v2 , w2 )
= Ω(v2 , w2 ).
is a scalar product:
• g(v, w) = g(w, v)
• g(v, v) ≥ 0 if v 6= 0.
Example. Consider (R2n , Ωcan ). Then the can. compatible complex structure is
defined as
J(ei ) = fi J(fi ) = −ei .
0 I
J= ,
−I 0
I 0
g(v, w) = v T w.
0 I
Exercise. Show all symplectic vector spaces admit a compatible complex struc-
ture.
Exercise. Show there are infinitely many complex structures on R2n that are
compatible with Ωcan . Idea: QT JQ with a Q in the symplectic linear group.
Exercise. If V is a vector space and J : V → V a complex structure and a scalar
product. Then we can define
Ω(v, w) = g(v, −Jw).
Show this is symplectic iff g(Jv, Jw) = g(v, w).
Exercise. Let J1 , J2 be compatible complex structure on (V, Ω). Assume Ω(v, J1 w) =
Ω(v, J2 w), then J1 = J2 .
So if we have two of (Ω, J, g), we can always determine the other one.
Exercise (Recommended!). U (n) = {A | A† A = 1}. Consider the isomorphism
of real vector spaces
Re ⊕ Im : Cn → R2n .
Show that the action of U (n) on R2n given by this isomorphism is by symplec-
tormorphism.
Lecture 2
wo 19 feb 11:06
Proposition 3 (Lagrangian Grassmannian). The set
Symplectic manifolds
Remark. The set of all derivations on an algebra A, Der(A) forms a Lie algebra.
[D1 , D2 ] = D1 ◦ D2 − D2 ◦ D1 .
9
Lecture 2
Definition 10. A Lie algebra (which is not an algebra in the above sense) is a
real vector space with an antisymmetric bilinear map [·, ·] : L×L → L called
the ‘Lie bracket’ which is not associative, but obeys the Jacobi identity:
or equivalently,
[a, [b, c]] = [[a, b], c] + [b, [a, c]],
which kinda looks like derivations.
Note that while this sum looks infinite, it is not in most of the cases.
Elements of Vn are called homogeneous elements of degree k.
a · b = (−1)kj (b · a).
LX (f ) = X(f ).
2. [iX , iY ] = iX iY + iY iX = 0a
3. [LX , LY ] = LX LY − LY LX = L[X,Y ]
6. [LX , iY ] = LX iY − iY LX = i[X,Y ]
a α(X, Y, . . .)= −α(Y, X, . . .)
b Useful for calculating LX α
Proof. Exercise.
1. Easy 4. Tedious
2. Easy 5. Easy
3. Fine 6. Easy
https://math.stackexchange.com/questions/2504737/proving-wikipedias-formula-on-interior
https://www.staff.science.uu.nl/~ban00101/lecnotes/lieder.pdf
Lecture 3
wo 26 feb 11:06
Example. Let M = R2n with the canonical basis with as coordinates (q1 , . . . , qn , p1 , . . . , pn ).
Then X
ω= dqi ∧ dpi
Then ωp = Ωcan .
Exercise. Show that there is a linear symplectomorphism (Tp R , ωp ) → (R , Ωcan )
2n 2n
Proposition 6. The tautological one form θ is the unique one form with
the property: For all one forms α ∈ Ω1 (Q), we can consider it as α : Q →
T ∗ Q : q 7→ αq , (α)∗ (θ) = α.
hτ, vi = hτ, α∗ wi
= hα∗ τ, wi
= hα, wi = hα∗ θ, wi = hθ, α∗ wi = hθ, vi ,
so these things agree on vectors that are not vertical. But these not-
vertical vectors form an open dense subset, so the forms agree.
Then
(q1 ◦ π, . . . , qn ◦ π, p1 , . . . , pn )
are coordinates for T ∗ Q|U . We often abuse notation and write (q1 , . . . , qn , p1 , . . . , pn ).
Example. When you find a manifold with H 2k = 0 for some k, the manifold
cannot have symplectic form. (If M is compact without boundary)
Exercise (Important). Proof this!
• Use stokes to show that ω n is not exact.
• Show that if [ω k ] = 0, then [ω k+1 ] = 0.
Proof. i[X,Y ] = LX iY − iY LX . So
In particular,
{pi , pj } = {qi , qj } = 0 {pi , qj } = δij .
X ∂f ∂g ∂f ∂g
{f, g} = − .
∂qi ∂pi ∂pi ∂qi
Exercise. Show that
• [Xf , Xg ] = X{f,g}
• {f, g} = −{g, f }
• {f, {g, h}} + {h, {f, g}} + {g, {h, f }} = 0
• {f, gh} = {f, g}h + g{f, h}.
?
• {f, ·} is a derivation on smooth functions. Note {f, ·} = ±LXf
• {f, g} = 0 if f is constant.
Exercise. Suppose ω1 , ω2 symplectic forms on M . Show that ω1 = ω2 iff Poisson
brackets agree.
Example. Consider R2n with any coordinates x1 , . . . , x2n . Suppose we have two
Poisson brackets of ω1 , ω2 . Show that the symplectic forms agree if the Poisson
brackets agree on the coordinate functions! Hint: use Taylors theorem.
LXH f = {f, H} = 0.
q̇ = p/m, q(0) = q0
ṗ = −∇V, p(0) = mv0 .
Then
1 2
H(q, p) = p + V (q).
2m
The previous equations can then we written as
Remark. Why are momenta covectors and not vectors? TODO. https://math.
stackexchange.com/questions/1324814/why-is-the-momentum-a-covector
Lecture 4
wo 04 mrt 11:04
Example. Let (Q, g) be a Riemannian manifold. Consider k · k2 : T Q → R :
v 7→ g(v, v). Let g [ : T Q → T ∗ Q : v 7→ g(v, ·). Let H : T ∗ Q → R, H = kg ] k =
k(g [ )−1 k. Then (T ∗ Q, Ωcan. , H) is Hamiltonian triple. Integral curves of this
system are geodescics.
LXf H = 0 ⇔ {f, H} = 0.
pends only on Xp , Yp !
Hint: show that it is C ∞ (M ) linear. (Hint: look at the proof that the torsion
tensor is a tensor!)
20
Lecture 4
γ p = (ρs )(q)
is injective. Not surjective: not all vector fields have a flow (only complete
vector fields.)
Exercise. Show this is surjective is M is compact. (Any vector field on a compact
manifold has a flow?)
Exercise (Bit tricky). Suppose ρt is an isotopy and ρt+s = ρt ◦ ρs for all 0 ≤
t, s, t + s ≤ 1. Then the corresponding X t is time independent.
Proposition 11. What if both forms are time dependent and we also have
an isotopy. Let αt be a smooth family of k-forms and ρt is an isotopy.a
d d
((ρt )∗ αt ) = ρ∗t LXt αt + αt .
dt dt
The left hand side: in a point, it is the usual derivative in a vector space.
a Actually this also holds when we just have a family of smooth functions that are
Proof. Idea of proof. Local coordinates: work, but lot of work. Better idea:
d d d
(ρ∗t αt ) = ρ∗t αt + ρ∗t αt .
dt t=t0 dt t=t0 dt t=t0
ρt (S 1 )
t t ∂
∂t
Xt
(ρt (p), t)
M = S1
p X0
https://math.stackexchange.com/questions/2608048/time-derivative-of-a-pullback-of-a-time
Homotopy operator
f1∗ − f0∗ = d ◦ Q + Q ◦ d : Ω• (N ) → Ω• (M ).
Lecture 5
wo 11 mrt 11:10
Aim of the upcoming sections:
• Mozer theorem
• Darboux theorem
• Weinstein Lagrangian neighbourhood theorem.
T M |X
Op ⊂ V ⊂ N X := of the zero section.
TX
such that for all p ∈ X, V ∩ Np X is convex (‘ball around zero section of all
fiber) and there exists a diffeomorphism φ : V → φ(V ) such that
φ
V M
i
0
X
commutes
1 1
Z
E(γ) = g(γ 0 (t), γ 0 (t)) dt.
2 0
expp : V ⊂ Tp M → M : v 7→ γv (1).
T M |X
E = {v ∈ Tp M | p ∈ X ⊂ M, v ∈ (Tp X)⊥ } ∼
= NX = .
TX
Define
Uε = {q ∈ M | dg(q, X) < ε}.
If ε is small enough, then we claim that Uε and Eε are diffeomorphic.
Indeed,
M
X
• i∗ : H • (U ) → H • (U ) is an isomorphism.
• If β ∈ Ωk (U ), dβ = 0 and i∗ β = 0 (so 0 on X) Then there exists
ξ ∈ Ωk−1 (U ) such that
dξ = β ξ|X = 0.
Q : Ω• (U ) → Ω•−1 (U )
Now
(π ∗ ◦ i∗ )β = 0,
so we get
β = dQ(β),
so we’ve shown already that β is exact. Now,
Z 1
Q(β) : = (It∗ ◦ i∂t ◦ H ∗ )(β) dt
0
Z 1
= ft∗ (iV β)dt,
0
Q(β)|X = 0 V |i(X) = 0.
Remark. Restriction of a form: β|X is not the same as pulling back with the
embedding. When pulling back, we also restrict tangent vectors to be along the
submanifold.
d ∗ d
ρt ωt = ρ∗t (LVt ωt + ωt ),
dt dt
where ρt is flow of Vt . We need to find Vt such that
d
LVt ωt + ωt = 0,
dt
as ρ0 = Id. Note that ωt − ωt0 is always exact by assumption. Fact (we
don’t prove this): ∃ηt ∈ Ω1 (M ) such that dηt = dt
d
ωt . a
Take Vt the unique time dependent vector field such that iVt ωt +ηt = 0.b
This implies that, using diX + iX d = LX and dω = 0,
d
0 = d(iVt ωt + ηt ) = LVt ωt + ωt .
dt
Then the flow of Vt , ρt satisfies ρ∗t ωt = ω0 .
a So we claim that when we take the limit t0 → t, the difference is still exact.
b Why does this exists? ω is non degenerate, we can invert it. So it is ωt−1 (−ηt ).
Lecture 6
wo 18 mrt 10:03
Theorem 5 (Local normal form). Let M be a manifold, i : X → M an
embedded compact submanifold. Let ω0 , ω1 be symplectic forms such that
ω0 |X = ω1 |X .
iVt ωt + η = 0.
f ∗ ω1 = ω0 .
Remark. Why local normal form. This show that if two symplectic manifolds
agree on a submanifold, then they will always agree in a neighbourhood around
that submanifold!
(q1 , . . . , qn , p1 , . . . , pn ) ∈ C ∞ (U ).
Remark. This means that every invariant in symplectic geometry must be global.
In Riemannian geometry, such thing, we cannot do! We can put the metric in
standard form at one single point, but not in a neighbourhood around a point!
The curvature prevents this.
Proof. As seen in one of the previous lectures, there exists a linear sym-
plectomorphism
Lp : (Tp M, ωp ) → (R2n , Ωcan ),
for all p. Let x1 , . . . , xn , y1 , . . . , yn ∈ C ∞ (U ) local coordinates around p.
such that under Lp ,
∂ ∂
= ei , = fi , Ω(ei , fj ) = δij .
∂xi p ∂y i p
This is just in one specific point p. Now we want to extend this to a full
neighbourhood. Write φ : R2n ⊃ V → U ⊂ M for a coordinate chart
corresponding to (xi , y i ). So
φ∗ ω|p = Ωcan .
dα∗ θ = dα
α∗ dθ = dα
∗
−α ωcan = dα,
Example. Consider (R2n , Ωcan ). You can view R2n = ∼ T ∗ Rn . Given a function
f ∈ R , then
n
∂f ∂f
df : Rn → R2n : (q1 , . . . , qn ) 7→ q1 , . . . , qn , (q), . . . , n (q) .
∂q1 ∂q
Proof. Exercise.
And dim M = dim graph. We need to show that graph f is isotropic iff
f ∗ ω2 = ω1 (symplectomorphism). Set Ω = π1∗ ω1 − π2∗ ω2 . Let f : M →
M × N : m 7→ (m, f (m)). This f is characterised by two properties:
• π1 ◦ f = IdM
• π2 ◦ f = f
So
∗ ∗ ∗
f ∗ ω2 = f π2∗ ω2 = f (π1∗ ω1 − Ω) = ω1 − f Ω.
∗
This equation gives the result. Indeed the graph being isotropic if f Ω = 0,
which will be precicely the case when f ∗ ω2 = ω1
Around any Lagrangian submanifold, the symplectic form can be brought
to a familiar standard form.
Proof. We’re doing the Riemannian geometry proof here, you can find other
proofs elsewhere. Choose a compactible almost structure J : T M → T M
for ω and write g = ω(·, J·).a Then we can a find a neighborhood U ⊂
(T L)⊥ ⊂ T M such that φ : exp |(T L)⊥ : U → M is an embedding.
Note, we can also embed U in T ∗ L as follows:
F : U ,→ T ∗ L : v 7→ ω(v, ·).
So this map is compatible with the splitting on the origin. We use this fact
to compute the pullback:
Remark. This seems lucky! Crucial that we choose the correct Riemannian
metric.
Lecture 7
za 28 mrt 14:51
Recall: if two symplectic forms agree on a submanifold, we can make them
agree in a neighbourhood with a symplectomorphism. Corollary of this is the
Darboux Theorem.
Today: local normal form stuff!
where we used iY ω = 0, and now writing the derivative with Cartan’s magic
formula, and using dω = 0, we get
−iY diX ω = 0.
Remark. Note that M/ ∼ where p ∼ q if they are in the same leaf, then this is
not necessary a smooth manifold!
Exercise. Suppose ∼ is an equivalence relation on smooth manifold M . Show
that there exists at most one smooth structure on M/ ∼ with makes M → M/ ∼
a smooth submersion.
ωred (π∗ (x), π∗ (w)) = ωred (π∗ (x − x + x0 ), π∗ (w)) = ωred (π∗ (x0 ), π∗ (w)).
Now we’ve showed that this is well defined and it fully determines ωred .
Now we show that ωred is a symplectic form.
Closed π ∗ : Ω0 (N ) → Ω0 (M ) is injective (because π is a surjective sub-
mersion) So
0 = dω = d(π ∗ ωred ) ⇒ dωred = 0.
Proof. (a) is an exercise, part (b) follows from the previous proposition.
• There exists (M̃ , ω̃) and embedding (N, ω) ,→ (M̃ , ω̃) such that i∗ ω̃ =
ω and N is embedded as a coisotropic submanifold.
• Given two such embeddings i1 , i2 there exists neighbourhoods U1 , U2
around the embedded N ’s a symplectomorphism φ such that i1 =
i2 ◦ φ.
So every presymplecic manifold embeds coisotropically into a symplectic
manifold. So the examples we’ve discussed before are the only examples in
some sense! This is like a neighbourhood theorem for coisotropics! But a little
weaker than Weinstein for examples, as that one gives a standard form. We
don’t yet have a standard form
T0p E ∗ ∼
= Ep ⊕ Gp ⊕ Ep∗ .
ωE (a ⊕ α, b ⊕ β) = β(a) − α(b).
T M |N ∼
= TN ⊕ E∗
|{z}
∼
= E ⊕ G ⊕ E∗,
∼
=N N viaΩ|N
(T M |X , (ω0 )|X ) ∼
= (T M |X , (ω1 )|X )
This is the local normal form theorem, but here, the symplectic forms
don’t have to agree, they just have to induce isomorphic symplectic
vector bundles!
This proves the theorem.
38
Lecture 8: Review of Lie groups, Lie-algebra cohomology, the adjoint
representation
Exercise. Covering space of a Lie group is also a Lie group. Suppose G is a Lie
group and G̃ is the covering space of G. Fix an element p ∈ G̃ which sits over the
identity of G. Show that G̃ has a unique Lie group structure for which p is the
identity and which makes the covering map G̃ → G a Lie group homomorphism.
(Remember that covering space can be tough of as homotopy classes of
paths, and what happens when you multiply two paths? How does the notion
of homotopy of a path interact with the group structure.)
Exercise. Suppose B ∈ gl(Rn ) is a matrix and X : GL(Rn ) → T GL(Rn ) is the
corresponding left-invariant vector field. Show that the flow φtX of X satisfies
Exercise. Recall that if X and Y are vector fields on a manifold, then the Lie
bracket [X, Y ] can be defined as follows:
d √ √ √ √
−
[X, Y ]p := φX t
◦ φ−
Y
t
◦ φXt ◦ φY t (p).
dt 0
dg : ∧g∗ → ∧g∗
X X
ddR (α)(. . .) = (−1)i LVi α(. . . V̂i ) + (−1)i+j+1 α([Vi , Vj ], . . . , V̂i , V̂j ).
i<j
Exercise. Let g be a Lie algebra and d the CE differential. This forms a complex!
Steps:
• Suppose η ∈ g∗ . Show that d2 η = 0. (Follows from Jacobi identity)
• Show that d is a graded derivation: η ∈ ∧k g∗ , ζ ∈ ∧l g∗ , then
d(η ∧ ζ) = dη ∧ ζ + (−1)k η ∧ dζ.
(Hint: Since both sides of the equations are linear maps, if suffices to show
this for basis vectors. And then induction)
• Use the previous parts to show that d2 = 0. (Hint: First apply this to
some basis vectors)
n {η ∈ ∧n g∗ | dη = 0} ker d
HCE (g) := = .
{dη | η ∈ ∧n−1 g∗ } im d
but we also mod out by everything, since that was the image, so HCE 2
(g) = 0.
Remark. A Lie group representation are a special kind of group actions. Take
A : G × V → V : (g, v) 7→ R(g)v.
ρ := de R : Te G → gl(V ) is a representation.
If you are given a Lie algebra representation, ρ, then we can look at forms
on the Lie algebra which take values in the vector space:
∧k g∗ ⊗ V := {antisymmetric maps gk → V }.
This is similar to the deRham differential, but we’ve replaced the Lie derivative
with the action by representation of vi .
One notable representation is the following:
• G acts on itself by
Cg (h) := ghg −1 .
This conjugation map is a Lie group homomorphism (whereas left
translation is not.)
• The Lie group homomorphism Ad : G → GL(g) defined by
Ad(g)v := (de Cg )v
Exercise. Let g be a Lie algebra. Using the formula ad(v)w = [v, w]. Show that
ad : g 7→ gl(g)
Lie groups
G×M →M (g, m) 7→ g · m.
Gp = {g ∈ G | g · p = p} ≤ G.
ρ : g → X(M ).
44
Lecture 9: Lie group actions, Moment maps
Remark. Why the minus sign? This is an artifact of some of the conventions
we chose (use of left-invariant vector fields).
Right invariant vector fields actually lie algebra anti homomorphism, which
is why we add a minus sign in fount of v. Because submersion (or some-
thing like that), (dΦ)∗ preserves Lie bracket. These two facts show that
(dΦ)∗ (X−v , 0) is a Lie algebra homomorphism.
To see this, first notice that for all g, h, p
Φ(h, Φ(g, p)) = Φ(Rg (h), p) ⇒ de,g·p Φ(w, 0) = d(g,p) Φ(de Rg (w), 0),
where the left hand side is the associativity, and the rhs is the derivative
w.r.t. h and p, leaving g fixed. Here w is a tangent vector at the identity,
and Rg is right multiplication with g.
Note that the rhs is de definition of a right invariant vector field. In-
serting w = −v, we get
which is to say
ρ(v) = (dΦ)∗ (X−v , 0).
Exercise. Suppose G has an action on M . Show that for all p ∈ M the set
G · p ⊂ M is an immersed submanifold of M , so its differential is injective.
Exercise. Suppose G has an action on M and ρ : g → X(M ) is the associated
Tp (G · p) = {ρ(v)p | v ∈ g}.
So not only immersion, but its tangent spaces arises as the image of the Lie
algebra under the infinitesimal action.
Example. Consider that action of (R, +) on C by rotation.
t · z := eit z.
Note that ρ(v) is clockwise rotation, while eit z is counter clockwise rotation!
This is a consequence of the minus sign in the definition of infinitesimal action.
µ : g → C ∞ (M ) : ci vi 7→ ci fi .
µ(v)(p) = hJ(p), vi v ∈ g, p ∈ M.
Therefore, we get a smooth function J : M → g∗ .
Remark. If you have an infinitesimal action with this property, all the informa-
tion is in J, i.e. you can completely recover ρ.
C ∞ (M )
µ f 7→−Xf
ρ
g X(M )
v · (q, p) = (q + v, p),
so the action is translating the position. Then the infinitesimal generator be-
comes (mind the sign swap):
∂
ρ(v1 , v2 , v3 ) = −v i .
∂q i
Recall that ∂
∂q i = Xpi . Then we can define:
µ : R3 → C ∞ (R3 ) : µ(v1 , v2 , v3 ) = v i pi .
Alternatively,
J : R3 × R3 → R3 : J(q, p) = p.
So the moment map is the projection to the momentum coordinates!
This example shows why the moment map is called the moment map. In
similar rotational context, the moment map is the angular momentum!
Lecture 10
di 28 apr 15:52
were we used ω = dα, Cartans magic formula. For the last step,
remember that the action preserved the one form, so Lρ(v) α = 0. So
we have that i−ρ(v) ω = dµ(v). This is the definition of hamiltonian
vector field.
• Now equivariance.
Now, we have
so this becomes
= idφ−1 φ∗ α(p)
g ρ(v) g
= iρ(Adg−1 v) α(p)
= J(p), Adg−1 v
= Ad∗g J(p)v .
J : S 2 → (g → R) : (θ, z) 7→ (a 7→ az)
µ : g → (S 2 → R) : a 7→ ((θ, z) 7→ az).
and
−d(µ(v)) = −d((θ, z) 7→ vz) = −vdz.
So when you contract ρ(v) with ω, we get dz, which is exact, so we find
that ρ(v) is Hamiltonian.
Here we see that the moment map is kinda like the angular momentum?
We think of angular momentum being orthogonal to the plane of rotation,
and here we project on the z-component.
Important example:
1 manifold is two dimensional, so automatically closed.
2 Equivariance for free because Lie algebra is abelian.
Remark. How should think of this? You want a two form, which is something
that when you give it two vectors, it gives you a vector. So a two form, evaluated
at some point in O ⊂ g∗ , you want to come up with a number given two vectors
tangent to the orbit. But a vector tangent to the orbit is exactly a vector that lies
in the image of the infinitesimal generator. So the ingredients we’re provided
with are two elements of the lie algbra and a point of the orbit, which is a
covector. The most natural thing you can do is hη, [v, w]i. The remarkable thing
with this is this is actually a symplectic form! Astounding! Deep connection
between symplectic geometry and Lie theory.
Given any Lie algebra, the coadjoint orbits are symplectic manifolds.
Exercise. Let G be a Lie group and suppose that g is its Lie algebra. Let
ρ : g → X(g) be the infinitesimal generator of the coadjoint action of G on g∗ .
Show that ρ satisfies the following equation:
Proof. Note that for all points η ∈ O and a vector ṽ ∈ Tη O, there exists
an element of you lie algebra, v ∈ g such that ρ(v)η = ṽ. So we can always
treat vectors as coming from the Lie algebras.
d
Lρ(v) ω(ρ(w), ρ(z))(η) = ω(ρ(w), ρ(z))(g(t) · η)
dt 0
Definition of KKS symplectic form
d
= hg(t) · η, [w, z]i
dt
0
d
= g(t) · η, [w, z]
dt 0
= hρ(v)η , [w, z]i
By previous exercise
= hη, [v, [w, z]]i .
Equip so(3)∗ with the dual metric to the one defined above. Show that
the coadjoint orbits of G are the unit spheres relative to this metric. Finally
compute the KKS symplectic form in the x, y, z coordinates. You should get
the standard symplectic form on the unit sphere in R3 .
Why is this in the section about moment maps?
Proposition 26. Consider the coadjoint action of a Lie group G on the dual
of its Lie algebra g∗ . Let O be an orbit of this action and ω the KKS form
on O. Then the inclusion O ,→ g∗ is a moment map for the transitive
action of G on (O, ω).
This is kind of amazing!
µ(v)(η) := hη, vi .
g ∈ G, f ∈ C ∞ (M ), p ∈ M (g · f )(p) := f (g −1 p).
• The fact that this mu is invariant under this action corresponds pre-
cisely to the fact that J is G-equivariant.
Note that because this is H 1 , there are no exact things. So going back to
the definition, being closed corresponds to annihilating commutators. So
the first property implies the second property.
Proof of the first property.
Firs, we know that for all v ∈ g, we have Xµ1 (v) = Xµ2 (v) . This implies
that µ1 (v) = µ2 (v) + c(v) for c : g → R ⊂ C ∞ (M ), where ⊂ is i : R →
C ∞ (M ) : c 7→ (p 7→ c). This is because the kernel of the map that sends
something to its Hamiltonian vector field is price cloy the locally constant
functions. Because the manifold is connected, these are globally constant
functions.
To finish, we need to show that c annihilates commutators, i.e.
Since H 1 (g) = [g, g]◦ = 0, we know that [g, g] = g. This means that
everything is the bracket of something. This means that ρ(u) = ρ([v, w])
for some v, w, which is a Hamiltonian vector field.
∀v ∈ g : ∃f ∈ C ∞ (M ) : ρ(v) = −Xf .
ρ = −Xµ ⇒ Xc(v,w) = 0,
Why?
• This new µ still satisfies the first moment map condition. Because
`(v) should be thought of as spitting out constant functions, so taking
the Hamiltonian of µ(v) is the same as taking that of µ(v) + `(v).
• Why Lie algebra morphism? Exactly because of the definition of `.
Check this yourself!
Symplectic reduction
t · (x, y, z, w) 7→ (z + t, y, z, w).
Then M/G = R3 which is not even even dimensional. And this action is even
Hamiltonian!
The goal of symplectic reduction is to use moment maps to construct new
symplectic manifolds out of group actions!
∀p ∈ M gp := {v ∈ g : ρ(v)p = 0} = ker ρp .
Gp := {g ∈ G : g · p = p}.
58
Lecture 11: Symplectic reduction theorem, examples
dp J : Tp M → TJp g∗ ∼
= g∗ .
Then
• ker dp J = (Tp (G · p))ω , so ω(ker dp J, Tp (G · p)) = 0.
Remark. The first bullet point implies that the tangent space to the fibers of J
are symplectic orthogonal with respect to the orbits of G.
(“dp Jx” = push the vector down to the Lie algebra) To see why the formula
is true, we refer to the following calculation:
ω(ρ(v), x) = ix iρ(v) ω
= ix d µ(v)
= Lx µ(v)
d
= µ(v)(p(t)) p0 (0) = x
dt t=0
d
= hJ(p(t)), vi
dt t=0
= hdp Jx, vi ,
in the last step we used that hJ(p(t)), ·i is a linear map and the derivative
of such a thing is its differential (?).
Exercise. Complete the proof of the previous Lemma, which should be pretty
straightforward.
dm f : Tm M → Tn N
i π
(M, ω) ←−− (J −1 (η), i∗ ω = π ∗ ωred ) −−→ (J −1 (η)/Gη , ωred ).
∀w ∈ Tp J −1 (η) ω(v, w) = 0 ⇔
• Closed? Submersion and you pull back closed two form, you get
a closed two form, because pullback operation is compatible with
differential.
Proof. The only thing to check is that 0 is a regular value of J. Since the
action on J −1 (0) is free, it follows that gp = 0. Therefore, dp J is surjective
for all p ∈ J −1 (0).
So S 1 is the stabiliser of the covector, but the Lie algebra is abelian, so the
stabilisor is everything, so we are modding out the whole unit circle. Another
way to think of it: what part of the group that’s acting preserves this level set?
And that’s everything.
Note that it was very important that we consider the fiber here. If we would
just consider Cn+1 /S 1 , we would end up with an odd-dimensional manifold.
Another natural example of symplectic reduction comes from any group
action.
J −1 (0)
h
quotient
J −1 (0)/G T ∗ (Q/G)
and the bottom left right arrow because the fibers are the same. Now we
need to check that these arrows are a symplectomorphism
Exercise. Recall the tautological 1-form α on T ∗ (Q/G). Let α̃ denote the
tautological 1-form on T ∗ Q. Show that:
α̃|J −1 (0) = h∗ α.
Lecture 12
di 12 mei 10:28
6.3 Reduction in stages
Definition 61. Suppose G1 , G2 are Lie groups that act on M . The actions
are said to commute if
∀g ∈ G1 , h ∈ G2 , p ∈ M g · (h · p) = h · (g · p).
Exercise. Suppose G1 and G2 have well defined actions on (M, ω) with moment
maps J1 : M → g1 and J2 : M → g2 resp. Show that there is a well defined
action of the product group G1 × G2 on M . Show that the dual of the Lie
algebra of G1 × G2 is canonically isomorphic to g∗1 ⊕ g∗2 . Finally, show that
J1 ⊕ J2 : M → g∗1 ⊕ g∗2 is a moment map for this action.
Given commuting Hamiltonian actions, one can define a Hamiltonian action
on the reduced spaces. So the following lemma says that when you first perform
Hamiltonian reduction w.r.t one of the action, then you still have a Hamiltonian
action on the reduced space with the other group.
J2 : J1−1 (0)/G1 → g∗ .
• Exercise.
1
The level set J1−1 (0) is the 5-dimensional unit sphere. As we saw inn a
previous lecture, J1−1 (0)/G1 = CP 2 . This way, we get:
1 1 z1 z1 −1
J2 ([z1 : z2 : z3 ]) = − P J2 (0)/G2 = CP 1 .
2 2 zi zi
moment maps can differ by constant. Here we tried to arrange it such that the level set of 0
is exactly the sphere of radius one! And on the right we have that the level set gives sphere
of norm 21 .
where
Remark. (So this way we can see in a convoluted way: acting by rotation on a
torus is not a Hamiltonian action, because it doesn’t have fixed points!
For a proof by Ana Rita Pires, see http://pi.math.cornell.edu/~apires/
convexitytalk.pdf
Remark. You should remember the statement of this theorem! The proof is
involved, uses a inversion of Darboux’s theorem, and also a bit of morse theory!
Example. Recall the action of S 1 on S 2 where the symplectic form in cylindrical
coordinates is given by dθ ∧ dz.
ψ · (θ, z) := (θ + ψ, z) J(θ, z) = z.
The fixed points of the action are the north and south pole pn , ps . So, as S 1 is
a torus, and M is compact,
[1 : 0 : 0] [0 : 1 : 0] [0 : 0 : 1].
(|z1 |2 , |z2 |2 )
J : CP 2 → R2 : [z1 : z2 : z3 ] 7→ − .
2|(z1 , z2 , z3 )|2
You can compute this by first looking at the action of θ1 and then, because sum
of, . . . , you get (. . . , . . .).
(− 12 , 0) = J([1 : 0 : 0])
(0, 0) = J([0 : 0 : 1])
Figure 6.1
This picture gives you a good look at the regular values of the moment map.
In this case, they will be the interior points of this triangle.
Now, an application of this:
The image of p is
Proof. Idea of the proof. Show that p is the moment map for a Hamiltonian
Tn action. How? The first step is to show that there is an isomorphism
∗
u(n) with Herm(n) and that under this identification the coadjoint repre-
sentation of U (n) on u(n)∗ is by conjugation.
And we know stuff about coadjoint actions! Orbits are symplectic man-
ifolds. And coadjoint action is Hamiltonian action, where moment map is
just the inclusion.
And every Hermitian matrix can be diagonalized by a unitary matrix.
It follows that
I1 0
{A ∈ Herm(n) : Eigval(A) = {I1 , . . . , In }} = U (n) · ..
. ,
0 In
so it equal to the coadjoint orbit of the diagonal matrix. Note that this
dot is not matrix multiplication, but conjugation! So the domain of p is a
symplectic manifold! And there is also a sorta moment map by including
it into the set of hermitian matrices.
But to apply the theorem, we need a torus action. How can we find
such a thing? Well, inside of U (n) there is a torus Tn with Lie algebra t
(diagonal matrices.)
Restricting the Lie group U (n) to just the torus, it’s no longer transitive,
but that action is Hamiltonian. The Hamiltonian action is inclusion in
hermitian matrices, and then projection to t∗ . The projection map g∗ →
t∗ = Rn corresponds to
A11 · · · A1n
.. .. .. 7→ (A , . . . , A ).
. . . 11 nn
An1 ··· Ann