Benn - The L2 Geometry of The Symplectomorphism Group
Benn - The L2 Geometry of The Symplectomorphism Group
A Dissertation
Doctor of Philosophy
by
James Benn,
James Benn
2015
All Rights Reserved
THE L2 GEOMETRY OF THE SYMPLECTOMORPHISM GROUP
Abstract
by
James Benn
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
CHAPTER 2: PRELIMINARIES . . . . . . . . . . . . . . . . . . . . . . . 8
2.1 Diffeomorphism Groups . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.1 Manifolds of Mappings . . . . . . . . . . . . . . . . . . . . 8
2.1.2 Groups of Diffeomorphisms . . . . . . . . . . . . . . . . . 11
2.1.3 Exponential Mappings . . . . . . . . . . . . . . . . . . . . 12
2.2 Weak Riemannian Structure on the Diffeomorphism Group . . . . 14
2.3 The Symplectic Diffeomorphism Group Dωs . . . . . . . . . . . . 19
2.3.1 The Manifold Structure of Dωs . . . . . . . . . . . . . . . . 19
2.3.2 Hodge Theory for Manifolds . . . . . . . . . . . . . . . . . 24
2.3.3 Weak Riemannian Structure for Dωs (M ) . . . . . . . . . . 33
2.4 L2 Geodesic Completeness . . . . . . . . . . . . . . . . . . . . . . 36
2.5 Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
ii
s
CHAPTER 5: CONJUGATE POINTS ON DHam (M n ) IN DIMENSIONS
n = 2 AND 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.2 A Conservation Law . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.3 A Characterization of Conjugate Points along Stationary Geodesics 120
5.4 Proof of Proposition 5.3.1 . . . . . . . . . . . . . . . . . . . . . . 122
BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
iii
ACKNOWLEDGMENTS
A very special thanks goes out to my friend and advisor, Gerard Misiolek,for
introducing me to the subject and for his endless ideas. Thankyou to Alex Hi-
monas for his ongoing support and to Steve Preston for his interest and useful
suggestions. My family, my friends; I thank all of you deeply.
iv
CHAPTER 1
Z
(u, v)L2 = g (u, v) dµ, u, v ∈ Te Dωs , (1.1)
M
defines a right-invariant metric on the group. This thesis is concerned with the
L2 geometry of the group Dωs , and it’s finite codimensional subgroup DHam
s
- the
group of Hamiltonian diffeomorphisms.
Diffeomorphism groups can be realized as the configuration spaces of a number
of equations in mathematical physics, which provides a strong motivation to study
their geometry. Perhaps the most famous example is the Euler equations of hy-
1
drodynamics, where Arnold, [A], noticed that a curve η(t) in the group of smooth
volume preserving diffeomorphisms (Dµs ) is a geodesic of the L2 metric (1.1) if and
only if the vector field v, defined by ∂t η = v ◦ η, solves the Euler equations of
hydrodynamics. The L2 metric is simply the kinetic energy of the fluid, and the
geodesic equation is a manifestation of Newton’s second law F = ma.
Analogously, a curve η(t) in Dωs (M ) is a geodesic of the L2 metric starting
from the identity in the direction vo if and only if the time dependent vector field
v = η̇ ◦ η −1 on M solves the Symplectic Euler equations
Lv ω = 0
v(0) = vo ,
where Peω is the orthogonal projection onto the space of Symplectic vector fields.
The subgroup of Hamiltonian diffeomorphisms plays a role in plasma dynamics
analogous to the role played by the volume preserving diffeomorphism group in
incompressible hydrodynamics, see Arnold and Khesin [A-Kh], Holm and Tronci
[H-T], Morrison [Mo], Marsden and Weinstein, [M-W], for details.
Chapter 2 contains a review of the manifold structure of mapping spaces and
diffeomorphism groups. We briefly describe the shortcomings of the group expo-
nential map on diffeomorphism groups and motivate the endowment of a weak
Riemannian structure (the L2 metric (1.1)) on these manifolds. Section 2.3 fo-
cuses our attention on the Symplectomorphism and Hamiltonian subgroups (and
submanifolds!) of the diffeomorphism group. Here we recall the Hodge decompo-
sition of forms and the fundamental results of Ebin and Marsden, [E-M], which
2
prove the existence of a smooth right-invariant connection and exponential map-
ping associated to the weak L2 metric. We define the weak Riemannian curvature
tensors on the Symplectomorphism group and indicate that they are trilinear op-
erators bounded in the strong Sobolev H s topology, as shown in Misiolek [M1].
Finally, in section 2.4 we review Ebin’s [Eb] and Khesin’s [Kh] proof that solutions
to the Symplectic Euler equations (1.2) exist globally in time for any Symplectic
manifold, so that the group Dωs (M ) is L2 geodesically complete.
Arnold ([A]) computed sectional curvatures of diffeomorphism groups and
found that they were mostly negative, although in some small regions they were
positive. He asked if there are conjugate points on Dµs and called for a description
of them. Much progress in understanding conjugate points on Dµs has been made
since the work of Misiolek. In [M1], some simple examples of conjugate points
in the Volumorphism group were constructed, answering Arnold’s first question
in the affirmative. More examples were later provided by Misiolek [M2], Preston
[P2], and Shnirelman [Sh2].
In contrast with finite dimensional geometry, two types of conjugate points can
occur in infinite dimensions. Grossman [Gro] gave the first examples of the two
types of conjugacies that may occur: on a sphere in Hilbert space the differential
of the exponential map may have infinite dimensional kernel corresponding to an
infinite-dimensional family of geodesics joining two antipodal points. In addition,
on an infinite dimensional ellipsoid the exponential map differential fails to be
surjective, even though it is injective, in certain directions.
A natural question to ask is whether conjugate points exist on Dωs . This is the
contents of chapter 3: Some simple examples of conjugate points are constructed
on the Symplectomorphism group of the complex projective plane. In particular,
3
n
Theorem. (3.4.3) Conjugate points exist on Dωs (CP n ), for s > 2
+ 1 and n ≥ 2.
We then show that geodesics which lie in the isometry subgroup of Dωs always
carry conjugate points, all of which have even multiplicity.
Theorem. (3.6.2, 3.6.4) Let η(t) = exp(tvo ) be a geodesic of the L2 metric (1.1)
generated by a Killing vector field vo . Let J(t) be a Jacobi field along η(t), with
initial conditions J(0) = 0, J 0 (0) = wo . Then
etKv0 − I
J(t) = Dη(t) · e−tKv0 w0 ,
Kv0
where Kvo (·) is a compact, skew self-adjoint operator on Te Dωs , and we have the
tKv R itλ
spectral representations e−tKvo = R eitλ dE(λ) and e Kv0 −I = R e iλ−1 dE(λ) which
R
0
4
This result provides a distinction between Symplectic diffeomorphisms and
Volume preserving diffeomorphisms, when equipped with the L2 metric. It holds
for any closed Symplectic manifold of dimension 2n, but fails for the group Dµs
of manifolds of dimension 3 and higher, cf. [E-M-P], [P2]. The relationship
between Fredholmness of the L2 exponential map and known classifications (e.g.
C 0 closure, Gromov’s non-squeezing Theorem) of Symplectic diffeomorphisms is,
at this point, unclear.
In chapter 5 we give a new geometric characterization of conjugate points along
a stationary geodesic, and relate their existence to the right-invariance of the L2
metric.
where Ad∗η(t) is the formal L2 adjoint of the push forward of vector fields operation,
Kvo (·) a compact skew self-adjoint operator on Te Dωs , and S(t) is the solution
operator of the linearized Symplectic Euler equations.
That is, conjugate points occur when a solution to the linearized Symplectic
Euler equations (with initial value v), which can be thought of as a curve in Te Dωs ,
intersects the coadjoint orbit of its initial value v which can also be thought of as
a curve in Te Dωs . Moreover, we are able to express solutions of the Jacobi equation
in terms of solutions to the linearized Symplectic Euler equation and coadjoint
5
orbits. Namely,
where {vi }i∈N is a complete orthonormal set of eigenvectors of Kvo spanning Te Dω0 ,
P
Kvo vi = λi vi , S(t)wo = i gi (t)vi solves the linearized Symplectic Euler equations
and Ad∗η−1 (t) wo =
P
ai (t)vi .
∂t v + ∇v v = 0
v(0) = vo
describes the motion of a collection of particles moving without any internal forces.
It is also the geodesic equation of the L2 metric on the full diffeomorphism group.
It has been shown by Khesin and Misiolek, [Kh-M], that the Burgers’ equation
has solutions in which particles begin colliding with one-another, forming shock-
waves, in finite time. That is, geodesics (i.e. particle trajectories) cease to be
diffeomorphisms (or reach the boundary of the group) in finite time. The first
conjugate point along a geodesic generated by such a solution signals the onset of
6
shock-waves in the material space. On the other hand, solutions of the 2D Euler-
equations of hydrodynamics (and the Symplectic Euler equations (1.2)) exist for
all time and yet some of the corresponding geodesics contain conjugate points.
7
CHAPTER 2
PRELIMINARIES
The basic idea of giving a manifold structure to mapping spaces was first laid
down by Eells [E1] in 1958 where he constructed a smooth manifold out of the
set of continuous maps between two manifolds. Constructing a manifold from C k -
diffeomorphisms of a compact manifold without boundary was done independently
around 1966 by Abraham, Eells and Leslie, [E2], [L]. The Sobolev H s case was
done by Ebin and Marsden [E-M] where they gave a manifold structure to the
H s diffeomorphism group, the Volume-preserving diffeomorphism subgroup and
the Symplectic diffeomorphism subgroup of a compact manifold with, or without,
boundary.
The construction is as follows. Let M and N be two compact manifolds each
endowed with a Riemannian metric, g and h, and suppose N is without boundary.
For an integer s, a map f : M → N is of Sobolev class H s (write f ∈ H s (M, N )) if
for any point p ∈ M and chart around p, (Up , ϕ), and any chart (V, ψ) around f (p)
the composite map ψ◦f ◦ϕ−1 : ϕ(U ) → Rn is in H s (ϕ(U ), Rn ). If the sobolev index
m
s satisfies s > 2
then by the Sobolev embedding Lemma, H s (M, N ) ⊂ C 0 (M, N ),
and the above notion is well-defined and independent of the charts chosen. We
8
refer the reader to appendix A for a review of Sobolev spaces.
In order to define charts on H s (M, N ) we need to determine the correct mod-
eling space for H s (M, N ). Just as in finite dimensions, we use the tangent space
as the model space for a manifold and we shall proceed similarly here. With
this in mind, we shall look for a good description of the tangent space at a point
f ∈ H s (M, N ).
Consider a curve c : (−, ) → H s (M, N ) such that c(0) = f . For a point
p ∈ M , the map t 7→ c(t)(p) is a curve in N . Now c(0)(p) = f (p) and so the
d
derivative of this curve at 0 is dt
c(t)(p)|t=0 and is an element of Tf (p) N . Therefore,
d
the map p 7→ dt
c(t)(p)|t=0 is a map from M to T N and such that the canonical
projection πN : T N → N covers f .
Making the identification
d d
c(t)|t=0 (p) = c(t)(p)|t=0 ,
dt dt
Tf H s (M, N ) = {X : M → T N : X ∈ H s (M, T N ), πN ◦ X = f } .
Here, H s (M, T N ) is the space of all sections from M to T N which have L2 deriva-
tives up to order s. Define the inner product
XZ
g ∇k V, ∇k W dµ
(V, W )s =
|k|≤s M
9
m
sections are continuous by the Sobolev Lemma (since s > 2
) and the topology of
H s (M, T N ) is stronger than that of uniform convergence.
In order to construct an f -centered chart for H s (M, N ) we use the Riemannian
exponential map of N . Since N is closed, it is geodesically complete and for each
x ∈ N , the exponential map expN x : Tx N → N is defined on the whole of Tx N .
Consequently, expN x can be extended to a map expN : T N → N , where for
vx ∈ Tx N , expN (vx ) = expN x (vx ). Since f (M ) is compact, there is a number
λf > 0 such that any point of N whose distance from f (x) is less than λf can be
joined by a unique geodesic arc of length less than λf . That is, for any point p of
N whose distance from f (x) is less than λf there is an X(x) ∈ Tf (x) N which lies
in the disk of radius λf centered at 0, such that expN
f (x) X(x) = p and the map
x 7→ expN
f (x) X(x)
Ψ : Tf H s (M, N ) → H s (M, N )
X 7→ expN X
10
2.1.2 Groups of Diffeomorphisms
The diffeomorphism groups have a very rich and complicated structure which
is still not very well understood. Let M be a compact manifold without boundary
n
and assume that s > 2
+ 1 so that the H s topology is stronger than C 1 . Let C 1 D
be the group of C 1 diffeomorphisms of M , to itself, and let Ds (M ) = H s (M, M ) ∩
C 1 D. According to Theorem 1.7 of Hirsch’s Differential Topology, [Hi], C 1 D is
open in C 1 (M, M ), so that Ds (M ) is open in H s (M, M ) and hence inherits its
manifold structure. Ds (M ) is also a topological group with composition as the
group operation.
Right multiplication is smooth:
Rη : Ds → Ds
ξ 7→ ξ ◦ η.
˙
Indeed, let t 7→ ξ(t) be a curve in Ds with ξ(0) = ξ, ξ(0) = X, then dξ Rη (X) =
d d
| (Rη (ξ(t)))
dt t=0
= | (ξ(t)
dt t=0
◦ η) = X ◦ η which is another right translation.
However, Left multiplication
Lη : D s → D s
ξ 7→ η ◦ ξ
11
If G is a Lie group and e ∈ G the identity element then the Lie algebra of G is
identified with Te G. Similarly, Te Ds (M ) = H s (T M ), which are H s vector fields
on M , serves as the Lie Algebra of Ds (M ). Since right multiplication is smooth
we are able to talk about right-invariant vector fields on Ds . Given any vector
field v on M , we define a right-invariant vector field vη on Ds (M ) by the formula
vη = v ◦ η.
Since these are fields of class C 1 we are able to define the Lie bracket. The Lie
bracket [uη , vη ] is calculated as
However, since u and v are both in H s , the bracket is only H s−1 and hence Te Ds
is not closed under the bracket operation.
v 7→ η1
12
is defined. Here, η1 is the value of the one-parameter subgroup ηt = exp (tv)
corresponding to t = 1.
However, the group exponential mapping has a number of shortcomings. It
is not even a homeomorphism in a neighborhood of the identity. There exist
diffeomorphisms arbitrarily close to the identity which are not embeddable in
a flow ([Ko]). This was shown for even the simplest case of M = S 1 . The
group exponential map is in fact much worse! In the work of Grabowski ([G1])
it was shown that for the group of compactly supported diffeomorphisms of a C k
manifold (k = 0, 1, 2, ...) there exist arcwise-connected, non-trivial free subgroups
of diffeomorphisms which embed in no flow.
It is shown in [E-M] that, just like a Lie group, the group of Sobolev H s dif-
feomorphisms of a closed, orientable manifold M admits an exponential mapping
which associates to every tangent vector at the identity a one parameter subgroup
of Ds (M ). Recall that such a tangent vector is an H s vector field on M and the
one parameter subgroup is the flow generated by the vector field.
n
Theorem 2.1.1. Let M be a compact manifold without boundary, s > 2
+ 2 and
Ds (M ) the group of H s diffeomorphisms.
We refer the reader to [E-M] for the proof of this Theorem. Note that the exponen-
tial mapping is not C 1 because it does not cover any neighborhood of the identity.
Since smooth diffeomorphisms embed densely in the set of H s diffeomorphisms the
13
above results regarding local injectivity and surjectivity of the exponential map-
ping apply to the exponential mapping of the H s diffeomorphisms: there exist
diffeomorphisms arbitrarily close to the identity which embed in no flow. More-
over, there exist arcwise-connected, non-trivial free subgroup of diffeomorphisms
which embed in no flow.
The solution to this problem is to use an exponential map associated to a
weaker metric defined on Ds (M ). It is not automatic that a weak metric admit
an exponential mapping but this turns out to be the case on Ds (M ).
Z
(uη , vη )L2 = g(u, v) (η −1 )∗ dµ. (2.1)
M
14
and uniqueness theorem of ODE’s in Banach space, if F is smooth in both it’s
arguments η and η̇ then there exists a unique solution η(t) defined on an open
interval around 0 ∈ R. It would then follow that around every point x ∈ N there
exists a neighbourhood U of x and a number ε > 0 so that for each p ∈ U and
each tangent vector v ∈ Tp N with length less than ε there is a unique geodesic
η : (δ1 , δ2 ) → N
η(0) = p η̇(0) = v.
expN
p (v)
and called the exponential of the tangent vector v. The geodesic η(t) can then be
described by
η(t) = expN
p (tv)
and expN
p is called the exponential map.
15
¯ associated to it; that is,
2. (2.1) has a unique torsion-free affine connection ∇
for smooth vector fields u, v and w on Ds (M ), we have
¯ u v, w ¯ uw
u (v, w)L2 = ∇ L2
+ v, ∇
¯ uv − ∇
∇ ¯ v u = [u, v] .
Z
(uη , vη )L2 = g(u, v) (η −1 )∗ dµ
M
16
hence its inverse) is a volume-preserving diffeomorphism.
If uη and vη are right-invariant vector fields on Ds (M ) then we have the follow-
ing formula for the covariant derivative in terms of of the fields u, v ∈ Te Ds (M ):
¯ uη vη = (∇u v) ◦ η
∇
This formula follows from the above covariant derivative formula, cf. [E-M] for
details.
Using the formula for the covariant derivative of right-invariant vector fields
and the formula for covariant derivative of vector fields along curves, we are able
¯ η̇ η̇ = 0. Letting v(t) be defined
to write down the geodesic equation in Ds (M ): ∇
by η̇(t) = v(t) ◦ η(t) we obtain
∂v
+ ∇v v = 0
∂t
17
Remark 2.2.2. The geodesic equation on Ds (M ) is the Burgers’ equation, which
describes the motion of a cloud of particles moving without any internal forces. It
is not always true that a weak metric yields geodesics (and hence an exponential
map). For if M has non-empty boundary, the weak metric may yield geodesics
which could try to cross the boundary of M :
Consider the unit disk in R2 , described by polar coordinates as
1 0
with metric g = . The christoffel symbols are given by
2
0 r
Γrθθ = −r Γθθr = r.
1
Let v(r, θ) = so that in components v r = 1 and v θ = 0, and v points in the
0
direction of the boundary of D2 . Then, in coordinates
X ∂v k
i k i i
∇v v = v i
+ Γij v v ∂k
i
∂x
X
Γkij v i v i ∂k
=
i
=0
18
Since v is constant and points in the direction of the boundary of D2 , particles
will cross the boundary of D2 in finite time. Particles initially on the boundary of
D2 leave the disk instantaneously.
Dωs (M ) = {η ∈ Ds (M ) : η ∗ ω = ω}
div ω = 0
19
Therefore, the Lie algebra of Dωs (M ) can be identified with
Te Dωs = {v ∈ H s (T M ) : div ω = 0} .
Dωs (M ) = {η ∈ Ds (M ) : η ∗ ω = ω}
The main ideas in the proof are as follows, see [E-M] for more details. If ω is
the Symplectic form on M then consider its H s cohomology class
[ω]s = ω + dα : α ∈ H s+1 (T ∗ M ) ,
Define a map
Ψ : Ds (M ) −→ [ω]s
η 7−→ η ∗ (ω).
If we can show that the map Ψ is a submersion onto [ω]s then the pre-image of ω
under Ψ is a submanifold. We do this in three steps:
20
3. Show that DΨ is surjective.
1. We write Ψ in the following way: For a curve η(t) in Ds (M ) with η(0) = e and
vector field v defined by η̇ = v ◦ η
Z t
∗ d
Ψ(η(t)) = η(t) (ω) = ω + η(t)∗ ωdt
0 dt
Z t
=ω+d η(t)∗ iv ωdt ∈ [ω]s
0
d
DΨ(X) = |t=0 η(t)∗ ω = Lv ω = div ω
dt
s
To each element in Te DHam there corresponds a unique element of H0s+1 . Also, each
element of H0s+1 uniquely determines an element of Te DHam
s
. So we have an iso-
21
s
morphism between the sets Te DHam and H0s+1 (M ) and we denote this isomorphism
by
s
T : H0s+1 (M ) → Te DHam (2.2)
F 7→ J∇F
The Lie bracket of any two globally Hamiltonian vector fields is again a glob-
ally Hamiltonian vector field. To the space of globally Hamiltonian vector fields
there corresponds the “Lie subgroup” of Hamiltonian diffeomorphisms denoted by
s
DHam (M ), see [E-M] or, alternatively the work of Ratiu and Schmid [R-S].
s
DHam (M ) = {η ∈ Ds (M ) : η ∗ ω = ω}
Adη v = Dη · v ◦ η −1 .
22
−1 T
= J Dη T ∇F ◦ η −1 = J Dη −1 ∇F ◦ η −1
= J∇ F ◦ η −1 ,
where in the second equality we used that Dη is a Symplectic matrix: DηJDη > =
J.
Define the Poisson bracket of two functions F, G : M → R by
If η is a Symplectomorphism, then
{F, G} ◦ η = ω(vF , vG ) ◦ η
= (η ∗ ω) (vF ◦η , vG◦η )
= ω(vF ◦η , vG◦η )
= {F ◦ η, G ◦ η}.
That is, Symplectomorphisms preserve the Poisson bracket. It follows that the
Poisson bracket satisfies the Jacobi identity:
23
Indeed, let η(t) be a geodesic in Dωs . Then, differentiating both sides of
satisfying
d2 = 0.
X
hα1 ∧ . . . ∧ αk , β1 ∧ . . . ∧ βk i = (sgnπ) gx α1 , βπ(1) · . . . · gx αk , βπ(k) ,
π
24
where π ranges over the set of permutations of {1, ...k}. Consequently, there is an
L2 inner product on k-forms (i.e. sections of Λk ) given by
Z
(α, β) = hα, βi ω n . (2.5)
M
On one forms, the inner product (2.5) is the same as the L2 metric (2.1).
There is a first order differential operator
?? = (−1)k(dim M −k) .
α ∧ ?β = hα, βi ω n
δ = (−1)dim M (k+1)+1 ? d ? .
25
The δ operator corresponds to the classical divergence operator div. Define a
map g [ : T M → T ∗ M which assigns to each vector field X the one form iX g,
where g is the metric on M . This is an isomorphism from the tangent bundle
to the cotangent bundle of M . The inverse is defined as g ] : T ∗ M → T M and
corresponds to contracting with the inverse components of the metric tensor. Let
X be a vector field on M with corresponding 1-form g [ (X). Let LX µ be the Lie
derivative of the volume form µ on M . Then, by definition
divX · µ = LX µ
and the divergence measures how much the volume form µ changes in the direction
of the flow of X. Cartan’s formula for the Lie derivative gives
26
is defined by
−4 = dδ + δd,
and is a self-adjoint (in the L2 metric), second order differential operator which
commutes with ?.
The Laplacian is a well-defined operator acting on forms. In the case of func-
tions
4 : H01 (M ) → H0−1 (M ).
Using the Poincare inequality, whose proof can be found in Evans [Ev],
we are able to define an inner product which is equivalent to the H 1 inner product
and involves only the derivatives of functions but not the functions themselves:
Consequently,
(−4F, F )L2 ≥ C kF k2H 1 (2.6)
0 0
so that
k4F kH −1 ≥ C kF kH 1 . (2.7)
0 0
27
Theorem 2.3.3. The Hodge Laplacian is an isomorphism from H01 (M ) to H0−1 (M ).
Proof. The estimate (2.7) implies that 4 is injective with closed range. Indeed,
If 4Fn is a Cauchy sequence in the range of 4, then the inequality
1
kFn − Fm kH 1 ≤ k4Fn − 4Fm kH −1
C
(4F, G)L2 = 0
28
Then F ∈ H k+2 and
kF kH k+2 ≤ C kGkH k ,
4 : H0k+2 (M ) → H0k (M )
is an isomorphism.
29
Then there is an α ∈ H s+1 (Λk−1 ), a β ∈ H s+1 (Λk+1 ) and γ ∈ C ∞ (Λk ) such that
ζ = dα + δβ + γ
so that dα, δβ and γ are uniquely determined. The projection maps onto the
first, second, and third summands are continuous; that is, each subspace is closed.
Moreover, the space of harmonic forms H is finite dimensional.
H s (T ∗ M ) = dH s+1 (T ∗ M ) ⊕ δH s+1 (T ∗ M ) ⊕ H.
X = ω ] (dδα + δdβ + γ)
where α and β are H s+2 1-forms and γ is a harmonic form. More generally, the
30
Lie algebra of Ds (M ) decomposes as
Recall that the Lie algebra Te Dωs is given by locally Hamiltonian vector fields
(vector fields X which satisfy diX ω = dω [ (X) = 0). Therefore,
Te Ds = Te Dωs ⊕ ω ] (δdH s (T ∗ M ))
Define a projection
Peω : Te Ds → Te Dωs
which makes P ω right-invariant and is correct since the metric (2.1) restricted to
Dωs is right-invariant.
Lemma 2.3.6. The projections Peω and Qωe , given by orthogonal projection onto
31
the first and second summands of (2.9), respectively, are given by
Qωe = −ω ] δ4−1 dω [ .
Proof. For any v ∈ Te Dωs we have ω [ (v) = dδβ + σ = d4−1 δω [ (v) + σ for some
β ∈ H s+2 (T ∗ M ) and σ ∈ H. Hence the orthogonal projection Peω : Te Ds → Te Dωs
can be written as
where PeH denotes the projection onto the finite-dimensional space of harmonic
forms which we henceforth neglect. Using the compatible structure J we may
further write
The projection Peω is a C ∞ bundle map (see [E-M] Appendix A). An alternative
reason for this can be seen by defining another metric on Ds by
s s
(u, v)s = (u, v)L2 + 4 2 u, 4 2 v L2
where 4 is the Hodge Laplacian. This metric is smooth and by regularity prop-
erties of the Laplacian can be shown to generate the H s topology on Ds . The
Hodge decomposition is orthogonal with respect to this stronger metric and it
32
follows that the projection P ω is smooth. A proof along these lines may be found
in [E].
We also obtain an L2 orthogonal splitting for the space of Hamiltonian vector
fields:
Tη Ds = Tη DHam
s
⊕ ω ] δdH s+2 (T ∗ M ) ◦ η ⊕ H ◦ η
(2.10)
PeH : Te Ds → Te DHam
s
Pω = PH + PH
where P H is the orthogonal projection onto the finite dimensional space of Har-
monic vector fields.
Theorem 2.3.7. ([E-M], Theorem 8.5 (i)(ii)) Let M be a closed Symplectic man-
ifold with Symplectic form ω. Then the metric (2.1) defined on Dωs (M ) is a smooth
Dωs -right-invariant weak Riemannian structure. On Dωs , (2.1) induces the affine
connection
¯
∇ω = P ω ◦ ∇,
33
¯ is the connection of Theorem 2.2.1, and the exponential mapping
where ∇
expω . Moreover, ∇ω and expω are invariant under right-translations.
s
The metric (2.1) restricted to DHam (M ) is a smooth Dωs -right-invariant weak
s
Riemannian structure. On DHam (M ), (2.1) induces the affine connection
¯
∇H = P H ◦ ∇
and the exponential mapping expH . Moreover, ∇H and expH are invariant
under right-translations.
s
Following Smolentsev, [Smo], define another inner product on Te DHam by
Z
hvF , vH i = F Hdµ. (2.11)
M
so that the weak L2 metric (2.1) on vector fields is related to the inner product
(2.11) by
(vF , vH )L2 = hv−4F , vH i . (2.12)
Define a map
T 4 : Te DHam
s s−2
→ Te DHam (2.13)
vF 7→ v4F
34
group of a closed Riemannian manifold M with volume form µ. Just like the
Symplectomorphism group, Dµs is a smooth submanifold of Ds whose tangent
space at the identity consists of divergence free vector fields on M . According to
the Hodge decomposition (Theorem 2.3.5),
For any v ∈ Te Ds we have g [ (v) = dδα + δdβ + σ for some α, β ∈ H s+2 (T ∗ M ) and
σ ∈ H. Then
δg [ (v) = δdδα = −4δα.
= ∇4−1 divv.
35
In complete analogy to Theorems 2.3.1 and 2.3.7 we have
Te Dµs = {v ∈ Te Ds : divv = 0} .
The metric (2.1) defined on Dµs (M ) is a smooth Dµs -right-invariant weak Rieman-
nian structure. On Dµs , (2.1) induces the affine connection
¯
∇µ = Peµ ◦ ∇,
We shall use the orthogonal projection Peω of Lemma 2.3.6 to write down the
equations describing geodesics in the submanifold Dωs . Recall that a geodesic on a
manifold is a curve η(t) such that the acceleration η̈(t) is identically zero. In the
case of Ds (M ) and a geodesic η(t) with flow equation
we have
0 = η̈(t) = ∂t η̇(t) = ∂t (v(t) ◦ η(t))
36
¯ η̇ (v(t) ◦ η(t)) = (∂t v + ∇v v) ◦ η(t)
= (∂t v(t)) ◦ η(t) + ∇
∂t v + ∇v v = 0. (2.15)
Peω (η̈(t)) = 0
or
∂t v + Peω (∇v v) = 0. (2.16)
Equation (2.16) is known as the Symplectic Euler equation and who’s study is
relatively new. By Lemma 2.3.6
Therefore, if η(t) is a geodesic in Dωs , with η̇(t) = v(t)◦η(t), it satisfies the equation
37
η̈(t) = ω ] δ4−1 dω [ ∇η̇◦η−1 η̇ ◦ η −1 = F (η, η̇). (2.19)
Theorem 2.4.2. For any vo ∈ Te Dωs there exists t1 and t2 and a unique curve
η : (t1 , t2 ) → Dωs such that η̇ is an integral curve of (2.19).
We now outline the main idea in showing that the C 1 norm of v remains
bounded in time.
Definition 2.4.4. [Kh] Define the Symplectic vorticity for a symplectic vector
field v to be the 2n form defined by ξ := dα ∈ Ω2n (M ), where 2n = dimM , v and
α are related by α = g [ (v) ∧ ω n−1 .
38
Lemma 2.4.5. The Symplectic vorticity, ξ = dg [ (v) ∧ ω n−1 , is given explicitly as
ξ(t) = 4F (t)ω n
g [ ω ] = −g [ J 2 ω ] = −g [ g ] ω [ g ] ω [ ω ] = −ω [ g ]
and therefore
g [ (v) = −ω [ g ] dF − ω [ g ] (h)
so that
g [ (v) ∧ ω n−1 = −i∇F ω n − ig] h ω n .
Lemma 2.4.6. The Symplectic Euler equation (2.18) has the vorticity formulation
or
∂t (4F (t) ◦ η(t)) = 0, (2.21)
where {·, ·} is the Poisson bracket and η(t) is the solution of η̇(t) = v(t) ◦ η(t),
39
and v(t) = vF (t) + h(t) solves (2.18).
Proof. Recall that g [ (∇v v) = Lv g [ (v) + d |v|2 , where Lv is the Lie derivative and
|v|2 = g(v, v). Applying dg [ to both sides of (2.18)
Wedging both sides with ω n−1 , and observing the Lv ω n−1 = 0 we get
To simplify the right hand side we use properties of the almost complex structure
J as in Lemma 2.4.5. Namely, g [ ω ] = −ω [ g ] so that
∂t ξ(t) + Lv ξ(t) = 0.
∂t 4F ω n + Lv 4F ω n = 0 (2.22)
40
Since ω is time-independent and v is divergence free, equation (2.22) becomes
∂t 4F + Lv 4F = 0. (2.23)
Therefore the vorticity and the vorticity function are both invariant under the
flow. This shows that the C 0 norm and all Lp norms of ξ(t) are constant in time.
In particular, since ξ = d g [ (v) ∧ ω n−1 we are able to obtain global estimates on
the derivatives of v.
We also mention a special type of geodesic called a stationary geodesic. A
geodesic is said to be stationary if its velocity vector v defined by η̇(t) = v ◦ η(t) is
independent of time. That is, v is a time independent solution of the Symplectic
Euler equations (2.16) and satisfies both
Peω ∇v v = 0 Qωe ∇v v = ∇v v.
41
solves the vorticity equation (2.21):
∂t 4Fo + {4Fo , Fo } = 0,
which reduces to
{4Fo , Fo } = 0. (2.25)
η(t)∗ Fo = Fo ◦ η(t) = Fo .
Indeed,
∂t η(t)∗ Fo = g(∇Fo , vFo ) ◦ η(t)
= 0.
2.5 Curvature
In view of Theorem 2.3.7 we are able to define the weak Riemann curvature
tensor of Dωs and DHam
s
:
42
ReH : Te Dωs × Te Dωs × Te Dωs → Te Dωs
Rηω (uη , vη )wη = (∇ωu ∇ωv w)η − (∇ωv ∇ωu w)η − ∇ω[u,v] w
η
.
For any v and w ∈ Te Dωs , we define the sectional curvature of the plane σ
spanned by v and w by
We are only interested in the sign of Keω (v, w) so we will ignore the normalizing
factor and compute only with (Reω (v, w)w, v)L2 . The following Lemma is from
[P4].
Lemma 2.5.2. Let v and w ∈ Te Dωs . Then the sectional curvature of the plane
spanned by v and w is given by
Z Z
Keω (v, w) = n
g(Aw v, v) ω + g(Peω ∇v w, Peω ∇v w) ω n , (2.27)
M M
43
where A is a linear operator depending on w.
we compute
Z Z
(Reω (v, w)w, v)L2 = g(∇ωv ∇ωw w, v) ω n − g(∇ωw ∇ωv w, v) ω n
M M
Z
− g(∇ω[v,w] w, v) ω n
M
Z Z
= g(∇v ∇ωw w, v) ω n − g(∇w ∇ωv w, v) ω n
M M
Z
− g(∇[v,w] w, v) ω n
M
Z Z
= g(∇v ∇w w − ∇v Qωe ∇w w, v) ω n − g(∇w ∇v w − ∇w Qωe ∇v w, v) ω n
M M
Z
− g(∇[v,w] w, v) ω n
M
Z Z Z
n
= g(R(v, w)w, v) ω − g(∇v Qωe ∇w w, v) ω n + g(∇w Qωe ∇v w, v) ω n
M M M
Z Z Z
= g(R(v, w)w, v) ω n + g(Qωe ∇w w, Qωe ∇v v) ω n − g(Qωe ∇v w, Qωe ∇w v) ω n
M M M
Z Z Z
= g(R(v, w)w, v) ω n + g(Qωe ∇w w, Qωe ∇v v) ω n − g(Qωe ∇v w, Qωe ∇v w) ω n
M M M
Note that the first two terms involve integrals of quantities which are bilinear in
v. This will be useful for their point wise analysis in later chapters. The last term
44
is non-positive and we write it as ∇v w = Peω ∇v w + Qωe ∇v w:
Z Z
(Reω (v, w)w, v)L2 = g(R(v, w)w, v) ω − n
g(∇v Qωe ∇w w, v) ω n
M M
Z Z
n
− g(∇v w, ∇v w) ω + g(Peω ∇v w, Peω ∇v w) ω n
M M
and the first three terms are integrals of quantities bilinear in v. Thus, we write
Z Z
(Reω (v, w)w, v)L2 = g(Au v, v) ω + n
g(Peω ∇v w, Peω ∇v w) ω n .
M M
X
Aij = i
wk wl + ∇j wj ∇j wi − gkl ∇j wk ∇i wl ,
Rjkl
k,l
45
Proof. The trace of the operator Aw may be computed:
i
Aii = Rikl wk wl + ∇i wj ∇j wi − gkl g mi ∇i wk ∇m wl
i
wk wl + ∇i wj ∇j wi + ∇i wk ∇i wk
= Rikl
i
wk wl + ∇i wj ∇j wi − wj ∇i ∇j wi − ∇i wj ∇i wj
= Rikl
i
wk wl + ∇i wj ∇j wi − wj ∇j ∇i wi + Rijk
i
w k − ∇ i w j ∇ i wj
= Rikl
= ∇i wj ∇j wi − ∇i wj ∇i wj
= −g ik g jl ∇i wj ∇l wk − g ik g jl ∇i wj ∇k wl .
1
= − g ik g jl (Sij + νij ) Skl
2
1
= − g ik g jl Sij Skl
2
since
g ik g jl νij Skl = g jl g ik νji Slk = −g ik g jl νij Skl
46
In particular, if Lv g(p) 6= 0 then the trace of A is strictly negative.
Proof. Let S be the symmetric part of B; then clearly g(w, Sw) = g(w, Bw) and
the trace of S is equal to the trace of B. Thus we may assume B is symmetric.
P
Choose an orthonormal basis {vi } such that B = diag{λ1 , . . . λn } with i λi =
0. Consider the set of 2n vectors defined by
1 1
J= √ (1 v1 + · · · + n vn ) : 2i = 1, ∀i = n
√ : ∈ {−1, 1} .
n n
n n
X 1 X X 1 X X
g(w, Aw) = aij i j = aij i j .
n n i,j=1
n i,j=1
w∈J ∈{−1,1} ∈{−1,1}n
X 2n
g(w, Aw) = Trace(A)
w∈J
n
and therefore the average value of g(w, Aw) over J is n1 Trace(A). So at least one
element of J must have
1
g(w, Aw) ≤ Trace(A).
n
47
CHAPTER 3
3.1 Introduction
D expe (w)
fails to be an isomorphism.
Definition 3.1.1. Let η(t) be a geodesic in Dωs with η(0) = e and η̇(0) = vo . The
point η(t∗ ) is said to be conjugate to η(0), t∗ ∈ (0, t], if the linear operator
48
gate point even if dim ker D expe (t∗ vo ) = 0. Therefore two types of singularities of
the exponential map can occur in infinite dimensions. Following Grosman ([Gro]),
we have
Let η(t) = expωe (tvo ) be a geodesic in Dωs with η̇(t) = v(t) ◦ η(t). Define
v(s, t) = tvo (s), where vo (s) is a variation of the initial condition vo depending on
s. Then, η(s, t) = expe (tvo (s)) is a variation of geodesics, with
∂v(s, t)
|s=0 = two ,
∂s
∂η(s, t) ∂
u(t) ◦ η(t) = |s=0 = |s=0 expe (tvo (s)) = D expe (tvo )two .
∂s ∂s
The vector field u(t) is called the variation field, or Jacobi field, of the variation
η(s, t) of geodesics.
The purpose of this chapter is to construct examples of conjugate points on Dωs
and describe their distribution along a certain class of geodesics. To find examples
of conjugate points, the above observation shows that it is enough to construct
a variation of geodesics such that the variation field is 0 at t = 0 and 0 at some
later time t∗ . The point η(t∗ ) will then be monoconjugate to e.
Section 3.2 we collect the properties of some useful operators defined on Te Dωs .
It is a classical result of Riemannian geometry that manifolds of non-positive
curvature do not contain conjugate points; in 3.3 we compute sectional curvatures
49
in Dωs to assist in locating conjugate points along geodesics. Explicit examples
of conjugate points are constructed in 3.4 and a complete characterization of
conjugate points along geodesics generated by Killing vector fields is given in
section 3.5 and 3.6.
In this section we will define and compute the adjoint and coadjoint operators
on Te Dωs which will be used throughout the rest of this text.
Every Lie group has two distinguished representations: the adjoint and coad-
joint representations. Any element g in a group G defines an automorphism cg of
the group G by conjugation:
cg : h 7→ ghg −1 .
The differential of cg at the identity e ∈ G maps the Lie algebra g to itself and
therefore defines an automorphism of g called the group adjoint operator
Adg : g → g.
Although the diffeomorphism group Ds is not exactly a Lie group we may still
define the group adjoint operator. Indeed, let h(t) be a curve of diffeomorphisms
d
with h(0) = e and | h(t)
dt t=0
= X ∈ Te Ds . Then, for any η ∈ Ds
d d
cη∗ X = |t=0 cη h(t) = |t=0 Rη−1 Lη h(t)
dt dt
= dLη dRη−1 X = Dη · X ◦ η −1
50
which is the usual pushfoward operation on vector fields. In the last inequality
we have used that dLη = Dη - the Jacobian matrix of η - and dRη is again a right
translation by η. Consequently
Adη : Te Ds → Te Ds (3.1)
X 7→ Dη · X ◦ η −1 .
The differential of Adg at the group identity g = e defines the algebra adjoint
representation
adu v : g → g
d
adu v = |t=0 Adη(t) v
dt
where η(t) is a curve on the group G issued from the identity η(0) = e with
d
velocity | η(t)
dt t=0
= u.
The algebra adjoint representation on Te Ds is given by the negative of the
usual Lie bracket of vector fields. Indeed, let η(t) be a curve in Ds with η(0) = e
d
and velocity | η(t)
dt t=0
= u. Also, let h(s) be a curve of diffeomorphisms with
d
h(0) = e and | h(s)
ds s=0
= v ∈ Te Ds . The diffeomorphisms η(t) and h(s) can be
expressed to first order in s and t (in local coordinates) as
51
As t → 0, we recognize the term sv (x − tu(x) + o(t)) as the directional derivative
∂v
s v(x) − tu(x) ∂x , so we write
−1 ∂v
h(s)η (t) : x 7→ x − tu(x) + s v(x) − tu(x) + o(s) + o(t).
∂x
Now
−1 ∂v
η(t)h(s)η (t) = x − tu(x) + s v(x) − tu(x) +
∂x
∂v
+tu x − tu(x) + s v(x) − tu(x) + o(t) + o(s) + o(s) + o(t)
∂x
∂v
= x − tu(x) + s v(x) − tu(x) +
∂x
∂v
+tu x + sv(x) − tu(x) − stu(x) + o(s) + o(t)
∂x
∂v
As s → 0, along with t → 0, we recognize the term tu x + sv(x) − tu(x) − stu(x) ∂x
as the directional derivative tu(x) + stv(x) ∂u
∂x
so that
∂u ∂v
= x + s v(x) + t v(x) − u(x) + o(t) + o(s).
∂x ∂x
Since
d d
adu v = |t=0 |s=0 cη(t) h(s)
dt ds
it follows that
∂u ∂v
adu v = v(x) − u(x) = −[u, v]. (3.2)
∂x ∂x
The group adjoint operator gives a representation of the group on its algebra.
The orbits of the group G in its algebra are called the adjoint orbits of G. The
52
algebra adjoint is given by the rate at which the group adjoint leaves the iden-
tity, the velocity of an adjoint orbit. In the next Proposition we will describe
the group Adjoint and algebra Adjoint operators on the Symplectomorphism and
Hamiltonian diffeomorphism groups.
∗
Adη v = ω ] d δα ◦ η −1 + g] η −1 ? h ∧ ω n−1
, (3.3)
where ? is the Hodge star operator, while the Lie algebra adjoint adu :
Te Dωs → Te Dωs is given by
53
where u = J∇G and {F, G} is the Poisson bracket defined by (2.3)
Proof. Let G be the Symplectomorphism group Dωs (M ) with Lie algebra Te Dωs of
locally Hamiltonian vector fields of the form v = ω ] (dδα + h) = J∇ (δα) + ω ] (h).
Then, using formula (3.1)
T −1
T
∇ (δα) ◦ η −1 + Adη ω ] (h) = J Dη −1 ∇ (δα) ◦ η −1 + Adη ω ] (h)
= J Dη
= J∇ δα ◦ η −1 + Adη ω ] (h) ,
where in the second equality we used that Dη is a Symplectic matrix (cf. Appendix
A). To compute Adη ω ] (h), let β be any one-form. Then
Z Z
]
β η∗ ω ] (h) ◦ ηdµ
β η∗ ω (h) dµ =
M M
Z Z
∗ [ ]
∗
η β, ω [ g ] (h) dµ
= η β, g ω (h) dµ = −
M M
since g [ ω ] = −g [ J 2 ω ] = −ω [ g ] .
Z Z
∗
η ∗ β ∧ ? ig] h ω dµ
=− η β, ig] h ω dµ = −
M M
Z Z
∗ n−1
β ∧ ? η −1 ? h ∧ ω n−1 dµ
=− η β∧h∧ω dµ = −
M M
Since this holds for any 1-form we obtain (3.3). The algebra adjoint (3.4) follows
from
ω [ (adv w) = −i[v,w] ω = div iw ω = ω [ J∇ω(v, w).
The result for the Hamiltonian diffeomorphism group follows from the result
54
on the Symplectomorphism group.
Recall that for any Hilbert space H with inner product (·, ·) there exists a unique
operator T ∗ on H such that
(T f, g) = (f, T ∗ g)
and the Lie algebra adjoint ad∗u : Te Dωs → Te Dωs is defined so that
From here on when we refer to the coadjoint operators we mean the operators
defined by (3.7) and (3.8).
• If G = Dωs (M ) with Lie algebra Te Dωs , then the group coadjoint action is
given by
Ad∗η = J∇4−1 (4δα ◦ η) − g ] ? η ∗ h ∧ ω n−1
(3.9)
55
The Lie algebra coadjoint action is given by
or
ad∗vG vF = v4−1 {4F,G} (3.13)
Proof. Let G be the Symplectomorphism group so that g is the set of vector fields
v = J∇F + h for a function F on M and a harmonic vector field h. For the Lie
algebra coadjoint
(ad∗u v, w)L2 = (v, adu w)L2
Z Z Z
= hv, J∇ω(u, w)i dµ = − ω(u, w)div(Jv)dµ = hdiv(Jv)Ju, wi dµ
M M M
= (div(Jv)Ju, w)L2
Since this holds for any w ∈ Te Dωs we conclude that ad∗u v = P ω (div(Jv)Ju),
proving (3.10).
Suppose G is the Hamiltonian diffeomorphism group. Letting vF and vG ∈
56
s
Te DHam
Z Z
∇F, ∇ G ◦ η −1 dµ = − (4F ) G ◦ η −1 dµ
=
M M
Z Z
=− (4F ◦ η) Gdµ = − 44−1 (4F ◦ η) Gdµ
M M
Z
J∇4−1 (4F ◦ η) , J∇G ,
=
M
ad∗vG vF , vH
L2
= (vF , advG vH )L2
Z
= − v4F , v{G,H} = − {G, H} · 4F dµ
M
Z Z
= ω (vH , vG ) · 4F dµ = g(vH , ∇G) · 4F dµ
M M
consequently
ad∗vG vF = Pe (4F · ∇G)
which yields (3.11). Using the formula for the projection P contained in Lemma
2.3.6 we have
ad∗vG vF = −J∇4−1 divJ (4F · ∇G)
57
= v4−1 {4F,G}
Kv : Te Dωs → Te Dωs
We will now record some properties of Kv which will be useful in later sections
and chapters.
= − (ad∗u v, w)L2
= − (Kv u, w)L2 .
58
Since this holds for any w ∈ Te Dω0 it follows that Kv is skew self-adjoint.
Lemma 3.2.4. Suppose v(t) solves the Symplectic Euler equations (1.2). Then
the operator Kv(t) satisfies
Proof. Since Kv X = adv X = ad∗v X = 0 for any harmonic vector field X, it suffices
s
to prove the Lemma for the subspace Te DHam . Let v = vF (t) solve the Symplectic
s
Euler equations (1.2) and vG ∈ Te DHam . In view of formula (3.13), (3.6) and (3.10)
The next two Lemmas show that the operator Kv is a compact operator for
any v ∈ Te Dωs . Since the space of Hamiltonian vector fields is of finite codimension
in Te Dωs , it suffices to prove the Kv is compact on Te DHam
s
. Let H = ∩s≥0 H0s+1 (M )
be the subspace consisting of smooth functions on M with zero mean and observe
that H is dense in H0s+1 (M ). Given F ∈ H0s+1 (M ), let Fk be a sequence of smooth
functions in H approximating F in the H s+1 norm. Consequently, vFk is a smooth
sequence of H s vector fields approximating vF in the H s norm.
For σ ≥ 0, let Te Dωσ denote the closure of the space of locally Hamiltonian
59
vector fields in the H σ norm. By the Hodge decomposition this is a closed subspace
dim M
in the space of all H σ vector fields ([M]). For σ > 2
+ 1 this coincides with
the actual tangent space to Dωσ . However, for smaller σ the group DHam
s
is not
necessarily a smooth manifold.
dim M
Lemma 3.2.5. Let s > 2
+ 1, s ≥ σ + 1 and let F and {Fk }k∈N be as above.
Then, KvFk → KvF in the H σ norm.
σ
Proof. Let vH ∈ Te DHam . We estimate
KvF vH − KvFk vH
= kPe (4F · ∇H − 4Fk · ∇H)kH σ
Hσ
. kF − Fk kH σ+2 · kvH kH σ
. kF − Fk kH s+1 · kvH kH σ
dim M s
Lemma 3.2.6. Let s > 2
+ 1, s ≥ σ + 1. For any vector field vF ∈ Te DHam
σ
the operator KvF defined in Proposition 4.2.2 is compact on Te DHam .
60
as
KvF (vH ) = P (4F · ∇H) = P (H · ∇4F )
s
since the projection of a gradient field vanishes. Then, for any vH ∈ Te DHam ,
. kvH kH σ
Therefore the map vH 7→ KvF (vH ), as a map from H σ vector fields to H σ+1
vector fields, is compact by the Rellich embedding Theorem (Appendix A Theorem
A.2.2).
Our goal in this section is to understand the sectional curvature of the group
Dωs and use these heuristics to locate examples of conjugate points. The Lemmas of
section 2.5 will be of use here. First we will describe the second fundamental tensor
of the Symplectomorphism group inside the Volume preserving diffeomorphism
group.
Let M be a closed Symplectic manifold with compatible Riemannian metric
g. It is useful to consider the group Dωs as a subgroup of the group of Volume-
preserving Diffeomorphism group, Dµs .
61
The Levi-Civita connection on Dµs = Dωs n at the identity is given by
¯
∇µ = Peµ (∇), (3.15)
For each η ∈ Dωs , the tangent space Tη Dµs = g ] (δH s+1 (Λ2 T ∗ M ) ⊕ H) admits
an orthogonal splitting
Tη Dµs = Tη Dωs ⊕ (Tη Dωs )⊥
where (Tη Dωs )⊥ is the normal space to Dωs at η, in Dµs . (Tη Dωs )⊥ consists of H s
sections X of the pull-back bundle η ∗ T M for which the corresponding vector field
u = X ◦η −1 on M is not locally Hamiltonian but is divergence free. Let (T Dωs )⊥ =
∪η∈Dωs (M ) (Tη Dωs )⊥ and π : (T Dωs )⊥ → Dωs where π(Tη Dωs ) = η, ∀η ∈ Dωs .
Any divergence free vector field Z : Dωs → T Dµs along Dωs decomposes uniquely
as Z = Z > + Z ⊥ , where Z > is an H s vector field tangent to Dωs and Z ⊥ is an H s
normal field to Dωs .
˜ µ to be the restriction of the Levi-Civita connection to T Dµs |Ds , π, Dωs .
Define ∇ ω
Then
∇ ˜ ωX Y + α(X, Y )
˜µ Y = ∇ (3.17)
X
where
⊥
α(X, Y ) = ∇˜ XY (3.18)
62
for all H s vector fields X and Y on Dωs . For η ∈ Dωs ,
⊥
αη (X, Y ) = ∇˜ XY ,
η
is because the projections are orthogonal so that by first projecting onto Dµs we
remove any gradient parts of the covariant derivative. Then using Qωe we remove
any locally Hamiltonian parts of the covariant derivative so that what we are left
with is an element of the complement of Te Dωs in Te Dµs .
¯ XY )
α(X, Y ) = Qω P µ (∇
¯ v w, for v, w ∈ Te Dωs . Then Peµ (u) = Peω (u) + α(, v, w). Recall
Proof. Let u = ∇
that
Qωe = ω ] δ4−1 dω [
63
We compute
α(v, w) = P µ (u) − P ω (u)
= Qω (u) − Qµ (u)
where we have used the formula δ (f α) = f δα + i∇f α, for any C 1 function f and
any k-form α, and noting that δω = 0. Continuing,
64
= −ω ] δ4−1 dω [ (I − Qµ ) (u)
= Qω P µ (u).
The next result is from [M1] and [P4] and gives a general criterion for a geodesic
in Dµs to have non-negative curvature in all plane sections along it which contain
its tangent vector.
Z Z
Keω (v, w) = n
g(Aw v, v) ω + g(Peω (∇v w), Peω (∇v w)) ω n
M M
65
generates a stationary geodesic η in Dωs (M ) such that the curvatures of the two-
dimensional planes spanned by η̇ and any vector field along η are non-negative.
In fact,
for any w ∈ Te Dµs (M ), where Rω is the curvature tensor of Dωs (M ) at the identity.
Proof. Let v ∈ Te Dωs be a Killing vector field. Let w be any other C ∞ vector field
on M . We compute
1
= −g(v, ∇w v) = − w · g(v, v)
2
1
g(− ∇ |v|2 , w)
2
1
∇v v = − ∇ |v|2 .
2
Since a gradient vector field is orthogonal to Te Dωs in the L2 metric by the Hodge
decomposition, Theorem 2.3.5, we have
1
Peω (∇v v) = Peω (− ∇ |v|2 ) = 0
2
66
so that v generates a stationary solution to the Symplectic Euler equations (1.2).
We use the Gauss equations; for any u, v, w, y ∈ Te Dωs ,
(Reω (u, v)w, y)L2 = (Reµ (u, v)w, y)L2 + (α(v, w), α(u, y))L2
(Reω (v, w)w, v)L2 = (Reµ (v, w)w, v)L2 + (α(w, w), α(v, v))L2
Proof. By Lemma 2.5.3, TraceAv (p) < 0 and by Lemma 2.5.4 we can find a
67
vector field u ∈ Tp M such that g(u, u) = 1, g(u, Av u) < 0, and g(u, ∇u v) = 0.
p
Let w = ∇u v ∈ Tp M , and a = g(w, w). We first define normal coordinates
(x1 , y1 , x2 , y2 , ..., xn , yn ) in a neighborhood Ω of the point p such that gij |p = δij ,
∂x1 (p) = up and ∂y1 (p) = wp and the Symplectic form ω takes its standard form
Pn i i
i=1 dx ∧ dy , by the Darboux Theorem.
y 2
x 2
1
y 2
1
x 2
n n
ξ(x1 , y1 , . . . , xn , yn ) = ψ + + ··· + +
2
1
z=√ 2 (− (∂y1 ξ) ∂x1 + (∂x1 ξ) ∂y1 )
det g
2ψ 0 (ρ2 )
1
z= − 2 y1 ∂x1 + x1 ∂y1 ,
1 + O(2 )
x1 2 y1 2 xn 2 yn 2
The term − 12 y1 is o(1) inside Ω0
where ρ2 =
+ 2 +···+
+
.
and the term x1 is O() in Ω0 . So to first order in we can write
1
z = −2ψ 0 (ρ2 ) 2 y1 ∂x1 + O().
68
Then we compute ∇z v to lowest order in :
1
∇z v = −2ψ 0 (ρ2 ) 2 y1 ∂x1 X + O()
1
= −2ψ 0 (ρ2 ) 2 y1 ∇u v + O()
1
== −2ψ 0 (ρ2 ) 2 y1 w + O()
a
= −2ψ 0 (ρ2 ) 2 y1 ∂y1 + O().
Then
2 0 2
y1
∇σ = 2a ψ (ρ ) x1 ∂x1 + 2 ∂y1 + . . . + xn ∂xn + yn ∂yn + O(2 )
Therefore we have
and so
Z Z
g(Peω (∇z v), Peω (∇z v)) ω n = g(O(), O()) ω n
M Ω0
69
On the other hand, the term (Av , z)L2 in
Z
ω
K (z, v) = (Av z, z)L2 + g(Peω (∇z v), Peω (∇z v)) ω n ,
M
a a
z = −2ψ 0 (ρ2 ) 2 y1 ∂y1 + O() = −2ψ 0 (ρ2 ) 2 y1 u + O(),
we have
a
Av z = −2ψ 0 (ρ2 ) 2 y1 Av u + O().
Therefore,
Z
(Av z, z)L2 = g(Av z, z) ω n
M
a2 y12
Z
= 4ψ 0 (ρ2 )2 g(Av u, u) ω n + O().
Ω0 4
The results of the previous section show that positive curvature is mostly
concentrated around the set of Killing vector fields on M . For this reason we look
for examples of conjugate points along geodesics which are generated by Killing
vector fields.
A manifold M is a Kahler manifold if it can be endowed with a compatible
Symplectic form and Riemannian metric such that ω, or equivalently the almost
complex structure J, is parallel with respect to the Levi-Civita connection of g.
70
That is, M is Kahler if ∇J = ∇ω = 0.
Let M be a Kahler manifold with compatible triple (J, g, ω) and let Dωs (M ) be
the group of diffeomorphisms of M preserving the Kahler form, ω. The form ω is
also a Symplectic form and is harmonic with respect to the metric. Thus Dωs (M )
is the group of Symplectomorphisms of M .
Remark 3.4.1. The isometry subgroup is wholly contained in the group of Sym-
plectomorphisms. For an isometry η and a harmonic form α, η ∗ α is also har-
monic. If G is any Lie group then the action of G on H ∗ (M, R) by α 7→ η ∗ α is
trivial. That is, if α is a representative of a cohomology class [α] then η ∗ α ∈ [α].
By the Hodge isomorphism, each cohomology class has a unique harmonic repre-
sentative and consequently harmonic forms are invariant under the action of the
isometry group of M . Since the Symplectic form ω is harmonic every isometry is
contained in the group of Symplectomorphisms of M .
Denote by Iso(M ) the isometry group of M . For more details on the Isometry
groups of Riemannian manifolds we refer the reader to Poor, [Po].
d
η(t)∗ g = η(t)∗ Lv g = 0
dt
71
Complex projective space CP n is a complex manifold that can be described
by n + 1 coordinates (zo , z1 , ..., zn ) where coordinates which differ by an over all
rescaling are identified:
The complex projective space CPn is a Kahler manifold with the Fubini-Study
metric, which is given in components as
where z = (z1 , ..., zn ) is a point in CPn , |z|2 = z12 + ... + zn2 . The isometry group
of CPn is given by P U (n + 1), the projective unitary group. P U (n + 1) is given
by the quotient of the unitary group, U (n + 1), by it’s center, U (1), embedded as
scalars. Thus, in terms of matrices, U (n + 1) consists of complex n + 1 × n + 1
matrices whose center consists of elements of the form eiθ I. Elements of P U (n+1)
correspond to equivalence classes of unitary matrices, where two matrices A and
B are equivalent if A = eiθ I × B and we write A ≡ B. Composition is given by
the usual matrix multiplication
72
where
i
0
A (s) 0
A(s) =
A0 (s)
...
0
A0 (s)
0
B (t)
0
B (t) 0
B(t) = ..
.
0 B 0 (t)
i
i cos s sin s i cos t sin t
with A0 (s) = and B 0 (s) = are 2 × 2 block
sin s i cos s sin t i cos t
matrices. Observe that γ(s, 0) = iI ≡ I for all s. We shall show that for each
s, γ(s, t) is a family of geodesics in Dωs (CP n ) and compute its variation field,
d
| γ(s, t),
ds s=0
along γ(0, t).
We first find that
d
γ(s, t) ◦ γ −1 (s, t) =
dt
73
0 −i cos s sin s
−i cos s 0 0 C1 (s)
− sin s 0 0
C 2 (s) 0 0 C 1 (s)
=
0 0
..
C2 (s) 0 0 .
0 0
... ...
with
2 2
− sin s cos s −i sin s sin s cos s −i cos s
C1 (s) = C2 (s) =
−i cos2 s sin s cos s −i sin2 s − sin s cos s
−1
where γ −1 (s, t) = (A(s)B(t)A−1 (s)) = A(s)B −1 (t)A−1 (s).
The matrices C1 and C2 are both skew hermitian matrices. Moreover, C1† =
−C2 , where † denotes the conjugate transpose. Therefore, the matrix associated
with the vector field V (s, t) = dtd γ(s, t) ◦ γ −1 (s, t) is a skew-hermitian matrix
and lies in the Lie algebra to P U (n + 1). Thus V (s, t) is a Killing vector field so
that V generates a stationary solution to the Symplectic Euler equations (2.16) by
Proposition 3.4.2. Hence γ(s, t) satisfies the geodesic equation (2.19) in Dωs (CP n ).
74
The variation field of this family of geodesics is given by
0 D1 (t)
−D1 (t) 0 D1 (t)
..
−D1 (t) 0 .
d
.. ..
J(t) = |s=0 γ(s, t) = . . D1 (t) ,
ds
...
D1 (t) 0
... ...
D2 (t)
−D2 (t) 0
0 0 − sin t 0 − sin t 0
0= D1 = D2 =
0 0 0 sin t 0 i(1 − cos t)
which clearly vanishes at t = 0 and t = 2π. Therefore, the point γ(2π) = B(2π)
is conjugate to the identity γ(0) = e.
The proof for odd n is exactly the same except we take
with
i
A1 (s)
A1 (s)
A(s) =
..
.
A1 (s)
i
75
B1 (s)
B1 (s)
...
B(t) =
B1 (s)
i 0
0 i
Having established the existence of conjugate points in the group Dωs we now
look for a a more general description of them. In this section we prove that
every geodesic of the L2 metric, which is generated by a Killing vector field, has
conjugate points. In section 3.6 we solve the Jacobi equation along a geodesic
generated by a Killing vector field and show that all conjugate points along such
a geodesic have even multiplicity.
76
η(t) in Dωs ,
Ad∗η(t) w = Adη−1 (t) w (3.19)
Z
g Dηt · u ◦ ηt−1 , Dηt · w ◦ ηt−1 dµ
(Adηt u, Adηt w)L2 =
M
Z
= (ηt∗ g) (u, w) dµ = (u, w)L2
M
since ηt ∈ Iso(M ).
To prove the formula (3.20) we differentiate both sides of the identity
Corollary 3.5.2. 1. Let v be a Killing field and u ∈ Te Dωs . Then the covariant
77
derivative of the L2 -metric (2.1) reduces to
1
∇ωv u = − ad∗u v.
2
1
∇ωu v = −adv u + ad∗u v
2
1
∇ωv u = − adv u
2
1
Rω (u, v) w = [w, [u, v]] (3.21)
4
Proof. This follows directly from the bi-invariance of the L2 metric when restricted
to Iso(M ). However, to see this another way let v be a Killing field and u ∈ Te Dωs .
We have
1
Peω (∇v u) = Peω (adv u + ad∗v u + ad∗u v)
2
1
Peω (∇v u) = Peω (adv u − adv u + ad∗u v)
2
1
= Peω (ad∗u v).
2
We also have
1
Peω (∇u v) = Peω (ad∗v u + ad∗u v − adv u)
2
1
= Peω (−adv u + ad∗u v − adv u)
2
78
1
= −adv u + ad∗u v
2
The expression for the curvature tensor follows from the expression for the
covariant derivative.
Recall that for a general finite dimensional Riemannian manifold the Ricci
tensor is defined as the contraction, with respect to the metric, of the full Riemann
curvature tensor. That is, the Ricci tensor Ric(u, v) evaluated at (u, v) is defined
as the linear map
Trace(w 7→ Reω (w, u)v)
In infinite dimensions, the Ricci tensor is not necessarily a well defined object:
the trace may not converge. However, when considering the restriction of the L2
metric to the subgroup Iso(M ), which is finite dimensional, the curvature tensor
behaves as in finite dimensions and the Ricci tensor is well defined when restricted
to this subgroup. Using the expression for the curvature tensor (3.21) we can write
1 1
Reω (w, u)v = [v, [w, u]] = − [v, [u, w]]
4 4
1
= − adv ◦ adu w
4
so that the Ricci tensor may be realized as the trace of the map
1
w 7→ − adv ◦ adu w
4
79
has conjugate points.
For any fixed Killing field v ∈ Te Dωs there must exist at least one vector u 6= v, also
tangent to Iso(M ), that does not commute with v. For otherwise v must lie in the
center of the Lie subalgebra to Iso(M ) but the center of any matrix group consists
only of scalar multiples of the identity matrix which does not belong to the Lie
subalgebra. Consequently at least one of the ai ’s is non-zero and the matrix of
the operator Rv,v = − 14 adv ◦ adv is given by
a21
a21
1
Rv,v = a22
4
a22
...
80
Let v be a Killing vector field. The Ricci tensor (3.22) is given by
dim Iso(M )
1 X
Trace(Rv,v ) = a2i ,
2 i=1
which is strictly greater than zero since not all ai ’s are zero by the above construc-
tion. The statement now follows directly from the classical theorem of Bonnet and
Myers, (Theorem 1.26, Cheegar and Ebin [C-E]).
We now solve the Jacobi equation along a geodesic of isometries explicitly and
use this solution to describe conjugate points explicitly. This will be done using
spectral integrals and spectral representations of self-adjoint operators.
Let B(R) be the Borel σ-algebra of R and H a Hilbert space. A spectral
measure on B(R) is a mapping E of B(R) into the orthogonal projections on H
such that E(R) = I and E is countably additive. Using E we can define operator
valued integrals
Z
I(f ) = f (λ) dE(λ)
R
Z
A= λ dEA (λ).
R
LetS ≡ S(R, B(R), EA ) be the set of all EA -a.e. finite Borel functions f : R →
81
C ∪ {∞}. For a function f ∈ S, we write f (A) for the spectral integral I(f ). Then
Z
f (A) = f (λ) dEA (λ),
so that we may define functions of an operator A. For the operator Kvo defined by
eitλ −1
(3.14), which is skew self-adjoint by Lemma 3.2.3, and the function f (λ) = iλ
we write
etKv0 − I eitλ − 1
Z
f (Kvo ) = = dE(λ)
Kv0 R iλ
We refer the reader to Appendix D for properties of spectral integrals and the
functional calculus.
Let vo be any vector in Te Dωs and let η be the geodesic of (2.1) starting from
the identity with initial velocity vo . The Jacobi equation along expωe (tvo ) is given
by
∇ωη̇ ∇ωη̇ J + Rω (J, η̇)η̇ = 0 (3.23)
82
expωe (tvo ), and v(s, t) denote the corresponding family of curves in Te Dωs . Define
∂v(s, t)
w(t) = |s=0
∂s
∂η(s, t)
|s=0 = y(t) ◦ η(t).
∂s
∂t y + [v, y] = w (3.28)
Plugging equation (3.28) into (3.27) yields the usual Jacobi equation (3.23). Using
the metric (2.1) and the algebra adjoint operator one can compute that
83
Using the definition of the operator Kv in (3.14) we have
∂t + ad∗v(t) + Kv(t)
∂t − adv(t) y(t) = 0. (3.32)
Proof. Expanding equations (3.32) and (3.33) and setting their difference equal
to zero gives the equation
That is, if we knew equation (3.34) holds then equations (3.32) and (3.33) would
be equal. Recall Lemma 3.2.4 which states that
Differentiating both sides of this identity yields (3.34) so that equations (3.32)
and (3.33) are equivalent.
(∂t + ad∗v ) w = 0
(∂t − adv + Kv ) y = w.
We can rewrite the operators adv and ad∗v in terms of the push-forward Adη and
84
its adjoint Ad∗η ,
d
Adη−1 = −Adη−1 adv (3.35)
dt
d
Ad∗−1 = −ad∗v Ad∗η−1 (3.36)
dt η
Using these equations, the factored, right-translated Jacobi equation can be writ-
ten as the pair of equations
d
Ad∗η(t)−1 Ad∗η(t) w(t) = 0
(3.37)
dt
d
Adη(t) Adη(t)−1 y(t) + Kv(t) (y(t)) = w(t) (3.38)
dt
The solution of (3.38) is obviously w(t) = Ad∗η(t)−1 wo , and from this we rewrite
(3.38) as
d
Adη(t)−1 y(t) + Adη(t)−1 Kv(t) (y(t)) = Adη(t)−1 Ad∗η(t)−1 wo .
(3.39)
dt
Theorem 3.6.2. Let η(t) = exp(tvo ) be a geodesic of the L2 metric (2.1) generated
by a Killing vector field vo . Let J(t) be a Jacobi field along η(t), with initial
conditions J(0) = 0, J 0 (0) = wo . Then
I − e−tKv0
J(t) = Dη(t) · w0 .
Kv0
Proof. We will solve the Jacobi equation in Te Dω0 . Let vo be any vector in Te Dωs
and let η be the geodesic of (2.1) starting from the identity with initial velocity
vo . By Theorem 2.4.3, the Cauchy problem for the Symplectic Euler equations
85
(1.2) is globally well posed and it follows that the corresponding geodesic in Dωs
can be extended indefinitely. From equation (3.39), the Jacobi equation along η
is equivalent to
d
Adη(t)−1 y(t) + Adη(t)−1 Kv(t) (y(t)) = Adη(t)−1 Ad∗η(t)−1 wo ,
(3.40)
dt
where J(t) = y(t) ◦ η solves the Jacobi equation (3.23) with initial conditions
(3.24). By Lemma 3.2.4
Using equation (3.41) and letting u(t) = Adη(t)−1 Y (t), equation (3.40) becomes
Now suppose that η(t) = expωe (tv0 ) is a geodesic generated by a Killing field
vo . Then Ad∗η(t)−1 = Adη(t) , by Lemma 3.5.1, and equation (3.42) reduces to
Since Kvo is skew self-adjoint by Lemma 3.2.3, the solution to the Homogeneous
equation is given by Stone’s Theorem , cf. Theorem E.3.3,
86
where e−tKvo has the spectral representation
Z
−tKvo
e = e−itλ dE(λ)
R
and E is the unique spectral measure associated to the operator Kvo . Therefore,
the solution to the inhomogeneous equation is given by Duhamel’s principle:
Z t Z t
−tKv0 −(t−s)Kv0
u(t) = e u(0) + e w0 ds = e−(t−s)Kv0 w0 ds
0 0
Z tZ
−tKv0
u(t) = e eisλ dE(λ) ds w0 = e−tKv0 U (t)wo ,
0 R
where
Z tZ
U (t) = eisλ dE(λ) ds w0
0 R
Z tZ Z tZ
isλ isλ
e d (E(λ)x, x) 2 ds = e dEx (λ) < ∞,
L
0 R 0 R
where Ex (λ) = (E(λ)x, x)L2 is a real, finite scalar measure. For any y ∈ Te Dω0 ,
there is a complex measure Ex,y on B(R): Ex,y (λ) = (E(λ)x, y)L2 which is a linear
combination of four positive finite measures, each of which yield a finite integral
as above, see the polarization formula (D.1) in Appendix D. We deduce that the
measure space (R, B(R), Ex,y ) is finite and
Z 1 Z
isλ
Z 1 Z
isλ
e d (E(λ)x, y) 2 ds = e dEx,y (λ) < ∞.
L
0 R 0 R
87
Making use of the Fubini-Tonelli Theorem we have
Z 1 Z Z Z 1
isλ
(S(t)wo , x)L2 = e d (E(λ)wo , x)L2 ds = eisλ ds d (E(λ)wo , x)L2
0 R R 0
eiλ − 1 eiλ − 1
Z Z
= d (E(λ)wo , x)L2 = dE(λ)wo , x .
R iλ R iλ L2
eiλ − 1 etKv0 − I
Z
S(t)wo = dE(λ)wo = wo .
R iλ Kv0
I − e−tKv0
J(t) = Dη(t) · wo .
Kv0
The spectrum, denoted by σ(A), is the complement of the resolvent set, i.e. σ(A) =
R \ ρ(A). A real number λ is said to be an eigenvalue of A if
88
field vo on M . Then η(t∗ ), t∗ > 0, is conjugate to the identity if and only if 2πik
t∗
is
an eigenvalue of Kvo for some non-zero integer k. Consequently, the multiplicity
of every conjugate point along η(t) is even.
etKvo − I
J(t) = Dη(t) · e−tKvo wo
Kvo
where
J(0) = 0 J 0 (0) = wo .
A point η(t∗ ) is conjugate to the identity if and only if there exists a Jacobi field
J(t), along η(t), such that J(t∗ ) = 0. Since both Dη(t) and e−tKvo are invertible
linear operators, conjugate points are determined by the operator
∗K ∗λ
et −I eit −1
vo
Z
= dE(λ).
Kvo R iλ
∗K
et vo −I
The spectrum of the operator Kvo
is determined by the essential range of
it∗ λ t∗ Kv
the function f (λ) = e iλ−1 , [Sch]. In particular, 0 is an eigenvalue of e Kvo −I if
o
∗
n o
eit λ −1
and only if E λ ∈ R : iλ = 0 6= 0. From the Taylor series of the function
∗λ
eit −1
f (λ) = iλ
we see that f (0) = 1. Therefore,
∗
eit λ − 1
∗
= 0 = λ ∈ R \ {0} : eit λ = 1 .
λ∈R:
iλ
∗K
et vo −I
This shows that 0 is an eigenvalue of Kvo
if and only if 1 is an eigenvalue of
∗K
et vo
. Since the function f (λ) = etλ is an analytic function, the semigroup etKvo
is an analytic semigroup. Applying the spectral mapping theorem (see Appendix
89
∗K ∗K
D) to the operator et vo
we have that 1 is an eigenvalue of et vo
if and only if
2πik
t∗
is an eigenvalue of Kvo for some non-zero integer k. We have the following
chain of equalities
∗λ
eit
−1 ∗ ∗
= E λ ∈ R \ {0} : eit λ = 1 = σ et Kvo
E λ∈R: =0
iλ
t∗ σ(Kvo ) 2πik
=e =E : k ∈ Z \ {0}
t∗
which shows that η(t∗ ) is conjugate to the identity if and only if 2πik
t∗
is an eigen-
value of Kvo for some non-zero integer k. By Lemma 3.2.6, the operator Kvo is
compact and skew self-adjoint in the L2 metric and therefore has a complete set
of orthonormal eigenvectors spanning Te Dω0 , see Theorem C.1.4. Since complex
eigenvalues always occur in conjugate pairs, whose associated eigenvectors are
orthonormal, the multiplicity of every conjugate point along η(t) is even.
90
CHAPTER 4
4.1 Introduction
Semi-Fredholm operators form an open set in the space of all bounded linear
operators and the index is a continuous function on this set into Z ∪ {±∞}. Kato
treats unbounded semi Fredholm operators in [K3], [K2] and shows that the index
is well defined for a semi-Fredholm operator.
The definition of a Fredholm map is due to Smale, [S]: A smooth map f : M →
N between Banach manifolds is called a Fredholm map if its Frechet derivative
91
df (p) is a Fredholm operator for each p. If the domain of f is connected then the
index of the operator df (p) is independent of p and by definition is the index of f .
In this chapter we answer Ebin’s question in the affirmative:
92
L2 metric. Theorem 4.1.1 holds for any closed Symplectic manifold of dimension
2n, while Theorem 4.1.1 fails for the Volume preserving diffeomorphism group of
manifolds of dimension 3 and higher, cf. [E-M-P], [P2]. The relationship between
Theorem 4.1.1 and known classifications (e.g. C 0 closure, Gromov’s non-squeezing
Theorem) of Symplectic diffeomorphisms is, at this point, unclear.
In section 4.2 we define relevant objects for the proof of Theorem 4.1.1 and
collect their useful properties. Section 4.3 begins an analysis of the Jacobi equation
and its solution operator, with the proof of Theorem 4.1.1 finally given in 4.4.
Let vo be any vector in Te Dωs and let η be the geodesic of (2.1) starting from
the identity with initial velocity vo . The Jacobi equation along expωe (tvo ) is given
by
∇ωη̇ ∇ωη̇ J + Rω (J, η̇)η̇ = 0 (4.1)
The fact that Rω is a bounded multi-linear operator implies the Jacobi fields exist,
are unique and are defined for as long as η is, cf. [M1]:
93
tion ∇ω . Let u, v ∈ Te Dωs and define
so that
kReω (u, v)wks = sup {(Reω (u, v)w, z)s : z ∈ C ∞ (T M ), diz ω = 0, kzks < 1}
94
The relationship between the Jacobi equation (4.1) and the exponential map is
as follows. Let η(t) = expωe (tvo ) be as above. Define v(s, t) = tvo (s), where vo (s) is
a variation of the initial condition vo depending on s. Then, η(s, t) = expe (tvo (s)),
∂v(s, t)
|s=0 = two ,
∂s
∂η(s, t) ∂
u(t) ◦ η(t) = |s=0 = |s=0 expe (tvo (s)) = D expe (tvo )two .
∂s ∂s
Combining this observation with the above, we see that the differential of the
exponential map yields solutions of the Jacobi equation (4.1) satisfying the initial
conditions (4.2).
If J is the Jacobi field along η with initial conditions (4.2), then
defines a family Φ(t) of bounded linear operators from Te Dωs to Tη(t) Dωs .
From section 3.6, equation (3.42) the Jacobi equation can be written as
d
Adη(t)−1 y(t) + Adη(t)−1 Kv(t) (y(t)) = Adη(t)−1 Ad∗η(t)−1 wo ,
(4.4)
dt
where J(t) = y(t) ◦ η(t) is the Jacobi field along η(t) with initial conditions (4.2).
For σ ≥ 0, let Te Dωσ denote the closure of the space of locally Hamiltonian
vector fields in the H σ norm. By the Hodge decomposition, Theorem 2.3.5, this
dim M
is a closed subspace in the space of all H σ vector fields ([M]). For σ > 2
+1
this coincides with the actual tangent space to Dωσ . However, for smaller σ, DHam
s
95
is not necessarily a smooth manifold.
We have the following convenient decomposition of the solution operator (4.3)
to the Jacobi equation (4.1). The decomposition loses one derivative (the equa-
tions are only defined on H σ if the initial velocity defining the geodesic is in H σ+1 ),
nevertheless we will be able to compensate for this later.
dim M
Proposition 4.2.2. If η is a geodesic in Dωs (M ), with s > 2
+ 1, s ≥ σ + 1
then the map Φ(t) defined in (4.3) extends to a continuous linear operator from
Te Dωσ to Tη(t) Dωσ . In addition, we have the formula
where
Z t
Ω(t) = Adη(τ )−1 Ad∗η(τ )−1 dτ (4.6)
0
Z t
Γ(t) = Adη(τ )−1 Kv(τ ) dRη−1 (τ ) Φ(τ ) dτ, (4.7)
0
d
Adη(t)−1 u(t) = Adη(t)−1 Ad∗η(t)−1 wo − Adη(t)−1 Kv(t) u(t).
dt
By definition, u(t) ◦ η(t) = J(t) = Φ(t)wo so that u(t) = dRη−1 (t) Φ(t)wo . Writing
this into (4.4) and integrating both sides in t yields
Z t Z t
Adη(t)−1 u(t) = Adη(s)−1 Ad∗η(s)−1 wo ds − Adη(s)−1 Kv(s) dRη−1 (s) Φ(s)wo ds.
0 0
96
Using the definition of Adη−1 (t) = dLη−1 (t) dRη(t) we arrive at
Z t
∗
= v, Adη(s)−1 Adη(s)−1 w ds
0 L2
Z t
v, Adη(s)−1 Ad∗η(s)−1 w
= L2
ds
0
Z t
Adη(s)−1 Ad∗η(s)−1 v, w
= L2
ds
0
= (Ω(t)v, w)L2 .
We are now in a position to prove Theorem 4.1.1 by showing that the operator
Ω(t) defined in (4.6) is invertible and the operator Γ(t) defined by (4.7) is compact.
s
As observed in chapter 1, Te DHam is of finite codimension in Te Dωs and thus it
s
suffices to prove that Φ(t) of Proposition 4.2.2 is Fredholm on Te DHam . We first
prove invertibility of the operator Ω(t) in a series of lemmas. Invertibility is first
established in the L2 topology; if Ω(t) has empty kernel in Te DHam
0
it will certainly
σ 0
have empty kernel in Te DHam ⊂ Te DHam , for all σ > 0. We have a very useful
criteria for determining when an operator is invertible: an operator T is invertible
on H if and only if it is bounded below and has dense range. An operator T on a
97
Hilbert space H is bounded below if there exists an ε > 0 such that kT xk ≥ ε kxk
for x ∈ H. Indeed, if T is invertible then the range of T is H which is dense.
Moreover,
1
−1
1
kT xk ≥
T T x
= kxk
kT −1 k kT −1 k
1
kxn − xm k ≤ kT xn − T xm k
ε
for any wo ∈ Te Dωs , which implies that Ω(t) has closed range and finite dimensional
kernel. To see that such an estimate implies closed range let Ω(t)xn → y in
H σ . We need an x such that Ω(t)x = y. Let L = ker Ω(t). If the distance
between xn and L remains bounded take zn = xn modL, kzn kH σ ≤ 2a < ∞;
98
Then Ω(t)zn = Ω(t)xn → y. Passing to a subsequence, znk → w in H σ−1 . Then,
applying the estimate (4.8) to zn − zm , implies that the sequence zn is Cauchy so
zn → z and Ω(t)z = y. If the distance between xn and L tends to infinity we can
assume that the distance is greater than two for all n. Pick un = xn modL such that
un 1
dist(xn , L) ≤ kun k ≤ dist(xn , L) + 1 and set wn = kun k
. Now 2
≤ dist(wn , L) and
since kwn k = 1 we can choose a subsequence converging in H σ−1 (by the Rellich
Lemma). Also, Ω(t)wn → 0 so that the estimate (4.8) applied to wm − wn implies
1
wn is a Cauchy sequence. Thus wn → w ∈ H σ with 2
≤ dist(wn , L) and Ω(t)w =
0, which is a contradiction. Thus this latter case is impossible and the claim is
proved. To prove finite dimensionality of the kernel, let wn be an H σ sequence on
the unit sphere in the kernel of Ω(t) so that
dim M
Lemma 4.3.1. For s > 2
+1 the operator Ω(t) defined by (4.6) is an invertible
0
linear operator on Te DHam with
99
0
for any vH ∈ Te DHam , and
Z t
1
C(t) = dτ.
0 kDη(τ )kL∞ kDη −1 (τ )kL∞
Proof. We compute
Z t
(vH , Ω(t)vH ) = (vH , Adη−1 Ad∗η−1 vH ) dτ
0
Z t Z t
Adη−1 vH
2 2 dτ ≥
∗
1 2
= L
∗
2 dτ kvH kL2
0 0
Adη−1 (t)
L2
Z t
1
≥ 2 dτ kvH k2L2 ,
T
kDη(t)Dη(t) kL∞
0
where
Adη−1 (τ )
2 =
Adη−1 (τ )
2 2 ≤
Dη(τ )T Dη(τ )
∞ .
∗
L2 L L
so that Ω(t) has empty kernel and closed range. Since Ω(t) is self-adjoint by
Proposition 4.2.2, and has closed range we have RanΩ(t) = (ker Ω(t))⊥ = {0}⊥ =
0 0
Te DHam and therefore Ω(t) is invertible on Te DHam .
100
(x1p , . . . , xnp ). The Sobolev topology can be defined locally by
X
kH ◦ ηkH σ (O) = k∂ α H(η)kL2 (O) ,
|α|≤σ
Pn
where α = (α1 , . . . αn ) is a multi-index with |α| = j=1 αj and Dα = ∂xα11 · · · ∂xαnn .
We compute (in coordinates)
X
kΩ(t)vH kH σ = kDα Ω(t)vH kL2
X X
≥ kΩ(t)Dα vH kL2 − k[Dα , Ω(t)] vH kL2
X
≥ C(t) kvH kH σ − k[Dα , Ω(t)] vH kL2 ,
where we have used Lemma 4.3.1 in the first inequality. To prove the estimate
(4.8) we must show that
X
k[Dα , Ω(t)] vH kL2 . kvH kH σ−1 .
This is the contents of the next Lemma. The proof of (4.8) will be completed in
Lemma 4.3.3.
α ω
∂ , Pη w
L2 (O)
≤ C kwkH σ−1 (O) (4.10)
101
Proof. Let w ∈ H σ . Then, according to the Hodge decomposition, Theorem 2.3.5,
we can write
w = v + Jg ] δα
We will prove
[∂xi , Pηω ]w ◦ η
2 ≤ bi kwk 2 . (4.11)
L L
∂η j
With Nij = ∂xi
◦ η −1 , Jlk the components of the almost complex structure, g lm the
components of the inverse metric, (δα)m the components of the 1-form δα,
X X X j ∂wk
k −1
∂xi (w ◦ η) ◦ η ∂k = Ni j
∂k
k k j
∂x
XX ∂v k ∂J k ∂g lm
∂(δα)m
= Nij j + Nij lj g lm (δα)m + Nij Jlk (δα)m + Nij Jlk g lm ∂k .
k j
∂x ∂x ∂xj ∂x j
∂(δα)m ∂α
Now ∂xj
= (δ ∂x j )m (since δ = − ? d?, with ? the Hodge star operator) and -
Nij ?d? ∂x
∂α
= −?Nij d? ∂x∂α j ∂α j ∂α
= δ(Nij ∂x
∂α ∂α
j j = −?d?(Ni ∂xj )+? (dNi ) ∧ ? ∂xj j )+i∇N j ∂xj
i
(which follows from the formula δ(f γ) = f δγ + i∇f γ for any function f and any
k-form γ). Consequently
X
∂xi (wk ◦ η) ◦ η −1 ∂k =
k
102
X X j ∂v k k
j ∂Jl lm j k ∂g
lm
∂α
k lm
Ni j + Ni j
g (δα)m + Ni Jl j
(δα)m + Jl g (i∇N j j )m ∂k
k j
∂x ∂x ∂x i ∂x
∂α
+Jg ] δ(Nij ).
∂xj
Since Peω = J∇4−1 divJ and J 2 = −I, the last term projects to zero. Thus the
expression ∂xi (wk ◦ η) ◦ η −1 ∂k contains only first derivatives of α and v so we have
essentially gained a derivative. Therefore
XX ∂v k
∂xi , Pηω (w ◦ η) = Qωe (Nij j ∂k ) ◦ η
k j
∂x
k lm
j ∂Jl lm j k ∂g ∂α
−Peω (Ni g (δα)m + Ni Jl k lm
(δα)m + Jl g (i∇N j j )m )∂k ◦ η.
∂xj ∂xj i ∂x
k lm
j ∂Jl lm j k ∂g ∂α
Peω (Ni g (δα)m + Ni Jl k lm
(δα)m + Jl g (i∇N j j )m )∂k ◦ η
∂xj ∂xj i ∂x
k lm
Pe (N j ∂Jl g lm (δα)m + N j Jlk ∂g (δα)m + Jlk g lm (i j ∂α )m )∂k ◦ η
ω
i i ∇Ni
∂xj ∂xj ∂xj
2
L
j ∂Jlk lm lm
j k ∂g k lm ∂α
≤
(N i g (δα) m + N i J l (δα) m + Jl g (i j
∇Ni )m )∂ k
∂xj ∂xj ∂xj
2
L
j ∂Jlk lm
j k ∂g lm
k lm ∂α
≤
(Ni
g (δα)m
+
Ni Jl
(δα)m
+
Jl g (i∇N j j )m
∂xj L 2 ∂x j
L2
i ∂x
2
L
k
k ∂g lm
∂Jl lm
+
g lm (i j ∂α )m
.
∂xj g (δα) m
+
Jl (δα) m ∇N
L2
∂xj
L2
i ∂xj
2
L
∂α
.
g lm (δα)m
L2 +
g,jlm (δα)m
L2 +
∇Ni ∂xj )m
2
(i j
L
103
∂α
. k(δα)m kL2 + k(δα)m kL2 +
( ∂xj )m
2
L
. kwkL2
To complete the proof it suffices to bound the first term by the L2 norm of w.
For any vector field w
by Lemma 2.3.6. Using the Leibniz rule and the fact that dω [ (v) = 0, the expres-
sion Qωe (Nij v,jk ∂k ) ◦ η only involves first derivatives of v. Hence
and we obtain (4.11). Inductively, the estimate (4.11) implies that for any multi-
index α, with no more than σ terms, we will have (4.10).
dim M
Lemma 4.3.3. For s > 2
+ 1, s ≥ σ + 1 the operator Ω(t) defined by (4.6) is
σ
an invertible linear operator on Te DHam with
σ
for some constant C and any vH ∈ Te DHam , and
Z t
1
C(t) = dτ,
0 kDη(τ )kL∞ kDη −1 (τ )kL∞
104
Proof. We have
X
kΩ(t)vH kH σ = k∂ α Ω(t)vH kL2
0≤|α|≤σ
X X
≥ kΩ(t)∂ α vH kL2 − k[∂ α , Ω(t)] vH kL2
0≤|α|≤σ 0≤|α|≤σ
X
≥ C(t) kvH kH σ − k[∂ α , Ω(t)] vH kL2 ,
0≤|α|≤σ
Z t
Ω(t) = Dη −1 dRη Peω dRη−1 (Dη −1 )T dτ
0
and use this to show that k[∂ α , Ω(t)] vH kL2 is bounded above by the H σ−1 norm
of vH . It is enough to show that, for each τ ,
X
∂ α , Dη −1 Pηω (Dη −1 )T vH
2 ≤ C kvH k σ−1 ,
L H (4.13)
X
X
∂ α , Dη −1 Pηω (Dη −1 )T vH
∂ α , Dη −1 Pηω (Dη −1 )T vH
L2
≤ L2
X
Dη −1 ∂ α , Pηω (Dη −1 )T vH
2
+ L
X
Dη −1 Pηω ∂ α , (Dη −1 )T vH
2 .
+ L
105
and therefore (4.12).
The estimate (4.12) shows that Ω(t) has empty kernel (since Ω(t) has empty
kernel in L2 ) and closed range on Te DHam
σ σ
. Choose any vH ∈ Te DHam . Since Ω(t) is
0 0
invertible on Te DHam we can find a vF ∈ Te DHam such that Ω(t)vF = vH . But now
the estimate (4.12) shows that vF is in H σ as well and Ω(t) is therefore invertible
σ
on Te DHam .
To prove that the operator Γ(t) is compact we will need the results from
chapter 3 which show that the operator Kv(t) is compact. For a geodesic η(t) with
η̇(t) = v(t) ◦ η(t) and initial velocity vo we have that
by Lemma 3.2.4. Lemma 3.2.6 states that for any vo ∈ Te Dωs the operator Kvo is
compact on Te Dωσ . Since the set of compact operators is a closed, two-sided Ideal
in the space of bounded linear operators, and the operators Ad∗η−1 (t) , Adη−1 (t) are
bounded it follows that Kv(t) is compact on Te Dωσ .
dim M
Lemma 4.3.4. Suppose M is a closed Symplectic manifold and let s > 2
+ 1,
s
s ≥ σ + 1. Let vo ∈ Te DHam and η(t) = expe (tvo ). Then, the operator Γ(t) defined
σ
by (4.7) is compact on Te DHam .
Proof. Since the operators Φ(t), dRη−1 and Adη are all continuous and Kv(t) is
compact, the composition appearing under the integral in (4.7) is compact. Thus
the integral, as a limit of Riemann sums of compact operators, is compact:
Z t
Γ(t) = Adη(τ )−1 Kv(τ ) dRη−1 (τ ) Φ(τ ) dτ
0
106
n
X
= lim Adη(t∗i )−1 Kv(t∗i ) dRη−1 (t∗i ) Φ(t∗i ) (ti − ti−1 ) ,
n→∞
i=1
where ti−1 ≤ t∗i ≤ ti and P = {[0, t1 ], [t1 , t2 ], . . . , [tn−1 , tn = t]} is a partition of the
interval [0, t].
0
Proof of Theorem 4.1.1 Since Ω(t) is invertible and Γ(t) is compact on Te DHam ,
0
the sum Ω(t) − Γ(t) is Fredholm on Te DHam . Since Dη(t) is continuous and invert-
0 0
ible on Te DHam , the operator Φ(t) = Dη [Ω(t) − Γ(t)] is Fredholm on Te DHam (see
Theorem C 3.3, Appendix C) and hence d expe (tvF ) is Fredholm aswell.
In order to prove the result in H s we need to approximate η by a smoother η̃
since if η is in H s the decomposition (4.5) only works in H s−1 due to the presence
of derivatives of η. We prove that Φ(t) has closed range and finite dimensional
kernel in H s . Then Φ(t) is semi-Fredholm and we can compute its index.
Let us calculate D expωe (0): For v ∈ Te DHam
s
d
D expωe (0)v = |t=0 expωe (tv)
dt
d
= |t=0 η(t)
dt
= v(t) ◦ η(t)|t=0
= v,
where η(t) = expωe (tv) is a geodesic with initial velocity v. From the relationship
(4.3) we see that Φ(0) is the identity with index zero so we can conclude that Φ(t)
is Fredholm of index zero for all time.
107
Finite dimensionality of the kernel in H s follows from finite dimensionality of
the kernel in L2 since the former is a subset of the latter. In order to prove closed
range it will suffice to establish an estimate of the form
h i
Φ̃(t) = Dη̃ Ω̃(t) − Γ̃(t)
with
Z t
Ω̃(t) = Adη̃(τ )−1 Ad∗η̃(τ )−1 dτ
0
Z t
Γ̃(t) = Adη̃(τ )−1 KṽF (τ ) dRη̃−1 (τ ) Φ̃(τ ) dτ
0
0
which is valid on Te DHam as well.
Since solutions to the geodesic and Jacobi equations depend continuously on
the initial conditions, Φ̃ is close to Φ in the H s operator norm and η̃ is close to η
108
in the H s norm. Consequently,
kΦ(t)u0 k ≥
Φ̃(t)u0
−
Φ(t) − Φ̃(t)
ku0 kH s
Hs Hs
≥
Dη̃(t) · Ω̃(t)u0
−
Dη̃(t) · Γ̃(t)u0
−
Φ(t) − Φ̃(t)
ku0 kH s
Hs Hs Hs
C̃(t)
≥ ku0 kH s −
Dη̃(t) · Γ̃(t)u0
−
Φ(t) − Φ̃(t)
ku0 kH s
kDη̃kL∞ Hs Hs
where we have used the estimate (4.12). Therefore, with κ(t) = Dη̃(t) · Γ̃(t) we
obtain the estimate
with !
C(t) C(t) C̃(t)
A= − − −
Φ(t) − Φ̃(t)
.
kDηkL∞ kDηkL∞ kDη̃kL∞ Hs
The number C(t) of Lemma 4.3.1 depends only on the C 1 norm of η, as does
kDηkL∞ . So if we choose ṽF0 close enough to vF0 in the H s norm then C̃ is close
to C, kDη̃kL∞ is close to kDηkL∞ . Also,
Φ(t) − Φ̃(t)
is close to zero so that
Hs
A can be made positive and the estimate (4.14) is satisfied. Thus Φ(t), and hence
D expe (tvF0 ), has closed range in the H s topology.
This completes the proof of Theorem 4.1.1.
Proof. Let η(t) be a geodesic in Dωs with η̇(0) = vo and suppose that η(t∗ ) is
epiconjugate to e. Since the range of D expe (t∗ v) is closed we have an orthogonal
109
splitting of the tangent space
Z a
u0η , vη0 + (Reω (uη , η̇)η̇, vη )L2 dt,
Ia (uη , wη ) = L2
0
where uη , wη ∈ V.
In general, given a symmetric bilinear form B over a space V, we define the
index of B as the maximal dimension of all subspaces of V on which the quadratic
form is negative definite. The nullity of B is defined to be the dimension of the
subspace of V formed by elements v ∈ V such that B(v, w) = 0 for all w ∈ V;
such a subspace is called the nullspace of B. We say that B is degenerate if its
nullity is strictly positive.
An element uη ∈ V belongs to the null space of Ia if and only if uη is a Jacobi
field along η. Therefore, Ia is degenerate if and only if η(0) and η(a) are conjugate
to one another. In this case the nullity of Ia is equal to the multiplicity of η(a)
as a conjugate point. Since the exponential map is a non-linear Fredholm map of
110
index zero, the multiplicty of any conjugate point is finite and hence the nullity
of Ia is finite dimensional.
Since η([0, a]) is compact and each point of Dωs is contained in a normal neigh-
bourhood, we can choose a subdivision
of [0, a] such that each η|[tj−1 ,tj ] is contained in a normal neighbourhood. Thus,
each η|[tj−1 ,tj ] is a minimal geodesic and doesn’t contain conjugate points. Let
V − (0, a) = V − be the subspace of V formed from the felds vη such that vη |(tj−1 ,tj ) ,
j = 1, ..., k, is a Jacobi field. The space V − has finite dimension because the
nullity of Ia has finite dimension. Let V + be the subspace of V consisting of
vector fields wη such that wη (t1 ) = wη (t2 ) = ... = wη (tk−1 ) = 0. . The spaces V −
and V + are orthogonal with respect to Ia : Given vη ∈ V, let uη ∈ V − be given by
uη (tj ) = vη (tj ). Since η|[tj−1 ,tj ] does not contain conjugate points, such a uη exists
and is unique. Hence vη − uη ∈ V + and, therefore, V = V + ⊕ V − . Also, if wη ∈ V +
and uη ∈ V − , we have
k Z
X tj
u00η , wη + (Reω (uη , η̇)η̇, wη )L2 dt = 0
Ia (uη , wη ) = L2
j=1 tj−1
since wη vanishes at the end points and uη |[tj−1 ,tj ] , j = 1, ..., k, is a Jacobi field.
We will now show that the index form is positive definite on V + so that the space
on which the index form is negative definite is contained in a finite dimensional
subspace .Since η|[tj−1 ,tj ] , j = 1, ..., k, are minimizing geodesics and do not contain
conjugate points, each η|[tj−1 ,tj ] has less energy than any other path between its
endpoints. Therefore, if wη ∈ V + then Ia (wη , wη ) ≥ 0. It remains to show that
111
Ia (wη , wη ) > 0 if wη ∈ V + \ {0}. On the contrary, suppose that Ia (wη , wη ) = 0
with wη 6= 0. If uη ∈ V − then Ia (wη , uη ) = 0. If uη ∈ V − , then we have
valid for all c ∈ R. This inequality says that there exist real numbers A ≥ 0 and
B such that Ac2 + 2Bc ≥ 0 for all c ∈ R. This is only possible if B = 0, i.e.
Ia (wη , uη ) = 0. This shows that wη lies in the null space of Ia and is therefore a
Jacobi field. But wη (tj ) = 0, j = 1, ..., k − 1, and each η|[tj−1 ,tj ] do not contain
conjugate points. Thus wη = 0.
Corollary 4.4.2. Any finite geodesic segment contains finitely many isolated con-
jugate points.
Proof. Let tk ∈ [0, a] be a sequence such that η(tk ) is conjugate to η(0) = 0 and
let Jk (t) be the corresponding Jacobi field vanishing at t = 0 and t = tk . For each
k define a vector field
Jk (t)
t ∈ [0, tk ]
Yk (t) = .
0 t ∈ [tk , a]
Then each Yk ∈ V and Ia (Yk , Yk ) = 0. Thus, Yk ∈ V − and there are only finitely
many linearly independent fields Jk (t) and consequently only finitely many t0k s.
Corollary 4.4.3. The set of points conjugate to the identity along any geodesic
(emanating from the identity) is of first Baire Category.
112
CHAPTER 5
s
CONJUGATE POINTS ON DHam (M n ) IN DIMENSIONS n = 2 AND 4
5.1 Introduction
vG (0) = vGo
cf. section 3.6, equation (3.30). Denote by S(t) the solution operator to this
equation defined by vG (t) = S(t)vGo . In this chapter we will show that a solution
to the linearized Symplectic Euler equation S(t)vGo = vG(t) , satisfies
where KvF (t) is the operator defined by (3.14), vF (t) = η̇(t) ◦ η −1 (t) and vH (t) is
the right translated Jacobi field J(t) = vH(t) ◦ η(t) along η(t) with J(0) = 0 and
J 0 (0) = vGo . A point η(t∗ ) is then conjugate to the identity if and only if
113
Recall that we had two factorizations of the Jacobi equation (cf. section 4.2). The
right factored Jacobi equation
or in system form
∂t + ad∗vF + KvF z = 0
(5.4)
Condition (5.2) tells us that conjugate points occur when a solution z(t) = vG (t)
to the linearized Symplectic Euler equation, which can be thought of as a curve in
s
Te DHam , intersects a solution z̃(t) = Ad∗η−1 (t) vGo to the left-factored equation (5.7)
at a time t∗ > 0, which is also a curve in Te DHam
s
. The fact that KvFo is compact
tell us that the solution operator to the linearized Symplectic Euler equations
differs from the coadjoint operator by a compact operator.
We also solve the Jacobi equation (4.1) explicitly in terms of solutions to the
linearized Symplectic Euler equations and coadjoint orbits. This shows that the
114
stability of geodesics (i.e. the growth of Jacobi fields in some norm) in the Sym-
plectomorphism group depends on how far a solution to the linearized Symplectic
Euler equations is from being a coadjoint orbit (measured in the same norm).
We will need the following definition
∂F ∂G ∂F ∂G
0 = {F, G} = −
∂x ∂y ∂y ∂x
∂ 2F ∂ 2F
∂ ∂F ∂ ∂F
= G − G− G + G
∂y ∂x ∂y∂x ∂x ∂y ∂x∂y
∂ ∂F ∂ ∂F
= G − G
∂y ∂x ∂x ∂y
∂F
− ∂y
= div G
∂F
∂x
= div (GJ∇F )
= ?LGJ∇F ω
= ?d Gω [ J∇F
115
= ? dG ∧ ω [ J∇F + Gdω [ J∇F
= ? dG ∧ ω [ J∇F ,
where in the third to last equality we have used Cartan’s magic formula for the
Lie derivative, and in the last equality we used that J∇F is Hamiltonian. In
particular,
dG ∧ dF = 0.
Now, the wedge product of two one forms is non-zero if and only if the forms
are linearly independent. Indeed, suppose dG is proportional to dF ; then the
wege product is zero since it contains repeated forms. If dG and dF are linearly
independent then ∇F and ∇G are linearly independent and the matrix A =
[∇F ∇G] has non-zero determinant. That is
0 6= det A = dF ∧ dG.
116
Hamiltonian function Fo solves the vorticity equation (2.21):
∂t 4Fo + {4Fo , Fo } = 0,
which reduces to
{4Fo , Fo } = 0. (5.9)
and reduces to
4Fo = 4Fo ◦ η −1 (t)
117
∂t vH (t) − advF (t) vH (t) = vG (t) (5.12)
where
vG (t) = ∂s |s=0 vF (s, t)
and vH (t) ◦ η(t) = ∂s |s=0 η(s, t) and (s, t) ∈ (−δ, δ) × R, δ > 0. Recall that
plugging equation (5.11) into equation (5.12) yields the usual Jacobi equation
(4.1), cf. section 3.6.
s
Proof. Let η(s, t) be a smooth variation of geodesics in DHam , and vF (s, t) the
corresponding variation of solutions to the Symplectic Euler equations (1.2). For
each s and t, the variation F (s, t) satisfies (5.10):
where {·, ·} is the Poisson bracket defined by (2.3). Composing both sides of this
equation, first with η −1 (t) and then applying 4−1 we obtain
118
where G(t) = ∂s |s=0 F (s, t). Applying J∇ to both sides of this equation yields the
Hamiltonian vector fields:
and using the formulas for Ad∗η−1 (t) and KvF (t) in Proposition 3.2.2 we recover
(5.13), where vH (t) ◦ η(t) solves the Jacobi equation (4.1) and vGo is the initial
condition of (4.2).
Let η(t) be a stationary geodesic with initial velocity vFo . The operator KvFo
0
is compact and skew self-adjoint on Te DHam , by Lemmas 3.2.3 and 3.2.6. Thus,
by the spectral theorem for compact (skew) self-adjoint operators (cf. Theorem
0
C.2.6) there exists an orthonormal basis of Te DHam consisting of eigenvectors of
KvF (t) , {vψi } and
0
⊥
Te DHam = ker KvFo ⊕ ker KvFo .
⊥
Let {vψi } denote the basis for ker KvFo and vψj the basis for ker KvFo . Given
s 0
a vector field w ∈ Te DHam ⊂ Te DHam we will write wker for the part of w in ker KvFo
⊥
and w⊥ for the part of w in ker KvFo .
The proof Proposition 5.3.1 will be postponed until the next section, and relies
119
critically on the Fredholm results of Chapter 4.
Proof. Suppose η(t∗ ) is a point conjugate to the identity along η(t). By Theorem
4.1.1 the exponential map is a non-linear Fredholm map of index zero so that
monoconjugate and epiconjugate points coincide. Therefore, there exists a Jacobi
field along η(t) vanishing at t∗ > 0. Let J(t) = vH (t) ◦ η(t), with initial conditions
J(0) = 0, J 0 (0) = vGo , be the Jacobi field along η(t) with J(t∗ ) = 0. By Propo-
ker
sition 5.3.1, vG o
= 0, for otherwise J(t) 6= 0 for all t. By Lemma 5.2.1 it follows
⊥
that (5.14) holds with vGo ∈ ker KvFo .
⊥
Conversely, suppose that (5.14) holds with vGo ∈ ker KvFo . Let J(t) =
vH (t) ◦ η(t) be the Jacobi field along η(t) with initial conditions J(0) = 0, J 0 (0) =
ker
vGo . Then vH (t) = 0 for all t by Lemma 5.3.1. Let vG (t) be a solution of the
linearized Symplectic Euler equations (3.30) with initial conditions vGo . Expand
vG (t), Ad∗η−1 (t) vGo and vH(t) in the orthonormal basis {vψi } of KvFo :
X
vH (t) = hi (t)vψi
i
X
vG (t) = gi (t)vψi
i
X
Ad∗η−1 (t) vGo = ai (t)vψi .
i
120
By Lemma 5.2.1 we have
gi (t) = ai (t) − λi hi (t)
hi (t∗ ) = 0
for each i. Thus vH (t∗ ) = 0 and η(t∗ ) is monoconjugate, and also epiconjugate, to
the identity.
Corollary 5.3.3. Let η(t) be a stationary geodesic. Then the Jacobi field J(t) =
⊥
vH (t) ◦ η(t) with initial conditions J(0) = 0, J 0 (0) = vGo ∈ ker KvFo is given
by
X 1
vH (t) = (gi (t) − ai (t)) vψi ,
i
λ i
P
where KvFo vψi = λi vψi , S(t)vGo = gi (t)vψi solves the linearized Symplectic
i
The proof of Proposition 5.3.1 will be broken up into a series of lemmas, which
rely critically on the Fredholm results of Chapter 4.
For a stationary geodesic with velocity η̇(t) = vFo ◦ η(t), equation (5.12) can
be written as
∂t vH (t) − advFo vH (t) = vG (t)
Since η(t) is a stationary geodesic, the operator Adη−1 (t) is a C0 semigroup with
infinitesimal generator −advFo , cf. Appendix D. Therefore −Adη−1 (t) advFo =
−advFo Adη−1 (t) . Multiplying both sides of this equation by Adη−1 (t) and using
121
the product rule we obtain
Adη−1 (t) vG (t) = Adη−1 (t) ∂t vH (t) − adv Adη−1 (t) vH (t) = ∂t Adη−1 (t) vH (t) .
Z t
vH (t) = Adη(t) Adη−1 (s) vG (s) ds.
0
ker ⊥
Write vG (t) for the part of vG (t) in ker KvFo and vG (t) for the part of vG (t) in
⊥
ker KvFo . Then
Z t Z t
ker ⊥
vH (t) = Adη(t) Adη−1 (s) vG (s) ds + Adη(t) Adη−1 (s) vG (s) ds. (5.15)
0 0
Z t
ker ker
vH (t) = Adη(t) Adη−1 (s) vG (s) ds (5.16)
0
Z t ker
= Adη(t) Adη−1 (s) Ad∗η−1 (s) vGo ds (5.17)
0
Z t
= Adη(t) Adη−1 (s) Ad∗η−1 (s) vG
ker
o
ds, (5.18)
0
with each equality established in a separate lemma. Recall that the operator
Z t
Ω(t) = Adη−1 (s) Ad∗η−1 (s) ds
0
is the invertible part of D expe (tvFo ) since any Fredholm operator of index zero is
122
the sum of an invertible operator and a compact operator, compare with Lemma
ker ker ker
4.3.3. Therefore, if vG o
6= 0 then vH(t) 6= 0 for all t and if vG o
= 0 then vH(t) =0
for all t. This proves (i) and (ii) of Proposition 5.3.1.
We now prove the first equality (5.16).
and
⊥ ⊥
Adη(t) ker KvFo ⊆ ker KvFo .
In particular, each eigenvector vψj of KvFo which lies in ker KvFo is invariant under
the adjoint action.
4−1 {4Fo , ψj } = 0.
Since 4 is an isomorphism, cf. Theorems 2.3.3 and 2.3.4, this condition is equiv-
alent to
{4Fo , ψj } = 0. (5.19)
123
But
4Fo = Φ(Fo )
by the chain rule. As a result, for each vψj ∈ ker KvFo we have
vψj ,
since η ∗ Fo = Fo ◦ η(t) = Fo for all t (cf. section 3.2) and in the second equality
we have used (3.5) of Proposition 3.2.1. This also shows that Adη(t) ker KvFo ⊆
⊥ ⊥
ker KvFo , and therefore Adη(t) ker KvFo ⊆ ker KvFo .
Lemma 5.4.2. Let vG (t) be a solution of (5.11) with initial condition vG (0) = vGo .
Then
ker
ker
vG (t) = Ad∗η−1 (t) vGo
for all t.
Proof. Let vG (t) be a solution of (5.11) with initial condition vGo . Write vG (t) =
P ∗ P
i gi (t)vψi and Adη −1 (t) vGo = i ai (t)vψi , using the orthonormal basis {vψi , λi }
124
of eigenvectors provided by KvFo . By Lemma 5.2.1
X X X
gi (t)vψi = ai (t)vψi − λi hi (t)vψi (5.20)
i i
where J(t) = vH(t) ◦η(t) is the Jacobi field along η(t) with initial conditions J(0) =
∗
ker
0, J 0 (0) = vGo . Write vG(t)
ker
P P
= j gj (t)vψj and Adη −1 (t) vGo = j aj (t)vψj .
Then by (5.20)
gi (t) = ai (t) − λi hi (t)
and therefore
gj (t) = aj (t)
ker
ker
vG(t) = Ad∗η−1 (t) vGo
Z t
ker ker
vH(t) = Adη(t) Adη−1 (s) vG(s) ds
0
Z t ker
= Adη(t) Adη−1 (s) Ad∗η−1 (s) vGo ds
0
Z t ker
= Adη(t) Adη−1 (s) Ad∗η−1 (s) vGo ds = Adη(t) (Ω(t)vGo )ker . (5.21)
0
In the third inequality we have used Lemma 5.4.1: since ker KvFo is invariant
under Adη(t) , applying Adη(t) to the part of Ad∗η−1 (s) vGo in ker KvFo is equivalent
125
to applying Adη(t) to Ad∗η−1 (t) vGo and then orthogonally projecting onto ker KvFo .
This gives the second equality (5.17) . The next Lemma will allow us to write
Z t ker Z t
Adη−1 (s) Ad∗η−1 (s) vGo ds = Adη−1 (s) Ad∗η−1 (s) vG
ker
o
ds.
0 0
⊥ ⊥
Ω(t) ker KvFo ⊆ ker KvFo
and
Ω(t) ker KvFo ⊆ ker KvFo ,
Rt
where Ω(t)vGo = 0
Adη−1 (s) Ad∗η−1 (s) vGo ds.
⊥ P
Proof. Suppose vGo ∈ ker KvFo and write vGo = gi (0)vψi . Using the formulas
for Adη−1 (t) , Ad∗η−1 (t) and KvFo found in Propositions 3.2.1 and 3.2.2 and using the
fact that each vψi is an eigenvector of KvFo with non-zero eigenvalue we write
X
Adη−1 (s) Ad∗η−1 (s) vGo = gi (0)vRη−1 4−1 Rη 4ψi 6= 0.
i
By Proposition 3.2.2,
λi ψi = 4−1 {4Fo , ψi }.
126
Since the Laplacian is an isomorphism and λi 6= 0:
1
4ψi = {4Fo , ψi },
λi
and therefore
1
vRη−1 4−1 Rη 4ψi = vR 4−1 Rη {4Fo ,ψi }
λi η−1
1
= vR 4−1 {4Fo ,Rη ψi }
λi η−1
1
= Adη KvFo Adη−1 vψi ,
λi
where in the second equality we have used that Rη(t) preserves the Poisson bracket
(see section 2.3) and vFo is a stationary solution.
⊥
By Lemma 5.4.1 Adη−1 vψi ∈ ker KvFo which implies that KvFo Adη−1 vψi ∈
⊥ ⊥
ker KvFo and thus Adη KvFo Adη−1 vψi ∈ ker KvFo . Therefore
X 1 ⊥
Adη−1 (s) Ad∗η−1 (s) vGo = gi (0) Adη KvFo Adη−1 vψi ∈ ker KvFo .
i
λi
Z t X Z t
⊥
Ω(t)vGo = Adη−1 (s) Ad∗η−1 (s) vGo ds = fi (s) ds vψi ∈ ker KvFo .
0 i 0
s ker
Let vGo ∈ Te DHam with vG o
6= 0. By Lemma 5.4.3 and (5.21) we have
ker
vH(t) = Adη(t) (Ω(t)vGo )ker
127
ker ⊥
ker
= Adη(t) Ω(t)vG o
+ Ω(t)vGo
ker
ker
= Adη(t) Ω(t)vG o
ker
= Adη(t) Ω(t)vG o
Remark 5.4.4. In Chapter 3 we saw that all solutions to the equation (5.11) were
constant. That is, every Jacobi field along a geodesic of isometries has constant
velocity. Moreover any Jacobi field whose initial velocity was an eigenvector of
KvFo vanished for some t∗ > 0. One may apply the above results to constant
speed Jacobi fields along general stationary geodesics. It is possible to prove that if
⊥
there exists an eigenvector of KvFo which lies in ker ad∗vFo + KvFo ∩ ker KvFo
then η(t) has conjugate points. In particular, the only constant speed Jacobi fields
along η(t) which vanish at some time t∗ > 0 are those fields whose velocity is an
eigenvector of KvFo .
If η(t) is a stationary geodesic generated by an eigenfunction of the Laplacian,
i.e. those functions on M satisfying
−4F = λF λ > 0,
then the constant speed Jacobi fields along η(t) have initial velocities given by
Moreover, if there exists a function Ψ such that (λI + (−4))−1 Ψ(Fo ) = vψi ∈
/
ker KvFo is an eigenfunction of KvFo then η(t) has conjugate points.
128
APPENDIX A
SOBOLEV SPACES
Our main reference for this appendix is Taylor, [T]. Let Ω ⊂ Rn be an open
bounded set with smooth boundary ∂Ω, and let Ω̄ denote the closure of Ω. Define
C ∞ (Ω, Rn ) to be the set of smooth functions from Ω to Rn that can be extended
to a smooth function on some open subset of Rn containing Ω̄. Let C0∞ (Ω, Rn ) =
{f ∈ C ∞ (Ω, Rn ) : The support of f is contained in a compact subset of Ω}.
An m multi index is a set of m ordered integers. If k = (k1 , . . . km ) is an m
multi index then set |k| = m ∞ n k
P
i=1 ki . If f ∈ C (Ω, R ) define D f to be
∂ |k|
Dk f =
∂xk11 . . . ∂xkmm
Z
Dk f (x)2 dx.
X
kf kH s =
Ω 0≤|k|≤s
Now H s (Ω, Rn ) (resp. H0s (Ω, Rn ))is defined to be the completion of C ∞ (Ω, Rn )
(resp. C0∞ (Ω, Rn )) in the norm k·kH s . The space H s (Ω, Rn ) is a Hilbert space
129
with the inner product
Z X
(f, g)H s = Dk f (x) · Dk g(x) dx.
Ω 0≤|k|≤s
k
∇ f = g(∇k f, ∇k f ) = g i1 j1 · · · g ik jk ∇k f ∇k f
i1 ···i j1 ···jk
.
k
∂f
For k = 1, (∇f )i = ∂xi
while
2 ∂ 2f ∂f
− Γkij
∇f ij
= .
∂xi ∂xj ∂xk
For f ∈ C ∞ (M ) define
X Z
kf k2H s = g(∇k f, ∇k f ) dµ,
0≤|k|≤s M
∇k f
2 2 .
X
= L
(A.1)
0≤|k|≤s
130
inner product
X Z
(f, g)H s = g(∇k f, ∇k g) dµ.
0≤|k|≤s M
If M is compact with two Riemannian metrics g and g̃ then the norms induced
by each of these metrics are equivalent so that the definition of H s (M ) does not
depend on the metric.
Theorem A.2.1. Let M be a compact manifold with boundary and suppose that
dim M
s> 2
+ k. Then,
1. H s (M ) embeds continuously into C k (M ).
2. H s (M ) is a ring under point wise multiplication.
Theorem A.2.2. For any integers s ≥ 0 and t > s, the space H t (M ) embeds
compactly into H s (M ).
131
order s are in L2 . This makes sense in view of the trivialization because a section
can be thought of locally as a map from Rn to Rm .
When E is the tangent bundle T M of M we define H s (T M ) to be the com-
pletion of C ∞ (T M ) in the norm defined by (A.1). The Sobolev space H s (T M )
is a ring under point-wise multiplication and embeds continuously in C k (E) for
dim M
s> 2
+ k. As before, compactness of M is used to show that the definition of
H s (T M ) is independent of the metric.
It is convenient to work in coordinates on M and for this we shall use the
trivialization. For a section X : M → T M of the tangent bundle of M and a
chart Ui ⊂ M we write
X = X l ∂xl
XX X
kXkH s =
Dk X l
2 ,
L (Ui )
i l 0≤|k|≤s
where
Z 21
2
kgkL2 (Ui ) = |g(x)| dx .
Ui
More generally, we can use the covariant derivative on M to give a coordinate free
description of this norm and inner product:
X Z
kXkH s = g(∇k X, ∇k X) dµ
0≤|k|≤s M
X Z
(X, Y )H s = g(∇k X, ∇k Y ) dµ,
0≤|k|≤s M
132
The Riemannian metric g(·, ·) on M gives rise to an inner product gx h·, ·i on
Tx∗ M for each x ∈ M , and then to an inner product on Λp Tx∗ M , via
X
hα1 ∧ . . . ∧ αk , β1 ∧ . . . ∧ βk i = (sgnπ) gx α1 , βπ(1) · . . . · gx αk , βπ(k) ,
π
where π ranges over the set of permutations of {1, ...k}. Consequently, there is an
L2 inner product on k-forms (i.e. sections of Λk ) given by
Z
(α, β) = hα, βi dµ. (A.2)
M
dim
XM
α= fi1 ···ip dxi1 ∧ · · · ∧ dxip ,
i1 ,...ip =1
where fi1 ...ip is a collection of functions and each dxij is basis element of T ∗ M . The
Sobolev norm is then defined analogously to the Sobolev norm of vector fields:
X dim
XM X
kαkH s =
Dk fi1 ···ip
2 .
L (Uj )
j i1 ···ip =1 0≤|k|≤s
X Z
∇k α, ∇k β dµ
(α, β)H s =
0≤|k|≤s M
where
X
∇k α, ∇k β = (sgnπ) gx ∇k α1 , ∇k βπ(1) · . . . · gx ∇k αp , ∇k βπ(p) .
π
133
APPENDIX B
Our main reference for this section is McDuff and Salamon, [M-S]. A Sym-
plectic structure on a smooth manifold M is a non-degenerate, closed two form
ω ∈ Ω2 (M ). That is, each tangent space (Tx M, ωx ) is a Symplectic vector space.
The manifold is necessarily of even dimensional and the n-fold wedge product
ω ∧ ... ∧ ω
for u, v ∈ Tx S 2 . Where h·, ·i is the euclidean inner product induced from the am-
bient space and × is the cross product of vectors. The non-degeneracy condition
is purely algebraic and means that there is an isomorphism between the tangent
and cotangent bundles of M :
ω] : T M → T ∗M
u 7→ iu ω = ω(x, ·).
134
The closedness of ω is a geometric condition.
x 7→ Jx : Tx M → Tx M
Jx2 = −I.
x 7→ gx : Tx M × Tx M → R
135
B.2 Symplectomorphisms
η ∗ ω = ω.
then η(t) ∈ Dω (M ) for every t if and only if X(t) is Symplectic for every t.
Moreover, if X, Y are Symplectic then [X, Y ], the Lie bracket of vector fields, is
Symplectic.
where we have used Cartan’s formula for the Lie derivative and that ω is closed.
Now let X and Y be Symplectic vector fields with corresponding flows η(t)
and γ(t). Then
[X, Y ] = LY X = ∂t |t=0 γ(t)∗ X
136
so that
show that the vector field XH is tangent to the level sets H = const. of H.
A smooth function F ∈ C ∞ (M ) is constant along the orbits of the flow of H
if and only if the Poisson bracket
vanishes. Because ω is closed, The Poisson bracket defines a Lie algebra structure
on the space of smooth functions on M .
137
Proposition B.2.2. Let (M, ω) be a Symplectic manifold.
(i) Wherever defined, the Hamiltonian flow ηH (t) is a Symplectomorphism,
which is tangent to the level surfaces of H.
(ii) For every Hamiltonian function H : M → R and every Symplectomor-
phism η ∈ Dω (M ) we have XH◦η = η ∗ XH .
(iii) The Lie bracket of two Hamiltonian vector fields XF and XG is the Hamil-
tonian vector field [XF , XG ] = X{F,G} .
Proof. Statement (i) was proved in A.4.1. Statement (ii) follows from the identity
Hence
The shows that the Hamiltonian vector fields form a Lie subalgebra of Sym-
plectic vector fields. The map X 7→ XH is a surjective Lie algebra homomorphism
from the Lie algebra of smooth functions on M with the Poisson bracket to the Lie
algebra of Hamiltonian vector fields. The kernel of this homormorphism consists
of constant functions.
138
APPENDIX C
FREDHOLM OPERATORS
The reader is referred to Schechter, [Sc], for details on this appendix. Let
X and Y be normed vector spaces. A linear operator K from X to Y is called
compact if the domain of K is X and for every sequence {xn } ⊂ X such that kxn kX ≤ C,
the sequence {Kxn } has a subsequence which converges in Y . The set of all compact
Theorem C.1.1. Let X be a normed vector space and Y a Banach space. Then,
K(X, Y ) is a closed, two-sided ideal in L(X, Y )
If K ∈ K(X, Y ) then K ∗ ∈ K(Y ∗ , X ∗ ).
139
is satisfied for any x ∈ H, and some compact operator K ∈ K(H). Then A has
closed range and finite dimensional kernel.
kwn kH σ . kKwn kH .
Now this estimate implies wn must contain a convergent subsequence which implies
that the unit sphere in the kernel of A is locally compact. Consequently, the kernel
is finite dimensional.
140
denoted by ρ(A), is defined by
The spectrum, denoted by σ(A), is the complement of the resolvent set, i.e. σ(A) =
R \ ρ(A). A real number λ is said to be an eigenvalue of A if
2. R(A) is closed in Y
141
The set of Fredholm operators will be denoted by Φ(X, Y ). If Y = X and K ∈
K(X) then by Theorem A.2.4 I − K is a Fredholm operator.
The index of a Fredholm operator is defined by
Theorem C.2.1. Let A ∈ L(X, Y ) and assume that there are operators A1 ,
A2 ∈ L(X, Y ), K1 ∈ K(X), K2 ∈ K(Y ) such that
1. A1 A = I − K1 on X
2. A2 A = I − K2 on Y .
Then A ∈ Φ(X, Y ).
i(A + K) = i(A).
142
In particular, it follows that the index of any Fredholm operator of the form
A = I + K is zero. More generally, if A is invertible then B = A + K, for some
compact operator K, is a Fredholm operator of index zero. A Fredholm operator
has index zero if and only if it is the sum of an invertible operator and a compact
operator.
One may relax the above restrictions somewhat, and consider a slightly more
general operator. An operator A is semi-Fredholm if the range of A is closed
and at least one of the other two conditions hold. Unbounded semi-Fredholm
operators were treated by Kato in [K2] and [K3] where it is shown that the index
of a semi-Fredholm operator is well-defined.
Denote the set of semi-Fredholm operators between two Banach spaces X and
Y , whose kernel is finite dimensional, by Φ+ (X, Y ).
Observe that this Theorem is analogous to Theorem C.2.4 and the estimate
characterizes semi-Fredholm operators with finite dimensional kernel. If we denote
by A∗ , the adjoint of a semi-Fredholm operator A, then A is semi-Fredholm with
finite dimensional cokernel if A∗ ∈ Φ+ (Y ∗ , X ∗ ). Define the set Φ− (X, Y ) to be
those bounded linear operators A with A∗ ∈ Φ+ (Y ∗ , X ∗ ). We now collect some
properties of semi-Fredholm operators which are analogous to the Fredholm case:
143
3. If A ∈ Φ+ (X, Y ) and K ∈ K(X, Y ), then A + K ∈ Φ+ (X, Y ).
4. If A ∈ Φ− (X, Y ) and K ∈ K(X, Y ), then A + K ∈ Φ− (X, Y ).
144
APPENDIX D
SPECTRAL INTEGRALS
Let B(R) be the Borel σ-algebra of subsets of R, and let H be a Hilbert space
with inner product h·, ·i.
Remark D.1.2. Let E be a spectral measure on the σ-algebra B(R). Each vector
x ∈ H gives rise to a scalar positive measure Ex on B(R) by
1
Ex,y = (Ex+y − Ex−y + iEx+iy − iEx−iy ) . (D.1)
4
145
Using E,one may define spectral integrals of the form
Z Z
I(f ) = f (t) dE(t) = f dE
Ω Ω
Z
I(χM ) ≡ χM (t) dE(t) := E(M )
Ω
Z
A= λ dE(λ).
J
R
If F is another spectral measure on B(R) such that A = R
λ dF (λ), then we have
E(M ∩ J ) = F (M ) for all M ∈ B(R).
LetS ≡ S(R, B(R), EA ) is the set of all EA -a.e. finite Borel functions f : R →
C ∪ {∞}. For a function f ∈ S, we write f (A) for the spectral integral I(f ). Then
Z
f (A) = f (λ) dEA (λ)
146
is a normal operator on H with dense domain
Z
2
D(f (A)) = x∈H: |f (λ)| d hEA (λ)x, xi < ∞
T (0) = I
147
T (t + s) = T (t)T (s)
lim T (t)x = x
t→0
Z
itA
e := eitλ dE(λ), t ∈ R.
R
148
unitary group U .
149
BIBLIOGRAPHY
E. Ebin, D., The manifold of Riemannian metrics, Proc. symp. Pure Math. xv,
Ameri. Math. Soc. 1970, 11-40 and Bull. Am. Math. Soc. 74 (1968) 1002-1004.
E1. Eells, J., On the Geometry of Function Spaces, Sympo. de Topoloqia Algebrica
Mexico (1958) 303-307.
E2. Eells, J., A Setting for Global Analysis, Bull. Amer. Math. Soc. 72 (1966),
751-807.
150
E-N. Engel, K-J., Nagel, R., One-Parameter Semigroups for Linear Evolution
Equations, Springer (2000).
E-M. Ebin, D., Marsden, J., Groups of Diffeomorphisms and the Motion of an
Incompressible Fluid, Ann. of Math. 92 (1970)
E-M-P. Ebin, D., Misiolek, G., Preston, S., Singularities of the Exponential Map
on the Volume-Preserving Diffeomorphism Group, Geom. Funct. Anal. 16
(2006).
Ev. Evans, L., C., Partial Differential Equations, Ameri. Math. Soc. (1998).
G1. Grabowski, J., Free Subgroups of Diffeomorphism Groups, Fund. Math. 313
(1988), 103-121
G2. Grabowski, J., Derivative of the Exponential Mapping for Infinite Dimen-
sional Lie Groups, Ann., Global Anal. Geom. 11 (1993), 213-220.
G-T. Gilbarg, D., Trudinger, N., Elliptic Partial Differential Equations of Second
Order, New York; Springer, (1983).
Gro. Grossman, N., Hilbert Manifolds without Epiconjugate Points, Proc. Amer.
Math. Soc. 16 (1965), 1365-1371.
H. Hamilton, R. S., The Inverse Function Theorem of Nash and Moser, Bull.
Amer. Math. Soc., Vol. 7 No. 1 (1982)
He. Hebey, E., Non-Linear Analysis on Manifolds: Sobolev Spaces and Inequali-
ties, Ameri. Math. Soc. (1999).
151
Hi. Hirsch, M. W., Differential Topology, Springer-Verlag (1988)
H-T. Holm, D. D., Tronci, C., The Geodesic Vlasov Equation and its Integrable
Moment Closures, J. Geom. Mech. 2, (2009), 181-208.
H-Z. Hofer, H., Zhender, E., Symplectic Invariants and Hamiltonian Dynamics,
Birkhauser-Verlag (1994)
K1. Kato, T., Perturbation Theory for Nullity, Deficiency and other quantities of
Linear Operators, J. d’Anal. Math, 6 (1958), 261-322
K2. Kato, T., Perturbation Theory for Linear Operators, Springer-Verlag (1966)
Kh. Khesin, B., Dynamics of Symplectic Fluids and Point Vortices, Geom. Funct.
Anal. 22 (2012) 1444-1459.
Kh-M. Khesin, B.,A., Misiolek, G., Shock Waves for the Burgers equation and
Curvatures of Diffeomorphism Groups, Proc. Steklov Math. Inst. 259 (2007),
116-144.
K-M. Kriegl, A., Michor, P. W., The Convenient Setting of Global Analysis, Amer.
Math. Soc. (1997)
152
Ko. Kopell, N., Commuting Diffeomorphisms, Proc. Symp. Pure Math. Amer.
Math. Soc. 14 (1970), 165 - 184.
M-Me. Markus, L., Meyer, R., Generic Hamiltonian Systems are neither Inte-
grable nor Ergodic, Memoirs of the AMS, vol. 144, Providence, RI, (1974).
M-S. McDuff, D., Salamon, D., Introduction to Symplectic Topology, Oxf. Uni.
Press. (1998).
M1. Misiolek, G., Stability of Flows of Ideal Fluids and the Geometry of the
Groups of Diffeomorphisms, Ind. Uni. Math. Jour., Vol. 42, No. 1 (1993)
215-235
M2. Misiolek, G., Conjugate Points in Dµ (T2 ), Proc. Amer. Math. Soc. 124
(1996), 977-892.
M3. Misiolek, G., The Exponential Map on the Free Loop Space is Fredholm,
Geom. Funct. Anal. 7 (1997), 954-696.
153
M-W. Marsden, J., Weinstein, A., The Hamiltonian Structure of the Maxwell-
Vlasov Equations, Physica D, 4, (1982), 394-406.
M-W-R-S. Marsden, J., Weinstein, A., Schmid, R., Spencer, R., Hamiltonian Sys-
tems and Symmetry Groups with applications to Plasma Physics, Atti Acad.
Sci. Torino Cl. Sci. Fis. Math. Natur. , 117 , (1983), 289-340.
Pa. Pazy, A., Semigroups of Linear Operators and Applications to Partial Differ-
ential Equations, Springer-Verlag (1983)
P1. Preston, S., Eulerian and Lagrangian Stability of Fluid Motions, Doctoral
Dissertation, SUNYSB, 2002
P2. Preston, S., For ideal fluids, Eulerian and Lagrangian instabilities are equiv-
alent, Geom. Funct. Anal. 14 (2004).
P4. Preston, S., On the Volumorphism Group, the First Conjugate Point is the
Hardest, Comm. Math. Phys. 276 (2006) 493-513.
R-S. Ratiu, T., Schmid, R., Three Remarkable Diffeomorphism Groups, Math. Z.
177 (1980) 81-100.
154
Sch. Schmudgen, K., Unbounded Self-adjoint Operators on Hilbert Space, Springer
(2012).
Sh1. Shnirelman, A., Generalized Fluid Flows, their Approximations and Appli-
cations, GAFA, Geom. Funct. Anal. 4 (1994) 586-620.
Sm. Smale, S., An infinite dimensional version of Sard’s theorem, Amer. J. Math.
87 (1965), 861-865.
Ta. Cohen-Tannoudji, C., Laser Cooling and Trapping of Neutral Atoms: Theory,
Phys. Rep. 219 (1992) 153.
T2. Taylor, M., Partial Differential Equations II, Qualitative Studies of Linear
Equations, Springer (2011).
155
W. Wolibner, W., Un Theoreme sur l’existence du mouvement plan d’un fluide
parfait, homogene, incompressible, pendant un temps infiniment longue, Math.
Z. 37 698-726 (1933)
This document was prepared & typeset with pdfLATEX, and formatted with
nddiss2ε classfile (v3.0[2005/07/27]) provided by Sameer Vijay.
156