Joakim Gasoy Master Thesis
Joakim Gasoy Master Thesis
Circle Manifolds
Joakim Gåsøy
Master’s Thesis, Spring 2020
This master’s thesis is submitted under the master’s programme Mathematics,
with programme option Mathematics, at the Department of Mathematics,
University of Oslo. The scope of the thesis is 60 credits.
The front page depicts a section of the root system of the exceptional
Lie group E8 , projected into the plane. Lie groups were invented by the
Norwegian mathematician Sophus Lie (1842–1899) to express symmetries in
differential equations and today they play a central role in various parts of
mathematics.
Abstract
i
Acknowledgements
First and foremost, I wish to thank my supervisor Kim Anders Frøyshov for
presenting me with this topic and helping me find valuable sources. Thank you
for your patience and assistance.
I also wish to dedicate a thank you to August and Haakon, who assisted me
in proofreading, and for their general support.
Finally I wish to thank my partner, friends and family for their invaluable
support throughout the years.
iii
Contents
Abstract i
Acknowledgements iii
Contents iv
1 Introduction 1
1.1 Terminology and notation . . . . . . . . . . . . . . . . . . . . 2
1.2 Thurston’s geometrization conjecture . . . . . . . . . . . . . . 2
4 Orbifolds 29
Bibliography 31
iv
CHAPTER 1
Introduction
Seifert manifolds were introduced and classified by Herbert Seifert in the 1930s.
Interest in these manifolds has had a resurgence due to the work of William
Thurston and his celebrated geometrization conjecture. This was proposed
by Thurston in 1992 and proved later by Grigorij Perelman in 2002. The
geometrization conjecture essentially states that any closed, oriented 3-manifold
can be canonically decomposed into something called geometric 3-manifolds. In
other words the classification of closed, oriented 3-manifolds can be reduced
to the classification of geometric 3-manifolds. Thurston further shows that
there are eight classes of geometric 3-manifolds (or eight geometries), seven of
which are completely classified. The only remaining case is the spaces with a
hyperbolic geometry.
The Seifert manifolds are special in that they occupy precisely six of these
Thurston geometries, and one can identify their geometry by calculating their
Seifert invariants. We give a short introduction to the geometrization conjecture
and its relevance in this first introductory chapter.
Seifert manifolds are a generalization of circle bundles where we allow certain
"singularities". One major part of this thesis to show that due to this circle
"bundle" we can consider most Seifert manifolds as special G-manifolds where
G is the circle group S 1 , with Lie group structure inherited from U (1). A
G-manifold is a manifold with a group action of the Lie group G. We introduce
Seifert manifolds and G-manifolds in general in the second chapter, and show
that most Seifert manifolds are S 1 -manifolds with a certain finiteness condition.
As the Seifert manifolds cover as many as six of the eight geometries, they
naturally include a lot of different examples. In the third chapter we introduce
some classes of manifolds and show that they are Seifert manifolds by identifying
them as certain S 1 -manifolds as explained in chapter 2.
In the fourth and final chapter we introduce the concept of an orbifold, a
generalization of manifolds. We prove that the orbit space of G-manifolds with
a certain finiteness condition are indeed orbifolds. In particular, all Seifert
manifolds that are S 1 -manifolds have this property. Since all the information
of the Seifert manifold is encoded into this orbifold and its projection map,
we conjecture that you can calculate the Seifert invariants by studying these
orbifolds further.
The main sources used for this thesis is [Sco83] and [Mar16], which summarize
the geometrization conjecture and the classification of seven of the geometries.
Further we have used a standard textbook about smooth manifolds ([Tu08]),
lecture notes by professors at the University of Oslo, and some articles for
1
1. Introduction
specific results.
2
1.2. Thurston’s geometrization conjecture
S 3 , R3 , H3 , S 2 × R, H2 × R, Nil, Sol, SL
g2 .
S 3 , R3 , S 2 × R, H2 × R, Nil, SL
g2
3
CHAPTER 2
for x ∈ D2 .
We say that two fibered solid tori T(p,q) and T(p0 ,q0 ) are isomorphic if there
exists a diffemorphism between them sending fibers to fibers. We must have
p = p0 for two tori to be isomorphic, since they are p-fold and p0 -fold covered
by a trivial fibered solid torus, respectively. However it is sufficient to have
q = ±q 0 (mod p) for them to be isomorphic. We can verify this by looking at
how the diffeomorphisms
for n ∈ Z and
5
2. Seifert manifolds as G-manifolds
6
2.1. Seifert manifolds
(S, (p1 , q1 ), . . . , (pk , qk )) where S is the base space of M and (p1 , q1 ), . . . , (pk , qk )
are the orbit invariants of the singular fibers.
Much like compact 2-manifolds can be classified into geometries based on
their Euler characteristic, there is a similar result for Seifert manifolds. We now
state the definition of the relevant invariants ([Mar16, p. 37, p. 159, p. 312]).
Definition 2.1.4. The Euler characteristic of an n-complex X is given by the
integer
X n
χ(X) = (−1)i Ci
i=1
is given by
k
X 1
χ(M ) = χ(S) − 1−
i=1
pi
is given by
k
X qi
e(M ) =
p
i=1 i
7
2. Seifert manifolds as G-manifolds
2.2 G-manifolds
In this section we introduce basic definitions of G-manifolds and the equivariant
tubular neighbourhood theorem from [Jän68, p. 1-4].
Definition 2.2.1. Let G be a compact Lie group and M a smooth manifold. By
an action of G on M we will mean a smooth map G × M → M , where the
image of (g, x) is written gx, with the following properties:
(i) g1 (g2 x) = (g1 g2 )x (Compatibility)
(ii) 1x = x (Identity)
A smooth manifold M together with a smooth action by G on M is called
a G-manifold, simply denoted by M when the G-action is understood.
For each g ∈ G we can define the (left-)action of g by
lg : M → M
x 7→ gx
By the previous definition this map is smooth and has a smooth inverse lg−1 ,
hence being a diffeomorphism.
Some examples of G-manifolds include any Lie group acting on itself, and
this also induces an H-manifold structure on G for any subgroup H ⊆ G.
We now define the concepts we need to know about G-manifolds.
Definition 2.2.2. Let M be a G-manifold and let x ∈ M . Then the set Gx =
{gx | g ∈ G} is the orbit of the point x, and Gx = {g | gx = x} is the stabilizer
group of the point x. Further we denote by M/G the orbit space, that is the set
of all orbits {Gx | x ∈ M }, equipped with the quotient topology, that is: A set
of orbits is open in M/G whenever their union is open in M .
Let M be a G-manifold and x ∈ M . The Lie group G is a G-manifold by
definition, and hence also a Gx -manifold, since Gx is a subgroup of G. The
orbit space G/Gx is by definition
G/Gx → Gx
gGx 7→ gx
showing the connection between the orbits and stabilizer groups in a different
way.
Definition 2.2.3. Let M and N be two G-manifolds. A map φ : M → N is said
to be equivariant if for every g ∈ G and x ∈ M we have
φ(gx) = gφ(x)
8
2.2. G-manifolds
G ×H V → E
which is indeed a G-vector bundle isomorphism. We conclude that any
G-vector bundle over G/H is determined by its H-module at the point 1H.
Now let M be a G-manifold and x ∈ M . We denote by Nx = Tx M/Tx Gx
the normal space of the orbit Gx at the point x. The identification of Gx
with G/Gx as we showed earlier and the identification of G-vector bundles over
homogenous spaces shows that we can idnetify the normal bundle of Gx in M
with the bundle G ×Gx Nx → G/Gx .
We can now formulate the equivariant tubular neighbourhood theorem:
Theorem 2.2.5. Let M be a G-manifold. There exists an equivariant diffeo-
morphism between a G-invariant open neighbourhood of the zero section in
G ×Gx Nx and a G-invariant open neighbourhood of Gx in M , which sends the
zero section G/Gx to the orbit Gx in a canonical way.
This is actually just a version of the tubular neighbourhood theorem found
in [Mil74, p.115-117] for G-manifolds:
Theorem 2.2.6. Let M be a Riemannian manifold and S be a smoothly embedded
submanifold of M . There exists an open neighbourhood of S in M which is
diffeomorphic to the total space of the normal bundle which maps each point
x ∈ S to the zero normal vector at x.
We will not reproduce all the details of the proofs, but we will sketch the
proofs.
For the tubular neighbourhood theorem one denotes the total space of the
normal bundle as E and defines the set
where x ∈ S and v is a normal vector at x. It is then shown that this set maps
diffeomorphically to a neighbourhood of S in M by the exponential map
Exp : E(ε) → M
(x, v) 7→ γ(1)
9
2. Seifert manifolds as G-manifolds
where γ : [0, 1] → M is a geodesic arc with γ(0) = x and velocity vector equal
to v at 0, which exists for sufficiently small ε. It is then shown that E(ε) is
indeed diffeomorphic to E, proving the result.
For the equivariant version it is sufficient to show that we can find an
equivariant Riemannian metric on M for the tubular neighbourhood to be
equivariant. Here it is shown that for any Riemannian metric p on M one
can construct an equivariant Riemannian metric p on M . Note that the
diffeomorphism between E(ε) and E is not generally equivariant, so the result
does not extend to all of E unless M is complete.
10
2.3. Circle manifolds
Proof. We start by noting that all orbits of M are circles. The orbit of S 1 is
homeomorphic to S 1 /Sx1 . Since the stabilizer Sx1 is finite by assumption, the orbit
S 1 x is homeomorphic to a finite quotient of S 1 , which is again homeomorphic
to S 1 .
By the equivariant tubular neighbourhood theorem there exists a S 1 -
invariant neighbourhood U of S 1 x that is diffeomorphic to E(ε) ⊆ E. Here
E is the total space of the normal bundle of S 1 x in M and E(ε) is the set of
normal vectors of length less than ε for some ε > 0. Since we only have finitely
many non-trivial stabilizer groups, we can choose ε > 0 small enough such that
U contains no orbits with non-trivial stabilizer group other than possibly S 1 x
itself.
The identification of S 1 ×Sx1 Nx with E from the equivariant tubular neigh-
bourhood theorem shows that U is indeed a fibered solid torus. Since S 1 is
11
2. Seifert manifolds as G-manifolds
Proving the converse requires a few lemmas which we will now state and
prove.
Lemma 2.3.4. Let X and Y be π-related vector fields on the manifolds M and
N , respectively, where π : M → N is a smooth map. If γ is an integral curve
of X then π ◦ γ is an integral curve of Y .
Lemma 2.3.5. There exists a smooth vector field X on the fibered solid torus
T(p,q) such that each integral curve is a fiber.
Proof. We use the fact that R is a Lie group to define the left-invariant vector
d
field Y generated by dt 0
∈ Te (R). This vector field is smooth by [Tu08, p.181].
We can use this to define a smooth vector field Xe on the infinite solid cylinder
2
R × D as the following composition:
π Y i
R × D2 − → T R ⊕ T D2 ' T (R × D2 ).
→ R −→ T R ,−
r(t, x) = (rt, x)
φ : R × D2 → S 1 × D2 ' T(p,q)
q
(t, x) 7→ (e2πit , e2πit p x)
12
2.3. Circle manifolds
is a covering map sending each fiber of the cylinder to a fiber of the torus. We
will now show that the pushforward X := φ(X) e is a well-defined smooth vector
field on T(p,q) . It will then have the fibers of T(p,q) as integral curves, since φ
sends integral curves to integral curves by Lemma 2.3.4.
Since φ is a covering map and R × D2 is simply connected, the infinite
cylinder is the universal cover of T(p,q) . The deck transformation group is
homomorphic to the fundamental group Z of T(p,q) . If n ∈ Z then this group
acts on R × D2 as n · (t, x) 7→ (t + pn, x). Since X e is invariant under the
previously defined R-action, it is also invariant under this Z-action. Hence if V
is any connected component of φ−1 (U ) where U ⊆ Tp,q is evenly covered and
ψ : U → V is the inverse of π|V , then the composition X e ◦ ψ is independent on
the choice of V . Thus the composition
ψ X π∗
→ R × D2 −→ T (R × D2 ) −→
T(p,q) − T (T(p,q) )
e
is well-defined.
q
−2πit p
The map ψ can be expressed explicitly as ψ(z, x) = ( log(z)
2πi , e x), where
the value of log(z) lies in V . This is clearly smooth as it is the composition of
smooth elementary functions. The map X e is smooth because it is a smooth
vector field, and the differential π∗ is smooth because all differentials are smooth.
This means that X = π∗ ◦ X e ◦ ψ is a smooth vector field on T(p,q) .
Note that the vector field X constructed in the above proof makes T(p,q)
fiber oriented. This is because it is smooth, hence continuous, and restricts
to a basis for each fiber on each fiber. Furthermore the vector field −X has
the opposite orientation of X, so we can always find a smooth vector field
corresponding to either orientation.
Proposition 2.3.6. Every Seifert manifold M with oriented fibers is a finitely
stabilized U (1)-manifold.
Proof. Let {Uj } be a cover of the Seifert manifold M where each Uj is a fibered
solid torus, and let {Xj } be a collection of smooth vector fields on {Uj } with
fibers as integral curves, as constructed in 2.3.5. We choose our Xj ’s such that
they all have the same fiber orientation. We can now find a partition of unity
subordinate to Ui to construct a smooth vector field X on M where all the
integral curves are fibers. We need to modify this vector field in such a way
that the period of all the integral curves of non-singular fibers is the same.
For any fibered solid torus Uj ' T(p,q) in our cover we can pull back the
vector field X|Uj to a vector field X ej on R × D2 . Since we know that the images
of the integral curves of this vector field are the fibers R × {x}, this vector
field must be given by (X ej )(t,x) = f ∂
∂t (t,x) for some smooth, strictly positive
function f . Now let γx : R → R be the integral curve of the fiber R × {x}
(evaluated only in the first coordinate) such that γx (0) = 0. By the definition
of integral curves we have γx0 (t) = (X ej )γ (t) . In terms of calculus this leads to
x
13
2. Seifert manifolds as G-manifolds
14
2.3. Circle manifolds
e2πi0 x = γx (0)
= x.
Every non-singular fiber has trivial stabilizer group, because they have period
1. Every singular fiber has finite stabilizer group of order p, because they have
period 1/p which means that
From 2.3.3 and 2.3.6 we see that a Seifert manifold admits a finitely stabilzied
S 1 -structure if and only if it is fiber orientable. From the proofs we see that
we can calculate the invariants p1 , . . . , pk of a fiber oriented Seifert manifold
M = (S, (p1 , q1 ), . . . , (pk , qk )) by finding the orders of the non-trivial stabilizer
groups. Moreover the base space of a Seifert manifold is the orbit space M/S 1 ,
since the points of the orbit space are the fibers of M by definition. We will
show that this is an orbifold in chapter 4.
15
CHAPTER 3
17
3. Examples of Seifert manifolds
Next we need to check that the circle action is well-defined by checking that uz
is contained in Σ(p, q, r).
m m m
(u p z1 )p + (u q z2 )q + (u r z3 )r ) = um z1p + um z1q + um z1r
= um (z1p + z2q + z3r )
=0
because um 6= 0.
m m m 2m 2m 2m
|u p z1 |2 + |u q z2 |2 + |u r z3 |2 = |u| p |z1 |2 + |u| q |z2 |2 + |u| r |z3 |2
= |z1 |2 + |z2 |2 + |z3 |2
=1
because |u| = 1.
To calculate the stabilizer groups of the circle action we consider two cases.
The first case we consider is when z1 , z2 , z3 6= 0. Then u is in the stabilizer
group of z if we have
uz = z
m m m
(u z1 , u z2 , u z3 ) = (z1 , z2 , z3 )
p q r
m m m
up =uq =ur =1
The last line shows that the order of u must divide m m m
p , q and r . From
elementary number theory we can find k1 , k2 and k3 such that
m m m m m m
k1 + k2 + k3 = gcd , , .
p q r p q r
Let d = gcd m , m m
, m m m
p q r . Then dp , dq and dr are integers, hence p, q and r
divide m
d . But since m is defined to be the least common multiple of p, q and r
this means that d = 1. It now follows that
m m m
u1 = uk1 p +k2 q +k3 r
m k1 m k2 m k
3
u= up uq ur
u=1
Hence the stabilizer group of z is trivial.
The remaining case is to check when either of z1 , z2 or z3 is equal to zero.
Note that we must have at least two non-zero coefficients in order to satisfy the
requirements to be a point on the Brieskorn manifold. We start by checking for
z3 = 0. Similar to before we get
uz = z
m m m
(u z1 , u z2 , u 0) = (z1 , z2 , 0)
p q r
m m
u p = u q = 1.
In this case the order of u must divide m m
p and q . We can find k1 and k2 such
that
m m m m
k1 + k2 = gcd , .
p q p q
18
3.1. Brieskorn manifolds
m m
To simplify notation we set d = gcd p, q . This leads to
m m
ud = uk1 p +k2 q
m k1 m k2
= up uq
=1
m m
Hence the order of u divides gcd p, q , which means that we have
2πin
u=e d .
with n ∈ Z. This shows that the stabilizer group of z is the cyclic group of
order d. In the particular case where p, q and r are relatively prime we get that
m m
d = gcd ,
p q
pqr pqr
= gcd ,
p q
= gcd (qr, pr)
= r.
Vk = {z ∈ Σ(p, q, r) | zk = 0}
for k = 1, 2, 3. We need to show that each set Vk splits into finitely many orbits.
Assume that z3 = 0. We then get the following relationship:
z1p + z2q = 0
z1p = −z2q
|z1 |p = |z2 |q
q
|z1 | = |z2 | p
p q
1 − |z2 |2 = |z2 | p
2q
1 − |z2 |2 − |z2 | p = 0
The final equation has a strictly decreasing function of |z2 | on the left-hand side.
This means that the equation has a unique solution |z2 |, and since the function
yields 1 for |z2 | = 0 and −1 for |z2 | = 1, the solution lies between 0 < |z2 | < 1.
This shows that the length of z2 (and hence z1 ) is only dependent on p and q.
19
3. Examples of Seifert manifolds
20
3.2. Spherical space forms
SU (2) ' {a + bi + cj + dk | a, b, c, d ∈ R, a2 + b2 + c2 + d2 = 1}
= {α + βj | α, β ∈ C, |α|2 + |β|2 = 1}
⊆ H.
Now S 3 inherits the Lie group structure of SU (2), and by letting it act
on itself it is therefore an SU (2)-manifold. Taking this further we know that
U (1) ⊆ SU (2) and that S 1 can be identified with U (1), so we also inherit an
S 1 -action on S 3 .
From [Mar16, p. 365] we have a group homomorphism
φ : SU (2) → SO(3)
with kernel {1, −1}. The preimage under φ of a subgroup of SO(3) is called
a binary group, and Martelli shows that all finite subgroups Γ ∈ SU (2) up
to conjugation are either cyclic or binary. The finite subgroups of SO(3) are
the cyclic groups Cn , the dihedral groups D2m , the tetrahedral group T12 , the
octahedral group O24 and the icosahedral group I60 . Thus the finite subgroups
of SU (2) up to conjugation are known to be the cyclic groups Cn , the binary
∗ ∗
dihedral groups D4n , the tetrahedral group T24 , the binary octahedral group
∗ ∗
O48 and the binary icosahedral group I120 . Thought of in terms of quaternions
they can be expressed as follows:
2πin
Cm = {e m | 0 ≤ n < m} (Cyclic groups)
πin πin
∗
D4m = {e , e j | 0 ≤ n < 2m}
m m (Binary dihedral groups)
∗ ±1 ± i ± j ± k
T24 = D8 ∪ (Binary tetrahedral group)
2
∗ ∗ a+b
O48 = T24 ∪ √ | a, b ∈ D8 , a 6= b (Binary octahedral group)
2
√ √
∗ ∗ a ( 5 − 1)b ( 5 + 1)c
I120 = T24 ∪ ± ± ± | (a, b, c, d) = σ(1, i, j, k)
2 4 4
(Binary icosahedral group)
for all even permutations σ and where D8 = Q8 = {±1, ±i, ±j, ±k} is the set
of Lipschitz units.
The key to define a circle action is to let Γ ⊆ SU (2) act on S 3 from the
right and let S 1 act on S 3 from the left to obtain a S 1 -action on S 3 /Γ. We do
this by defining
g[x] := [gx]
21
3. Examples of Seifert manifolds
In the case where −1 ∈ Γ (that is, Γ having even order) all the stabilizer groups
would contain −1, so in that case we use the fact that the quotient group
S 1 /{±1} is diffeomorphic to S 1 by the map
φ : S 1 /{±1} → S 1
[eθi ] 7→ e2θi
and define a circle action on S 3 /Γ by using the quotient group instead. We do
this by setting
g[x] := [gx]
for [g] ∈ S 1 /{±1}. Note that this definition does not depend on the repre-
sentative g, since we either have g = eiθ or g = −eiθ , and [gx] = [−gx] when
1
the order of Γ is even. It is easy to see that if S[x] is the stabilizer group of
3 1 1 1
[x] ∈ S /Γ under the S -action, then (S /{±1})[x] = (S[x] )/{±1}. Thus we
can calculate the stabilizer groups of the S /{±1}-action from the S 1 -action.
1
1
In particular if S[x] has finite order n, then (S 1 /{±1})[x] has order n/2.
For the remainder of this section let G denote S 1 for Γ of odd order and
denote S 1 /{±1} for Γ of even order. We will now prove that this G-action yields
a Seifert structure on S 3 /Γ. We start by proving the following proposition.
Proposition 3.2.1. Let Γ ∈ SU (2) be a finite subgroup. The number of orbits
of the set
Im(h)
Γ0 = | h ∈ Γ, Im(h) 6= 0
|Im(h)|
under the action of Γ-conjugation is the same as the number of orbits of S 3 /Γ
with non-trivial stabilizer group under the previously defined G-action.
Proof. The stabilizer group of [x] ∈ S 3 /Γ by the G-action is given by
G[x] = {g ∈ G | [gx] = [x]}
= {g ∈ G | ∃h ∈ Γ, gx = xh}
= {g ∈ G | ∃h ∈ Γ, g = xhx}.
In other words, g is in the stabilizer group of [x] if and only if there exists h ∈ Γ
such that g = xhx. To find the elements [x] with non-trivial stabilizer group we
need to solve the equation g = xhx for g with non-zero imaginary part, that is
g = cos(ψ) + i sin(ψ) for sin(ψ) 6= 0. We note that any quaternion h ∈ H can be
written as h = cos(θ) + q sin(θ) where q ∈ H0 is the normalized imaginary part
of h (see for instance [Frø19]). The set Γ0 is the set of normalized imaginary
parts of elements h ∈ Γ, so we can write h = cos(θ) + q sin(θ) for q ∈ Γ0 . We
then get that
g = xhx
cos(ψ) + i sin(ψ) = x(cos(θ) + sin(θ)q)x
cos(ψ) + i sin(ψ) = x cos(θ)x + x sin(θ)qx
cos(ψ) + i sin(ψ) = cos(θ) + sin(θ)xqx
22
3.2. Spherical space forms
23
3. Examples of Seifert manifolds
g = cos(θ) + i sin(θ) ∈ G.
Thus we have x = gyh for g ∈ G and h ∈ Γ, showing that [x] and [y] are on the
same G-orbit in S 3 /Γ. Therefore for every orbit in Γ0 there exists a G-orbit in
S 3 /Γ.
Proof. From the previous proposition we know that the number of singular
G-orbits of S 3 /Γ is the same as the number of orbits of Γ0 under Γ-conjugation.
Since Γ0 is finite, hence has finitely many orbits, there are only finitely many
singular orbits of S 3 /Γ.
Now recall that g ∈ G is in the stabilizer group of [x] if and only if there
exists h ∈ Γ such that g = xhx. Since there are only finitely many h ∈ Γ, there
are only finitely many g ∈ G[x] , hence all stabilizer groups are finite.
Since the G-action on S 3 /Γ is finitely stabilized, S 3 /Γ is a Seifert manifold.
Cyclic groups
Let Γ = Cn . Then Γ0 = {±i}. Since Γ consists of rotations about the axis
through i and −i, both i and −i are left fixed, so the orbits are Cl(i) and
Cl(−i).
Every non-real element of Cn has normalized imaginary part ±i, hence all
non-real elements are sent to both conjugacy classes. Therefore if n is odd we
have |Cl(±i)∗ | = n−1, so both the orbit of Cl(i) and Cl(−i) have stabilizer group
24
3.2. Spherical space forms
i. Rotating 2π 4π
3 and 3 about the axis through a vertex and the centroid of
the opposing face.
ii. Rotating π about the axis through the midpoints of opposing edges.
ii. Rotating π about the axis through the midpoints of opposing edges.
25
3. Examples of Seifert manifolds
2π 4π
iii. Rotating 3 and 3 about the axis through opposing vertices.
i. Rotating 2π 4π
5 , 5 ,
6π
5 and 8π
5 about the axis through the centroid of
opposing faces.
ii. Rotating π about the axis through the midpoints of opposing edges.
2π 4π
iii. Rotating 3 and 3 about the axis through opposing vertices.
and
∗ ∗
i+j+k −i − j − k ±1 ± i ± j ± k
Cl √ √= Cl = .
3 3 2
∗ ∗
−i−j−k
Thus we have |Cl(i)∗ | = 6 and |Cl i+j+k
√
3
| = |Cl √
3
| = 16. So the
orbit of Cl(i)∗ has stabilizer group of order 6/2+1 = 4, and the orbits Cl( i+j+k
√
3
)
and Cl( −i−j−k
√
3
) have stabilizer groups of order 16/2 + 1 = 9.
∗
Next let Γ = O48 . We now have the orbits of normalized faces Cl(i),
normalized vertices Cl( i+j+k
√
3
) and of normalized edges Cl( i+j
√ ) of the cube,
2
with
|Cl(i)∗ | = 18
∗
i+j+k
Cl √ = 16
3
∗
i+j
Cl √ = 12
2
So the orbit of Cl(i) has stabilizer group of order 18/2 + 1 = 10, the orbit of
Cl( i+j+k
√
3
) has stabilizer group of order 16/2 + 1 = 9, and the orbit of Cl( i+j
√ )
2
has stabilizer group of order 12/2 + 1 = 7.
∗
Finally let Γ = I120 . We now have the orbits of normalized
q √edges q Cl(i), nor-
√
i+j+k 5+ 5 5− 5
malized vertices Cl( 3 ) and of normalized faces Cl 10 i + 10 j
√
26
3.2. Spherical space forms
|Cl(i)∗ | = 30
∗
i+j+k
Cl √ = 40
3
∗
√ √
s s
5 + 5 5 − 5
Cl i+ j = 48
10 10
So the orbit of Cl(i) has stabilizer group of order 30/2 + 1 = 16, the orbit
of Cl( i+j+k
√
3
) has stabilizer group of order 40/2 + 1 = 21, and the orbit of
q √ q √
5+ 5 5− 5
Cl 10 i + 10 j has stabilizer group of order 48/2 + 1 = 25.
∗
In summary, S 3 /T24 has three singular orbits, one with stabilizer group C4
∗
and two with stabilizer group C9 . S 3 /O48 has three singular orbits with stabilizer
∗
groups C10 , C9 and C7 . S 3 /I120 has three singular orbits with stabilizer groups
C16 , C21 and C25 .
27
CHAPTER 4
Orbifolds
29
4. Orbifolds
Now let V be the image of U (ε) under the projection π. From the projection
π : U (ε) → V ⊂ M
π=ψ◦θ
The map ψ sends each equivalent class [y] = {gy | y ∈ U (ε), g ∈ Gx } to
the orbit Gy. This is clearly injective, because if Gy 6= Gz then [y] 6= [z] since
[y] ⊂ Gy and [z] ⊂ Gz. ψ is also surjective because π is a projection. The
continuity of ψ and its inverse is clear from the topologies of U (ε)/Gx and
V : A subset {[y] ∈ U (ε)/Gx | y ∈ W } of U (ε)/Gx is open if and only if W
is open in U (ε). A subset {Gy ∈ V | y ∈ W } is open if and only if the set
{gy ∈ P | g ∈ G, y ∈ W } is open in P . However, W is open in U (ε) if and only
if it is the restriction of an open set in P , showing that a subset of M is open
if and only if it is the image of a subset of U (ε)/Gx under ψ. Hence ψ is a
homeomorphism.
The triple (V, Gx , φ ◦ ψ) an oribfold chart, because U (ε) is homeomorphic
to U ⊂ Rk , which means that U (ε)/Gx is homeomorphic to U/Gx . Since Gx is
distance preserving it is orthogonal, and since it is also finite it is homeomorphic
to a finite subgroup of O(n).
By our construction all the transition maps between charts are trivially
compatible, since none of the charts containing singular points intersect. We
complete the orbifold atlas by adding every finite intersection of every chart.
Finally, from [Lee12, p.543] we have that M is Hausdorff and second count-
able since G is compact and acts smoothly on P .
Final remarks
We have shown that we can find the invariants (p1 , . . . , pk ) and base space S
of a fiber oriented Seifert manifold M by considering it as a finitely stabilized
S 1 -manifold. This allows us to calculate the Euler characteristic χ of M .
To calculate the Euler number e of M we need to also know the invariants
q1 , . . . , qk . These invariants are decided by the particular way S 1 acts on M ,
which means that it should be able to recover them from the projection map to
M/S 1 somehow.
By developing a theory of smooth orbifolds, similar to smooth manifolds, we
might be able to calculate the Euler number e of M directly by looking at the
projection M → M/S 1 . We leave others to investigate this perspective further.
30
Bibliography
31