153
Normal congruence subgroups of the Bianchi groups and related groups by Robert M Scarth A thesis submitted to the Faculty of Science at the University of Glasgow for the degree of Doctor of Philosophy July 13, 2003 c Robert M Scarth 1998

Normal congruence subgroups of the Bianchi groups and related groups

Embed Size (px)

DESCRIPTION

Maths PhD thesis.

Citation preview

Normal congruence subgroups of the

Bianchi groups and related groups

by

Robert M Scarth

A thesis submitted to

the Faculty of Science

at the University of Glasgow

for the degree of

Doctor of Philosophy

July 13, 2003

c©Robert M Scarth 1998

2

Summary

Let O be an order in an imaginary quadratic number field. This thesis is mainly concerned

with normal subgroups of SL2(O) and of PSL2(O). Suppose that O is a maximal order then

O is the ring of integers in the number field and the group PSL2(O) is a Bianchi group.

In chapter one we discuss the geometric background of these groups and introduce some

fundamental algebraic concepts; those of order and level. We also discuss the Congruence

subgroup problem. Chapter two is a discussion of the fundamental theorem of Zimmert

[93]. In chapter three we discuss PSL2(O) where O is not a maximal order. We derive a

formula for their index in the Bianchi groups and presentations for some of these groups.

In particular we derive a presentation for PSL2(Z[√−3]) and using this presentation get a

partial classification of the normal subgroups of PSL2(Z[√−3]).

Chapter four generalizes a result of Mason and Pride [62] about SL2(Z) to all but finitely

many SL2(O). This result shows that for an arbitrary normal subgroup of N SL2(O)

there is no relationship between the order and level of N . This is in distinction to the groups

SLn(O), n > 3, where the order and level of a normal subgroup coincide. This answers a

question of Lubotzky’s.

Let O be an order in an imaginary quadratic number field. Then O is a Noetherian

domain of Krull dimension one and has characteristic zero. Chapter five discusses SL2 over

the class of all Noetherian domains of Krull dimension one, including those of non-zero

characteristic. In particular we generalize the work of Mason [58] and derive a relationship

between the order and level of a normal congruence subgroup of SL2(K) for any Noetherian

domain of Krull dimension one, K. In chapter six we apply this work to SL2(O) and

i

construct a new and vast class of normal non-congruence subgroups of SL2(O). Finally we

take a closer look at some particular PSL2(O).

ii

“Results! Why man, I have gotten a lot of results. I know several thousand things that

don’t work.”

- Thomas Edison [50] p.121.

iii

List of Notation

R A Ring.

R∗ The group of units in the ring R.

charR The characteristic of R.

L A Local Ring.

u A unit in a ring R.

Fq The field of q elements.

d A positive square-free integer. p.4

m A positive integer. p.4

Od The ring of integers in the imaginary quadratic number field Q(√−d). p.4

Od,m The Order of index m in Od. p.4

ω

1+i√

d2

if d ≡ 3 (mod 4)

i√d else

p.4

D The discriminant of Q(√−d). p.21

q An ideal in a ring R.

eij The matrix with a 1 in the (i, j)th position and zero’s elsewhere. p.9

Eij(r) Denotes I + reij.

Eij An elementary matrix; Eij(1).

E(x) Denotes the matrix

x 1

−1 0

. p.43

D(µ) Denotes the matrix

µ 0

0 µ−1

. p.43

iv

SLn(R) The group of n× n matrices over R with determinant 1.

PSLn(R) SLn(R) with the centre factored out.

GLn(R) The group of n× n matrices over R with non-zero determinant.

PSL2(Z) The Modular group. p.2

PSL2(Od) The Bianchi groups. p.4

A The matrix

0 1

−1 0

∈ SL2(R). p.4

T The matrix

1 1

0 1

∈ SL2(R). p.4

U The matrix

1 mω

0 1

∈ SL2(Od,m). p.4

a, t, u The images of A, T , and U in PSL2(R), or PSL2(Od,m). p.4

M An arbitrary 2× 2 matrix

a b

c d

.

trM The trace of the matrix M , ie a+ d. p.5

detM The determinant of M , ie ad− bc.

En(R) The subgroup of SLn(R) generated by the elementary matrices. p.9

Un(R) The normal subgroup of SLn(R) generated by the unipotent matrices. p.24

En(R, q) The subgroup of En(R) generated by the q-elementary matrices. p.10

Hn(R, q) The subgroup of SLn(R) consisting of the q-scalar matrices. p.10

SLn(R, q) The principal congruence subgroup of level q. p.11

PSLn(R, q) The principal congruence subgroup of level q in PSL2(R). p.108

H(q) Alternative notation to Hn(R, q).

Γ(q) Alternative notation to SLn(R, q).

∆(q) Alternative notation to En(R, q). p.70

Un(R, q) The normal subgroup of SLn(R) generated by the unipotent matrices

in SLn(R, q). p.30

o(S) The order of the subgroup S. p.11

v

l(S) The level of the subgroup S. p.10

E0 (2, R; q) The set of normal subgroups of SL2(R) of level zero and order q. p.55

Z(d,m) The Zimmert Set. p.21

r(d,m) The cardinality of Z(d,m). p.21

ρ(d,m) The largest rank of a free quotient of Γd,m. p.32

Nd,m(n) The number of normal congruence subgroups of Γd,m of index exactly n. p.110

G An arbitrary group.

Fr The free group of rank r.

[x, y] The commutator of x and y, ie x−1y−1xy.

G′ The commutator subgroup of G.

Gab The abelianization of G, ie G/G′.

N An arbitrary normal subgroup of G.

N(x1, . . . ) The normal closure in G of the elements x1, . . . . p.97

an(G) The number of subgroups of G of index exactly n. p.97

gcd(x, y) The greatest common divisor of x and y.

H2 Hyperbolic 2-space. p.1

H3 Hyperbolic 3-space. p.3

vi

Contents

Summary i

Contents vii

Acknowledgements viii

Statement x

1 Introduction 1

1.1 Geometric Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Hyperbolic geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.2 Hyperbolic 3-space . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 The normal subgroups of SLn(R) . . . . . . . . . . . . . . . . . . . . . . . . 8

1.2.1 SLn over a local ring . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.2.2 SLn over an arbitrary ring . . . . . . . . . . . . . . . . . . . . . . . . 10

1.2.3 The Congruence Subgroup Problem . . . . . . . . . . . . . . . . . . . 15

1.3 Some Number Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 Zimmert’s Theorem 21

2.1 Hyperbolic 3-space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2 Zimmert’s Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.3 Explicit result about the free generators . . . . . . . . . . . . . . . . . . . . 25

2.4 Unipotent matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

vii

2.5 Computation of Zimmert Sets . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3 The groups PSL2(Od,m) 39

3.1 Computation of |PSL2(Od) : PSL2(Od,m)| . . . . . . . . . . . . . . . . . . . 39

3.2 A miscellany of results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.3 A presentation for SL2(Z[√−3]) . . . . . . . . . . . . . . . . . . . . . . . . 46

3.4 The Group PSL2(O3,2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.4.1 HNN and amalgam decomposition . . . . . . . . . . . . . . . . . . . . 50

3.4.2 Some Consequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

3.5 Presentations for some other PSL2(Od,m). . . . . . . . . . . . . . . . . . . . 58

4 Non-standard normal subgroups 60

4.1 Non-standard normal subgroups of SL2(Z) . . . . . . . . . . . . . . . . . . . 61

4.2 Non-standard normal subgroups of SL2(Od,m) . . . . . . . . . . . . . . . . . 62

5 Order and level... 72

5.1 Primary decomposition in Noetherian domains . . . . . . . . . . . . . . . . . 72

5.2 Wohlfahrt’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

5.3 Preliminary results about SL2(L) . . . . . . . . . . . . . . . . . . . . . . . . 79

5.4 Order and level of a normal congruence subgroup . . . . . . . . . . . . . . . 94

6 Non-congruence subgroups 106

6.1 Counting finite index subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 106

6.2 The growth of non-congruence subgroups in the groups SL2(Od,m) . . . . . . 109

6.3 SL2(L) where L is a local image of Od,m . . . . . . . . . . . . . . . . . . . 110

6.4 Non-congruence subgroups of small index in the Bianchi groups . . . . . . . 117

6.5 The groups PSL2(Od), d = 2, 7, 11 and PSL2(O3,2) and PSL2(O1,2) . . . . 124

6.5.1 Normal congruence subgroups of PSL2(O2) . . . . . . . . . . . . . . 127

131

Bibliography 134

viii

Acknowledgements

I would like to thank my supervisor, Dr. Alec Mason, for his patience, kindness and en-

couragement during the last three and a half years. I would also like to thank everyone at

the Department of Mathematics at the University of Glasgow for contributing to a friendly

and stimulating environment. Thanks also to the EPSRC for giving me financial support

between October 1995 and September 1998, the Department of Mathematics here at Glas-

gow for giving my funding to attend various conferences, and my parents for paying my rent

while I completed this work.

I would like to thank Andy, Iain, Mark Brightwell, Mark Ratter, Mohammad, and

Stephen for the many interesting and stimulating discussions we had about Politics, Foot-

ball, Philosophy, Theology, and even some Mathematics. I’d have learnt a lot less without

you.

Finally, I would like to thank my family,especially my parents, for their unfailing support

and belief throughout 21 years of education.

ix

Statement

This thesis is submitted in accordance with the regulations for the degree of Doctor of

Philosophy in the University of Glasgow.

Chapter one provides some motivation and introduces some fundamental concepts. Chapter

two is a discussion of Zimmert’s theorem [93]. Chapters three and six and the independ-

ent work of the author and chapters four and five are the joint work of the author and his

supervisor. References are given throughout.

x

Chapter 1

Introduction

1.1 Geometric Background

We are interested in groups acting on Hyperbolic space. Hyperbolic geometry is the best

known example of a non-Euclidean geometry. The importance of the parallel axiom and the

development of non-Euclidean geometry in the history of Mathematics, indeed the history

of Western thought cannot be overstated. See [41, 67, 82] for brief and accesible accounts.

1.1.1 Hyperbolic geometry

Let

H2 = z ∈ C : Im(z) > 0

and equip H2 with the following metric

ds =

√dx2 + dy2

y

We now have the Poincare half plane model for two dimensional hyperbolic geometry. The

geodesics in H2 are straight lines and semicircles orthogonal to the real axis. The historical

point is that given a geodesic L and a point P not on L there are infinitely many geodesics

passing through P which do not intersect L. That is Euclids parallel axiom does not hold.

1

We are interested in distance preserving maps, or rigid motions. The hyperbolic distance

between z1, z2 ∈ H2 shall be denoted ρ(z1, z2).

Definition. A function from H2 onto itself which preserves hyperbolic distance is called an

isometry. The group of isometries is denoted Isom(H2).

Consider the following set

M =

z 7→ az + b

cz + d: a, b, c, d ∈ R, ad− bc = 1

this is the set of Moebius transformations of C. They map H2 onto itself and are all isometries.

M can be identified with the group PSL2(R) via the obvious map.

Theorem. ([38] theorem 1.3.1) Isom(H2) is generated by PSL2(R) and the map z 7→ −z.

PSL2(R) is of index 2 in Isom(H2).

Let X be a metric space and G a group acting on X.

Definition. A family Mα : α ∈ A of subsets of X is called locally finite if for any compact

set K ⊂ X we have Mα ∩K 6= ∅ for only finitely many α ∈ A.

Definition. We say that a group G acts properly discontinuously on X if the G-orbit of any

point x ∈ X is locally finite.

Definition. A closed connected F ⊂ X, with int(F ) 6= ∅, is a fundamental region for G if

1. GF = X.

2. int(F ) ∩ g(int(F )) = ∅ ∀ 1 6= g ∈ G.

The existence of a fundamental region allows us, in particular to find a presentation for

the group (see [49]).

Definition. G 6 PSL2(C) is said to be discrete if it contains no sequence of matrices conver-

ging elementwise to the identity. Discrete subgroups of PSL2(R) are called

Fuchsian groups.

2

Example 1.1.1. PSL2(Z) is obviously discrete and therefore a Fuchsian group.

Suppose that G 6 PSL2(R) is not discrete. So G contains a sequence gn such that

gn → I. Suppose that G has a fundamental region F ⊂ H2 and let x ∈ int(F ). Then

gnx → x as n → ∞. So ∃N such that ∀ n > N int(F ) ∩ gn(int(F )) 6= ∅. Contradiction.

Hence G cannot have a fundamental region.

Let Γ be a Fuchsian group. Let p ∈ H2 be not fixed by any non-trivial element of Γ. The

Dirichlet fundamental polygon for Γ centred at p is

Dp(Γ) =z ∈ H2 : ρ(z, p) 6 ρ(Tz, p) ∀T ∈ Γ

Theorem. ([38] theorem 3.2.2) For every Fuchsian group Γ and every p ∈ H2 not fixed by

a non-trivial element of Γ, Dp(Γ) is a connected convex fundamental region for Γ.

Example 1.1.2. Let Γ = PSL2(Z). The set

F =

z ∈ H2 : |z| > 1, |Re(z)| 6 1

2

is a Dirichlet region for Γ centred at ki, some k > 1.

The theory of Fuchsian groups is of great interest and has been extensively studied. See

[6, 7, 38].

1.1.2 Hyperbolic 3-space

The upper half-space in Euclidean three-space gives a convenient model of 3-dimensional

hyperbolic space

H3 = (z, r) ∈ C× R : r > 0

which we equip with the hyperbolic metric

ds2 =dx2 + dy2 + dr2

r2

The group PSL2(C) acts on H3 in the following way. Let M ∈ PSL2(C) where

M =

α β

γ δ

3

Then

M(z, r) =

((δ − γz)(αz − β)− r2γα

τ,r

τ

)where

τ = |γz − δ|2 + r2|γ|2

Under this action the hyperbolic metric is PSL2(C)-invariant. SL2(C) acts on H3 in exactly

the same way. As above we want a class of discrete subgroups of PSL2(C). We make use of

the following

Proposition. [19] Let A be a discrete subring of C with a one then SL2(A) is a discrete

subgroup of SL2(C).

Proposition. [19] The discrete subrings of C with a one are

1. Z.

2. The ring of integers Od = Z + ωZ in an imaginary quadratic number field.

3. The orders Od,m = Z +mωZ in an imaginary quadratic number field.

where d is a positive square-free integer, m is a positive integer, and

ω =

1+i√

d2

if d ≡ 3 (mod 4)

i√d else

This gives us a class of discrete subgroups of SL2(C) and of PSL2(C). We have already

met the group PSL2(Z). It is known as the Modular group. It was Picard who, in 1883,

first studied PSL2(Z [i]), and this group is known as the Picard group [67, 73]. The groups

PSL2(Od) are called the Bianchi groups. They were first studied by Bianchi in the 1890s

[8, 9] as a natural class of discrete subgroups of PSL2(C). See [23] chapter 7 for a discussion

of their action on H3 and [25] for a discussion of their algebraic properties. The groups

PSL2(Od,m) are of finite index in the Bianchi groups. See [23] for a general treatment of

discrete subgroups of PSL2(C) acting on H3. The Modular group and the Bianchi groups

are the first arithmetic examples of such groups and are of great interest in number theory.

4

We take the opportunity here to introduce three matrices which will be very important in

what follows. Let R be any commutative ring with a one and let

A =

0 1

−1 0

, T =

1 1

0 1

∈ SL2(R)

We will also need the following

U =

1 mω

0 1

∈ SL2(Od,m).

We denote the corresponding matrices in PSL2 by a, t, u. This will be a general convention.

A description of a fundamental region for the Bianchi groups in H3 can be found in

[23] section 7.3. Swan [84] has used this to derive presentations for the group SL2(Od) for

d = 1, 2, 3, 7, 11, 5, 6, 15, 19. It is then easy to derive a presentation for PSL2(Od).

We now look at matrices of finite order in SL2(C) and SL2(Od,m). The results are well

known but the presentation is our own. Let

M =

a b

c d

∈ SL2(C)

and let tr = a+ d be the trace of M . Recall that conjugate matrices have the same trace.

Lemma 1.1.1. If tr = ±2 and M 6= I then M is conjugate to ±T and so is of infinite

order.

Proof. We can suppose that

M =

1 + α β

γ 1− α

Now a b

c d

1 1

0 1

d −b

−c a

=

1− ac a2

−c2 1 + ac

So letting a =

√β, and c = i

√γ and then choosing b, d, so that ad − bc = 1 we get the

result.

5

Lemma 1.1.2. If tr2 6= 4 (ie tr 6= ±2) then M is conjugate to α 0

0 α−1

for some α ∈ C.

Proof. Suppose that z ∈ C is a fixed point of M ie

az + b

cz + d= z

So cz2 + (d − a)z − b = 0. The discriminant of this quadratic is (d − a)2 + 4bc = (a +

d)2 − 4(ad− bc) = tr2 − 4. Now tr2 6= 4 so M has two distinct fixed points z1, and z2. Let

w = (z2 − z1)−1 and consider 1 −z2

w −z1w

a b

c d

1 −z2

w −z1w

−1

Using the fact that cz2i +(d−a)zi− b = 0, for i = 1, 2, we can show that this matrix is equal

to α 0

0 α−1

for some α ∈ C.

Lemma 1.1.3. If |tr| > 2 then M is of infinite order.

Proof. Clearly tr2 6= 4, so M is conjugate to α 0

0 α−1

for some α ∈ C. Suppose that |α| = 1, so |α−1| = 1 and |tr| = |α + α−1| 6 |α| + |α−1| = 2.

So |tr| 6 2. Contradiction. So |α| 6= 1, so |α| > 1, or |α−1| > 1. So |αn| → ∞, or |α−n| → ∞

as n→∞. Now α 0

0 α−1

n

=

αn 0

0 α−n

.

So Mn 6= I for every n ∈ Z.

6

Lemma 1.1.4. If |tr| = 2 and tr2 6= 4 then M is of infinite order.

Proof. Now tr2 6= 4 so M is conjugate to α 0

0 α−1

for some α ∈ C. As above if |α| 6= 1 then M is of infinite order. So suppose that |α| = 1, so

α−1 = α. So α + α−1 = α + α = 2Re(α). So 2 = |tr| = |α + α−1| = 2|Re(α)|, so Re(α) = 1.

Hence α = α−1 = 1 and M = I. Contradiction.

Lemma 1.1.5. Suppose that |tr| < 2. Then if tr /∈ R then M is of infinite order.

Proof. tr2 6= 4 so M is conjugate to α 0

0 α−1

for some α ∈ C and where α + α−1 = tr. Now if |α| 6= 1 then, as above, M is of infinite

order. So suppose that |α| = 1 then α−1 = α so tr = α+ α = 2Re(α) ∈ R.

Lemma 1.1.6. If tr = 0 then M2 = −I.

Proof. Now M is conjugate to

X =

α 0

0 α−1

some α ∈ C such that α+ α−1 = 0. So α2 + 1 = 0, so α = ±i. It is then easy to verify that

X2 = −I.

Lemma 1.1.7. If tr = 1 then M3 = −I and if tr = −1 then M3 = I.

Proof. Now M is conjugate to

X =

α 0

0 α−1

for some α ∈ C such that α + α−1 = tr. First suppose that tr = −1. Then α2 + α + 1 = 0,

so α(α2 + α + 1) = 0 and so α3 = 1. Thus X3 = I and so M3 = I. Similarly if tr = 1 then

α3 = −1 and so X3 = −I. Hence result.

7

Theorem 1.1.8. Let I 6= M ∈ SL2(Od,m). Then M is of finite order if and only if tr = 0,

or ±1.

Proof. Now tr ∈ Od,m so |tr|2 ∈ Z. First, if |tr|2 > 4 then M is of infinite order. Suppose

that |tr|2 < 4, so |tr|2 = 0, 1, 2, or 3, but if M is of finite order tr ∈ Od,m ∩ R = Z, so

|tr| = 0, or 1. Now if |tr| = 0 then tr = 0 and so M2 = −I. If |tr|2 = 1 then tr = ±1, so

M6 = I.

1.2 The normal subgroups of SLn(R)

Let R be a ring, with a one. Let n ∈ N, n > 2, and form the group SLn(R). It is natural to

ask the following

Question. What are the normal subgroups of SLn(R)?

The case where R is a field is simple; literally:

Theorem. [21] Let F be any field, n ∈ N, n > 2. Then PSLn(F) is simple with two

exceptions:

PSL2(F2) ∼= S3 and PSL2(F3) ∼= A4

where Fd denotes the field of d elements.

By another result of Dieudonne [21] SL2(F)′ = SL2(F), and so the only non-trivial

normal subgroup of SL2(F) is ±I. Note that the exceptions are both two dimensional

linear groups. We shall see that the two dimensional case is, in general, more complicated

than the higher dimensional cases. Further, when investigating normal subgroups of SL2(R)

one finds that ideals of index 2 or 3 in R play an important role. When we pass from fields

to rings we no longer have simplicity.

Example 1.2.1. Let m ∈ Z. Consider the following subgroup

Γ(m) = M ∈ SL2(L) : M ≡ I (mod m)

8

This is clearly a non-trivial normal subgroup of SL2(L). It is called the Principal congruence

subgroup of level m, and is a member of a very important class of normal subgroups, as we

shall see later on.

1.2.1 SLn over a local ring

The next easiest case after that of a field is a local ring,so let R = L, be a commutative

local ring, with maximal ideal m, and let Nm = |L : m|. We introduce two classes of normal

subgroup of SL2(L). Let q L be an ideal in L, then define

Γ(q) = M ∈ SL2(L) : M ≡ I (mod q)

this is the kernel of the natural map SLn(L) −→ SLn(L/q) and let

H(q) = M ∈ Γ : M ≡ kI (mod q), k ∈ L

H(q)/Γ(q) is the centre of SL2(L)/Γ(q). Let S 6 SL2(L), by the level of S, denoted l(S), we

mean the largest ideal, q of L such that Γ(q) 6 S, and by the order of S, denoted o(S), we

mean the smallest ideal, q of L such that S 6 H(q). Since Γ(q) 6 H(q) we have l(S) 6 o(S).

The first attempt to classify the Normal subgroups of SL2(L) was in 1961 by Klingenberg.

Theorem. [42] Let N 6 SL2(L) be of order q. Then N SL2(L) ⇔ l(N) = o(N) where,

for n = 2 we assume Nm 6= 3, and 2 is a unit.

Lacroix, in 1966 [43], dropped the condition that 2 was a unit and included the case

where Nm = 3 but was unable to deal with the case Nm = 2:

Theorem. Suppose that Nm > 2, and let N 6 SL2(L) be of order q. Then N SL2(L)⇔

o(N) = l(N) unless Nm = 3, and o(N) = L. If Nm = 3 and N SL2(L), o(N) = L then

N = SL2(L), or N = SL2(L)′.

However Lacroix provided examples of non-normal subgroups of SL2(L) of order L.

The case where Nm = 2 appears to be very complicated in general (see [43]). Mason [57]

has investigated the case where Nm = 2, m is principal and every ideal of L is a power of

9

m. In section (5.3) we investigate the normal subgroups of SL2(L) and introduce techniques

which allow us to deal with the case Nm = 2 and m nilpotent.

Klingenberg showed (roughly) that N SLn(L) ⇔ l(N) = o(N). This leads to the

following

Definition. Let S 6 SLn(L). If l(S) = o(S) we say that S is standard.

This gives rise to the standard criterion:

N SLn(L)⇔ N is standard

With slight modification these concepts carry over to the case of an arbitrary ring. We

remark that Costa and Keller [18] have characterized the normal subgroups of GL2(A) for an

arbitrary commutative local ring A in terms of certain commutator groups. Their solution

reduces to that of Klingenberg and Lacroix in the relevant cases. We mention it only in

passing here because we are mainly interested in the standard criterion, or, where that fails

the relationship between the order and level of a normal subgroup.

1.2.2 SLn over an arbitrary ring

We introduce the following subgroup of SLn(R):

En(R) =< I + reij : r ∈ R, i 6= j >

It is well known that when R = F is a field, SLn(F ) = En(F ). This remains true for some

rings. Firstly, in light of the previous section

Proposition. ([3] corollary 5.9.2) Let L be a semilocal ring. Then ∀n > 2 En(L) = SLn(L).

Proposition. ([33] proposition 2.4) Let R be a euclidean ring. Then ∀n > 2 En(R) =

SLn(R).

As Z is euclidean we have

Proposition. ([51] lemma 3.1) SL2(Z) = E2(Z).

10

Let Od be the ring of integers of Q(√−d). The Bianchi groups are the groups PSL2(Od),

we have already mentioned them in section (1.1.2). Now Od has a euclidean algorithm

⇔ d = 1, 2, 3, 7, 11 ([13] p.21), so if d = 1, 2, 3, 7, 11 then SL2(Od) = E2(Od). In fact

Proposition. ([14] Theorem 6.1)

SL2(Od) = E2(Od)⇔ d = 1, 2, 3, 7, 11

The groups PSL2(Od), d = 1, 2, 3, 7, 11 are called the Euclidean Bianchi groups. We can

now ask the following which is obviously related to our original question

Question. What are the En(R)-normalized subgroups of SLn(R)?

We now introduce an important concept, the SRn-condition. Let R be a ring. If

a1, . . . , an ∈ R such that∑Rai = R then ∃b1, . . . , bn−1 ∈ R such that

∑R(ai + bian) = R

then we say that R has stable range n, and we write SRn(R) or say R has SRn. The SR2-

condition is particularly important so we describe it separately: If Ra1 + Ra2 = R then

∃t ∈ R such that a1 + ta2 is a unit. See [16, 29, 85, 87, 88] for examples of rings with SR2.

Proposition. ([3] Proposition 5.3.4) Semilocal rings are SR2-rings.

Proposition. [3] If R is an SR2-ring then SL2(R) = E2(R).

We remark that Dedekind domains have SR3 ([3] theorem 3.5 page 239). We must also

modify our concept of level because, famously, in general not every normal subgroup of

SLn(R) contains a principal congruence subgroup. Let q R. Let

En(R, q) =< I + αeij : α ∈ q, 1 6 i, j 6 n, i 6= j >En(R)

Hn(R, q) = M ∈ SLn(R) : M ≡ kI (mod q), k ∈ R

Let S 6 SLn(R). By the level of S, denoted l(S), we mean the largest q R such that

En(R, q) 6 S. This is well defined because En(R, q1)En(R, q2) = En(R, q1 + q2). By the

order of S, denoted o(S) we mean the smallest q R such that S 6 Hn(R, q). As before,

we say S is standard if l(S) = o(S). We define a Principal congruence subgroup as before

SLn(R, q) = M ∈ SLn(R) : M ≡ I (mod q)

11

clearly En(R, q) 6 SLn(R, q). As in the case of a local ring Hn(R, q)/SLn(R, q) is the centre

of SLn(R)/SLn(R, q). In the case of a local ring our two concepts of level are the same

because:

Proposition. ([3] corollary 5.9.2) Let L be a local ring. Then ∀n > 2 En(L, q) = SLn(L, q).

Any subgroup of SLn(R) which contains a principal congruence subgroup is known as a

congruence subgroup. The question of whether every subgroup of finite index in SLn(R) is

a congruence subgroup or not is of great interest and is known as the Congruence Subgroup

Problem. We discuss it in the next section.

Most attempts to understand the normal subgroups of SLn(R) are centred round the

standard criterion, and we now describe some of these attempts.

Theorem. Let H 6 GLn(A). Then for n > 3, if A has SR2 [3], or is commutative [86], or

is a Banach algebra [88], or is von-Neumann regular [89] then

H is En(A)-normalized⇔ H is standard

We remark that there exist examples of rings for which the standard criterion fails for

n > 3 (See [28, 90] ). We now focus exclusively on commutative rings and ask how are

the E2(R)-normalized subgroups and the standard subgroups related? It turns out that the

answer depends very much on R.

Theorem. [3] IF R has SR2 and S 6 SL2(R) then S standard ⇒ S is E2(R)-normalized.

Costa and Keller ([17] theorem 2.6) have provided a partial converse

Theorem. Let R be an SR2 ring with 6 ∈ R∗. Then N SL2(R)⇒ N is standard.

The cases R = F2, or F3 show that 6 ∈ R∗ is necessary. Suppose now that A is a

Dedekind domain of arithmetic type (see the section at the end of this chapter on number

theory) so A has SR3, and suppose that A has infinitely many units. Serre has shown ([80]

Prop 2 p. 492)

12

Theorem. Let A be a Dedekind domain of arithmetic type and suppose that A∗ is infinite.

Then SL2(A) has no E2(A)-normalized subgroups of level zero and non-zero order.

Mason has proved the following

Theorem. [55] Let A be a Dedekind domain of arithmetic type with infinitely many units

then every standard subgroup of SL2(A) is E2(A)-normalized.

we also have

Theorem. [75] Let A be a Dedekind domain of arithmetic type with infinitely many units.

Then the E2(A)-normalized subgroups of SL2(A) are precisley the standard subgroups if and

only if the following three conditions hold for A:

1. All residue class fields of A have more than 3 elements.

2. 2 ∈ A∗ or 2 is unramified in A.

3. E2(A, a) = [E2(A), E2(A, a)] for every a A.

However Mason has shown that when A∗ is infinite the order and level of an E2(A)-

normalized subgroup are closely related.

Theorem. ([58]) Let N 6 SL2(A) be E2(A)-normalized. Let q = o(N), and q∗ = l(N).

Then

1. If A is contained in a number field and is not totally imaginary then 12q 6 q∗.

2. If A is contained in a number field and is totally imaginary then 12u0q 6 q∗.

3. If A is contained in a function field in one variable over a finite field then q3 6 q∗.

where u0 is defined as follows. Let m be the total number of roots of unity in A, and let u be

the A-ideal generated by u2 − 1 where u ∈ A∗. If m = 2 Let u0 = u + 2A and if m > 2 then

let u0 = u.

13

So we can see that in the two dimensional case the unit structure becomes important.

What happens to the standard criterion when the unit group is finite? Consider first the

group SL2(Z), here we see that not only does the standard criterion fail but it fails badly:

Theorem. [56, 62] The group SL2(Z) has 2ℵ0 non-normal standard subgroups and 2ℵ0 non-

standard normal subgroups.

In fact a more precise result is achieved. Let let E0(2,Z; q) denote the set of normal

subgroups of SL2(Z) of level zero, and order q.

Theorem. [62] Let 0 6= q Z then |E0(2,Z; q)| = 2ℵ0.

The situation in the groups SL2(Od) is similar

Theorem. [60] For every positive square-free integer d, the group SL2(Od) has 2ℵ0 non-

normal standard subgroups.

The obvious question now is

Question. Does every group SL2(Od) have 2ℵ0 non-standard normal subgroups?

Mason has shown

Theorem. [59] For every positive square-free integer d, the group SL2(Od) has 2ℵ0 normal

subgroups of level zero.

The only normal subgroup of SL2(Od) with order 0 is the trivial subgroup. Thus there

are uncountably many non-standard normal subgroups in the groups SL2(Od). Later we ex-

tend this to show that for all but finitely many (d,m), and all 0 6= qOd,m, |E0(2,Od,m; q)| =

2ℵ0 . The exceptions are almost certainly due to an inadequacy in our proof. Mason has ob-

tained similar results in the final case. Let C be a Dedekind domain of arithmetic type

contained in a function field and with finitely many units. Let Γ = SL2(C).

Theorem. ([53] theorem 3.1) There exist 2ℵ0 normal subgroups of finite index in Γ which

have level zero.

Theorem. ([53] theorem 3.2) Let q C be such that Nq > c0, some constant c0. Then Γ

contains 2ℵ0 non-normal standard subgroups of level q.

14

1.2.3 The Congruence Subgroup Problem

Let R be a commutative ring with a one.

Congruence Subgroup Problem. Does every subgroup of SLn(R) of finite index contain

a principal congruence subgroup?

This question can, with some care, be made to make sense in PSLn(R), the details can

be found in section (6.4). The Congruence Subgroup Problem has a long history and has its

origins in the work of Fricke and Klein, and Pick in the 19th century (see [34] section 4.3 and

[11] chapter I.6 section B). Klein [39] pointed out, at a meeting of the Munich Academy on

6th December 1879, that the Modular Group, PSL2(Z), contains subgroups of finite index

which do not contain a principal congruence subgroup. This was later proven simultaneously

and independently by Fricke [27] and Pick [74]. We outline the proof, which can be found

in [51]. Chandler and Magnus [11] assert, without any evidence, that it may be pre-1914.

Lemma.PSL2(Z)

PSL2(Z, nZ)∼= PSL2(Zn)

Where Zn denotes the ring of integers mod n.

Lemma. The only non-abelian quotient groups that can appear in a composition series of

PSL2(Zn) are the groups PSL2(Zp), where p is prime.

Lemma. A11 is a quotient of PSL2(Z), and A11 is not isomorphic to any of the groups

PSL2(Zp).

The kernel of the map PSL2(Z) A11 is then a normal non-congruence subgroup of

PSL2(Z). It is then a simple matter to see that SL2(Z) must also contain non-congruence

subgroups. The positive solution of the problem in the n > 3 case was proved simultaneously

in 1965 by Mennicke [66] and Bass-Lazard-Serre [4] in the context of SL2(Z):

Theorem. Let n > 3. If H 6 SLn(Z) is of finite index then SL2(Z, nZ) 6 H for some

n 6= 0.

15

Again we see that in the two dimensional case the normal subgroup structure of SLn(Z)

is much more complicated than the higher dimensional cases. In fact the situation is a lot

more complicated than may at first appear because “most” subgroups of the Modular group

are non-congruence subgroups. We now outline two different ways in which this idea is made

precise.

The Modular group has the following presentation PSL2(Z) =< a, t; a2, (at)3 >∼= C2 ∗C3

[25, 51, 70, 81]. Newman [71] has derived an asymptotic formula for the number of subgroups

of a given index in the free product of finitely many cyclic groups. Applied to the Modular

group we get

Theorem. [71] Let an(G) denote the number of subgroups of G of index n. Then

an(PSL2(Z)) ∼ (12πe1/2)−1/2 exp

(n log n

6− n

6+ n

12 + n

13 +

log n

2

)Now let γn(PSL2(Z)) denote the number of congruence subgroups of PSL2(Z) of index

at most n. A special case of a theorem of Lubotzky [47] is

Theorem. There exists positive constants C1, C2 such that

nC1 log n/ log log n 6 γn(PSL2(Z)) 6 nC2 log n/ log log n.

So it can be seen that

γn(PSL2(Z))

an(PSL2(Z))−→ 0 as n −→∞

and in this sense most subgroups of the modular group are non-congruence subgroups. Again

it is now simple to see that in SL2(Z) most subgroups are non-congruence. We now outline

another interpretation. Let

F = S 6 SL2(Z) : |SL2(Z) : S| <∞

and

C = C 6 SL2(Z) : C a congruence subgroup

16

These constitute bases for neighbourhoods of the identity for two topologies on SL2(Q). Let

SL2(Q) be the completion relative to F and let SL2(Q) be the completion relative to C.

Since C ⊆ F we have a natural surjection

Π : SL2(Q) −→ SL2(Q)

We denote the kernel of this map by C(SL2,Z) and call it the non-congruence kernel. The

Congruence Subgroup Problem now becomes the following

Question. Is C(SL2,Z) trivial?

Theorem. [45]

C(SL2,Z) = Fω

where Fω is the free profinite group of countable rank.

That is the congruence kernel is enormous, and so again most subgroups of SL2(Z) are

non-congruence. It is simple to see that the same is true in the modular group PSL2(Z).

So far we have looked at the congruence subgroup problem in SLn(Z). What about more

general rings?

Let K be a global field (see section on number theory at the end of this chapter), S∞ ⊆

S ⊆ Ω, S 6= ∅ and O = OS the ring of S-integers. Form the group SLn(O). Clearly the

congruence subgroup problem makes sense in this group and we can construct the congruence

kernel C(SLn,O) in exactly the same way as C(SL2,Z). The congruence subgroup problem

then becomes to determine the structure of C(SLn,O). Bass-Milnor-Serre [5] proved the

following

Theorem. With the above notation suppose that O is a Dedekind domain of arithmetic type,

and n > 3. Then C(SLn,O) = 1 unless K is a number field which is totally complex and

O is the ring of integers, in which case C(SLn,O) ∼= µ(K), the (finite cyclic) group of all

roots of unity in K.

The n = 2 case was dealt with by Serre [80].

17

Theorem. With the above notation and n = 2 then

1. If |S| > 2 and it is not the case that K is a totally complex number field and O its ring

of integers then C(SL2,O) = 1.

2. If |S| > 2 and K is a totally complex number field and O is its ring of integers then

C(SL2,O) ∼= µ(K).

3. If |S| = 1 then C(SL2,O) is infinite.

Thus the Congruence Subgroup Problem fails completely only in the case where n = 2

and |S| = 1. So again we see that the two dimensional case is more complicated than the

higher dimensional cases, and in the two dimensional case the unit structure of the ring

becomes important. There are three families of rings O for which |S| = 1: Z , Od, and C (see

section on number theory). We have already seen Lubotzky’s characterization of C(SL2,Z).

He also proved

Theorem. [45] With the above notation let O = Od, or C. Then C(SL2,O) contains Fω,

the free profinite group of countable rank, as a closed subgroup.

So the Modular group and the Bianchi groups contain a great many non-congruence

subgroups. We have already mentioned that Fricke [27] and Pick [74] gave examples of non-

congruence subgroups of the modular group (see also [40]). Reiner [77] generalized their

construction and many authors have produced classes of non-congruence subgroups (see [37]

and [70]). Stothers has shown [83] that the minimal index of a non-congruence subgroup in

the Modular group is 7. McQuillan has classified the normal congruence subgroups of the

Modular group [65]. Drillick [22] has adapted the approach in Magnus [51] outlined earlier to

the Bianchi group PSL2(O1), known as the Picard Group. Britto [10] generalized Drillicks

arguments to construct an infinite family of non-congruence subgroups in PSL2(Od) for

d = 1, 2, 3, 7, 11, 5, 6, 15. Since the basis of this construction is a surjection of PSL2(Od)

onto An, n > 7, all of the normal non-congruence subgroups constructed by these methods

are of index 6k, for some k ∈ N. It is also the case that these normal non-congruence

subgroups are torsion free.

18

In chapter five we extend the work of Mason in [58] and derive a relationship between

the order and level of a normal congruence subgroup of PSL2(Od,m). We then use this

relationship in chapter six to show that nearly every normal subgroup in PSL2(Od,m) of

index not divisible by 6 is a non-congruence subgroup. Further they all have torsion. Thus

our normal non-congruence subgroups are all different from those constructed by Drillick

and Britto.

1.3 Some Number Theory

For more information see [34] section 2.2E and the references therein.

Let K be any field. Let

v : K −→ R

Consider the following four conditions:

1. v(a) > 0, and v(a) = 0⇔ a = 0.

2. v(ab) = v(a)v(b).

3. v(a+ b) 6 v(a) + v(b).

4. v(a+ b) 6 max(v(a), v(b)).

Clearly 1 and 4⇒ 3. If v satisfies 1, 2, and 3 we say v is a valuation on K. If v also satisfies

4, v is non-Archimedean, and if not it is Archimedean. If v(a) = 1 ∀a ∈ K − 0 then v is

the trivial valuation. We say that a field is global if it is a finite separable extension of Q or

of the quotient field of a polynomial ring Fd [X], where Fd is a finite field of order d. In the

first case K is a number field, in the second a function field. We say that two valuations v1,

and v2 are equivalent if for every a ∈ K we have v1(a) < 1⇔ v2(a) < 1. Let K be a global

field. Let Ω be a complete set of inequivalent non-trivial valuations on K and let S∞ ⊆ Ω

be the set of Archimedean valuations. Suppose S∞ ⊆ S ⊆ Ω and S 6= ∅ then

OS = x ∈ K : v(x) 6 1 ∀v /∈ S

19

is called the ring of S-integers of K. OS is a Dedekind domain. If S is finite then we say

OS is a Dedekind domain of arithmetic type. If K is a number field and S = S∞ then OS is

the ring of integers of K.

The completion of K with respect to an Archimedean valuation, v, is isomorphic (as a

topological field) to R, or C and we say v is real or complex accordingly. The number field

K is totally real if all valuations in S∞ are real, and totally complex if they are all complex.

Theorem. Let O = OS be a Dedekind domain of arithmetic type. Suppose |S| = 1 then one

of the following is the case

1. O = Z.

2. O = Od the ring of integers in Q(√−d), d a positive square free integer.

3. O = C, the coordinate ring of an affine curve obtained by removing a point from a

projective curve defined over a finite field.

farther these are precisely the Dedekind domains of arithmetic type with finitely many units.

Example 1.3.1. Let d ∈ N be square free. Let O be the ring of integers in Q(√d). Then

O is a Dedekind domain of arithmetic type, and by [13] theorem 11.4, O∗ is infinite.

20

Chapter 2

Zimmert’s Theorem

In this chapter we describe a topological method invented by Zimmert [93] and its extension

due to Grunewald and Schwermer [32]. It concerns the action of SL2(R) on hyperbolic 3-

space H3, where R is an order in an imaginary quadratic number field. See also [23] chapter

7 section 5 for a discussion of this method.

Recall that the Bianchi groups are PSL2(Od) where Od is the ring of integers in the

imaginary quadratic number field Q(√−d), and recall that

Od = Z + ωZ where ω =

1+i√

d2

if d ≡ 3 (mod 4)

i√d else

and where d is a positive square-free integer. Now let m ∈ N and let

Od,m = Z +mωZ

where ω is as above. The Od,m are the orders in the imaginary quadratic number field

Q(√−d). It is clear that Od = Od,1 and that |Od : Od,m| = m. We can form the groups

SL2(Od,m). Since mOd ⊆ Od,m we have SL2(Od,mOd) 6 SL2(Od,m) so that SL2(Od,m) is

of finite index in SL2(Od). Similar comments can be made to show that PSL2(Od,m) is of

finite index in PSL2(Od). The aim of Zimmert’s method is to prove the following

Theorem. For every d there exists m such that SL2(Od,m) has a free non-abelian quotient.

As PSL2(Od,m) = SL2(Od,m)/ < −I > the following theorem is clear

21

Theorem. For every d there exists an m such that PSL2(Od,m) has a free non-abelian

quotient.

We say that a group G is SQ-Universal if every countable group can be embedded in a

quotient of G. It follows from a result of P. M. Neumann [69] and the above theorem that

Theorem. [32] Every Bianchi group PSL2(Od) is SQ-universal.

Thus the Bianchi groups may be considered “large” [76]. It follows from SQ-Universality

that every Bianchi group has 2ℵ0 normal subgroups [69] and it is this that is the source of

the extremely complicated normal subgroup structure described in the previous chapter.

2.1 Hyperbolic 3-space

The upper half-space in Euclidean three-space gives a convenient model of 3-dimensional

hyperbolic space

H3 = (z, r) ∈ C× R : r > 0

which we equip with the hyperbolic metric

ds2 =dx2 + dy2 + dr2

r2

The group SL2(C) acts on H3 in the following way. Let M ∈ SL2(C) where

M =

α β

γ δ

Then

M(z, r) =

((δ − γz)(αz − β)− r2γα

τ,r

τ

)where

τ = |γz − δ|2 + r2|γ|2

Under this action the hyperbolic metric is SL2(C)-invariant. The group SL2(C) is generated

by the matrices 1 a

0 1

,

0 −1

1 0

22

where a ∈ C. These generators operate on H3 in the following way 1 a

0 1

(z, r) = (z + a, r) ,

0 −1

1 0

(z, r) =

(−z

|z|2 + r2,

r

|z|2 + r2

).

2.2 Zimmert’s Method

Let D be the discriminant of Q(√−d). It is well known that [52]

D =

−d if d ≡ 3 (mod 4)

−4d else

Definition. The Zimmert Set Z(d,m) is the set of all n ∈ N such that

1. 4n2 6 m2|D| − 3.

2. D is a quadratic non-residue modulo all the odd prime divisors of n, and if D 6≡ 5

(mod 8) then n is odd.

3. n > 0, gcd(n,m) = 1 and n 6= 2.

It is easy to prove that Z(d,m) = ∅ ⇔ (d,m) = (1, 1), or (3, 1), and if Z(d,m) 6= ∅ then

1 ∈ Z(d,m). We let r(d,m) = |Z(d,m)|. Zimmert’s theorem is

Theorem. [32] SL2(Od,m) has a free quotient of rank r(d,m).

Zimmert [93] proved the m = 1 case. We now outline the proof. Let

Bd,m =(z, r) ∈ H3 : τ > 1 ∀ coprime γ, δ ∈ Od,m

where, γ, δ coprime means that γOd,m + δOd,m = Od,m, and, as before

τ = |γz − δ|2 + r2|γ|2

Every point of H3 is equivalent to a point of Bd,m under the action of SL2(Od,m) and so the

natural map

Bd,m −→ SL2(Od,m)\H3

23

is surjective. Let

D = (s1 + s2mω, r) ∈ Bd,m : s1, s2 ∈ [0, 1]

D is a fundamental domain for SL2(Od,m) [84, 93].

Proposition 2.2.1. ([84] Proposition 3.9) Every h ∈ H3 has a neighbourhood U such that

σU ∩D 6= ∅ for only finitely many σ ∈ SL2(Od,m).

Let n ∈ Z(d,m), t ∈ Z such that (n, t) = 1. Let

Fn,t =

(z, r) ∈ Bd,m :

∣∣∣∣Im(z − tmω

n

)∣∣∣∣ 6 1

m4|D|2

By condition 1 in the definition of the Zimmert set we have Fn1,t2 ∩ Fn2,t2 6= ∅ ⇔ n1 = n2,

and t1 = t2.

Lemma 2.2.2. ([93] Hilfssatz 1) Let (z, r) ∈ Fn,t, let σ ∈ SL2(Od,m) such that σ(z, r) =

(z′, r′) ∈ Bd,m. Then ∃t′ ∈ Z such that gcd(n, t′) = 1 and

Im

(z − tmω

n

)= Im

(z′ − t′mω

n

).

Further

r >5

2m2|D|

Now for each n ∈ Z(d,m) define ϕn : Bd,m → S1, where S1 = z ∈ C : |z| = 1, by

ϕn(z, r) =

1 if (z, r) /∈⋃

(n,t)=1 Fn,t

exp2πi(

12

+ m4|D|22

Im(z − tmω

n

))if (z, r) ∈ Fn,t

There is a unique factorization of ϕn over SL2(Od,m)\H3 by a continuous map

fn : SL2(Od,m)\H3 −→ S1

This is well defined by lemma (2.2.2) and continuous by proposition (2.2.1). Suppose that

Z(d,m) = n1, . . . , nr. Let Y denote the one point union of r(d,m) copies of S1 with base

point 1 ie

Y =(z1, . . . , zr) ∈ S1 × · · · × S1 : zi 6= 1 for at most one i

24

Now define

f : SL2(Od,m)\H3 −→ Y

by

(z, r) 7−→ (f1(z, r), . . . , fr(z, r))

where fi denotes fni. Now f induces a homomorphism

f∗ : π1

(SL2(Od,m)\H3

)−→ π1 (Y, 1)

Now let g ∈ SL2(Od,m) and let h0 ∈ H3. Let P be any path from h0 to gh0. The image of P

in SL2(Od,m)\H3 is a loop and therefore represents some αg ∈ π1 (SL2(Od,m)\H3). Define

θ : SL2(Od,m) −→ π1

(SL2(Od,m)\H3

)by g 7→ αg. This map is well defined. We now have

σ : SL2(Od,m) −→ π1

(SL2(Od,m)\H3

)−→ π1(Y ) ∼= Fr

where r = r(d,m). Now consider the natural map

ψ : Fr Zr

Zimmert shows ([93] Satz 2(i)) that σψ is surjective. So 1 6= imσ 6 Fr, also SL2(Od,m) is

finitely generated, so imσ is a free group of rank s, where r 6 s < ∞ and so maps onto a

free group of rank r. Thus

Theorem 2.2.3. SL2(Od,m) maps onto a free group of rank r(d,m).

so clearly

Theorem 2.2.4. PSL2(Od,m) maps onto a free group of rank r(d,m).

2.3 Explicit result about the free generators

We have just seen that

σ : SL2(Od,m) π1(Y ) ∼= Fr

where r is the order of the Zimmert Set.

25

Lemma 2.3.1. The image, under σ, of the matrix

U =

1 mω

0 1

can be taken as a free generator of π1(Y ).

Proof. Let z0 = −im4|D|2 , z1 = z0 + mω, and let hi = (zi, 1). So uh0 = h1. Now define a

path P from h0 to h1 by P = (z(s), 1) : s ∈ [0, 1], where z(s) = (1−s)z0 +sz1 = z0 +smω.

Now

P ∩

(⋃t∈Z

F1,t

)= (P ∩ F1,0) ∪ h1

and h1 ∈ F1,1. So f1(P ) = S1. Now recall that

Y =(z1, . . . , zr) ∈ S1 × · · · × S1 : zi 6= 1 for at most one i

Let S1

i denote the ith circle of Y . Then S1i gives rise to gi ∈ π1(Y ) and g1, . . . , gr is a set

of free generators of π1(Y ) ∼= Fr. So

f∗(σ(u)) = g1g0 = x

where g0 ∈< g2, . . . , gr >. By Tietze transformations we can take x, g2, . . . , gr to be a set

of free generators of π1(Y ), as required.

Lemma 2.3.2. σ(u) is a free generator of imσ.

Proof. Clearly σ(u) ∈ imσ. We can then apply Proposition 2.10 on page 8 of [48] to get the

result.

2.4 Unipotent matrices

Let R be any commutative ring with a one. Recall that a matrix M ∈ GL2(R) is unipotent if

(M − I)2 = 0. Let U2(R) denote the normal subgroup of SL2(R) generated by the unipotent

matrices. Clearly E2(R) 6 U2(R) and so NE2(R), the normal subgroup generated by the

elementary matrices is contained in U2(R).

26

Lemma 2.4.1. Let M ∈ GL2(R). Then

M is unipotent ⇔ detM = 1 and trM = 2

⇔M =

1 + α γ

β 1− α

where α2 + βγ = 0, α, β, γ ∈ R.

Proof. Let

I 6= M =

a b

c d

∈ GL2(R)

be unipotent. Let t =trM = a+ d. Then

0 = (M − I)2 =

(a− 1)2 + bc b(t− 2)

c(t− 2) (d− 1)2 + bc

Now suppose that t 6= 2, so b = c = 0, so (a − 1)2 + bc = (a − 1)2 = 0, so a = 1, similarly

d = 1, so M = I. Contradiction. So t = 2, and d = 2 − a, so detM = a(2 − a) − bc =

−(a2 − 2a+ 1) + 1− bc = −((a− 1)2 + bc) + 1 = 1.

Conversely suppose that detM = 1 and trM = 2. So

M =

a b

c 2− a

so

(M − I)2 =

(a− 1)2 + bc 0

0 (1− a)2 + bc

and (a− 1)2 + bc = a2− 2a+1− bc = −(a(2−a)− bc)+1 = −1+1 = 0. So M is unipotent.

Now suppose that detM = 1 and tr M = 2. It is clear that M can be written in the

form 1 + α γ

β 1− α

where α2 + βγ = 0, some α, β, γ ∈ R. So conversely suppose that M is of this form. Then

(M − I)2 =

α2 + βγ αγ − αγ

αβ − αβ α2 + βγ

= 0

so M is unipotent.

27

Lemma 2.4.2. Let δ ∈ R. Then every conjugate of 1 δ

0 1

in SL2(R) is unipotent.

Proof. a b

c d

1 δ

0 1

d −b

−c a

=

1− acδ a2δ

−c2δ 1 + acδ

Lemma 2.4.3. Suppose that R is an integral domain and let α, β ∈ R then αR + βR is

principal if and only if ∃α′, β′ ∈ R such that α′, and β′ are coprime and αα′ + ββ′ = 0.

Proof. Suppose first that αR+βR = δR. So α = δr1, β = δr2, δ = αr3 +βr4, where ri ∈ R,

i = 1, 2, 3, 4. So δ = δ(r1r3 + r2r4), and because R is an integral domain and δ 6= 0, we have

r1r3 + r2r4 = 1. Let α′ = r2, and β′ = −r1. So α′, and β′ are coprime, and αα′ + ββ′ = 0.

Conversely suppose that ∃ coprime α′, β′ ∈ R such that αα′+ββ′ = 0. Now αα′ = −ββ′,

and α′ and β′ are coprime, so α = β′β′′, and β = α′α′′. So 0 = αα′ + ββ′ = α′β′(α′′ + β′′),

and as R is an integral domain and α′ 6= 0 6= β′ we have α′′ = −β′′. So αR + βR =

β′β′′R + α′α′′R = α′′(α′R + β′R) = α′′R.

Lemma 2.4.4. Suppose that R is an integral domain and let

M =

1 + α γ

β 1− α

be unipotent. Then M is conjugate in SL2(R) to 1 δ

0 1

if and only if αR + βR is principal.

28

Proof. Suppose that 1 + acδ a2δ

−c2δ 1− acδ

=

1 + α γ

β 1− α

.

So α = acδ, β = −c2δ, γ = a2δ. Clearly αR + βR ⊆ cδR. Now let x ∈ R and consider cδx.

cδx = cδx(ad− bc) as ad− bc = 1

= acδxd− c2δbx

= αxd+ βbx as α = acδ, β = −c2δ

∈ αR + βR

So αR + βR = δcR, ie is a principal ideal.

Conversely suppose that αR + βR is principal. Now a b

c d

1 + α γ

β 1− α

d −b

−c a

=

1 + 1β

(aα+ bβ) (cα + dβ) −1β

(aα+ bβ)2

(cα + dβ)2 −1β

(aα+ bβ) (cα + dβ)

= Y

and αR + βR is principal so, by (2.4.3), ∃α′, β′ ∈ R such that α′, and β′ are coprime and

α′α+ β′β = 0. Let c = α′, and d = β′ so ∃a, b such that ad− bc = 1 and cα + dβ = 0, so

Y =

1 δ

0 1

where δ = −1

β(aα+ bβ)2.

Having set out the basics that we require in the context of an arbitrary commutative

ring we now turn to the case where R = Od,m is an order in an imaginary quadratic number

field. Let r be the order of the Zimmert Set and

σ : SL2(Od,m) −→ Fr

be the map given in the previous section. Let K = kerσ.

29

Lemma 2.4.5. ([93] Hilfssatz 2) Let O = Od,m and let M ∈ SL2(O), and α, β ∈ O such

that αO + βO is not a principal ideal. Let ε, t ∈ R such that 0 < ε 6 1, and |t| 6 ε. Let

h′ = (z, r) =

(−α+ t

β,

1

|β|√

2ε− ε2)

and suppose that Mh′ = (z′, r′). Then

r′ 6 |β|√

2ε− ε2

Proof. Suppose that

M =

a b

c d

so r′ = r/τ , where τ = |cz − d|2 + r2|c|2. Now

τ =

∣∣∣∣c−α+ t

β− d∣∣∣∣2 +

(1

|β|√

2ε− ε2)2

|c|2 >1

|β|2

For equivalently

|(c− α− βd) + ct|2 +(2ε− ε2

)|c|2 > 1

Now c and d are coprime and βO+αO is not a principal ideal so, by (2.4.3), 0 6= −cα−βd ∈

O, so | − cα− βd| > 1. Thus

|(−cα− βd) + ct|2 +(2ε− ε2

)|c|2 > (1− |ct|)2 +

(2ε− ε2

)|c|2

Suppose that |ct| > 1. Clearly (1 − |ct|)2 > 0. Also |c| > 1/|t| > 1/ε and so (2ε −

ε2)|c|2 =(

2ε− 1)|c|2ε2 > 2

ε− 1 > 1. Now suppose that |ct| 6 1. First of all observe that

2ε − ε2 > 2|t| − |t|2 ⇔ (ε − |t|)(2 − ε − |t|) > 0, and as 0 < |t| 6 ε 6 1, this is true, so

2ε− ε2 > 2|t| − |t|2. Now

(1− |ct|)2 + (2ε− ε2)|c|2 = 1− 2|ct|+ |ct|2 + (2ε− ε2)|c|2 > 1

⇔ |ct|2 − 2|ct|+ (2ε− ε2)|c|2 > 0

30

Now if |c| = 0 then this is certainly true, so suppose that |c| 6= 0, so as c ∈ O we have |c| > 1.

Now

|ct|2 − 2|ct|+ (2ε− ε2)|c|2

=|c|(|c||t|2 − 2|t|+ (2ε− ε2)|c|

)>|c|

(|c||t|2 − 2|t|+ (2|t| − |t|2)|c|

)=2|c||t|(|c| − 1) > 0.

So r′ = r/τ 6 |β|2(1/|β|)√

2ε− ε2 = |β|√

2ε− ε2.

Theorem 2.4.6. ([93] Satz 2(ii)) Let α, β, γ ∈ Od,m such that α2 + βγ = 0 and

αOd,m + βOd,m is not principal. Then

M =

1 + α γ

β 1 + α

∈ KProof. Let ε ∈ R and let

h1 = (z1, r1) =

(−α+ ε

β,

1

|β|√

2ε− ε2)

and suppose that Mh1 = h2 = (z2, r2). Now

τ =

∣∣∣∣β−α+ ε

β− (1− α)

∣∣∣∣2 +1

|β|2(2ε− ε2

)|β|2

= | − α+ ε− 1 + α|2 + 2ε− ε2

= (ε− 1)2 + 2ε− ε2

= 1

So r2 = r1, and

z2 =

(1− α− β−α+ ε

β

)((1 + α)

−α+ ε

β+α2

β

)− 1

|β|(2ε− ε2

)β(1 + α)

After some algebraic manipulation we see that

z2 =−α− εβ

31

so

h2 =

(−α− εβ

,1

|β|√

2ε− ε2)

Now let

P =

h′ =

(−α+ t

β,

1

|β|√

2ε− ε2)∈ H3 : t ∈ R,−ε 6 t 6 ε

so P is a path in H3 from h1 to h2. Let h′ ∈ P and M2 ∈ SL2(O). Then, by (2.4.5), if

M2h′ = (z′, r′) we have r′ 6 |β|

√2ε− ε2. Now choose ε so that |β|

√2ε− ε2 < 5/(2m2|D|).

So, by (2.2.2), M2h′ /∈ Fn,t for every n ∈ Z(d,m) and t ∈ Z coprime to n. Thus ϕn(h′) = 1

and it follows that M is in the kernel of the Zimmert map.

Theorem 2.4.7. With the above notation, there exists an epimorphism

τ :SL2(Od,m)

U2(Od,m)−→ Fs

where s = r − 1.

Proof. We have already seen, in the previous section that SL2(Od,m) has a free quotient of

rank s+ 1 where

U =

1 mω

0 1

can be taken as a free generator. Now consider a unipotent matrix

M =

1 + α γ

β 1− α

where α2 + βγ = 0. Now either αR + βR is principal or it isn’t. If it is not principal then,

by theorem (2.4.6), M ∈ K. If αR+βR is principal then, by lemma (2.4.4), M is conjugate,

in SL2(Od,m) to 1 δ

0 1

for some δ ∈ Od,m. Nowδ = z1 +mωz2, zi ∈ Z, but T ∈ SL2(Z) 6 K, so we can suppose that

M is conjugate to some power of U . So either M is in the kernel of the map SL2(Od,m) Fr

or it is conjugate to a power of U , which can be taken as a free generator of Fr. So clearly

SL2(Od,m)

U2(Od,m)

< u, x2, . . . , xr;>

< u >∼=< x2, . . . xr;>= Fs

32

Corollary 2.4.8.SL2(Od,m)

NE2(Od,m)

has a free quotient of rank s = r − 1.

Let R be any commutative ring and let q R. Then let

SLn(R, q) = ker(SLn(R)→ SLn(R/q))

and let Un(R, q) be the normal subgroup of SLn(R) generated by all unipotent matrices in

SLn(R, q), let En(R, q) denote the normal subgroup of En(R) generated by all q elementary

matrices and NEn(R, q) denote the normal subgroup of SLn(R) generated by q elementary

matrices.

Corollary 2.4.9. Let 0 6= q Od,m. Then

SL2(Od,m, q)

U2(Od,m, q)

has a free quotient of rank s = r − 1.

Proof. Now

SL2(Od,m, q)U2(Od,m) > SL2(Od,m, q)E2(Od,m) = SL2(Od,m)

soSL2(Od,m, q)

U2(Od,m) ∩ SL2(Od,m, q)∼=SL2(Od,m)

U2(Od,m) Fs

thusSL2(Od,m, q)

U2(Od,m, q) Fs

33

2.5 Computation of Zimmert Sets

Given the definition of the Zimmert Set Z(d,m) it is a simple matter to write a computer

program to calculate any Zimmert Set. We have written such a program in GAP [24]. Since

we are interested in free non-abelian quotients it would be useful to know values of (d,m)

such that r(d,m) 6 1. Consider first the Zimmert sets Z(d, 1) which we denote Z(d), let

r(d) = r(d, 1). Mason, Odoni and Stothers [61] have proved

Theorem. For all but finitely many d r(d) > 2.

Obviously we would like a list of the d such that r(d) = 1. By means of a computer

search Mason et al [61] establish that up to 2× 1013 the only values of d for which r(d) = 1

are: 2, 5, 6, 7, 11, 14, 15, 17, 19, 21, 23, 26, 29, 30, 31, 35, 39, 41, 47, 51, 59, 66, 69, 71,

87, 89, 95, 101, 105, 110, 111, 119, 129, 159, 161, 191, 194, 209, 215, 230, 231, 255, 285,

311, 321, 335, 341, 374, 399, 426, 455, 479, 546, 591, 615, 671, 831, 959, 1095, 1119, 1239,

2415 and they conjecture that these are the only values of d for which r(d) = 1. Recall that

r(1) = r(3) = 0.

We extended this work by using our computer program to calculate pairs (d,m) such

that r(d,m) = 1. For every square free d ∈ N and every m ∈ N we checked every pair (d,m)

such that m2|D| 6 107 and found 104 pairs (d,m) such that r(d,m) = 1. Excluding those

mentioned above these are : (1, 2), (1, 3), (1, 6), (2, 2), (2, 3), (3, 2), (3, 3), (3, 4), (3, 10),

(5, 2), (5, 3), (5, 4), (6, 2), (6, 3), (6, 4), (6, 5), (7, 2), (7, 3), (10, 3), (11, 2), (11, 4), (14, 2),

(15, 2), (15, 3), (21, 2), (23, 2), (26, 2), (31, 3), (35, 2), (35, 4), (35, 6), (39, 2), (41, 2), (55, 3),

(59, 2), (66, 2), (111, 2), (119, 2), (131, 2), (195, 2), (231, 2), (341, 2). Notice that, perhaps

surprisingly, r(d, 1) > 1 does not imply r(d,m) > 1, for m > 1; for example r(10, 1) = 2, yet

r(10, 3) = 1. It seems very reasonable to make the following

Conjecture. The only pairs (d,m) such that r(d,m) = 1 are those listed above.

It can be seen from the data above that for most d such that r(d) = 1, r(d, 2) > 1, so that

Zimmert’s theorem only just fails to give a free non-abelian quotient. However Zimmert’s

34

theorem is not best possible; for r(5) = 1 yet PSL2(O5) has a free quotient of rank 2. With

this in mind make the following

Definition. Let ρ(d,m) be the largest rank of a free quotient of SL2(Od,m).

ρ(d,m) is well defined as PSL2(Od,m) is finitely generated. Let ρ(d) = ρ(d, 1). Clearly

ρ(d,m) > r(d,m). If we have a presentation for PSL2(Od,m) we can compute ρ(d,m):

Theorem. ([61] theorem 6, and [78])

ρ(d) 0 1 2 3 5 7 8

d 1, 3 2, 7, 11, 19 5, 6, 15, 43 10, 13, 67 22 21, 163 37

Mason et al [61] have the following

Conjecture. ρ(d) > 1 for all d > 19.

We have no reason to doubt this. The only square free d 6 19 missing from the above

are 14, and 17, we partially close this gap

Proposition 2.5.1. 4 6 ρ(14) 6 5

Proof. The following presentation for PSL2(O14) can be found in [30] Proposition 3.5.

< g1, g2, g3, g4, g5, g6, g7 ; g21, (g1g2)

3, g1g3g−11 g−1

3 , g2g4g−12 g−1

4 ,

g5g−16 g7g

−13 g6g

−15 g3g

−17 , g6g4g1g

−13 g6g1g

−16 g3g

−14 g1g

−16 g1,

g1g6g1g2g−16 g1g

−16 g3g1g

−12 g−1

3 g6,

g−12 g−1

7 g6g−12 g1g

−16 g1g5g2g

−16 g7g2g1g

−13 g6g1g

−15 g3 >

From this it is easy to compute that PSL2(O14)ab ∼= Z6×Z5, and so ρ(14) 6 5. It is also easy

to see that by setting g1 = g2 = g5g−16 = 1 we get the free group of rank 4. So ρ(14) > 4.

Proposition 2.5.2. ρ(1, 2) = ρ(3, 2) = 1.

35

Proof. We have the following presentations from (3.5.1) and (3.3.1)

PSL2(O1,2) =< a, t, z, w; a2, z2, (at)3, (atz)2, [t, w] , (atw−1zw)2 >

PSL2(O3,2) =< a, t, w; a2, (at)3, (w−1awa)3, [t, w] >

from which it is easy to see PSL2(O1,2)ab = Z2 × Z, and PSL2(O3,2)

ab = Z6 × Z. So

ρ(1, 2) = ρ(3, 2) = 1.

Proposition 2.5.3. ρ(7, 2) = 2.

Proof. We have from section (3.5)

PSL2(O7,2) =< a, t, w, x, y ; a2, x2, [t, w] , (ax)3, (at)3, y = txyxt−1, (ytay−1x)2,

xyat−1y−1wt−1aw−1t >

Now suppose that PSL2(O7,2)/N is free then N contains the normal closure of all elements in

PSL2(O7,2) of finite order, soN(a, x, t) 6 N . It is easy to calculate that PSL2(O7,2)/N(a, x, t) ∼=

F2. So ρ(7, 2) = 2.

Proposition 2.5.4. 3 6 ρ(11, 2) 6 4.

Proof. We have from section (3.5)

PSL2(O11,2) =< a, t, w, k, l,m, n ; a2, k2, [m,n] , [t, w] , (at)3, tklkt−1l−1,

km−1ltat−1l−1m,nlat−1l−1n−1mwaw−1tm−1,

m−1tklt−1l−1mwt−1aw−1t >

so the kernel of a free quotient of PSL2(O11,2) must contain a, k, t. It is easy to compute

that PSL2(O11,2)/N(a, k, t) =< z, l,m, n; [m,n] >. This maps onto the free group of rank

3, so ρ(11, 2) > 3, and it abelianization is Z4 so ρ(11, 2) 6 4.

Given the evidence presented above it seems reasonable to make the following

Conjecture. The only values of (d,m) for which PSL2(Od,m) does not have a free non-

abelian quotient are: (1, 1), (1, 2), (2, 1), (3, 1), (3, 2), (7, 1), (11, 1), (19, 1).

36

There are results later on in this thesis which depend on the fact that PSL2(Od,m) has

a free non-abelian quotient. Usually we then attempt to deal with the cases mentioned in

the conjecture. We are quite certain that the conjecture is true and so the idea is that

we have a proved a theorem (or a version of it) for all Bianchi groups. For example it is

not yet proven that PSL2(O2,2) has a free non-abelian quotient, so strictly speaking our

theorem which depends on PSL2(Od,m) having a free non-abelian quotient has not been

proved for PSL2(O2,2), although we are quite certain that it is true and all that is required

is a presentation of PSL2(O2,2) to verify it. Contrast this with PSL2(O2) where a slightly

different technique may be required or with PSL2(O3) where a radically different technique is

perhaps needed. For this reason we do not consider PSL2(O2,2) to be a “true” exception and

we refer to the pairs mentioned in the conjecture as the “true exceptions”. The conjecture

could (and probably will) be proved by means of a computer search. See the comments at

the end of section (3.5).

Recall theorem (2.4.7). To get a free non-abelian quotient we require that r(d,m) > 3.

In exactly the same way as above we have used our GAP [24] program to get a list of pairs

(d,m) such that r(d,m) = 2. Again for every square free d ∈ N and every m ∈ N we

checked every pair (d,m) such that m2|D| 6 107 and found 215 pairs (d,m) such that

r(d,m) = 2. These are: (1, 4), (1, 5), (1, 7), (1, 9), (2, 4), (2, 5), (2, 7), (3, 5), (3, 6), (3, 8),

(3, 12), (5, 5), (6, 6), (7, 5), (7, 6), (7, 9), (10, 1), (10, 2), (10, 6), (10, 9), (11, 3), (11, 6), (13, 1),

(13, 3), (13, 6), (14, 3), (14, 4), (15, 4), (15, 5), (17, 2), (17, 3), (17, 4), (17, 5), (19, 2), (19, 3),

(19, 4), (19, 6), (19, 12), (21, 3), (21, 4), (21, 5), (21, 6), (22, 1), (26, 3), (29, 2), (29, 3), (30, 2),

(30, 3), (30, 7), (31, 2), (33, 1), (33, 2), (33, 5), (34, 1), (38, 1), (38, 5), (39, 3), (39, 4), (39, 5),

(41, 1), (43, 1), (46, 1), (47, 2), (47, 3), (51, 2), (51, 4), (51, 6), (51, 8), (55, 1), (55, 2), (55, 6),

(61, 1), (62, 1), (62, 2), (65, 1), (65, 2), (66, 3), (69, 2), (69, 3), (70, 1), (71, 2), (74, 1), (77, 1),

(77, 2), (79, 1), (79, 3), (83, 1), (83, 2), (86, 1), (87, 2), (87, 3), (87, 4), (94, 3), (95, 2), (95, 3),

(101, 2), (105, 2), (110, 2), (114, 1), (129, 2), (131, 1), (134, 1), (138, 1), (143, 1), (143, 2),

(145, 3), (146, 1), (149, 1), (151, 1), (155, 1), (159, 2), (165, 1), (167, 1), (167, 2), (173, 1),

(179, 1), (182, 1), (183, 1), (185, 1), (186, 1), (195, 1), (195, 4), (199, 1), (206, 1), (215, 2),

37

(222, 1), (230, 2), (231, 3), (237, 1), (239, 1), (251, 1), (251, 2), (255, 2), (266, 1), (269, 1),

(271, 1), (285, 2), (287, 1), (290, 1), (299, 2), (314, 1), (327, 1), (329, 1), (339, 2), (359, 1),

(383, 1), (390, 1), (395, 2), (399, 2), (431, 1), (446, 1), (447, 1), (455, 2), (458, 1), (471, 1),

(494, 1), (497, 1), (503, 1), (506, 1), (519, 1), (545, 1), (551, 1), (563, 2), (569, 1), (623, 1),

(654, 1), (659, 2), (689, 1), (699, 2), (705, 1), (719, 1), (755, 2), (759, 1), (770, 1), (789, 1),

(791, 1), (815, 1), (831, 2), (854, 1), (887, 1), (935, 1), (1031, 1), (1055, 1), (1151, 1), (1169, 1),

(1190, 1), (1199, 1), (1209, 1), (1223, 1), (1271, 1), (1326, 1), (1335, 1), (1407, 1), (1511, 1),

(1551, 1), (1599, 1), (1751, 1), (1767, 1), (1823, 1), (1895, 1), (1959, 1), (1991, 1), (2015, 1),

(2159, 1), (2435, 2), (2639, 1), (2679, 1), (2735, 1), (3119, 1), (3311, 1), (3471, 1), (4479, 1),

(6215, 1), (6815, 1), (8655, 1). Again it seems reasonable to make the following

Conjecture. The only pairs (d,m) such that r(d,m) = 2 are those listed above.

38

Chapter 3

The groups PSL2(Od,m)

In chapter two we saw that almost all of the groups PSL2(Od,m) had a free non-abelian

quotient. In this chapter we take a closer look at these groups. For completeness we restate

the definition.

Let m, d ∈ N, d square-free then the orders in the imaginary quadratic number field

Q(√−d) are

Od,m = Z +mωZ where ω =

1+i√

d2

if d ≡ 3 (mod 4)

i√d else

The maximal order Od,1 is the ring of integers in Q(√−d) and is denoted Od.

| Od : Od,m | = m. We are considering the groups PSL2(Od,m). Since mOd ⊆ Od,m, we

have PSL2(Od,mOd) 6 PSL2(Od,m), so PSL2(Od,m) is of finite index in PSL2(Od). We

now compute this index.

3.1 Computation of |PSL2(Od) : PSL2(Od,m)|

First observe that

|PSL2(Od) : PSL2(Od,m)| = |PSL2(Od) : PSL2(Od,p1)| . . . |PSL2(Od,p1...pr−1) : PSL2(Od,m)|

where m = p1 . . . pr and pi are not necessarily distinct primes. So for m, p ∈ N, p prime, we

compute |PSL2(Od,m) : PSL2(Od,mp)|. We need the following well known lemma.

39

Lemma 3.1.1. Let R be any commutative ring. Let q R such that R/q is an SR2-ring.

Then the natural map

SL2(R) −→ SL2(R/q)

is onto.

Proof. R/q is SR2 so SL2(R/q) = E2(R/q) and clearly 1 α

0 1

7−→ 1 + q α+ q

q 1 + q

.

So the map is onto.

Lemma 3.1.2. For any d,m, p ∈ N such that d is square free, and p is prime

|SL2(Od,m) : SL2(Od,mp)| = |PSL2(Od,m) : PSL2(Od,mp)|

and so

|SL2(Od) : SL2(Od,m)| = |PSL2(Od) : PSL2(Od,m)|

Proof. First observe that

|PSL2(Od,m) : PSL2(Od,mp)| =|PSL2(Od,m) : PSL2(Od,m, pOd,m)||PSL2(Od,mp) : PSL2(Od,m, pOd,m)|

Now

|PSL2(Od,m) : PSL2(Od,m, pOd,m)| =∣∣∣∣PSL2

(Od,m

pOd.m

)∣∣∣∣Now, for any ring R and any ideal q of R such that R/q is finite, we have |PSL2(R/q)| =

ρ|SL2(R/q)|, where ρ = 1 if 2 ∈ q and ρ = 1/2 if 2 /∈ q. So∣∣∣∣PSL2

(Od,m

pOd,m

)∣∣∣∣ = ρ

∣∣∣∣SL2

(Od,m

pOd,m

)∣∣∣∣where ρ = 1 if p = 2 and ρ = 1/2 if p 6= 2. Also, as |Od,mp : pOd,m| = p,

|PSL2(Od,mp) : PSL2(Od,m, pOd,m)| = |PSL2(Fp)| = ρ|SL2(Fp)|

40

where ρ is as above (this follows from [79] theorem 8.14). Thus

|PSL2(Od,m) : PSL2(Od,m, pOd,m)||PSL2(Od,mp) : PSL2(Od,m, pOd,m)|

=ρ|SL2(Od,m/pOd,m)|

ρ|SL2(Fp)|

=|SL2(Od,m/pOd,m)|

|SL2(Fp)|

=|SL2(Od,m) : SL2(Od,m, pOd,m)||SL2(Od,mp) : SL2(Od,m, pOd,m)|

= |SL2(Od,m) : SL2(Od,mp)|

In what follows we work with SL2 as it makes the proofs a little simpler. Observe that

|SL2(Od,m) : SL2(Od,mp)| =|SL2(Od,m) : SL2(Od,m, pOd,m)||SL2(Od,mp) : SL2(Od,m, pOd,m)|

Lemma 3.1.3. |SL2(Od,mp) : SL2(Od,m, pOd,m)| = |SL2(Fp)| = p(p2 − 1).

Proof. Observe that |Od,mp : pOd,m| = p so

SL2(Od,mp)

SL2(Od,m, pOd,m)∼= SL2(Fp).

This group has the required order by [79] theorem 8.8 and 8.14

Now

|SL2(Od,m) : SL2(Od,m, pOd,m)| =∣∣∣∣SL2

(Od,m

pOd,m

)∣∣∣∣So we need to have some understanding of the structure of Od,m/pOd,m.

Lemma 3.1.4. If p | m thenOd,m

pOd,mis local with maximal ideal m = (mω).

Proof. Let R = Od,m/pOd,m. We compute R∗. Now

R = r +mωs : r, s = 0, 1, . . . , p− 1

Let ui = ri +mωsi ∈ R, i = 1, 2. Then

u1u2 = r1r2 + (r1s2 + r2s1)mω +m2ω2s1s2

41

suppose first that d ≡ 1, 2 (mod 4), so ω2 = −d. Then

u1u2 = r1r2 − dm2s1s2 + (r1s2 + r2s1)mω

= r1r2 + (r1s2 + r2s1)mω

since p | m. Now suppose that d ≡ 3 (mod 4), so ω2 = ω − (d+ 1)/4. Then

u1u2 = r1r2 −m2d+ 1

4s1s2 + (r1s2 + r2s1 +ms1s2)mω

= r1r2 + (r1s2 + r2s1)mω

again as p | m. Now suppose that r1 6= 0 and let r2 = r−11 , s2 = −r−2

1 s1, so u1u2 = 1. So

r1 6= 0⇒ u−11 exists. Hence r +mωs : r 6= 0 ⊆ R∗. But if r1 = 0 then u1u2 = r2s1mω 6= 1.

So r1 = 0⇒ u1 is not a unit. Hence

R∗ = r +mωs : r 6= 0

We now show that R − R∗ = mωs : s = 0, 1, . . . , p− 1 = (mω). Clearly R − R∗ ⊆ (mω).

Now if d ≡ 1, 2 (mod 4) then

mω(r +mωs) = rmω +m2ω2s

= rmω − dm2s

= rmω ∈ R−R∗

similarly if d ≡ 3 (mod 4) then

mω(r +mωs) = rmω +m2ω2s

= rmω +m2s

(ω − d+ 1

4

)= (r +ms)mω −m2s

d+ 1

4

= rmω ∈ R−R∗

Hence R−R∗ = (mω). So R is a local ring with maximal ideal m = (mω) of index p.

42

Lemma 3.1.5. If p - m thenOd,m

pOd,m

∼=Od

pOd

and the isomorphism is given by

r + (mω)s 7−→ r + (ms)ω

Proof. It is a trivial matter to verify that the map is a homomorphism. Now let r + sω ∈

Od/pOd. As p - m, m−1 exists and r + (m−1s)mω 7→ r + sω. So the map is onto. Now

suppose that r + mωs 7→ 1, so r + (ms)ω = 1 in Od/pOd, so r = 1 and ms ≡ 0 (mod p).

But p - m, so s = 0. So r +mωs = 1. So the map is injective. Hence result.

The structure of Od/pOd is well known and can be found in any book on algebraic number

theory. See, for example, [52] page 108. First recall the definition of the discriminant of

K = Q(√−d):

D =

−d if d ≡ 3 (mod 4),

−4d else.

Now χK , the quadratic character of K is defined as follows. If p | D then χK(p) = 0,

χK(2) =

1 if D ≡ 1 (mod 8),

−1 if D ≡ 5 (mod 8).

and if p 6= 2 then

χK(p) =

1 if D ≡ x2 (mod p),

−1 else.

Theorem. [52] Let p ∈ Z be prime then

pOd =

p1p2 for χK(p) = 1,

p for χK(p) = −1,

p2 for χK(p) = 0.

where p, pi are prime ideals of Od.

Lemma 3.1.6. If p | m then |SL2(Od,m/pOd,m)| = p3|SL2(Fp)|.

43

Proof. Now Od,m/pOd,m is local with maximal ideal m, of index p, so∣∣∣∣SL2

(Od,m

pOd,m

)∣∣∣∣ =

∣∣∣∣SL2

(Od,m

m

)∣∣∣∣ ∣∣∣∣SL2

(Od,m

pOd,m

,m

pOd,m

)∣∣∣∣ .and |SL2(Od,m/m)| = |SL2(Fp)|, and by lemma (4.2.9)

∣∣∣SL2

(Od,m

pOd,m, m

pOd,m

)∣∣∣ = p3.

Thus

Lemma 3.1.7. If p | m then

|SL2(Od,m) : SL2(Od,mp)| = p3

Lemma 3.1.8. If p - m then

∣∣∣∣SL2

(Od,m

pOd,m

)∣∣∣∣ =

p2(p2 − 1)2 if χK(p) = 1,

p4(p2 − 1) if χK(p) = 0,

p2(p4 − 1) if χK(p) = −1.

Proof. If χK(p) = −1 then Od,m/pOd,m∼= Fp2 and |SL2(Fp2)| = p2(p4 − 1), by [79] theorem

8.8 and the comments at the bottom of p.157. Now suppose that χK(p) = 1, so that R =

Od,m/pOd,m has two ideals m1, and m2 of index p. So, by (6.4.4), SL2(R) = SL2(R/m1) ×

SL2(R/m2) and |SL2(R)| = |SL2(Fp)|2 = p2(p2−1)2, by [79] theorem 8.8 and the comments

at the bottom of p.157. Finally suppose that χK(p) = 0, so L = Od,m/pOd,m is a local ring

and, as before |SL2(L)| = p3|SL2(Fp)| = p4(p2 − 1).

Thus

Lemma 3.1.9. Suppose that p - m then

|SL2(Od,m) : SL2(Od,mp)| =

p(p2 − 1) if χK(p) = 1,

p3 if χK(p) = 0,

p(p2 + 1) if χK(p) = −1.

Now observe that, from the above, if p - m

|SL2(Od,m) : SL2(Od,mpα)| = |SL2(Od : SL2(Od,pα)|

44

so that

|SL2(Od) : SL2(Od,m)| =r∏

i=1

|SL2(Od) : SL2(Od,pαii

)|

where m = pα11 . . . pαr

r and

|SL2(Od) : SL2(Od,pα)| = |SL2(Od) : SL2(Od,p)||SL2(Od,p) : SL2(Od,pα)|

= p3(α−1)|SL2(Od) : SL2(Od,p)|.

Thus

Theorem 3.1.10.

|PSL2(Od) : PSL2(Od,m)| = m3

r∏i=1

1

p3i

|PSL2(Od) : PSL2(Od,p)| ,

where m = pα11 . . . pαr

r , and

|PSL2(Od) : PSL2(Od,p)| =

p(p2 − 1) if χK(p) = 1,

p3 if χK(p) = 0,

p(p2 + 1) if χK(p) = −1.

3.2 A miscellany of results

Theorem 3.2.1. Let m1,m2 ∈ N be distinct, and suppose m2 | m1, so PSL2(Od,m1) 6

PSL2(Od,m2). Then PSL2(Od,m1) 6 PSL2(Od,m2).

Proof. Let

X =

α β

γ δ

∈ PSL2(Od,m2),

and let

M =

1 1 +m1ω

0 1

∈ PSL2(Od,m1)

so

XMX−1 ≡

1− αγ α2

−γ2 1 + αγ

( mod m1ωM2(Od)) .

45

Now suppose that d ≡ 1, 2 (mod 4). Let p be a rational prime such that p - d and p - m1,

so p - dm22. So ∃x, y ∈ Z such that px+ dm2

2y = 1 ie px+m2ω(−ωm2y) = 1. Let α = m2ω,

γ = p then

X =

m2ω −x

p −m2ωy

∈ PSL2(Od,m2)

and XMX−1 /∈ PSL2(Od,m1)

Now suppose that d ≡ 3 (mod 4), so ω2 = ω− (d+ 1)/4, so (d+ 1)/4 = ω(1− ω). Let p

be a rational prime such that p - (d+ 1)/4 and p - m1, so p - m22(d+ 1)/4. So ∃x, y ∈ Z such

that px+m22y(d+ 1)/4 = 1 ie px+m2ω(m2y(1− ω)) = 1. Then

X =

m2ω −x

p (1− ω)m2y

∈ PSL2(Od,m2)

and XMX−1 /∈ PSL2(Od,m1). Hence result.

Theorem 3.2.2. Let N PSL2(Od,m) be of index n. Then if 6 - n then o(N) = Od,m and

N has torsion.

Proof. Suppose that 6 - n, so 2 - n or 3 - n. If 2 - n then

a =

0 −1

1 0

∈ N,and a2 = 1 so N has torsion. Also a ≡ kI (mod q) ⇒ 1 ∈ q ie q = Od,m, so o(N) = Od,m.

If 3 - n then

at =

0 −1

1 1

∈ Nso, as above N has torsion and o(N) = Od,m.

3.3 A presentation for SL2(Z[√−3])

Dennis has shown [19]

Theorem. SL2(Z[√−3]) = E2(Z

[√−3]).

46

Following Fine [25] we apply the following, due to Cohn [15]

Theorem. Let R be a subring of C with the usual absolute valuation. Suppose that if α ∈ R,

and |α| 2 then α = 0, or |α| = 1, or |α| = √p, where p = 2, 3. Then E2(R) is generated

by E(x) with the following complete set of relations:

1. E(x)E(0)E(y) = −E(x+ y)

2. E(x)D(µ) = D(µ−1)E(µxµ)

3. (E(α)E(α))p = −I ∀α ∈ R such that |α| = √p, where p = 2, 3.

4. E(µ)E(µ−1)E(µ) = D(−µ)

Where x, y ∈ R, µ ∈ R∗, and

E(x) =

x 0

−1 0

, D(µ) =

µ 0

0 µ−1

.

Let O = Z + i√

3Z. Now O∗ = ±1, and the only elements of norm between 1 and 2

are ±i√

3. So we get for SL2(Z[√−3])

Generators:

E(x), where x ∈ R

Relators:

E(x)E(0)E(y) = −E(x+ y), (3.1)

−I central, (−I)2 = I, (3.2)

(E(i√

3)E(−i√

3))3 = −I, (3.3)

E(1)3 = −I, E(−1)3 = I. (3.4)

Using (3.1) we can reduce the generators to E(0), E(±1), E(±i√

3), as follows. Let z ∈ Z,

and suppose z > 0. Then

E(z) = −E(z − 1)E(0)E(1),

47

and

E(iz√

3) = −E(i(z − 1)√

3)E(0)E(i√

3).

We get similar formulas for z 6 0. Putting these together we get

E(z1 + iz2

√3) = −E(z1)E(0)E(iz2

√3).

If x, y = 0 then it is easy to see

E(O)2 = −I ⇔ E(x)E(0)E(y) = −E(x+ y) (3.5)

and (3.1) certainly implys the following

E(1)E(0)E(i√

3) = E(i√

3)E(0)E(1) (3.6)

E(1)E(0)E(−i√

3) = E(−i√

3)E(0)E(1) (3.7)

E(−1)E(0)E(i√

3) = E(i√

3)E(0)E(−1) (3.8)

E(−1)E(0)E(−i√

3) = E(−i√

3)E(0)E(−1) (3.9)

and we also get

E(−1) = −E(0)E(1)−1E(0)−1 (3.10)

E(−i√

3) = −E(0)E(i√

3)−1E(0)−1 (3.11)

Using (3.10), and (3.11) we can eliminate the generatorsE(−1), and E(−i√

3), so SL2(Z[√−3])

is generated by E(0), E(1), and E(i√

3). We can show that (3.7), (3.8), (3.9) can be deduced

from (3.6) as follows. RTP

E(1)E(0)E(−i√

3) = E(−i√

3)E(0)E(1)

now, using (3.10), and (3.11)

LHS = E(1)E(0)(−E(0)E(i√

3)−1E(0)−1)

= E(1)E(i√

3)−1E(0)−1

48

and

RHS = −E(0)E(i√

3)−1E(0)−1E(0)E(1)

= −E(0)E(i√

3)−1E(1)

now

E(1)E(i√

3)−1E(0)−1(−E(0)E(i√

3)−1E(1))−1

=− E(1)E(i√

3)−1E(0)−1E(1)−1E(i√

3)E(0)−1

=− E(1)E(1)−1E(0)−1E(i√

3)−1E(i√

3)E(0)−1

=− E(0)−2 = I

and so LHS = RHS. The others are done in a similar manner.

And so a presentation for SL2(Z[√−3]) is:

Generators:

E(0), E(1), E(i√

3),−I

Relators:

E(1)3 = −I

E(0)2 = −I

−I central, (−I)2 = I

(E(i√

3)E(0)E(i√

3)−1E(0)−1)3 = I

E(1)E(0)E(i√

3) = E(i√

3)E(0)E(1)

Letting

A = E(0) =

0 1

−1 0

, T = E(0)E(1)−1 =

1 1

0 1

,

U = −E(0)E(1)−1E(0)E(i√

3)−1 =

1 2ω

0 1

.

49

We get

Generators:

A, T, U, J

Relators:

A2 = (AT )3 = J,

J central, J2 = I

(U−1TAT−1UA−1)3 = [T, U ] = I

Let W = T−1U and set J = I to get

Theorem 3.3.1.

PSL2(O3,2) = PSL2(Z[√−3]) =< a, t, w; a2, (at)3, (w−1awa)3, [t, w] >

3.4 The Group PSL2(O3,2)

First of all observe

Theorem 3.4.1.

PSL2(O3,2)ab =< t,w; t6, [t, w] >∼= Z6 × Z

Letting N = N(w), the normal closure of w in PSL2(O3,2), we see that l(N) = 0 and

PSL2(O3,2)

N=< a, t; a2, (at)3 >= M.

We aim to decompose PSL2(O3,2) as a non-trivial free product with amalgamation and as

an HNN group. We do here for PSL2(O3,2) what Fine did in [25] for PSL2(O2), PSL2(O7),

PSL2(O11).

3.4.1 HNN and amalgam decomposition

Theorem 3.4.2. PSL2(O3,2) is an HNN extension of K3,2 with the Modular group asso-

ciated, where K3,2 = S3 ∗Z3 D(3, 3, 3), and D(3, 3, 3) =< x, y;x3, y3, (xy)3 > is one of von

Dyck’s groups (see [36]).

50

Proof. We start with the presentation

< a, t, w; a2, (at)3, (w−1awa)3, [t, w] >

Letting v = w−1aw, the presentation becomes

< a, t, v, w; a2, (at)3, (av)3, v2, (tv)3, t = w−1tw, v = w−1aw >

Let

K3,2 =< a, t, v; a2, (at)3, (av)3, v2, (tv)3 >

We claim that PSL2(O3,2) is an HNN extension of K3,2 with < a, t >∼= M ∼=< v, t >

associated. First of all we show that < a, t >∼= M . Now as a2 = (at)3 = 1 it is clear that

M < a, t >. But < a, t >6 K3,2, and K3,2 M via the map t 7→ t, a 7→ a, and

v 7→ a. We see from this that < a, t > M , and so that M ∼=< a, t >. We can show that

< v, t >∼= M in exactly the same way. Let θ be the automorphism of K3,2 given by a 7→ v,

v 7→ a, and t 7→ t. We must show that θ(< a, t >) =< v, t >. Now θ(a) = v ∈< v, t >, and

θ(t) = t ∈< v, t >, so clearly θ(< a, t >) 6< v, t >. Similarly θ(< v, t >) 6< a, t >. Now

< v, t >= θ(θ(< v, t >)) 6 θ(< a, t >), and so θ(< a, t >) =< v, t >.

Now consider K3,2 and let s = at, m = av, so t = as, and v = am. Thus

K3,2 =< a, s,m; a2, s3,m3, (am)2, (sm2)3 >

Recall that

< a,m; a2,m3, (am)2 >∼= S3

and

< s,m; s3,m3, (sm−1)3 >∼= D(3, 3, 3)

So

K3,2 = S3 ∗m=m D(3, 3, 3)

We have the following theorem

51

Theorem 3.4.3. [48] If G maps onto a free product then G can be decomposed as an amal-

gamated free product. More precisely if θ : G A ∗ B and H = θ−1(A) ∩ θ−1(B), GA and

GB are disjoint copies of θ−1(A) and θ−1(B) respectively and HA, HB are copies of H in GA

and GB respectively Then G is isomorphic to the amalgamated free product GA ∗HA=HBGB.

As PSL2(O3,2) maps onto PSL2(Z) which is the free product of a 2-cycle and a 3-cycle,

we immediately get that PSL2(O3,2) is an amalgamated free product. However if we do

some direct calculations we get a nice decomposition.

Theorem 3.4.4.

PSL2(O3,2) = G1 ∗H G2

where G1 is an HNN extension of S3, G2 is an HNN extension of D(3, 3, 3), and

H = Z ∗ Z3.

Proof. We start with the presentation

< a, t, w; a2, (at)3, (w−1awa)3, [t, w] >

Letting t = as, x = w−1sw, m = w−1aw the presentation becomes

< a,w, s, x,m; a2, s3, (am)3, x = w−1sw,m = w−1aw, am = sx−1, x3,m2, (sx−1)3 >

Now let

G1 =< a,m,w; a2,m2, (am)3,m = w−1aw >

and

G2 =< w, s, x; s3, x3, (sx−1)3, x = w−1sw >

and

H1 =< w, am >6 G1

and

H2 =< w, sx−1 >6 G2

52

Using the normal form theorem for HNN extensions ([12] chapter 1 theorem 31) it can be

shown that H1∼= H2

∼= Z∗Z3. We can now combine the presentation for G1 and G2, adding

the relations w = w and am = sx−1, to get

< a,m,w, s, x, w; a2,m2, (am)3,m = w−1aw, s3, x3, (sx−1)3, x = w−1sw >

after applying Tietze transformations to eliminate w we get the earlier presentation for

PSL2(O3,2). Thus

PSL2(O3,2) = G1 ∗H G2

Now recall that

S3 =< x, y;x2, y2, (xy)3 >

It is clear that

< a,m; a2,m2, (am)3 >6 G1

and that < a >∼= Z2∼=< m >. It is very easy to see that under the automorphism, θ, of S3

where θ : a 7→ m and θ : m 7→ a, we have θ(< a >) =< m > and so G1 is an HNN extension

of S3 with two 2-cycles associated. Again it is easy to see that

D(3, 3, 3) =< s, x; s2, x3, (sx−1)3 >6 G2

and that under that automorphism of D(3, 3, 3) given by ϕ : s 7→ x, and ϕ : x 7→ s, we

have ϕ(< s >) =< x >∼= Z3 and so G2 is an HNN extension of D(3, 3, 3) with two 3-cycles

associated.

Now recall that PSL2(O3) does not decompose as a non-trivial free product with amal-

gamation or as an HNN group. However the above gives us a natural example of a subgroup

of finite index in PSL2(O3) which does. Hence

Theorem 3.4.5. PSL2(O3) is virtually a non-trivial product with amalgamation and vir-

tually an HNN group. More precisely |PSL2(O3) : PSL2(O3,2)| = 10 and PSL2(O3,2)

decomposes as a non-trivial free product with amalgamation and as an HNN group.

53

Contrast this with Fine’s [25] rather artificial construction of a subgroup of index 144

which decomposes as a non-trivial free product with amalgamation. We can pose the fol-

lowing

Question. Is 10 the least index of a subgroup of PSL2(O3) which decomposes as a free

product with amalgamation or as an HNN group?

3.4.2 Some Consequences

First recall that in an HNN group a torsion element must be conjugate to a torsion element

in the base ([12] chapter 1 exercise 22). We need the following

Lemma 3.4.6. Every element of finite order in D(3, 3, 3) is conjugate to one of:

x, x2, y, y2, xy, (xy)2

all of which have order 3.

Proof. This is easy to see by making D(3, 3, 3) act on the infinite lattice in Euclidean 2-space

made up of regular hexagons and triangles as given in [36] p.93.

Theorem 3.4.7. In PSL2(O3,2) every element of finite order is conjugate to one of:

a, at, (at)2, aw−1aw, (aw−1aw)2, w−1awt, w−1awt−1

where a has order 2 and the rest have order 3

Proof. Recall that PSL2(O3,2) is an HNN extension of K3,2 and K3,2 = S3 ∗Z3 D(3, 3, 3). Let

g ∈ PSL2(O3,2) be of finite order. So g is conjugate to an element of finite order in K3,2 and

so g is conjugate to an element of finite order in S3 or D(3, 3, 3). Now suppose that g has

order 2, then, as all the elements of finite order in D(3, 3, 3) have order 3, g is conjugate to

an element of order 2 in S3, recall

S3 =< a,m; a2,m3, (am)2 >

54

and S3 has exactly one conjugacy class of order 2. So g is conjugate to a.

Now suppose that g has order 3, so as S3 has only one conjugacy class of order 3 and it

is represented by m we have that g is conjugate to one of:

s, s2,m,m2, sm−1, s−1m

The result follows, as s = at, and m = aw−1aw.

Theorem 3.4.8. Let τ ∈ PSL2(O3,2) be a torsion element. Write x ∼ y if x and y are

conjugate. Then

1. If τ ∼ at, (at)2, w−1awt, w−1awt−1 then PSL2(O3,2)

N(τ)∼= Z2 × Z

2. If τ ∼ a then PSL2(O3,2)

N(τ)∼= Z3 × Z

3. If τ ∼ aw−1aw, (aw−1aw)2 then PSL2(O3,2)

N(τ)∼= Z×M

Proof. Recall

PSL2(O3,2) =< a, t, w; a2, (at)3, [t, w] , (w−1awa)3 >

set a = 1, so t3 = 1, so we get

PSL2(O3,2)

N(a)=< t,w; t3, [t, w] >∼= Z3 × Z

The others are similar

Theorem 3.4.9. Let T be the set of all torsion elements in PSL2(O3,2). Then N(T ) =

N(M).

Proof. N(T ) = N(a, at, aw−1aw,w−1awt, w−1awt−1), andN(M) = N(a, t). ClearlyN(M) ⊆

N(T ). Now suppose that a = t = 1. So a = at = aw−1aw = w−1awt = w−1awt−1 = 1. So

N(T ) ⊆ N(M).

Lemma 3.4.10. Let N Zp × Z =< x, y >, be of index n , where p is a rational prime.

Then if p - n then

N = N(x, yn)

55

if p | n then

N = N(x, yn), N(xmyk)

where n = pk, and m = 0, 1, . . . , p− 1.

Proof. Suppose p - n. Then x ∈ N and it is clear that N = N(x, yn).

Suppose that p | n, so n = pk, say. If, as above x ∈ N then N = N(x, yn) so suppose

that x /∈ N . Let l be the order of y mod N , so l | pk. Suppose that l | k, and l 6= k, so

yl = 1, and l < k, but then there are < pk distinct cosets of N . Contradiction. So if l | k

then l = k, and as l | pk we must have l = k, or pk.

Suppose that l = k. Then N(yk) 6 N . But N(yk) is of index pk = n, so N =

N(yk) = N(x0yk). Now suppose that l = pk. So y, y2, . . . , ypk are pk = n distinct cosets

representatives of N . Now x /∈ N , so x = yr some r, 1 6 r 6 pk, and 1 = xp = ypr, so

pk | pr so k | r, so r = k, 2k, . . . , pk. If r = pk then x = ypk = 1. Contradiction, as x /∈ N .

So r 6= pk, so x = yk, y2k, . . . , y(p−1)k. Hence, using the fact that Z/pZ is a field yk = xm

where m = 1, . . . , p− 1. So N(xmyk) 6 N . But N(xmyk) is of index n, so N = N(xmyk).

We now show that these groups are distinct. Suppose that

N(x, yn) = N(xmyk)

then yk ∈ N(x, ypk). Contradiction. Suppose that

N(xm1yk) = N(xm2yk) = N

where 0 6 m1,m2 6 p− 1. Then xm1yky−kx−m2 ∈ N , so xm1−m2 ∈ N , so m1 ≡ m2 (mod p)

so m1 = m2. Hence all the groups are distinct.

Theorem 3.4.11. Let N PSL2(O3,2) be of index n. Then if (n, 6) = 1

N = N(a, t, wn)

if 2 | n, and 3 - n then N is one of

N(a, t, wn), N(at, wn/2), N(at, awn/2)

56

if 2 - n, and 3 | n then N is one of

N(a, t, wn), N(a, wn/3), N(a, twn/3), N(a, t2wn/3)

Proof. Suppose that (n, 6) = 1, so working mod N , a = at = w−1awa = 1, so a = t = 1, so

N = N(a, t, wn). Now suppose that 2 - n, and 3 | n, so a = 1. So PSL2(O3,2)/N is a factor

of < t,w; t3, tw = wt >∼= Z3 × Z. So, by (3.4.10), N = N(a, t, wn), N(a, wn/3), N(a, twn/3),

N(a, t2wn/3). Now suppose 2 | n, and 3 - n, so at = w−1awa = 1. So PSL2(O3,2)/N a

factor of < a,w; a2, aw = wa >∼= Z2 × Z. So, by (3.4.10), N = N(a, t, wn), N(at, wn/2),

N(at, awn/2).

Theorem 3.4.12. Let n ∈ N. Let H = PSL2(O3,2)/PSL2(O3,2)n. Then

1. If (n, 6) = 1 then H ∼= Zn and PSL2(O3,2)n = N(a, t, wn).

2. If 3 - n and 2|n then H ∼= Z2 × Zn and PSL2(O3,2)n = N(at, wn).

3. If 2 - n and 3|n then H ∼= Z3 × Zn and PSL2(O3,2)n = N(a, wn).

Proof. Recall

PSL2(O3,2) =< a, t, w; a2, (at)3, [t, w] , (w−1awa)3 >

Suppose that (n, 6) = 1 then, working mod PSL2(O3,2)n, a = at = 1, so a = t = 1, and

wn = 1. So H ∼= Zn. Now suppose that 3 - n and 2 | n then at = 1, so a = t, and

(w−1awa) = 1 so [a, w] = 1, and wn = 1, so H ∼= Z2×Zn. Suppose that 2 - n and 3 | n then

a = 1 so t3 = 1, and wn = 1. So H ∼= Z3 × Zn.

Theorem 3.4.13.

PSL2(O3,2)′ < PSL2(O3,2)

2 ∩ PSL2(O3,2)3

Proof. Obviously PSL2(O3,2)′6 PSL2(O3,2)

2 ∩PSL2(O3,2)3. Consideration of index shows

that the inclusion is strict.

Remark. It is well known [70] that, in the Modular group PSL2(Z)′ = PSL2(Z)2 ∩

PSL2(Z)3. Fine ([25] Corollary 4.5.4.3) shows that if d = 2, 7, 11 then PSL2(Od)′ <

PSL2(Od)2 ∩ PSL2(Od)

3.

57

3.5 Presentations for some other PSL2(Od,m).

Using GAP [24] we find the following presentation for a subgroup of PSL2(O1) of index 8

< a, t, w, z; a2, z2, (at)3, (atz)2, [t, w] , (atw−1zw)2 >

where

a =

0 1

−1 0

, t =

1 1

0 1

, w =

1 2i

0 1

,

z = (uat)a(uat)−1 =

1− 2i 1− 2i

−2 2i− 1

.

Since all the generators lie in PSL2(O1,2), and, by (3.1.10), |PSL2(O1) : PSL2(O1,2)| = 8,

the group given by this presentation must be PSL2(O1,2). Hence

Theorem 3.5.1.

PSL2(O1,2) =< a, t, w, z; a2, z2, (at)3, (atz)2, [t, w] , (atw−1zw)2 >

We can show that PSL2(O1,2) decomposes as an HNN group and as a non-trivial amal-

gamated free product in the same way as we did for PSL2(O3,2).

Using GAP [24] we find the following presentation for a subgroup of PSL2(O11) of index

10

< a, t, w, k, l,m, n; a2, (at)3, k2, [m,n] , [t, w] , tklkt−1l−1, km−1ltat−1l−1m,

nlat−1l−1n−1mwaw−1tm−1,m−1tklt−1l−1mwt−1aw−1t >

Where

w =

1 2ω

0 1

, k = (uat)a(uat)−1 =

−1 + 2ω 5

2 1− 2ω

,

l = (ua)u2(ua)−1 =

−7 + 2ω 6 + 4ω

2ω 5− 2ω

,

m = (uaua)u(uaua)−1 =

−8 + 6ω 21 + 6ω

3 + 2ω 10− 6ω

,

58

n = (ua)uuatat−1au−1t−1(ua)−1 =

7 + 6ω 26− 14ω

6− 2ω −5− 6ω

.

Since all the generators lie in PSL2(O11,2), and, by (3.1.10), |PSL2(O11) : PSL2(O11,2)| =

10, the group given by this presentation must be PSL2(O11,2).

Again using GAP [24] wefind the following presentation for a subgroup of index 6 in

PSL2(O7):

< a, t, w, x, y ; a2, x2, [t, w] , (ax)3, (at)3, y = txyxt−1, (ytay−1x)2,

xyat−1y−1wt−1aw−1t >

where

w =

1 2ω

0 1

, x = (uat)a(uat)−1 =

1− 2ω −3

−2 −1 + 2ω

,

y = (ua)u2(ua)−1 =

−5 + 2ω −4− 2ω

2ω 3− 2ω

.

Since all the generators lie in PSL2(O7,2) and, by (3.1.10), |PSL2(O7) : PSL2(O7,2)| = 6

the group given by this presentation must be PSL2(O7,2).

The method of this section ie getting GAP [24] to list all subgroups of Γd of the relevent

index and then labouriously checking whether the generators for each lay in PSL2(Od,m) is

very unsophisticated and consequently of use in only very few cases. Swan [84] has developed

a method for computing presentations for PSL2(Od) which can easily be extended to the

PSL2(Od,m) ([78] page 628). However the computations rapidly become unwieldy as d

becomes large. Riley [78] developed a computer package (written in Fortran) called the

Poincare File which he used to find presentations for PSL2(Od) for 10 < d < 37 and

d = 43, 67, 163. Such an approach can be used to find presentations for PSL2(Od,m). The

implementation of this in GAP [24] could be an interesting future project.

59

Chapter 4

Non-standard normal subgroups of

SL2(Od,m)

Let R be a commutative ring with a one. Let n ∈ N such that n > 2. Let 1 6 i, j 6 n then

eij denotes the n× n matrix with a 0 in every position except (i, j), where it has a 1.

En(R) =< I + reij : 1 6 i, j 6 n, i 6= j >

Let q R

En(R, q) =< I + αeij : α ∈ q, 1 6 i, j 6 n, i 6= j >En(R)

the largest q such that En(R, q) 6 S is the level of S and is denoted l(S). This is well defined

because En(R, q1)En(R, q2) = En(R, q1 + q2).

Hn(R, q) = M ∈ SLn(R) : M ≡ kI (mod q), some k ∈ R

the smallest q such that S 6 Hn(R, q) is the order of S and is denoted o(S). S 6 SLn(R)

is said to be standard if l(S) = o(S).

E0(2, R; q) = N SL2(R) : o(N) = q, l(N) = 0

60

4.1 Non-standard normal subgroups of SL2(Z)

Here we describe the proof of the following theorem, mentioned in chapter one. This proof

appeared in [62].

Theorem. Let 0 6= q Z then |E0(2,Z; q)| = 2ℵ0.

First recall that the Modular group, PSL2(Z), has the presentation < x, y;x2, y3 > (see

[81]).

Theorem. [62] Every countable group can be embedded in an infinite simple group generated

by x, y of order 2, 3 respectively, and where xy has infinite order.

Corollary. [62] ∃2ℵ0 N SL2(Z) such that SL2(Z)/N is simple, and l(N) = 0.

Lemma 4.1.1. [62] Let X,Y be sets, X infinite. Let f : X −→ Y be a surjection. For

y ∈ Y let cy = |f−1(y)|. Then if ∃c0 such that cy 6 c0 |X| ∀y ∈ Y then |X| = |Y |.

Proof. Obviously |Y | 6 |X|. Now |X| =∑

y∈Y cy 6 |Y |cy 6 |Y |c0, and co |X| so

|X| 6 |Y |.

Theorem. [62] Let 0 6= q Z then |E0(2,Z; q)| = 2ℵ0.

Proof. Let

S = N SL2(Z) : l(N) = 0, SL2(Z)/N simple

so |S| = 2ℵ0 . We show that if N ∈ S then o(N) = Z and so |E0(2,Z,Z)| = |S| = 2ℵ0 . Let

N ∈ S, let q0 = o(N). So, since N 6 H(q0) SL2(Z) H(q0) = N or SL2(Z). Suppose

H(q0) = N then q0 = l(H(q0)) = l(N) = 0 so o(N) = 0, but o(N) = 0⇔ N = 1 or I,−I

and PSL2(Z) is not simple, so o(N) 6= 0. Hence H(q0) = SL2(Z) so q0 = l(H(q0)) =

l(SL2(Z)) = Z i.e. o(N) = Z.

Now suppose q 6= Z. Let S1 = H(q) ∩N : N ∈ S. Define ρ : S −→ S1 by ρ(N) = H(q)∩

N . Let X ∈ S1, and let S2 = ρ−1(X) ⊆ S. Suppose Y ∈ S2 so SL2(Z)/Y is simple and

61

o(Y ) = Z. Now Y H(q) = Y , or SL2(Z). Suppose that Y H(q) = Y , so H(q) 6 Y . So

q 6 l(Y ) = 0. Contradiction. Hence Y H(q) = SL2(Z), so

|Y : X| = |Y : H(q) ∩ Y | = |SL2(Z) : H(q)| ∞

so |S2| 6 ℵ0. Now S is infinite, ρ is a surjection and |S2| 6 ℵ0 2ℵ0 = |S|, so by the lemma

|S1| = |S| = 2ℵ0 .

We now show that if M ∈ S1 then l(M) = 0 and o(M) = q, and so |E0(2,Z; q)| = 2ℵ0 , as

required. Let M ∈ S1, so M = H(q) ∩N , some N ∈ S. Now M 6 N , so l(M) 6 l(N) = 0,

so l(M) = 0. Let q1 = o(M). Now M 6 H(q) so o(M) = q1 6 q. Let N be the image of N

in PSL2(Z), so, as −I ∈ NSL2(Z)

N∼=PSL2(Z)

N

since N is of infinite index in PSL2(Z), N is free and so contains elements of infinite order.

So N contains elements of infinite order. Now, as before

|N : M | = |N : H(q) ∩N | = |SL2(Z) : H(q)| <∞

so M contains elements of infinite order, so q1 = o(M) 6= 0. Now, as before, NH(q1) =

SL2(Z), so

H(q) = H(q) ∩NH(q1)

= H(q1)(N ∩H(q))

= H(q1)M

= H(q1).

so q1 = q. Hence result.

4.2 Non-standard normal subgroups of SL2(Od,m)

In this section we aim to extend the results proved above to the groups SL2(Od). This

answers a question of Lubotzky’s (MR92c : 20088). In fact we extend it to all SL2(Od,m)

62

with a free non-abelian quotient, more precisely what we require is that the group maps onto

the Modular group, M in a “nice” way. Recall from chapter two that it is conjectured that

the only SL2(Od,m) which do not have a free non-abelian quotient are SL2(O1), SL2(O1,2),

SL2(O2), SL2(O3), SL2(O3,2), SL2(O7), SL2(O11), and SL2(O19). We show that in fact

SL2(O1,2), SL2(O2), and SL2(O3,2) map onto the Modular group in a “nice” way and so

the result goes though for these groups as well, leaving only 5 true exceptions. We prove an

unsatisfactory version of the result for SL2(O7), SL2(O11), and SL2(O19) (which relies on

the fact that they have an infinite cyclic quotient) but we were unable to make any progress

with SL2(O1) and SL2(O3).

Lemma 4.2.1. Let N SL2(Od,m). Then

l(N) = 0⇔

1 1

0 1

or

1 mω

0 1

has infinite order mod N.

Proof. Suppose that 1 1

0 1

and

1 mω

0 1

have finite order mod N i.e. 1 s

0 1

,

1 tmω

0 1

∈ N, some s, t ∈ N.

Let M = st, let q = (M) Od,m, so q 6= 0. Let α ∈ q, so

α = (z1 +mωz2)M = z1M +mωz2M,

some z1, z2 ∈ Z. So 1 α

0 1

=

1 M

0 1

z1 1 mωM

0 1

z2

∈ N.

Hence E2(Od,m, q) 6 N , so l(N) 6= 0.

Conversely suppose l(N) 6= 0. So E2(Od,m, q) 6 N , some q 6= 0. Now |Od,m : q| ∞,

and so 1 + q has finite (additive) order i.e. s ∈ q, some s ∈ N. Similarly tmω ∈ q, some

63

t ∈ N. So 1 s

0 1

, and

1 tmω

0 1

∈ Ni.e. 1 1

0 1

, and

1 ω

0 1

have finite order mod N .

Recall the following result from lemma (2.3.2).

Theorem 4.2.2. Suppose r = r(d,m) > 1 then there exists a surjective homomorphism

ρ : SL2(Od,m) −→ Fr =< x1, . . . , xr >

such that

ρ

1 mω

0 1

= x1

Now consider the case r(d,m) > 1.

Lemma 4.2.3. If r(d,m) > 1 then ∃2ℵ0 N SL2(Od,m) such that SL2(Od,m)/N is simple,

l(N) = 0, and N contains elements of infinite order.

Proof. We have

ϕ : SL2(Od,m) Fr < x1, x2;x22, (x1x2)

3 >∼= PSL2(Z)

and

u =

1 mω

0 1

7→ x1 7→ t

Now by a result of Mason and Pride [62] mentioned in the previous section the Modular group

PSL2(Z) has 2ℵ0 normal subgroups N such that PSL2(Z)/N is simple and t has infinite

order mod N . Pulling these subgroups back to SL2(Od,m) we get 2ℵ0 normal subgroups N

of SL2(Od,m) such that SL2(Od,m)/N is simple and u has infinite order mod N . Further, as

kerϕ contains SL2(Z), N contains elements of infinite order. Hence result.

64

We can now generalize almost word for word the corresponding proof for the Modular

group.

Theorem 4.2.4. Suppose r(d,m) > 1 and let 0 6= q Od,m then |E0(2,Od,m; q)| = 2ℵ0.

Proof. Let

S = N SL2(Od,m) : l(N) = 0, SL2(Z) 6 N,SL2(Od,m)/N simple

so, by (4.2.3), |S| = 2ℵ0 . We show that ifN ∈ S then o(N) = Od,m and so |E0(2,Od,m,Od,m)| =

|S| = 2ℵ0 . Let N ∈ S, let q0 = o(N). So, since N 6 H(q0) SL2(Od,m) we have

H(q0) = N or SL2(Od,m). Suppose H(q0) = N then q0 = l(H(q0)) = l(N) = 0 so o(N) = 0,

but o(N) = 0 ⇔ N = 1 or I,−I and PSL2(Od,m) is not simple, so o(N) 6= 0. Hence

H(q0) = SL2(Od,m) so q0 = l(H(q0)) = l(SL2(Od,m)) = Od,m i.e. o(N) = Od,m.

Now suppose q 6= Od,m. Let S1 = H(q) ∩N : N ∈ S. Define ρ : S −→ S1 by

ρ(N) = H(q) ∩ N . Let X ∈ S1, and let S2 = ρ−1(X) ⊆ S. Suppose Y ∈ S2 so

SL2(Od,m)/Y is simple and o(Y ) = Od,m. So Y H(q) = Y , or SL2(Od,m). Suppose that

Y H(q) = Y , so H(q) 6 Y , so q 6 l(Y ) = 0. Contradiction. Hence Y H(q) = SL2(Od,m).

Now

|Y : X| = |Y : H(q) ∩ Y | = |SL2(Od,m) : H(q)| ∞

so |S2| 6 ℵ0. Now S is infinite, ρ is a surjection and |S2| 6 ℵ0 2ℵ0 = |S|, so by lemma

(4.1.1), |S1| = |S| = 2ℵ0 .

We now show that if M ∈ S1 then l(M) = 0 and o(M) = q, and so |E0(2,Od,m; q)| = 2ℵ0 ,

as required. Let M ∈ S1, so M = H(q)∩N , some N ∈ S. Now M 6 N , so l(M) 6 l(N) = 0,

so l(M) = 0. Let q1 = o(M). Now M 6 H(q) so o(M) = q1 6 q. Now N contains elements

of infinite order. So, as before,

|N : M | = |N : H(q) ∩N | = |SL2(Od,m) : H(q)| <∞

soM contains elements of infinite order, so q1 = o(M) 6= 0. As before, NH(q1) = SL2(Od,m),

65

so

H(q) = H(q) ∩NH(q1)

= H(q1)(N ∩H(q))

= H(q1)M

= H(q1).

so q1 = q. Hence result.

We have already seen in chapter two that when r(d,m) = 1 we may still have a free

non-abelian quotient. We now verify that in certain cases the mapping is “nice”

Lemma 4.2.5. For (d,m) = (5, 1), (6, 1), (15, 1), (14, 1), (7, 2), (11, 2), SL2(Od,m) has a

free quotient of rank 2 and the image of 1 mω

0 1

can be taken as a free generator. Further the kernel of this map contains an element of

infinite order.

Proof. Here we work with the presentations for PSL2(Od,m), as these are a homomorphic

images of SL2(Od,m) this is sufficient. In the following the presentations for d = 5, 6, 15 are

taken from [84], and for d = 14 from [30]. The presentations for PSL2(O7,2) and PSL2(O11,2)

can be found in (3.5). a, t, u have the usual meanings.

PSL2(O5) =< a, t, u, b, c; a2, (at)3, [t, u] , b2, (ab)2, (aubu−1)2, aca = tct−1, ubu−1cb = tct−1 >

so PSL2(O5)/N(a, t, b) =< u, c;>∼= F2.

PSL2(O6) =< a, t, u, b, c; a2, (at)3, [t, u] , b2, [a, c] , (atb)3, t−1ctubu−1 = bc, (atubu−1)3 >

so PSL2(O6)/N(a, t, b) =< u, c;>∼= F2.

PSL2(O15) =< a, t, u, c; a2, (at)3, [t, u] , [a, c] , ucuat = taucu >

66

so PSL2(O15)/N(a, t) =< u, c;>∼= F2.

PSL2(O14) =< a, t, u, b, c, d, e;a2, (at)3, [t, u] , [a, b] , cd−1eb−1dc−1be−1,

dau−1b−1dad−1baud−1a,

adtad−1ad−1bt−1ab−1d,

at−1ae−1dat−1d−1acatad−1eatb−1dac−1b >

after some calculation we see that PSL2(O14)/N(a, t, b, e, d) =< u, c;>∼= F2.

PSL2(O11,2) =< a, t, w, k, l,m, n; a2, (at)3, k2, [m,n] , [t, w] , tklkt−1l−1, km−1ltat−1l−1m,

nlat−1l−1n−1mwaw−1tm−1,m−1tklt−1l−1mwt−1aw−1t >

so PSL2(O11,2)/N(a, k, t,m, n) =< w, l;>∼= F2.

PSL2(O7,2) =< a, t, w, x, y ; a2, x2, [t, w] , (ax)3, (at)3, y = txyxt−1, (ytay−1x)2,

xyat−1y−1wt−1aw−1t >

so PSL2(O7,2)/N(a, x, t) =< w, y;>∼= F2.

It is clear from the proof of theorem (4.2.4) that it is sufficient for SL2(Od,m) to map

onto the Modular group with kernel of level zero. We remark that PSL2(O5)/N(b, c, u) ∼=

PSL2(O6)/N(b, c, u) =

< a, t; a2, (at)3 >= PSL2(Z).

Lemma 4.2.6.PSL2(O2)

N(u)=< a, t; a2, (at)3 >∼= PSL2(Z)

PSL2(O3,2)

N(w)=< a, t; a2, (at)3 >∼= PSL2(Z)

PSL2(O1,2)

N(a, t, (zw)3)=< z,w; z2, (zw)3 >∼= PSL2(Z)

Clearly the kernels contain elements of infinite order.

67

Proof. This is obvious given the following presentations

PSL2(O2) =< a, t, u; a2, (at)3, (u−1aua)2, [t, u] >

PSL2(O3,2) =< a, t, w; a2, (at)3, (w−1awa)3, [t, w] >

PSL2(O1,2) =< a, t, w, z; a2, z2, (at)3, (atz)2, [t, w] , (atw−1zw)2 >

The presentation for PSL2(O2) can be found in [25]. The presentation for PSL2(O1,2) can

be found in (3.5.1) and for PSL2(O3,2) in (3.3.1).

Thus we have

Proposition 4.2.7. Let (d,m) = (1, 2), (2, 1), (3, 2), (5, 1), (6, 1), (14, 1), (15, 1). Let

0 6= q Od,m then |E0(2,Od,m; q)| = 2ℵ0.

The only “true exceptions” not covered above are SL2(O1), SL2(O3), SL2(O7), SL2(O11),

and SL2(O19). The results we can achieve are very unsatisfactory. When we have an infinite

cyclic quotient we get the following result

Theorem 4.2.8. Suppose SL2(Od,m)/K ∼= Z and l(K) = 0 then ∀ 0 6= q Od,m ∃N

SL2(Od,m) such that l(N) = 0 and o(N) = q.

The following lemma was proved in [63] as Theorem 4.1 for any n > 2 and any Dedekind

domain of arithmetic type, A.

Lemma 4.2.9. Let R be a commutative ring with a one, q1, q2 R such that R/q2 is an

SR2-ring q2 6 q1. ThenSL2(R, q1)

SL2(R, q2)is abelian ⇔ q2

1 6 q2

Furthermore, if SL2(R, q1)/SL2(R, q2) is abelian then SL2(R, q1)/SL2(R, q2) ∼= a3, where

a = q1/q2, and the isomorphism is given by: 1 + z x

y 1− z

←→ (x, y, z).

68

Proof. Denote SL2(R, qi) by Γ(qi). Now suppose that q21 6 q2 and let X, Y ∈ Γ(q1). Then

X − I, Y − I ≡ 0 (mod q1) and so we have

(X − I)(Y − I) ≡ (Y − I)(X − I) ≡ 0 (mod q21)

It follows that XY ≡ Y X (mod q2) and so Γ(q1)′ ⊆ Γ(q2). Conversely suppose that

Γ(q1) / Γ(q2) is abelian. Let x, y ∈ q1 then

[I + xe12, I + ye21] =

∗ ∗

∗ 1− xy

≡ I (mod q2)

and so xy ∈ q2 hence q21 ⊆ q2.

Recall that by lemma (3.1.1), ϕ : SL2(R) −→ SL2(R/q2) is surjective, and Γ(q2) =Kerϕ.

Let γ(q1) = ϕ(Γ(q1)) and observe that γ(q1) ∼= Γ(q1)/Γ(q2). We show that γ(q1) ∼= a3.

Define θ : a3 −→ SL2(R/q2) by

(x, y, z) 7−→

1 + z x

y 1− z

.

We first show that θ is a homomorphism. Let xi, yi, zi ∈ a where i = 1, 2. Then since q21 6 q2,

θ (x1, y1, z1) θ (x2, y2, z2) =

1 + z1 + z2 x1 + x2

y1 + y2 1− z1 − z2

= θ (x1 + x2, y1 + y2, z1 + z2)

So θ is a homomorphism.

We now show that θ is injective. Suppose that θ (x, y, z) = 1 i.e. 1 + z x

y 1− z

≡ I (mod q2)

so (x, y, z) = 0 and so θ is injective.

We now show that imθ = γ(q1). Let X ∈ imθ. Since X ∈ SL2(R/q2) and ϕ is surjective

∃Y ∈ SL2(R) such that ϕ (Y ) = X. We show Y ∈ Γ(q1) and so X = ϕ ( Y ) ∈ γ( q1 ).

Now X ∈ imθ so

X =

1 + z x

y 1− z

69

some x, y, z ∈ a. Let

Y =

a b

c d

where a, b, c, d ∈ R. So a ∈ 1 + z, b ∈ x, c ∈ y, d ∈ 1− z, so a, d ≡ 1 (mod q1) and b, c ≡ 0

(mod q1), so Y ≡ I (mod q1) i.e. Y ∈ Γ(q1). Hence imθ ⊆ γ(q1).

Now we show that γ(q1) ⊆ imθ. Let X ∈ γ(q1), so X = ϕ(Y ) some Y ∈ Γ(q1). RTP

X =

1 + z x

y 1− z

some x, y, z ∈ a. Let

Y =

a b

c d

where a, b, c, d ∈ R. Now Y ≡ I (mod q1) so a, d ≡ 1 (mod q1) and b, c ≡ 0 (mod q1). Let

x = b + q2, y = c + q2 so that x, y ∈ a. Now let z = ζ + q2 = (a − 1) + q2 and since a ≡ 1

(mod q1) we have z ∈ a. We show that d+ q2 = 1− z. Now (a− 1)(d− 1) ∈ q21 6 q2 and so

a + d ≡ 2 (mod q2). So now 1 − z = 1 − ζ + q2 = 1 − (a − 1) + q2 = 2 − a + q2 = d + q2.

Hence

X = ϕ(Y ) =

a+ q2 b+ q2

c+ q2 d+ q2

=

1 + z x

y 1− z

some x, y, z ∈ a. Hence γ(q1) ⊆ imθ. Hence result.

Theorem 4.2.10. Suppose that SL2(Od,m)/K ∼= Z and l(K) = 0. Then ∀ 0 6= q Od,m

∃N SL2(Od,m) such that l (N) = 0 and o (N) = q.

Proof. Let 0 6= q1 Od,m and consider H (q1) ∩K. Let q2 = o (H (q1) ∩K), so q2 6 q1.

Clearly l (H (q1) ∩K) = 0. We show q1 = q2. Suppose not. Now Γ/K ∼= Z so

H (q1) / H (q2) is cyclic, so Γ (q1) /Γ (q1) ∩ H (q2) is cyclic. Also Γ (q1) ∩ H (q2) /Γ (q2)

is central in Γ (q1) /Γ (q2). So Γ (q1) /Γ (q2) is central by cyclic, and so abelian. Hence

q21 6 q2. Let a = q1/q2. So Γ (q1) /Γ (q2) ∼= a3. Where 1 + z x

y 1− z

←→ (x, y, z)

70

We show a2 → Γ (q1) /Γ (q1) ∩H (q2). Define

ϕ : a2 −→ Γ (q1)

Γ (q1) ∩H (q2)

by

ϕ(x, y) =

1 x

y 1

.

ϕ is clearly a homomorphism. Now suppose that ϕ(x, y) = 1. So 1 x

y 1

∈ Γ (q1) ∩H (q2) ,

so x, y ∈ q2, so (x, y) = 0 in a2, hence ϕ is injective. So a2 is cyclic, so a is trivial i.e.

q1 = q2.

The only (d,m) excluded from the above theorem are (1, 1), and (3, 1). Neither SL2(O1)

nor SL2(O3) have infinite cyclic quotients and so using the above techniques we are unable

to say anything about E0(2,Od; q) for d = 1, 3.

However despite the difficulties mentioned above we feel confident in making the following

Conjecture. ∀(d,m) and ∀ 0 6= q Od,m, |E0(2,Od,m; q)| = 2ℵ0 .

71

Chapter 5

Order and level of a normal

congruence subgroup

We have seen that in general there is no relationship between the order and level of a normal

subgroup of SL2(Od,m). However if we restrict ourselves to normal congruence subgroups

we do find a nice relationship. Mason [58] has proved

Theorem. Let NSL2(Od) be a congruence subgroup of level q∗ and order q. Then 12q 6 q∗.

Suppose that N SL2(Od) is a congruence subgroup of order Od then SL2(Od, 12Od) 6

N . We use this fact to show that a large class of subgroups of the SL2(Od) are non-congruence

subgroups. First of all we aim to extend the above results to SL2(Od,m). In fact we extend

it to SL2 over a larger class of rings. The material in this chapter is the most technically

complex and perhaps the hardest to follow in this thesis, so we apologize to the reader for

the Cimmerian night which is about to descend.

5.1 Primary decomposition in Noetherian domains

The material in this section has been lifted from chapters four, seven, and eleven of Atiyah

and MacDonald [2]. Let A be any commutative ring.

Definition. q A is primary if q 6= A and xy ∈ q⇒ x ∈ q or yn ∈ q, some n > 0.

72

Definition. Let q A. Then the radical of q is

r(q) = x ∈ A : xn ∈ q, some n > 0

Proposition 5.1.1. ([2] Proposition 1.14) The radical of q A is the intersection of all

prime ideals which contain q.

Proposition 5.1.2. ([2] Exercise 1.13) Let a, b A. Then

r(ab) = r(a ∩ b) = r(a) ∩ r(b)

Proposition 5.1.3. ([2] Proposition 4.1) Let q A be primary. Then r(q) is the smallest

prime ideal containing q.

Definition. Let a A. A primary decomposition of a is

a =n⋂

i=1

qi

where qi A is primary.

Definition. If p = r(q) we say that q is p-primary.

Lemma 5.1.4. ([2] Lemma 3.1) Let qi A, 1 6 i 6 n be p-primary. Then q = ∩ni=1qi is

p-primary.

If in a primary decomposition, all the r(qi) are distinct and ∩i6=jqj * qi, i = 1, . . . , n

then we say that the decomposition is minimal. All primary decompositions discussed here

shall be minimal.

Recall that a ring A is Noetherian if it satisfies three equivalent conditions:

1. Every nonempty set of ideals in A has a maximal element.

2. Every ascending chain of ideals in A is stationary.

3. Every ideal in A is finitely generated.

From now on let A be Noetherian.

73

Definition. We say that a A is irreducible if a = b ∩ c⇒ a = b, or c.

Lemma 5.1.5. ([2] Lemma 7.11) Every ideal in A is a finite intersection of irreducible

ideals.

Definition. Let x ∈ A. Then the annihilator of x is Ann(x) = a ∈ A : ax = 0.

Lemma 5.1.6. ([2] Lemma 7.12) Every irreducible ideal in A is primary.

Thus we have proved the following

Theorem 5.1.7. ([2] Theorem 7.13) In a Noetherian ring, every ideal has a primary de-

composition.

Proposition 5.1.8. ([2] Proposition 7.14) In a Noetherian ring A, every ideal contains a

power of its radical.

Let R be a commutative ring. A proper chain of prime ideals

0 p1 · · · pn R

is of length n. The Krull dimension of R is the maximum length of chains of prime ideals

in R. Thus a domain is of Krull dimension one, if and only if all non-zero prime ideals are

maximal. From now on K shall denote a Noetherian domain of Krull dimension one.

Theorem 5.1.9. Let K be a Noetherian domain of Krull dimension one. Let 0 6= q K.

Then

q = p1 . . . pt

where pi is a primary ideal. Furthermore, for each i, r(pi) = mi, a maximal ideal; mnii 6 pi,

for some ni; and pi + pj = K, for i 6= j.

Proof. The radical of a primary ideal is prime, and so, as K is of Krull dimension one, is

maximal. Suppose that

q =t⋂

i=1

pi

74

is a minimal primary decomposition. Now r(pi) = mi is maximal and, as the decomposition

is minimal, mi 6= mj if i 6= j. Further mnii 6 pi, as every ideal contains a power of its radical.

Thus pi + pj = K if i 6= j. So q is an intersection of coprime ideals. Hence a product

q = pi . . . pt

Lemma 5.1.10. Let K be a Noetherian domain of Krull dimension one. Let q∗ 6 q be

ideals in K such that 0 6= q∗ 6= K. Suppose that q∗ = p∗1 . . . p∗t is a primary decomposition.

Then q = p1 . . . pt where pi = K or r(pi) = r(p∗i ) and p∗i 6 pi.

Proof. If q = K then the result is trivial, so suppose that K 6= q = p1 . . . ps is a primary

decomposition. We can suppose that r(p1) = r(p∗1) = m1. Let q∗0 = p∗2 . . . p∗t , q0 = p2 . . . ps.

So q∗ = p∗1q∗0, and q = p1q0. Now p∗1 + q∗0 = K = p1 + q0. Since r(p∗1) = r(p1) we have

p1 + q∗0 = K. So p∗1(p1 + q∗0) = p∗1, so p∗1p1 + p∗1q∗0 = p∗1. Now p∗1q

∗0 = q∗ 6 q 6 p1, and

p∗1p1 6 p1, so p∗1 6 p1. Now for those p∗i which have a radical distinct from all the r(pj) we

let pi = K.

Example 5.1.1. Let D be a Dedekind domain. Then by definition ([2] chapter 9) D is

a Noetherian domain of Krull dimension one. In this case every primary ideal is a prime

power.

Example 5.1.2. Let Od,m = Z +mωZ be an order in an imaginary quadratic number field.

Then Od,m is a Z-module of finite rank and so, by Hilbert’s basis theorem (see [2] Theorem

7.5) is Noetherian. Now Od,m is of finite index in Od and Od is a Dedekind domain and so a

Noetherian domain of Krull dimension one. Thus by theorem 20(1) on page 81 of [64], Od,m

is a Noetherian domain of Krull dimension one, it is clearly of characteristic zero. Note that

maximal orders are Dedekind domains. However in non-maximal orders not every primary

ideal is a power of a prime ideal.

For example consider the order O3,2 = Z + i√

3Z. Let m = (2, 1 + i√

3). Then m is a

maximal ideal of O3,2 of index 2 and m2 < 2O3,2 < m. So the ideal 2O3,2 is a primary ideal

which is not a power of a prime ideal.

75

Example 5.1.3. Let p ∈ Z be a rational prime and let K = Z/pZ. Let R = K [X] and let

q R be of finite index. R is Noetherian by Hilbert’s basis theorem. Consider the subring

S = K + q, of R. We show that S is a Noetherian domain of Krull dimension one. By

theorem 20(1) on page 81 of [64], if S is Noetherian then it is of Krull dimension one, so it

suffices to show that S is Noetherian. Now R is a PID so q = (f) where deg f = d. Let

g =∑aiX

i ∈ R and consider fg = b0f + · · ·+ btXtf , which is in the K-module generated

by fK,XfK, . . . , X tfK. Suppose that t > d, so degX tf > 2d = deg f 2. So, as R has a

Euclidean algorithm, X tf = hf2+r, where deg r < deg f 2 = 2d, and deg h 6 degX tf−2d <

degX tf . So that fg is in the K-module generated by fK,XfK, . . . , X t−1fK. Repeat the

above until we see that fg is in the K-module generated by fK,XfK, . . . , Xd−1fK. So

S = K + q is generated as a K-module by K, fK,XfK, . . . , Xd−1fK and so is Noetherian,

by Hilbert’s basis theorem. S is thus an example of a Noetherian domain of Krull dimension

one and non-zero characteristic.

5.2 Wohlfahrt’s Theorem

Wohlfahrt [91] showed that if S is a subgroup of SL2(Z), of level m then S is a congruence

subgroup if and only if SL2(Z,mZ) 6 S. That is he extended Klein’s concept of the level of

a congruence subgroup of the Modular group to an arbitrary subgroup of the Modular group.

This concept of level and Wohlfahrt’s theorem have been of great use in the construction of

non-congruence subgroups of the Modular group. We have seen how the concept of level can

be extended to SLn(R) where R is any ring with a one. In this section we extend Wohlfahrt’s

theorem to all Noetherian domains of Krull dimension one.

Definition. R is said to be an SR2-ring if a, b ∈ R such that gcd(a, b) = 1⇒ ∃t ∈ R such

that a+ tb ∈ R∗.

Lemma 5.2.1. Let A be any commutative ring with a one. Let a, b ∈ A such that gcd(a, b) =

1. Let q A such that A = A/q is an SR2-ring. Then ∃t ∈ A such that (a+ bt)A+ q = A.

76

Proof. As a, b are coprime ∃s, t ∈ A such that as + bt = 1. Now 1 = ϕ(1) = ϕ(as + bt) =

as + bt, where ϕ : A → A is the natural homomorphism. So a, b are coprime. So, as A is

an SR2-ring ∃t ∈ A such that (a + bt) = u, where u ∈ A∗, so (a + bt) ≡ u (mod q), so

u = (a+ bt)− q, some q ∈ q. Hence (a+ bt)A+ q = A.

Theorem 5.2.2. Let A be any commutative ring. Let qiA, i = 1, 2. Suppose that A = A/q1

is an SR2-ring. Then

SL2(A, q2) 6 E2(A, q2)SL2(A, q1)

Proof. Denote SL2(A, qi) by Γ(qi) and E2(A, qi) by ∆(qi). Let

X =

a b

c d

∈ Γ(q2).

Now ad− bc = 1 so a, b are coprime, and A is an SR2-ring, so by the lemma (5.2.1) ∃t ∈ A

such that (a+ bt)A is prime to q1. Let

T =

1 0

t 1

∈ ∆(A)

so

T−1XT =

a+ bt ∗

∗ ∗

∈ Γ(q2)

so wlog aA and q1 are coprime. Let x ∈ A such that ax ≡ 1 (mod q1). Now let

T1 =

1 0

1 1

, X1 =

1 x(1− a− b)

0 1

,

X2 =

1 a− 1

0 1

.

a ≡ 1 (mod q2), so X2 ∈ ∆(q2), and 1− a, b ∈ q2 so X1 ∈ ∆(q2). Now

T−11 XX1T1X2 ≡

1 0

q 1

(mod q1)

where q ∈ q2. So T−11 XX1T1X2 ∈ ∆(q2)Γ(q1). Hence X ∈ ∆(q2)Γ(q1), as X1, X2 ∈ ∆(q2),

and T1 ∈ ∆(A), and ∆(q2) ∆(A).

77

The following corollary is the classical form of Wohlfahrt’s theorem and is equivalent to

the theorem.

Corollary 5.2.3. Let G 6 SL2(A) have level q. Then G is a congruence subgroup if and

only if SL2(A, q) 6 G.

Corollary 5.2.4. Let A be an SR2-ring. Let q A. Then SL2(A, q) = E2(A, q). So in

particular SL2(A) = E2(A).

Proof. Take q1 = 0 in the theorem.

Corollary 5.2.5. Let A be any commutative ring. Let qi A, i = 1, 2. Suppose that

A = A/q2 is an SR2-ring. Then

SL2(A, q1 + q2) = E2(A, q1)SL2(A, q2)

Proof. Obviously Γ(q1)Γ(q2) 6 Γ(q1 + q2). Then

Γ(q1 + q2) 6 ∆(q1 + q2)Γ(q2) by theorem (5.2.2)

= ∆(q1)∆(q2)Γ(q2)

6 ∆(q1)Γ(q2)

Lemma 5.2.6. [3] Semilocal rings are SR2 rings.

Proof. Let R be a semilocal ring,and let m1, . . . ,mt be the maximal ideals of R. Let a, b ∈ R

such that gcd(a, b) = 1. RTP ∃t ∈ R such that a+ tb ∈ R∗.

∀i = 1, . . . , t suppose

∀s ∈ R, a+ bs ≡ 0 (mod mi)

Then, taking s = 0, we get a ∈ mi, and, taking s = 1, we get b ∈ mi. But a, b are coprime.

Contradiction. So ∀i ∃ti such that a + tib 6≡ 0 (mod mi). Then, by the Chinese Remainder

Theorem, ∃t ∈ R such that t ≡ ti (mod mi) ∀i. So

a+ tb ≡ a+ tib 6≡ 0 (mod mi) ∀i

So a+ tb ∈ R∗, as required.

78

Lemma 5.2.7. Let K be a Noetherian domain of Krull dimension one. Let 0 6= q K.

Then K/q is semilocal and therefore an SR2-ring.

Proof. AsK is Noetherian we have a primary decomposition q = ∩ti=1pi. AsK has dimension

one this is a product pi . . . pt and r(pi) is maximal. Let mi = r(pi). So ∀i ∃ni ∈ N such that

mnii 6 pi. So mn1

1 . . .mntt 6 q. Now suppose that m K/q is a maximal ideal. So m = m/q,

some maximal ideal m > q. Suppose that m 6= mi ∀i. Then m + mn11 . . .mnt

t = K, but

mn11 . . .mnt

t 6 q, so m + q = K. Contradiction. Hence K/q has only finitely many maximal

ideals and so is semilocal and therefore SR2.

Thus

Theorem 5.2.8. Let K be a Noetherian domain of Krull dimension one. Let qiK, i = 1, 2,

q1 6= 0. Then

SL2(K, q1 + q2) = E2(K, q2)SL2(K, q1)

SL2(K, q1 + q2) = SL2(K, q1)SL2(K, q2)

Let O be an order of an imaginary quadratic number field. So O is a Noetherian domain

of Krull dimension one, so we have a Wohlfahrt theorem in these rings. We remark that

maximal orders are Dedekind domains and in this case the result has already been proved

[63].

5.3 Preliminary results about SL2(L)

Let L be a commutative local ring with maximal ideal m. Let N SL2(L). Let o(N) = q,

and l(N) = q∗. We are interested in how q and q∗ are related. It will turn out that whether

2 is a unit or not is of critical importance. The case 2 ∈ L∗ is easier to handle than 2 /∈ L∗.

When 2 /∈ L∗ we have two cases, o(N) = L and o(N) 6= L both of which split into two

subcases, |L : m| > 2 and |L : m| = 2, with the latter case being the most difficult. Here we

present a complete account; the results up to lemma (5.3.15) are known.

79

Lemma 5.3.1. [3] Let L be a commutative local ring and let q L. Then

SL2(L, q) = E2(L, q)

and so, in particular

SL2(L) = E2(L)

Proof. This follows from (5.2.4) and (5.2.6)

Lemma 5.3.2. Let N SL2(L). Then o(N) is generated byc ∈ L :

∗ ∗c ∗

∈ N

Proof. o(N) is generated by b, c, a− d for all a b

c d

∈ N.The result follows as 0 −1

1 0

a b

c d

0 1

−1 0

=

d −c

−b a

∈ Nand

V T−1XTV −1 =

d −c

d− a+ c− b a− c

∈ Nwhere

T =

1 1

0 1

, V =

0 −1

1 0

Corollary 5.3.3. Let L be any local ring. Let N SL2(L) such that o(N) = L. Then

∃X =

∗ ∗c ∗

∈ N , such that c is a unit.

80

We make use of the following notation.

E12(x) =

1 x

0 1

, E21(y) =

1 0

y 1

The following lemma is a slight generalization of lemme 3.3 part (ii) of [44].

Lemma 5.3.4. Let A be a commutative ring. Let t ∈ A, and u ∈ A∗, and let

Y =

u t

0 u−1

∈ N SL2(A).

Then E12((u− u−1)α) ∈ N ∀α ∈ A.

Proof. 1 (u− u−1)α

0 1

=

1 −u−1α

0 1

u t

0 u−1

1 u−1α

0 1

u−1 −t

0 u

∈ N

Lemma 5.3.5. ( [54] lemma 1.1 ) Let A be a commutative ring. Let N SL2(A) and

suppose that

M =

a b

c d

∈ N.Then ∀u ∈ A∗ such that u2 ≡ 1( mod c), u4 − 1 ∈ l(N).

Proof. Let

M2 =

u t

0 u−1

where u ∈ A∗ such that u2 ≡ 1( mod c), and t ∈ A. Then

[M2,M ] = M−12 M−1M2M =

α β

γ δ

,

where

α = (u−1d+ ct)(ua+ tc)− u−1c(bu−1 + at),

81

γ = ac− u2ac− utc2,

δ = ad− u2bc− utcd.

Now if a−u2a−utc = 0 then γ = 0 and a−u2a−utc = 0⇔ a(1−u2) = utc. Since u2 ≡ 1(

mod c), u2 − 1 = kc, some k ∈ A. So let t = u−1ak, so γ = 0. Now tc = u−1a(1 − u2),

substituting this into the expressions for α and δ and simplifying we see that α− δ = u4− 1

and so by (5.3.4) we get u4 − 1 ∈ l(N).

Lemma 5.3.6. Let L be a commutative local ring with maximal ideal m. Let N SL2(L),

o(N) = q 6 m. Then 8q 6 l(N).

Proof. Let

M =

a b

c d

∈ Nby (5.3.2) it is enough to show that 8c ∈ l(N). Let u1 = 1 + c, u2 = 1 − c. As o(N) 6= L,

ui ∈ L∗, and u2i ≡ 1( mod c), so by (5.3.5) u4

i −1 ∈ l(N), so (u41−1)−(u4

2−1) = 8c(c2+1) ∈

l(N), and 1 + c2 ∈ L∗ so 8c ∈ l(N).

Lemma 5.3.7. Let L be a commutative local ring with maximal ideal m. Let N SL2(L).

Suppose that 2 ∈ L∗ and o(N) 6= L. Then o(N) = l(N).

Proof. As 2 ∈ L∗, 8 ∈ L∗ and so this follows from (5.3.6).

Lemma 5.3.8. Let L be a commutative local ring with maximal ideal m. Suppose that 2 = 0.

Let N SL2(L) and suppose that o(N) = q 6 m. Then

q(4) =< α4 : α ∈ q >6 l(N)

Proof. Consider ∗ ∗c ∗

∈ Nas o(N) 6= L, c ∈ m. Let u1 = 1 + xc, u2 = 1 + yc, some x, y ∈ L. Now, as 2 = 0,

c4(x4 − y4) = u41 − u4

2 ∈ l(N), by (5.3.5). Pick x, y ∈ L such that x = y + 1, so x4 − y4 = 1.

Hence result.

82

When |L : m| > 2 we can improve the above results. First we need Whitehead’s lemma

Lemma 5.3.9. Let R be any commutative ring with a one. Let u ∈ R∗. Then u 0

0 u−1

∈ E2(R).

Proof. Let u ∈ R∗. Then 1 u

0 1

1 0

−u−1 1

1 u

0 1

=

0 u

−u−1 0

∈ E2(R).

Now u2,−u ∈ R∗, so 0 u2

−u2 0

0 −u

u−1 0

=

u 0

0 u−1

∈ E2(R)

Lemma 5.3.10. ([43] Proposition 1.3.6 ) Let A be any commutative ring. Let q A. Let

Γ = SL2(A), ∆ = E2(A), ∆(q) = E2(A, q), Γ(q) = SL2(A, q), H(q) = H2(A, q). Then

[Γ, H(q)] 6 Γ(q)

[∆,∆(q)] 6 ∆(q)

and if ∃u ∈ A∗ such that u2 − 1 ∈ A∗ then

[∆,∆(q)] = ∆(q)

Proof. Let M1 ∈ Γ, M2 ∈ H(q). So [M1,M2] = M1M2M−11 M−1

2 ≡ M1kIM−11 k−1I ≡ I

mod q. So [Γ, H(q)] ∈ Γ(q). Now, as ∆,∆(q) are both normal in ∆, [∆,∆(q)] 6 ∆∩∆(q) =

∆(q).

Now let t ∈ A and suppose that u ∈ A∗ such that u2 − 1 ∈ A∗ and consider u 0

0 u−1

,

1 t

0 1

=

1 (u2 − 1)t

0 1

83

By Whitehead’s lemma (5.3.9) u 0

0 u−1

∈ ∆

Now let α ∈ q and let t = (u2 − 1)−1α ∈ q, so that 1 α

0 1

=

1 (u2 − 1)t

0 1

∈ [∆,∆(q)]

Lemma 5.3.11. ( [54] lemma 1.2 ) Let L be a commutative local ring with maximal ideal

m. Suppose that |L : m| > 2 and 2 /∈ L∗. Let N SL2(L) have order q 6 m. Let

M =

a b

c d

∈ N.Then 2c2, c4 ∈ l(N)⇒ 2c, c2 ∈ l(N).

Proof. Now

M2 = MT−1M−1T =

e f

g h

∈ Nwhere e = 1 + ac and g = c2. Note that 2g, and g2 ∈ l(N) = q. Now let

M3 = M2T (−t)M−12 T (t) =

1 + get t(1− e2 + get)

g2t 1− get+ g2t2

where t ∈ L. Let r = 1 + get and s = t(1− e2) + get2. So

D(r, r)T (r−1s) =

r 0

0 r

1 r−1s

0 1

=

r s

0 r

and g2, 2g ∈ q so

M3 ≡ D(r, r)T (r−1s)( mod q)

Let q = 1− r2 ∈ q and let

M4 =

r(1 + q) q

−q r

84

then detM4 = r2(1+q)+q2 = r2(2−r2)+(1−r2)2 = 1, so M4 ∈ SL2(L). Clearly M4 ∈ H(q).

Now

M4M3 ≡

r 0

0 r

r s

0 r

=

r2 rs

0 r2

( mod q)

and q = 1− r2 ∈ q, so r2 ≡ 1( mod q). Now rs = s+ tegs = s+ t2eg(1− e2) + t3e2g2, and

g2 ∈ q, so t3e2g2 ∈ q. Also 1− e2 = −(2ac+ a2g), so t2eg(1− e2) = −t2e(2gac+ a2g2), and

2g, g2 ∈ q, so rs ≡ s( mod q). Hence SR ≡ T (s)( mod q). So T (s) = T (t(1−e2))T (t2eg) ∈

NH(q). Now u−1 0

0 u

e f

g h

u 0

0 u−1

=

e u−2f

u2g h

∈ Nand repeat the above argument with this matrix to get T (t(1 − e2))T (t2eu2g) ∈ NH(q),

∀u ∈ L∗ ∀t ∈ L.

Now, as |L : m| > 2 we can find v ∈ L∗ such that v − 1 ∈ L∗. Now consider the above

with u = v, v − 1. So

T (t(1− e2))T (t2ev2g) ∈ NH(q)

and

T (t(1− e2))T (t2eg)T (t2egv2)T (−t2e2gv) ∈ NH(q)

Now 2g ∈ q, so T (−2t2e2gv) ∈ NH(q) and so it follows from the above that

T (t(1− e2)) ∈ NH(q)

∀t ∈ L. Hence e2−1 ∈ l(NH(q)). Now e2−1 = (1+ac)2−1 = a(2c+ac2), and ad− bc = 1,

c ∈ m, so a, d ∈ L∗, so 2c+ ac2 ∈ l(NH(q)). Now, again using the fact that |L : m| > 2, let

w ∈ L∗ such that w2 − 1 ∈ L∗, so w−1 0

0 w

a b

c d

w 0

0 w−1

=

a w−2b

w2c d

Repeating the above argument we see that 2w2c+ aw4c2 = w2(2c+ aw2c2) ∈ l(NH(q)). So,

as w2 ∈ L∗, 2c+ aw2c2 ∈ l(NH(q)). So 2c+ aw2c2 − (2c+ ac2) = ac2(w2 − 1) ∈ l(NH(q)),

and a, w2 − 1 ∈ L∗, so c2 ∈ l(NH(q)), and 2c ∈ l(NH(q)).

85

So Γ(2c+ c2) 6 NH(q). Let q0 = (2c) + (c2). Now

Γ(q0) = [Γ,Γ(q0)] by (5.3.10) and (5.3.1)

6 [Γ, NH(q)] as Γ(q0) 6 NH(q0)

6 [Γ, N ] [Γ, H(q)]

6 NΓ(q) by (5.3.10)

6 N as Γ(q) 6 N

Lemma 5.3.12. ( [54] theorem 1.3 ) Let L be a commutative local ring with maximal ideal

m. Suppose that |L : m| > 2, 2 /∈ L∗, and 2 6= 0. Let N SL2(L), o(N) = q 6 m. Then

2q 6 l(N).

Proof. We show that 2c4, c8 ∈ l(N) and apply (5.3.11). Let

M =

a b

c d

∈ Nso c ∈ m. Let x, y ∈ L and consider 1 + xc, 1 + yc ∈ L∗, so by (5.3.5),

(1 + xc)4 − (1 + yc)4 = 4c(x− y) + 6c2(x2 − y2) + 4c3(x3 − y3) + c4(x4 − y4)

= (x− y)c (2 + (x+ y)c)(2 + 2(x+ y)c+ (x2 + y2)c

)∈ l(N)

Now, using the fact that |L : m| > 2 choose x = u, y = v such that u, v ∈ L∗, u + v = 1, so

u− v = 1− 2v ∈ L∗, as 2 /∈ L∗. So

c(2 + c)(2 + 2c+ (1− 2uv)c2) ∈ l(N)

Choosing x = 1, y = 0, we see that

c(2 + c)(2 + 2c+ c2) ∈ l(N)

86

so that

2c3(2 + c) ∈ l(N)

Now, again using the fact that |L : m| > 2 choose u ∈ L∗ such that u2−1 ∈ L∗. Conjugating

X by D(u, u−1) and repeating the above argument shows that

2c3(2 + u2c) ∈ l(N)

so, since u2 − 1 ∈ l(N), we see that

2c4 ∈ l(N)

Now, with x = 1, y = 0, we see that

c4 + 4c3 + 6c2 + 4c ∈ l(N)

so, multiplying by c4 we get

c8 + 4c7 + 6c6 + 4c5 = c8 + 2c4(2c3 + 3c2 + 2c) ∈ l(N)

so

c8 ∈ l(N)

Hence result.

Lemma 5.3.13. ( [54] theorem 1.3 ) Let L be a commutative local ring with maximal ideal

m. Suppose that |L : m| > 2, and 2 = 0. Let N SL2(L) and suppose that o(N) = q 6 m.

Then q(2) 6 l(N).

Proof. By (5.3.8) 2c4, c4 ∈ l(N). Hence by (5.3.11), 2c, c2 ∈ l(N). Clearly 2c = 0. Hence

result.

Lemma 5.3.14. Let L be a commutative local ring with maximal ideal m. Let N SL2(L).

Suppose that o(N) = L and |L : m| > 3. Then l(N) = L.

Proof. Let ∗ ∗c ∗

∈ N87

such that c ∈ L∗. Now let u ∈ L∗, so u2 − 1 = cc−1(u2 − 1), so u2 ≡ 1( mod c). So by

(5.3.5) u4 − 1 ∈ l(N). Now if ∀u ∈ L∗ u4 − 1 ∈ m, |L : m| 6 5. So if |L : m| > 5 then we are

done.

Suppose that |L : m| = 5. Let α ∈ m, and let u1 = 1 + α, u2 = 1 − α, so ui ∈ L∗ and

u2i ≡ 1 (mod c), so, by (5.3.5), 8α(1 + α2) = u4

1 − u42 ∈ l(N). Now |L : m| = 5 so 2 ∈ L∗, so

8 ∈ L∗, and 1+α2 ∈ L∗, so α ∈ l(N), hence m 6 l(N). Now let MSL2(F5) be the image of

N under the natural map SL2(L) SL2(F5), so o(M) = F5, M 6= ±I, and as PSL2(F5)

is simple, M ±I = SL2(F5). So M is of index 1 or 2 in SL2(F5) but SL2(F5)′ = SL2(F5)

[72], so M = SL2(F5). So N = SL2(L) and l(N) = L. Now suppose that |L : m| = 4. Let

u ∈ L∗ − 1. Then u4 − 1 = u− 1 /∈ m, so l(N) = L.

Lemma 5.3.15. Let L be a commutative local ring with maximal ideal m. Let N SL2(L).

Suppose that o(N) = L and |L : m| = 3. Then m 6 l(N).

Proof. As o(N) = L, by (5.3.3),

∗ ∗c ∗

∈ Nsuch that c ∈ L∗. Letα ∈ m and let u1 = 1 +α, u2 = 1−α, so ui ∈ L∗ and u2

i ≡ 1( mod c).

So 8α(1 + α2) = u41 − u4

2 ∈ l(N) by (5.3.5). As |L : m| = 3, 2 is a unit and, as α ∈ m, 1 + α2

is also a unit. So α ∈ l(N). Hence m 6 l(N).

Lemma 5.3.16. Let L be a commutative local ring with nilpotent maximal ideal m of index

2. Then |SL2(L)| = 2α3.

Proof. Let n ∈ N be minimal such that mn = 0. Let Γ = SL2(L), and Γ(i) = SL2(L,mi).

Then

|Γ| = |Γ : Γ(1)||Γ(1) : Γ(2)| . . . |Γ(n− 2) : Γ(n− 1)||Γ(n− 1) : Γ(n)|

First, as m of index 2, Γ/Γ(1) ∼= PSL2(F2) ∼= S3, so |Γ : Γ(1)| = 6. Now, Γ(i− 1)/Γ(i) ∼= Z32

by lemma (4.2.9), so |Γ(i− 1) : Γ(i)| = 8. Hence |Γ| = 23(n−1)+13. Hence result.

88

Lemma 5.3.17. Let L be a commutative local ring with nilpotent maximal ideal m of index

2. Let N SL2(L). Then

o(N) = L⇔ |SL2(L) : N | = 2α

Proof. Let Γ denote SL2(L). Suppose first that o(N) = L. If Γ(m) 6 N then, as Γ/Γ(m) ∼=

S3, |Γ : N | = 1, or 2. So suppose that Γ(m) N and suppose that 3 | |Γ : N |. Now

NΓ(m) = Γ, or is of index 2 in Γ so 3 | |NΓ(m) : N | and 3 | |NΓ(m) : Γ(m)| so 9 | |Γ :

N ∩ Γ(m)|. So 9 | |Γ|, contradicting lemma (5.3.16). So 3 - |Γ : N | ie |Γ : N | = 2α.

Now suppose that o(N) 6= L. So N 6 H(m) = Γ(m) and |Γ : Γ(m)| = 6, so 3 | |Γ : N |.

Hence result.

Now observe the following consequence of (5.3.17)

Corollary 5.3.18. Let L be a commutative local ring with maximal nilpotent ideal m of

index 2. Let N SL2(L), o(N) = L and M ∈ SL2(L) of order 3. Then M ∈ N .

Lemma 5.3.19. Let L be a commutative local ring with nilpotent maximal ideal m of index

2. Let N SL2(L), o(N) = L. Then T 4 ∈ N , and T 2 ∈ N ⇔ −I ∈ N .

Proof. AT is of order 3 so AT ∈ N , so A ≡ T−1( mod N), and A4 = I, so T 4 ∈ N . Also

A2 = −I so T 2 ≡ −I( mod N).

Lemma 5.3.20. Let a ∈ L, u ∈ L∗. Then a u

−cu−1 −(1 + a)

is of order 3, where c = a2 + a+ 1 ∈ L∗.

Proof. Tedious calculation.

Lemma 5.3.21. Let L be a commutative local ring with nilpotent maximal ideal m, of index

2. Let α ∈ m and Let N SL2(L) be of order L. Then 1 4α

0 1

∈ N.89

Proof. By lemma (5.3.20) 1 u

−3u−1 −2

∈ Nfor every unit u. Let α ∈ m , u = 1 + α, and suppose that u−1 = 1 + β, some β ∈ m. Now 1 u

−3u−1 −2

=

1 u

0 1

A

1 u−1

0 1

A−1

1 3u

0 1

A

1 −u−1

0 1

A−1

=

1 1 + α

0 1

A

1 1 + β

0 1

A−1

1 3 + 3α

0 1

A

1 −1− β

0 1

A−1

= T

1 α

0 1

AT

1 β

0 1

A−1T 3

1 3α

0 1

AT−1

1 −β

0 1

A−1

and AT ∈ N , A−1T 3 ≡ A−1T−1 ≡ I( mod N), AT−1 ≡ ATT−2 ≡ T−2( mod N), so

T

1 4α+ β

0 1

T−2

1 −β

0 1

A−1 ∈ N.

so 1 4α

0 1

A−1T−1 ∈ N

and A−1T−1 ∈ N . Hence 1 4α

0 1

∈ Nas required.

Theorem 5.3.22. Let L be a commutative local ring with nilpotent maximal ideal m of index

2. Let N SL2(L) be of order L. Then 4L 6 l(N).

Proof. We show E2(L, 4L) 6 N . Recall that E2(L, 4L) is generated by conjugates of

E12(x) =

1 x

0 1

90

where x ∈ 4L. Let x ∈ 4L, so x = 4y, some y ∈ L. Now y ∈ L∗ or y ∈ m. Suppose that

y ∈ L∗, so y = 1 + α, where α ∈ m. So x = 4y = 4 + 4α, so

E12(x) =

1 x

0 1

=

1 4

0 1

1 4α

0 1

∈ N.If y ∈ m then similarly e12(x) ∈ N . So E2(L, 4L) 6 N , as required.

Since we are also interested in PSL2(R), for commutative rings R we consider normal

subgroups of SL2 which contain −I.

Lemma 5.3.23. Let L be a commutative local ring with nilpotent maximal ideal m of index

2. Let N SL2(L) have order L and let α ∈ m. Suppose that −I ∈ N . Then 1 2α

0 1

∈ Nand so 2L 6 l(N).

Proof. −I ∈ N so by (5.3.19), T 2 ∈ N . Now, by (5.3.20), ∀u ∈ L∗, 0 u

−u−1 −1

=

1 u

0 1

A

1 u−1

0 1

A−1

1 2u

0 1

∈ NLet α ∈ m, let u = 1 + α ∈ L∗ and suppose that u−1 = 1 + α′, α′ ∈ m. So that 1 1 + α

0 1

A

1 1 + α′

0 1

A−1

1 2 + 2α

0 1

∈ Nso

T

1 α

0 1

AT

1 α′

0 1

A−1T 2

1 2α

0 1

∈ Nbut, as AT , T 2 ∈ N , 1 α+ α′

0 1

1 2α

0 1

∈ N

91

Now, by (5.3.20), and since −I ∈ N , 1 u

−u−1 0

= A

1 u−1

0 1

A−1

1 u

0 1

∈ Nso

A

1 1 + α′

0 1

A−1

1 1 + α

0 1

∈ Nie

AT

1 α′

0 1

A−1T

1 α

0 1

∈ Nhence, as AT, T 2 ∈ N , we have 1 α+ α′

0 1

∈ N.Hence 1 2α

0 1

∈ N.Now using the fact that T 2 ∈ N and arguing in exactly the same way as in theorem (5.3.22),

we see that E2(L, 2L) 6 N , ie 2L 6 l(N).

Lemma 5.3.24. Let L be a commutative local ring with maximal nilpotent ideal m of index

2 and suppose that 2 = 0 in L. Let N SL2(L) have order L. Then

m(2) =< α2 : α ∈ m >6 l(N)

Proof. By lemma (5.3.20), and since 2 = 0, a u

−cu−1 1 + a

∈ Nand this matrix equals 1 u

0 1

A

1 u−1

0 1

A−1

1 au

0 1

A

1 −au−1

0 1

A−1

92

Now let α, β ∈ m, let u = 1 + α, a = 1 + β, and suppose that u−1 = 1 + α′, where α′ ∈ m.

Then au = 1 + α+ β + αβ, and au−1 = 1 + α′ + β + α′β, and so 1 1 + α

0 1

A

1 1 + α′

0 1

A−1

1 1 + α+ β + αβ

0 1

A

1 −(1 + α′ + β + α′β)

0 1

A−1

lies in N , and so, as AT ∈ N , and 2 = 0, we get 1 (α− α′)β

0 1

∈ NNow (1 +α)(1 +α′) = 1, so α+α′ +αα′ = 0, so, as 2 = 0, α−α′ = αα′, and α′ = α(1 +α′),

so α− α′ = α2u−1. So 1 α2u−1β

0 1

∈ NNow let x ∈ L, if x ∈ m then 1 α2(1 + α)−1x

0 1

∈ N.Now if x /∈ m then x = 1 + β, some β ∈ m, and 1 α2(1 + α)−1x

0 1

=

1 α2(1 + α)−1

0 1

1 α2(1 + α)−1β

0 1

Now 1 u

−u−1 0

= A

1 u−1

0 1

A−1

1 u

0 1

∈ Nby lemma (5.3.20). So using the fact that u = 1 + α, u−1 = 1 + α′, and AT ∈ N we get 1 α+ α′

0 1

∈ Nand so, as α+ α′ + αα′ = 0 we get 1 α2(1 + α)−1

0 1

∈ N93

and so 1 α2(1 + α)−1x

0 1

∈ N.Hence α2(1 + α)−1 ∈ l(N) ∀α ∈ m, and so α2 ∈ l(N) ∀α ∈ m. Hence result.

5.4 Order and level of a normal congruence subgroup

Throughout this section K shall be a Noetherian domain of Krull dimension one, so K has

primary decomposition and a Wohlfahrt theorem. Let N SL2(K), and suppose N is a

congruence subgroup. Let q∗ = l(N) 6= 0, and q = o(N). We ask how q∗ and q are related.

We are particularly interested in the case where q = K. The following 3 lemmas can be

proved using elementary group theory.

Lemma 5.4.1. Let G be any group and let A,B,C G. Then [BA,CA] 6 [B,C]A.

Lemma 5.4.2. Let A,B G. Then [A,B] 6 A ∩B.

Lemma 5.4.3. Let G be a group and let A,B,C G where C 6 A. Then A ∩ BC =

(A ∩B)(A ∩ C).

The following lemma is a generalization of [58] lemma 3.3.

Lemma 5.4.4. Let N SL2(K) = Γ, l(N) = q∗ = q∗1q∗2 6= 0, where q∗1 + q∗2 = K, and q1 is

primary. Then by (5.1.10), o(N) = q = q1q2, where q∗1 6 q1. Let N0 = (N ∩ Γ(q∗2))Γ(q∗1),

q′ = l(N0), and N = NΓ(q∗1). Then

1. o(N) = q1

2.[Γ, N

]6 N0

3. q′ = q∗1

94

Proof. Now

o(N) = o(NΓ(q∗1))

= o(N) + q∗1

= q + q∗1

= q1q2 + q∗1

= (q1 ∩ q2) + (q1 ∩ q∗1)

= q1 ∩ (q2 + q∗1) by the modular law (see [2] p.6)

= q1 ∩K = q1.

Now, as K has a Wohlfahrt theorem and q∗1 + q∗2 = K, we have, by (5.2.8), Γ = Γ(q∗1 + q∗2) =

Γ(q∗1)Γ(q∗2), so

[Γ, N

]= [Γ(q∗1)Γ(q∗2), NΓ(q∗1)]

6 [Γ(q∗2), N ] Γ(q∗1) by lemma (5.4.1)

6 N0 by lemma (5.4.2)

Now Γ(q∗1) 6 N0, so q∗1 6 l(N0) = q′. RTP q′ 6 q∗1. Now N0 is a congruence subgroup of

level q′, so by Wohlfahrt (5.2.3) Γ(q′) 6 N0, so

Γ(q′q∗2) = Γ(q′) ∩ Γ(q∗2) as q′ + q∗2 = K

6 N0 ∩ Γ(q∗2)

= ((N ∩ Γ(q∗2))Γ(q∗1)) ∩ Γ(q∗2)

= (N ∩ Γ(q∗2))Γ(q∗) by (5.4.3)

6 NΓ(q∗)

= N

So q′q∗2 6 l(N) = q∗1q∗2. Now q′ + q∗2 = q∗1 + q∗2 = K, so q′ ∩ q∗2 = q′q∗2 6 q∗1q

∗2 = q∗1 ∩ q∗2, and

95

q∗1 6 q′, so q∗1 ∩ q∗2 6 q′ ∩ q∗2. Hence q′ ∩ q∗2 = q∗1 ∩ q∗2. So

q′ = q′ ∩ (q∗1 + q∗2)

= (q′ ∩ q∗2) + q∗1

= (q∗1 ∩ q∗2) + q∗1

= q∗1.

The following lemma is a generalization of [58] lemma 3.4.

Lemma 5.4.5. Let NSL2(K), l(N) = q∗ = q∗0p∗ 6= 0, where p∗ is primary and q∗0+p∗ = K,

so by (5.1.10), o(N) = q = q0p, where p∗ 6 p. Let N0 = (N ∩Γ(q∗0))Γ(p∗), and N = NΓ(p∗).

Let L denote the local ring K/p∗. Let ϕ : SL2(K)→ SL2(L) be the natural homomorphism.

Let M0 = ϕ(N0), and M = ϕ(N). Then

1. M0 SL2(L)

2. l(M0) = 0

3. o(M) = p/p∗

4.[E2(L),M

]6 M0

Proof. 1 is obvious. For 4 recall that[SL2(K), N

]6 N0, on applying ϕ we see that[

SL2(L),M]

6 M0. Now as L is local so an SR2-ring we have E2(L) = SL2(L). For

3, as M = ϕ(N), and o(N) = p we get o(M) = p/p∗. Now suppose that l(M0) = a 6= 0. So

E2(L, a) = SL2(L, a) 6 M0, so E2(K, a) 6 N0. So, as l(N0) = p∗, we have a 6 p∗, so a = p,

so a = 0. Contradiction. Hence l(M0) = 0.

Lemma 5.4.6. ([57] lemma 2.1) Let L be any local ring. Let NSL2(L) such that o(N) = L.

Then ∃X ∈ N of the form ∗ ∗1 1

96

Proof. By (5.3.3)

∃M1 =

a b

c d

∈ Nsuch that c is a unit. Now 1 c−1d

0 1

a b

c d

1 −c−1d

0 1

=

∗ −c−1

c 0

= M2 ∈ N

Let

T =

1 −t

0 1

Then

[T,M2] = T−1M−12 TM2 =

1 + t2c2 t

tc2 1

∈ NChoose t = c−2 to get the result.

Let L be a local ring, with maximal ideal m, and residue field K = L/m. The following

theorem is a slight generalization of [58] theorem 2.2.

Theorem 5.4.7. Let L be a commutative local ring. Let N SL2(L),M 6 GL2(L) such

that N 6 M , and [E2(L),M ] 6 N . Let q = o(N), q∗ = l(N), q0 = o(M), so q∗ 6 q 6 q0.

Then q = q0, unless K = F2, and q0 6= L in which case mq0 6 q 6 q0.

Proof. Let

X =

a b

c d

∈MLet δ = ad−bc ∈ L∗. Then, as [E2(L),M ] 6 N , [T (r), X] ∈ N and so (d2−δ)rδ−1+r2cdδ−1 ∈

q ∀r ∈ L. So ∀u ∈ L∗, (d2−δ)δ−1+ucdδ−1 ∈ q, so (u−1)cd ∈ q. Suppose that q0 6 m. Then

c ∈ m, as o(M) = q0. Now ad − bc = δ ∈ L∗ and c ∈ m, so bc ∈ m. So if a or d ∈ m then

δ ∈ m. Contradiction. Hence a, d /∈ m ie a, d ∈ L∗. So, as (u− 1)cd ∈ q, (u− 1)c ∈ q. When

|K| > 2 we can choose u ∈ L∗ such that u − 1 ≡ 1 (mod q). So c ∈ q. Hence, by (5.3.2),

q0 = o(M) 6 q. So q0 = q. If |K| = 2 then u = 1 + α, where α ∈ m, so u − 1 = α ∈ m, so

αc ∈ q, so, by (5.3.2), mq0 6 q.

97

Now suppose that q0 = L, so ∃X0 ∈ N ⊆M of the form ∗ ∗1 1

So(u− 1)c = u− 1 ∈ q ∀u ∈ L∗. Thus if |K| > 2 then q = L = q0. Suppose that |K| = 2, let

α ∈ m, so u = 1 +α ∈ L∗, so α = u− 1 ∈ q, so m 6 q. Suppose that m = q and consider the

natural homomorphism f : SL2(L) → SL2(F2). Now o(N) = q so f(N) = 1, and X0 ∈ M

and [SL2(L),M ] 6 N , so f(X0) is central. But SL2(F2) has trivial centre, so f(X0) = I.

But X0 =

∗ ∗1 1

. So, under the map L→ F2, 1 7→ 0. Contradiction. Hence q = L.

We are now in a position to address the main problem of this chapter; to derive a

relationship between the order and level of a normal congruence subgroups of SL2 over a

Noetherian domain of Krull dimension one. We deal first with the case of a normal subgroup

of order K. This is for two reasons, first we are able to obtain a better bound in this case, in

fact our bound is best possible, and second, when we apply these results to the construction

of non-congruence subgroups of the Bianchi groups we do so by showing a large class of

normal subgroups has order Od,m.

Theorem 5.4.8. Let K be a Noetherian domain of Krull dimension one and suppose that

charK 6= 2, or 3. Let N SL2(K) be a congruence subgroup. Suppose that o(N) = K and

that l(N) = q∗. Then 12K 6 q∗.

Proof. Let q∗ = p∗1 . . . p∗t be a primary decomposition. Fix i ∈ 1, . . . , t and consider p∗i .

Let q∗0 = ∩j 6=ip∗j , p∗ = p∗i , and so we can apply (5.4.5) let q0 = p = K. Let Li = K/p∗i and

denote the maximal ideal of Li by mi. By an abuse of notation we also use mi to denote the

corresponding maximal ideal of K, ie r(p∗i ) = mi. Let ϕi : SL2(K)→ SL2(Li) be the natural

homomorphism. Let N0 = (N ∩ Γ(q∗0))Γ(p∗), N = NΓ(p∗). Let M0 = ϕi(N0), M = ϕi(N).

Then by lemma (5.4.5) l(M0) = 0, o(M) = p/p∗ = Li, and[E2(Li),M

]6 M0. Now by

theorem (5.4.7), as o(N) = K then o(M0) = o(M) = Li.

98

Suppose that 2 ∈ L∗i . Then, by (5.3.14) o(M0) = l(M0) unless |Li : mi| = 3, in which

case, by (5.3.15), mi 6 l(M0). If o(M0) = l(M0) then Li = 0 and so p∗i = K. If |Li : mi| = 3

and mi 6 l(M0) then Li is the field of 3 elements and 3K 6 p∗i .

Now suppose that 2 /∈ L∗i . If |Li : mi| > 3 then, by (5.3.14), l(M0) = o(M0) and so

p∗i = K. If |Li : mi| = 3 then, by (5.3.15), mi 6 l(M0) and so Li is the field of 3 elements

and 3K 6 p∗i . If |Li : mi| = 2 and 2 6= 0 in Li then, by (5.3.22), 4o(M0) 6 l(M0) and so

4K 6 p∗i . If |Li : mi| = 2 and 2 = 0 in Li then 2 ∈ p∗i so 2K 6 p∗i . So, in all cases 12K 6 p∗i .

Now q∗ = p∗1 . . . p∗t =

⋂ti=1 p∗i > 12K. Hence result.

Note from the proof above that the 12K comes from ideals of index 2 and ideals of index

3. When one or both of these are not present we get a better bound.

Corollary 5.4.9. Let K be a Noetherian domain of Krull dimension one. Let N SL2(K)

be a congruence subgroup. Suppose that o(N) = K and l(N) = q∗. Then

1. If K has no ideals of index 2, then 3K 6 q∗.

2. If K has no ideals of index 3 then 4K 6 q∗.

3. If K has no ideals of index 2 or 3 then q∗ = K.

Corollary 5.4.10. Let K be a Noetherian domain of Krull dimension one and suppose that

charK = 0. Let N SL2(K) be a congruence subgroup which contains −I. Suppose that

o(N) = K, and l(N) = q∗. Then 6K 6 q∗.

Proof. This follows from (5.4.8) and (5.3.23).

Lemma 5.4.11. Let K be a Noetherian domain of Krull dimension one. Let p ∈ Z be a

rational prime. Then K has only finitely many maximal ideals of index p.

Proof. Let m K be a maximal ideal of index p. Now ∀x ∈ K xp − x ∈ m. So consider the

ideal

q =∑x∈K

(xp − x)K

99

So q 6 m. Now, by (5.1.9), q = p1 . . . pt, where pi is primary, r(pi) = mi, a maximal

ideal; mnii 6 pi, for some ni; and pi + pj = K, for i 6= j. Suppose that m is distinct

from mi, i = 1, . . . , t. Now mnii 6 pi so K = mni

i + m 6 pi + m. So ∀i pi + m = K, so

m + q = m + (p1 . . . pt) = K. But q 6 m. Contradiction. So m = mi, some i = 1, . . . , t.

Hence result.

In light of this lemma make the following

Definition. Let K be a Noetherian domain of Krull dimension one. Let p ∈ Z be a rational

prime. Let

Mp =⋂m K : |K : m| = p

and if K has no ideals of index p letMp = K.

Lemma 5.4.12. Let R be a commutative ring of characteristic 2. Let q1, q2 R. Then

(q1q2)(4) = q

(4)1 q

(4)2 .

Proof. First recall that qi =< α4 : α ∈ qi > and q(4)1 q

(4)2 =< α4β4 : α ∈ q1, β ∈ q2 >. Let

α ∈ q1, and β ∈ q2 so α4β4 = (αβ)4 ∈ (q1q2)(4). So q

(4)1 q

(4)2 6 (q1q2)

(4).

Conversely let γ ∈ q1q2, so γ =∑αiβiri where αi ∈ q1, βi ∈ q2, ri ∈ R. So γ4 =

(∑αiβiri)

4 =∑α4

iβ4i r

4i as R is of characteristic 2. So γ4 ∈ q

(4)1 q

(4)2 . Hence result.

Recall that if charK = p, some prime p then any other rational prime, q is a unit in K.

Using this fact we get the following

Theorem 5.4.13. Let K be a Noetherian domain of Krull dimension one and suppose that

charK = p, for some prime p. Let N SL2(K) be a congruence subgroup. Suppose that

o(N) = K and l(N) = q∗. Then

1. If charK > 3 then q∗ = K.

2. If charK = 3 then M3 6 q∗ and if K has no ideals of index 3 then q∗ = K.

3. If charK = 2 then M22 6 q∗ and if K has no ideals of index 2 then q∗ = K.

100

Proof. Part 1 follows from (5.4.8). For part 2 consider the setup in the proof of theorem

(5.4.8). As charK = 3 then 2 is a unit in K and Li so, by (5.3.14), o(M0) = l(M0) unless

|Li : mi| = 3, in which case, by (5.3.15), mi 6 l(M0). If o(M0) = l(M0) then p∗i = K. If

|Li : mi| = 3 and mi 6 l(M0) then mi 6 p∗i . The result follows.

For part 3 we use the same setup as theorem (5.4.8) but in this case, as charK = 2 then

2 /∈ L∗i . Now if |Li : mi| > 3 then p∗i = K, as before. Also |Li : mi| 6= 3 for if |Li : mi| = 3

then 3 ∈ mi but as charK = 2 then 3 is a unit. Now suppose that |Li : mi| = 2, as 2 = 0 we

have, by (5.3.24) m(2)i 6 l(M0) and so m

(2)i 6 p∗i . The result follows.

We now deal with case of a normal congruence subgroup of order di stinct from K.

Theorem 5.4.14. Let K be a Noetherian domain of Krull dimension one and suppose that

charK 6= 2, or 3. Let N SL2(K) be a congruence subgroup. Suppose that o(N) = q 6= K,

and l(N) = q∗. Then 48q 6 q∗.

Proof. Let q∗ = p∗1 . . . p∗t be a primary decomposition. Then, by (5.1.10) q = p1 . . . pt, where

p∗i 6 pi. Fix i ∈ 1, . . . , t and consider p∗i . Let q∗0 = ∩j 6=ip∗j , p∗ = p∗i , q0 = ∩j 6=ipj, and

let p = pi. Let Li = K/p∗i and let ϕi : SL2(K) → SL2(Li) be the natural homomorphism.

Let N0 = (N ∩ Γ(q∗0))Γ(p∗), N = NΓ(p∗). Let M0 = ϕi(N0), M = ϕi(N). Then by lemma

(5.4.5) l(M0) = 0, o(M) = p/p∗, and[E2(Li),M

]6 M0.

First suppose that |Li : mi| > 2, so by theorem (5.4.7) o(M0) = o(M). Now suppose that

2 ∈ L∗i . So, by (5.3.7), and (5.3.14), o(M0) = l(M0) = 0 unless |Li : mi| = 3 and o(M0) = Li,

in which case, by (5.3.15), mi 6 l(M0). If o(M0) = l(M0) then pi = p∗i . If |Li : mi| = 3

and o(M0) = Li then mi 6 p∗i ie 3pi 6 p∗i . Now suppose that 2 /∈ L∗i . If o(M0) = Li

and |Li : mi| > 3 then, by (5.3.14), l(M0) = Li and so p∗i = K. Also |Li : mi| 6= 3 for if

|Li : mi| = 3 then 3 ∈ mi and, as 2 ∈ mi, 1 = 3−2 ∈ mi. Suppose that o(M0) 6= Li and 2 6= 0

in Li. Then as |Li : mi| > 2, by (5.3.12), 2o(M0) 6 l(M0) = 0, and so 2pi 6 p∗i . Finally if

2 = 0 in Li then 2 ∈ p∗i , so 2pi 6 p∗i .

Now suppose that |Li : mi| = 2. So if o(M) = Li ( ie pi = K ) then, by (5.4.7),

o(M0) = o(M) and if o(M) 6= Li ( ie pi 6= K ) then, by (5.4.7), mio(M) 6 o(M0). Now

|Li : mi| = 2 so 2 /∈ L∗i and 2 6= 0. If o(M) = Li then, by (5.4.7), o(M) = o(M0) and, by

101

(5.3.22), 4o(M0) 6 l(M0) so 4pi 6 p∗i . If o(M) 6= Li ( and so o(M0) 6= Li ) then, by (5.4.7),

mio(M) 6 o(M0) and, by (5.3.6), 8o(M0) 6 l(M0) so 16pi 6 p∗i . If 2 = 0 in Li then 2 ∈ p∗i ,

so 2pi 6 p∗i .

So in all cases 48pi 6 p∗i and so, as before 48q 6 q∗.

Again, when ideals of index 2 or 3 are not present we can improve the bound

Corollary 5.4.15. Let K be a Noetherian domain of Krull dimension one and suppose that

charK 6= 2, or 3. Let N SL2(K) be a congruence subgroup. Suppose that o(N) = q 6= K,

and l(N) = q∗. Then

1. If K has no ideals of index 2 then 6q 6 q∗. In particular if 6 is a unit then q∗ = q.

2. If K has no ideals of index 3 then 16q 6 q∗.

3. If K has no ideals of index 2 or 3 then 2q 6 q∗.

Example 5.4.1. Let K = Z[

16

]. So charK = 0, and 6 is a unit in K further, K is a

Noetherian domain of Krull dimension one, in fact K is a Dedekind domain of arithmetic

type. Recall the material in the section on number theory (1.3). Let p ∈ Z be prime and

define

vp : Q→ R

as follows. Let x ∈ Q, so x = pα ab

where p - a, and p - b. Then let vp(x) = p−α. Let v∞

denote the usual Archimedean valuation and let S = v∞, v2, v3 and form OS as indicated

in section (1.3). We claim that K = OS and so is a Dedekind domain of arithmetic type.

Let x ∈ OS, so x = a/b where a, b ∈ Z are coprime. Suppose that a prime p 6= 2, 3 divides b,

b = pγb′, γ > 0, p - b′ say. Now a and b are coprime so p - a, so x = p−γ ab′, so vp(x) = pγ > 1,

as γ > 0, so x /∈ OS. Contradiction. Hence OS ⊆ K. Now consider a2α3β ∈ K and let p 6= 2, 3

be prime. So vp(a

2α3β ) = vp(pγa′

2α3β ) = p−γ 6 1, as γ > 0, so a2α3β ∈ OS. So K ⊆ OS and so K

is a Dedekind domain of arithmetic type.

102

Theorem 5.4.16. Let K be a Noetherian domain of Krull dimension one and suppose that

charK = p, for some prime p. Let N SL2(K) be a congruence subgroup. Suppose that

o(N) = q 6= K, and l(N) = q∗. Then

1. If charK > 3 then q∗ = q.

2. If charK = 3 then qM3 6 q∗.

3. If charK = 2 then q(4)M(4)2 6 q∗.

Proof. Part 1 is obvious. For part 2 we use the same set up as before. As charK = 3

then 2 ∈ K∗ and so, by examining the above proofs it is obvious that pi = p∗i unless

|Li : mi| = 3 and o(M0) = Li, in which case, by (5.3.15), mi 6 l(M0). In this case mi 6 p∗i .

Hence q∗ = p∗1 . . . p∗t =

⋂p∗i > qM3.

Now suppose that charK = 2 so that 2 = 0 and 2 /∈ L∗i . Again we use the usual setup.

First suppose that |Li : mi| > 3 so, by theorem (5.4.7), o(M0) = o(M). If o(M0) = Li

( ie pi = K ) then, by (5.3.14), l(M0) = o(M0) and so p∗i = pi = K. If o(M0) 6= Li (ie

pi 6= K) then, by (5.3.13) o(M0)(2) 6 l(M0) and so p

(2)i 6 p∗i . Now |Li : mi| 6= 3 for if

|Li : mi| = 3 then 3 ∈ mi and 3 is a unit. Now suppose that |Li : mi| = 2. If o(M) = Li (ie

pi = K) then, by (5.4.7), o(M0) = o(M) = Li, so by (5.3.24), m(2)i 6 l(M0), so m

(2)i 6 p∗i .

If o(M) 6= Li (so o(M0) 6= Li) then, by (5.4.7), mio(M) 6 o(M0) and so, as o(M0) 6= Li,

by (5.3.8) o(M0)(4) 6 l(M0) so, applying (5.4.12), m

(4)i p

(4)i = (mipi)

(4) 6 p∗i . It follows that

q∗ =⋂

p∗i 6M(4)2

⋂p

(4)i = q(4)M(4)

2 (again using (5.4.12)).

In [57] Mason dealt with the case of a normal subgroup of SL2(L) where L is a com-

mutative local ring with principal maximal ideal m of index 2. This meant that in [58] he

was only able to derive a relationship between the order and level of a normal congruence

subgroup of SL2 over a Dedekind domain. By dropping this condition in section (5.3) we

were able to generalize Mason’s work to a normal congruence subgroup of SL2 over any

Noetherian domain of Krull dimension one. For comparison we state Mason’s results

103

Theorem. ([58] theroem 3.6) Let A be a Dedekind domain of arithmetic type (see section

(1.3)) and suppose that A is contained in a number field and is not totally imaginary. Let

N SL2(A) be a congruence subgroup. Suppose that o(N) = q, and l(N) = q∗ 6= 0. Then

1. If A has no ideals of index 2 or 3 and 2 is unramified in A then q = q∗.

2. If A has no ideals of index 2 and 2 is unramified in A then 3q 6 q∗.

3. If A has no ideals of index 2 or 3 then 2q 6 q∗.

4. If A has no ideals of index 2 then 6q 6 q∗.

5. If A has no ideals of index 3 then 4q 6 q∗.

6. Otherwise, 12q 6 q∗.

Theorem. ([58] theorem 3.14) Let A be a Dedekind domain of arithmetic type and suppose

that A is contained in a function field in one variable over a finite field k. Let N SL2(A)

be a congruence subgroup. Suppose that o(N) = q, and l(N) = q∗ 6= 0. Then

1. If charK > 3 or |k| = 3α, where α > 1 then q = q∗.

2. If |k| = 3 then qM3 6 q∗.

3. If |k| = 2α, where α > 1 then q2 6 q∗.

4. If |k| = 2 then q2M22 6 q∗.

So when N has order K we get the same results as Mason. However when N has order

distinct for K our results are not as good as Mason’s. It is clear from comparing the above

with our theorems that the problems arise when K has an ideal of index 2 and when the

characteristic is 2. It is not known whether our results are best possible or not. Our results

are definitely an extension of Mason’s for consider the following argument

Lemma 5.4.17. Let L be a commutative local ring with principal maximal ideal m. Then if

q L such that mi+1 6 q 6 mi then q = mi+1, or mi.

104

Proof. Suppose that m = αL, so mi = αiL. Let k = L/m, so mi+1/mi is a k-vector space.

Let x + mi+1 ∈ mi/mi+1, where x ∈ mi, so x = αiy, where y ∈ L. Let y = y + m ∈ k. So

y(αi + mi+1) = x + mi+1. So mi/mi+1 is generated by αi + mi+1 as a k-vector space and so

is one dimensional.

Suppose that mi+1 6 q 6 mi. So q/mi+1 is a subspace of mi/mi+1. So, as mi/mi+1 is one

dimensional q = mi, or mi+1.

Lemma 5.4.18. Let L be a commutative local ring with principal maximal ideal m. Suppose

that⋂∞

i=1 mi = 0. Then ∀ 0 6= q L, q = mα, some α > 1.

Proof. ∃α ∈ N such that q 6 mα, otherwise q = 0. Consider q + mα+1. By (5.4.17),

q + mα+1 = mα+1, or mα. If q + mα+1 = mα+1 then q 6 mα+1, contradicting the maximality

of α. So q + mα+1 = mα. Now let M = mα/q, so M is a finitely generated L-module. Then

mM = (q + mα+1)/q = mα/q = M . So, by Nakayama’s lemma ( see [2] page 21 ), M = 0, ie

q = mα.

Now, again consider O = O3,2 = Z + i√

3Z. Let m = (2, 1 + i√

3). Then, as in example

(5.1.2), m is a maximal ideal of O of index 2 and m2 < 2O < m. Let L = O/mn, any n > 2,

so in L the ideal 2L is not a power of the maximal ideal and so L has a maximal ideal of

index 2 which is not principal. In a Dedekind domain ideals are 32-generated. That is let q

be an ideal in a Dedekind domain and let x ∈ q, then ∃y ∈ q such that q is generated by

x and y. As a consequence of this every local image of a Dedekind domain has a principal

maximal ideal.

Example 5.4.2. In this chapter we have derived a relationship between the order and level

of a normal congruence subgroup. What about non-normal congruence subgroups? Here

unfortunately there is no relationship. Let d, p ∈ Z such that p is a rational prime and

d > 0 is square free. Consider the groups SL2(Od,p). By (3.2.1), SL2(Od,p) 6 SL2(Od),

and clearly SL2(Od, pOd) 6 SL2(Od,p) so SL2(Od,p) is a non-normal congruence subgroup

of SL2(Od). However A, T ∈ SL2(Od,p), so o(SL2(Od,p)) = Od and Od,p/pOd∼= Fp, so

l(SL2(Od,p)) = pOd.

105

Chapter 6

Non-congruence subgroups of the

Bianchi groups

We now apply the results of the previous chapter to the Bianchi groups PSL2(Od,m). We take

two view points. First we look at the growth of non-congruence subgroups. Then we look at

normal subgroups of small index, showing that all but finitely many normal congruence sub-

groups are of index divisible by 6. Further we classify the normal congruence subgroups with

torsion in PSL2(Od), d = 7, 11, 19, and PSL2(O1,2) and the normal congruence subgroups of

index not divisible by 6 in PSL2(O2) and PSL2(O3,2). Normal subgroups of PSL2(Od,m) of

index not divisible by 6 all have torsion. In [25] Fine shows that for d = 2, 7, 11 all principal

congruence subgroups of PSL2(Od) are torsion free with the exception of PSL2(O2, ωO2).

Let xi ∈ G, we use the notation N(xi) to denote the normal closure of the elements xi in G.

6.1 Counting finite index subgroups

In this section we introduce some basic notions about counting subgroups of finite index.

The material has been lifted from the survey article of Lubotzky [46]. Let G be any group.

Let an(G) denote the number of subgroups of G of index exactly n. Ideally we would like to

get a formula for an(G). The first attempt to do this was by M. Hall in 1949 [35] in which

106

a recursive formula for the number of subgroups of finite index in a free group of finite rank

was given. This method was extended and simplified by Dey [20] and Wohlfahrt [92].

Proposition 6.1.1. [46] Let tn(G) be the number of transitive permutation representations

of G on n objects. Then

an(G) =tn(G)

(n− 1)!

Proof. Let H 6 G be a subgroup of index n. G acts in an obvious way on G/H. Identify

G/H with the set 1, . . . , n such that H is identified with 1. There are (n−1)! ways of doing

this. Each of these gives rise to a homomorphism ϕ : G→ Sn such that ϕ(G) is transitive and

Stab(1) = γ ∈ G : ϕ(γ)(1) = 1 = H. Conversely every transitive permutation representa-

tion of G on n objects give a subgroup Stab(1) of index n. Thus (n− 1)!an(G) = tn(G).

Let hn(G) denote the number of homomorphisms from G to Sn.

Lemma 6.1.2. [46]

hn(G) =n∑

k=1

(n− 1

k − 1

)tk(G)hn−k(G)

Proof. Suppose that the orbit of 1 is of length k, 1 6 k 6 n. There are

(n− 1

k − 1

)ways

of choosing such an orbit, tk(G) ways of acting on it, and hn−k(G) ways of acting on its

complement.

Theorem 6.1.3. [46] Let G be any group then

an(G) =1

(n− 1)!hn(G)−

n−1∑k=1

1

(n− k)!hn−k(G)ak(G).

107

Proof.

hn(G) =n∑

k=1

(n− 1

k − 1

)tk(G)hn−k(G) by (6.1.2)

=n∑

k=1

(n− 1

k − 1

)(k − 1)!ak(G)hn−k(G) by (6.1.1)

=n∑

k=1

(n− 1)!

(n− k)!ak(G)hn−k(G)

= (n− 1)!an(G) +n−1∑k=1

(n− 1)!

(n− k)!ak(G)hn−k(G).

Therefore

an(G) =1

(n− 1)!hn(G)−

n−1∑k=1

1

(n− k)!hn−k(G)ak(G).

Now consider the free group on r generators. Since each generator of Fr can be mapped

to one of n! elements of Sn we see that hn(Fr) = (n!)r. Thus

Theorem 6.1.4. [35] Let Fr be the free group on r generators. Then

an(Fr) = n(n!)r−1 −n−1∑k=1

[(n− k)!]r−1 ak(Fr).

From this it follows [71] that

Theorem 6.1.5.

an(Fr) ∼ n(n!)r−1

In [20] Dey extended this work to free products

Theorem 6.1.6. Let G = ∗ni=1Ai be a free product of groups. Let hin = hn(Ai). Then

an(G) =1

(n− 1)!

(r∏

i=1

hin

)−

n−1∑k=1

1

(n− k)!ak(G)

(r∏

i=1

hin−k

).

108

Proof. This follows from (6.1.2) and the fact that hn(G) =∏n

i=1 hin.

Recall that the Modular group PSL2(Z) = Z2 ∗ Z3. So we can apply the above theorem

from which we get

Proposition 6.1.7. [71]

an(PSL2(Z)) ∼ (12πe1/2)−1/2 exp

(n log n

6− n

6+ n1/2 + n1/3 +

log n

2

).

6.2 The growth of non-congruence subgroups in the

groups SL2(Od,m)

Recall the following results

Theorem. (2.2.3) There exists a surjective homomorphism

SL2(Od,m) −→ Fr

where r = r(d,m) the rank of the Zimmert Set and a, t lie in the kernel.

Theorem. (5.4.8) Let N SL2(Od,m) be a congruence subgroup of order Od,m. Then

SL2(Od,m, 12Od,m) 6 N .

Recall that Zimmert’s theorem was not best possible; r(5) = 1, but SL2(O5) F2.

So, as before, let ρ(d,m) denote the largest rank of a free quotient of SL2(Od,m). Let FK

denote the kernel of the map from SL2(Od,m) onto Fρ(d,m). Let N SL2(Od,m) containing

FK. As t ∈ FK then N has order Od,m. Therefore if N is a congruence subgroup it must

contain SL2(Od,m, 12Od,m). Observe that S 6 SL2(Od,m) is a congruence subgroup if and

only if coreS is a congruence subgroup, where by the core of S we mean the largest normal

subgroup contained in S. Thus

Proposition 6.2.1. Let S 6 SL2(Od,m) and suppose that FK 6 S. Then with finitely many

exceptions S is a non-congruence subgroup.

109

Using Newmans result (6.1.5) we get

Theorem 6.2.2. Asymptotically, the number of non-congruence subgroups in the group,

SL2(Od,m), of index precisely n is at least n(n!)ρ−1 where ρ = ρ(d,m) > r(d,m), the rank of

the Zimmert Set.

The vast number of non-congruence subgroups in the above theorem all come from the

largest free quotient, all have order Od,m and all contain elements of finite order. Are most

non-congruence subgroups like this? It would be interesting to know about the growth of

non-congruence subgroups of order q 6= Od,m, or of torsion free non-congruence subgroups.

We now take the opposite view and look at normal subgroups of the Bianchi groups

PSL2(Od,m) of small index with the aim of determining whether they are congruence or

non-congruence. First, we have to take a look at SL2(L) where L is a local image of Od,m.

6.3 SL2(L) where L is a local image of Od,m

Let L be a finite local homomorphic image of Od,m with maximal ideal of index 2. Then we

can apply (5.3.22) to get

Theorem 6.3.1. Let L be a finite local homomorphic image of Od,m with a maximal ideal

m of index 2. Let N SL2(L) be of order L. Then 4L 6 l(N).

We now classify all normal subgroups of SL2(L) with order L by means of the commutator

subgroup viz

Theorem 6.3.2. Let N SL2(L). Then o(N) = L⇔ SL2(L)′ 6 N .

Now o(SL2(L)′) = L, as

[A, T ] =

0 −1

1 0

,

1 1

0 1

=

1 −1

−1 2

so (⇐) is obvious. First we describe the ideals of Od,m of index 2.

110

Theorem 6.3.3. If m ≡ 0 (mod 2) then Od,m has exactly one ideal of index 2, namely

(2,mω). If m ≡ 1 (mod 2) then the ideals of index 2 are:

(2, 1 +mω) if d ≡ 1 (mod 4)

(2,mω) if d ≡ 2 (mod 4)

(2,mω), (2, 1 +mω) if d ≡ 7 (mod 8)

none if d ≡ 3 (mod 8)

Proof. The m ≡ 0 (mod 2) case follows from lemma (3.1.4). If m ≡ 1 (mod 2) then we use

the isomorphism given in lemma (3.1.5) ie Od,m/2Od,m∼= Od/2Od where mω ↔ ω. So we

can assume that m = 1. The generators of the maximal ideals can now be found using the

theorem on page 107 of [52]. Let

fω(X) =

X2 + d if d ≡ 1, 2 (mod 4)

X2 −X + d+14

if d ≡ 3 (mod 4)

We decompose fω (mod 2) then if gi is a factor of fω (mod 2), an ideal of index 2 in Od

is given by (2, gi(ω)) (see p. 107 of [52]). First, if d ≡ 3 (mod 8) then D ≡ 5 (mod 8), so

χK(2) = −1 (see chapter 3 for the definition) so Od/2Od is a field and Od,m has no ideals of

index 2. Now suppose that d ≡ 1, 2 (mod 4) so

fω(X) = X2 + d ≡

(X + 1)2 (mod 2) if d ≡ 1 (mod 4)

X2 (mod 2) if d ≡ 2 (mod 4)

So if d ≡ 1 (mod 4) then (2, 1 + mω) is the only ideal of index 2 in Od,m and if d ≡ 2

(mod 4) then (2,mω) is the only ideal of index 2 in Od,m. Now suppose d ≡ 7 (mod 8), so

(d+ 1)/4 ≡ 0 (mod 2), so fω(X) ≡ X2 −X ≡ X(X + 1) (mod 2). So Od,m has exactly two

ideals of index 2, namely (2,mω), and (2, 1 +mω).

We compute SL2(L)ab using a method developed by P. M. Cohn in [15] which we now

describe.

Theorem 6.3.4. Let L be a commutative local ring. Then

SL2(L)ab ∼= (L/M)+

111

where M is the additive subgroup of L generated by (u2 − 1)x : x ∈ L, u ∈ L∗ and 3(u+ 1)(v + 1) : u, v ∈ L∗,

and (L/M)+ is understood to be an additive group. The isomorphism is given by

E(x) =

x 1

−1 0

7−→ x− 3 (mod M)

Proof. This follows from [15] theorem 2, [14] theorem 4.1, and the fact that SL2(L) = E2(L)

[3].

First, ignore the case d ≡ 7 (mod 8) and m ≡ 1 (mod 2) and let L denote the local ring

Od,m/4Od,m.

Lemma 6.3.5. If m ≡ 0 (mod 2) then SL2(L)ab ∼= Z2 × Z4.

Proof. L = r +mωs : r, s = 0, 1, 2, 3, and it is easy to show that

L∗ = r +mωs : r ≡ 1 (mod 2). Let u ∈ L∗ then u2 − 1 = 0, or 2mω. Now 2mω(r +

mωs) = 0, or 2mω. So (u2 − 1)x = 0, 2mω. Let u, v ∈ L∗, it is trvial to show that

3(u+ 1)(v + 1) ∈ 0, 2mω. So M = 0, 2mω, so (L/M)+ ∼= Z2 × Z4. Hence result.

Similarly we can prove:

Lemma 6.3.6. Suppose that d ≡ 1, 2 (mod 4) and m ≡ 1 (mod 2). Then

SL2(L)ab ∼= Z2 × Z2.

Consider the following two matrices

T =

1 1

0 1

, U =

1 mω

0 1

Under Cohn’s map SL2(L) SL2(L)ab they behave as follows

T = E(−1)E(0)−1 7→ (−1− 3)− (0− 3) = −1

U = E(−mω)E(0)−1 7→ (−mω − 3)− (0− 3) = −mω

and so, from the proofs of the above lemmata we see that SL2(L)ab is generated by the images

of T and U ie if m ≡ 0 (mod 2) then SL2(L)ab ∼= Z2×Z4 =< U, T ;U2, T 4, TU = UT > and

112

if d ≡ 1, 2 (mod 4) and m ≡ 1 (mod 2) then SL2(L)ab ∼= Z2 × Z2 =< U, T ;U2, T 2, TU =

UT >. It follows from this that only one of the normal subgroups of SL2(L) of index 2

contains T . We will use this observation in the next section and we record it as a lemma.

Lemma 6.3.7. Suppose m ≡ 0 (mod 2) or d ≡ 1, 2 (mod 4) and m ≡ 1 (mod 2). Then

∃N SL2(L) of index 2 which does not contain T .

We now deal with the case d ≡ 7 (mod 8) and m ≡ 1 (mod 2). In this case Od,m has

two ideals of index 2, so Od,m/4Od,m is not local. Instead we let m be one of the ideals of

index 2 and L = Od,m/m3. We need to describe m3.

Lemma 6.3.8. If m = (2,mω) then m3 = (8,mω), (8, 4 +mω), or (8,±2 +mω).

Proof. First we calculate generators for m2. Clearly m2 = (4, 2mω, (mω)2). Now (mω)2 =

m(mω) −m2 d+14

. Now d ≡ 7 (mod 8) so (d + 1)/4 ≡ 0, 2 (mod 4) and m ≡ 1 (mod 2) so

m2 ≡ 1 (mod 4) so

(mω)2 =

mω (mod m2) if d ≡ 15 (mod 16)

2 +mω (mod m2) if d ≡ 7 (mod 16)

and 2(2 +mω) ≡ 2mω (mod 4) so m2 = (4,mω), or (4, 2 +mω).

Now m3 = mm2. First suppose that m2 = (4,mω), so d ≡ 15 (mod 16). So m3 =

(8, 2mω, (mω)2). Again (mω)2 = m(mω)−m2(d+ 1)/4. Now m ≡ 1 (mod 2) so m(mω) ≡

mω (mod m3). Since m2 ≡ 1 (mod 8) and (d + 1)/4 ≡ 0, 4 (mod 8), so m2(d + 1)/4 ≡

0, 4 (mod 8). So (mω)2 ≡ mω, or 4 + mω (mod m3). So m3 = (8,mω), or (8, 4 + mω).

Now suppose that m2 = (4, 2 + mω). So m3 = (8, 4mω, 2(2 + mω),mω(2 + mω)). Now

2×2(2+mω) ≡ 4mω, and we can show that (mω)2 ≡ ±2+mω. Also 2(±2+mω) ≡ 4+2mω.

So m3 = (8,±2 +mω). Hence result.

In a similar way we can prove

Lemma 6.3.9. If m = (2, 1 +mω) then m3 = (8,±1 +mω).

Lemma 6.3.10. Suppose that d ≡ 7 (mod 8) and m ≡ 1 (mod 2). Let mOd,m be an ideal

of index 2, so m = (2,mω), or (2, 1 +mω). Let L = Od,m/m3. Then SL2(L)ab ∼= Z4.

113

Proof. Suppose first that m = (2,mω), so m3 = (8,mω), (8, 4 + mω), or (8,±2 + mω).

Suppose r+mωs ∈ L, we can assume that 0 6 r, s 6 7. Then r+mωs ≡ r+mωs−s(mω) = r

(mod (8,mω)), or r+mωs ≡ r+mωs−s(4+mω) = r−4s (mod (8, 4+mω)), or r+mωs ≡

r + mωs − s(±2 + mω) = r ± 2s (mod (8,±2 + mω)). So L = 0, 1, 2, 3, 4, 5, 6, 7 ∼= Z8.

Applying Cohn’s method we see that SL2(L)ab ∼= Z4.

Now suppose that m = (2, 1 +mωs), so m3 = (8,±1 +mω). As above we can show that

L = 0, 1, 2, 3, 4, 5, 6, 7 and so SL2(L)ab ∼= Z4.

We now present the proof of theorem (6.3.2) as a series of lemmas.

Lemma 6.3.11. Suppose that d ≡ 7 (mod 8) and m ≡ 1 (mod 2). Let N SL2(L). Then

o(N) = L⇔ SL2(L)′ 6 N .

Proof. Let Γ denote SL2(L). Now o(Γ′) = L, so (⇐) is obvious. Conversely suppose that

o(N) = L and Γ′ N . So, by (5.3.17), |Γ : N | = 2α and Γ/N is non-abelian, so α > 3. So

Γ/N is a 2-group of order > 8, so ∃MΓ such that N 6 M and |Γ : M | = 8, and o(M) = L.

Now Γab = Z4, so Γ/M is non-abelian, so Γ/M ∼= Q8, or D8, the only non-abelian groups

of order 8 [68]. Now Qab8∼= Dab

8∼= Z2 × Z2, so (Γ/M)ab ∼= Z2 × Z2. But Γ Γ/M , so

Γab (Γ/M)ab, so Z4 Z2 × Z2. Contradiction. Hence result.

Unfortunately this technique does not work in the other cases. Instead we use the ob-

servation (5.3.18) that if N SL2(L) is of order L then N contains every matrix in SL2(L)

of order 3. Let N(3) denote the normal closure of all elements of order 3. We attempt to

show that SL2(L)′ = N(3). Note that because we have shown (5.3.22) that if o(N) = L

then SL2(L, 4L) 6 N we can work mod 4. Recall that SL2(L) = E2(L) and let

E12(x) =

1 x

0 1

, E21(y) =

1 0

y 1

.

So SL2(L) is generated by E12(x) and E21(y) where x, y = 1,mω. It is easy to compute that

[E12(x), E21(y)] =

1 + xy + (xy)2 −x2y

xy2 1− xy

114

we want to show that [E12(x), E21(y)] ∈ N(3) for (x, y) = (1, 1), (1,mω), (mω, 1), (mω,mω).

We make frequent use of lemma (5.3.20) without comment.

Lemma 6.3.12. [E12(1), E21(1)]3 = I.

Lemma 6.3.13. Suppose that d ≡ 1, 2 (mod 4) and m ≡ 0 (mod 2). Then SL2(L)′ = N(3).

Proof. In this case (mω)2 = −dm2 = 0 as m ≡ 0 (mod 2) and we are working mod 4. This

shows that [E12(mω), E21(mω)] = I. Now

[E12(1), E21(mω)] =

1 +mω −mω

0 1−mω

=

−1 1

−1 0

0 −(1−mω)

1 +mω −1

.

Similarly for [E12(mω), E21(1)].

Lemma 6.3.14. Suppose that d ≡ 3 (mod 4) and m ≡ 0 (mod 2). Then SL2(L)′ = N(3).

Proof. First note that if m ≡ 0 (mod 4) then (mω)2 = 0, and if m ≡ 2 (mod 4) then

(mω)2 = 2mω. This renders the m ≡ 0 (mod 4) case identical to the d ≡ 1, 2 (mod 4) cases

above. So suppose m ≡ 2 (mod 4). Then

[E12(1), E21(mω)] =

1−mω −mω

2mω 1−mω

and this equals −1 −1

1 0

2mω 1−mω

−(1 +mω) −1 + 2mω

∈ N(3)

Similarly [E12(mω), E21(1)] ∈ N(3). Now

[E12(mω), E21(mω)] =

1 + 2mω 0

0 1 + 2mω

and this equals −1 +mω −1

1 +mω −mω

−1 −1 +mω

1−mω 0

−1 −1 + 2mω

1 + 2mω 0

∈ N(3)

115

Lemma 6.3.15. Suppose that d ≡ 1 (mod 4) and m ≡ 1 (mod 2). Then SL2(L)′ = N(3).

Proof. In this case (mω)2 = −1, so

[E12(mω), E21(1)] =

mω 1

mω 1−mω

and this equals −1 1

−1 0

2 0 −mω

−mω −1

∈ N(3).

similarly [E12(mω), E21(1)] ∈ N(3). Also

[E12(mω), E21(mω)] =

1 −mω

mω 2

so [E12(mω), E21(mω)]3 = I. Hence result.

Lemma 6.3.16. Suppose that d ≡ 2 (mod 4) and m ≡ 1 (mod 2). Then T 2 ∈ N(3), and

so −I ∈ N(3).

Proof. By lemma (5.3.20) 0 1 +mω

3 +mω −1

∈ N(3)

and this equals 1 1 +mω

0 1

A

1 3 +mω

0 1

A−1

1 2 + 2mω

0 1

= T

1 mω

0 1

AT

1 2 +mω

0 1

A−1T−1

1 3 + 2mω

0 1

so, as AT,A−1T−1 ∈ N(3), and 4 = 0, we see that T 2 ∈ N(3).

Lemma 6.3.17. Suppose that d ≡ 2 (mod 4) and m ≡ 1 (mod 2). Then SL2(L)′ = N(3).

116

Proof. In this case (mω)2 = 2, so

[E12(1), E21(mω)] =

−1 +mω −mω

2 1−mω

and this equals −1 1−mω

1 +mω 0

2 + 2mω 1− 2mω

1 1 + 2mω

∈ N(3)

Similarly [E12(mω), E21(1)] ∈ N(3). However

[E12(mω), E21(mω)] =

−1 2mω

2mω −1

and this equals

−I

2mω 1

−1 + 2mω −1 + 2mω

−1 −1 + 2mω

1 + 2mω 0

∈ N(3)

Hence result.

We have now proved theorem (6.3.2). From the above we can get an improved lower

bound for the level in some cases

Theorem 6.3.18. Suppose that d ≡ 1, 2 (mod 4) and m ≡ 1 (mod 2). Let L = Od,m/4Od,m.

Let N SL2(L), o(N) = L. Then 2L 6 l(N).

6.4 Non-congruence subgroups of small index in the

Bianchi groups

Here we need to be very clear about the differences between SL2 and PSL2 so we restate

the relevant definitions. Let O = Od,m be an order in an imaginary quadratic number field.

Let 0 6= q O.

SL2(O, q) = M ∈ SL2(O) : M ≡ I (mod q)

117

SL2(O, q) is the principal congruence subgroup of level q. Let S 6 SL2(O). We say that S

is a congruence subgroup if SL2(O, q) 6 S for some q O. If q is the largest O-ideal such

that SL2(O, q) 6 S we say that S is of level q. Otherwise S is a non-congruence subgroup.

Let

ϕ : SL2(O) −→ PSL2(O)

be the natural homomorphism. Let q O, then

PSL2(O, q) = ϕ(SL2(O, q))

PSL2(O, q) is the principal congruence subgroup of level q. Let S 6 PSL2(O). We say that

S is a congruence subgroup if PSL2(O, q) 6 S, for some q O. If q is the largest O-ideal

such that PSL2(O, q) 6 S we say that S is of level q. Otherwise S is a non-congruence

subgroup.

Lemma 6.4.1. Let N SL2(O) such that −I ∈ N . Let N = ϕ(N). Then

N is a congruence subgroup ⇔ N is a congruence subgroup

Further N and N have the same level.

Proof. Suppose that N is a congruence subgroup. So SL2(O, q) 6 N , some 0 6= q O. So

PSL2(O, q) = ϕ(SL2(O, q)) 6 ϕ(N) = N ie N is a congruence subgroup.

Conversely suppose that N is a congruence subgroup. So PSL2(O, q) 6 N , some q O.

Suppose that SL2(O, q) N , so ∃M1 ∈ SL2(O, q) such that M1 /∈ N . But ϕ(M1) ∈

ϕ(SL2(O, q)) = PSL2(O, q) 6 N = ϕ(N). So ∃M2 ∈ N such that ϕ(M1) = ϕ(M2). So

M1M−12 ∈ kerϕ = ±I, so M1 = ±M2. Now M1 6= M2, as M1 /∈ N , and M2 ∈ N .

So M1 = −M2 = −IM2, but −I ∈ N and M2 ∈ N , so M1 ∈ N . Contradiction. So

SL2(O, q) 6 N ie N is a congruence subgroup.

So the normal congruence subgroups of SL2(O) of level q that contain −I are in one to

one correspondence with the normal congruence subgroups of PSL2(O) of level q and they

clearly have the same index. So we attempt to classify the normal congruence subgroups of

118

PSL2(O) of index n by classifying the normal congruence subgroups of SL2(O) of index n

which contain −I.

Lemma 6.4.2. Let O = Od,m and suppose that N SL2(O). Let n ∈ N. Then if T n, and

Un ∈ N then E2(O, nO) 6 N .

Proof. Recall that

E2(O, nO) =< I + αeij : α ∈ nO, i 6= j >E2(O) .

Let α ∈ nO, so α = n(z1 +mωz2), where z1, z2 ∈ Z. Now

I + αe12 =

1 nz1 + nmωz2

0 1

= (T n)z1(Un)z2 ∈ N.

Similarly I + αe21 ∈ N . So, as N is normalized by E2(O) then E2(O, nO) 6 N .

Recall the following results

Theorem. (3.2.2) Let O = Od,m. Let N SL2(O) be of index n and suppose that 6 - n.

Then N is of order O.

Theorem. (5.4.10) Let O = Od,m. Let N SL2(O) be a congruence subgroup of order O

that contains −I. Then SL2(O, 6O) 6 N .

So that any normal congruence subgroup of SL2(Od,m) of index not divisible by 6 and

containing −I must contain SL2(Od,m, 6Od,m). This gives us a method to classify the normal

congruence subgroups of SL2(Od,m) of index not divisible by 6. In fact we can be more

precise.

Lemma 6.4.3. Let O = Od,m. Let N SL2(O) be of index n that contains −I and suppose

that 6 - n. Then

1. If 2 | n and 3 - n then N is a congruence subgroup ⇔ SL2(O, 2O) 6 N .

119

2. If 2 - n and 3 | n then N is a congruence subgroup ⇔ SL2(O, 3O) 6 N .

3. If (6, n) = 1 then N is a non-congruence subgroup.

Proof. Now, N is a normal congruence subgroup of index n, so using (6.4.2) and (5.2.5),

SL2(O, nO) 6 N . Applying Wohlfahrts theorem (5.2.3) we see that

SL2(O, nO)SL2(O, 6O) = SL2(O, gcd(6, n)O) 6 N from which the result follows.

Let Nd,m(n) denote the number of normal congruence subgroups of PSL2(Od,m) of index

precisely n. So by the above lemma if (6, n) = 1 then Nd,m(n) = 0. If SL2(Od,m) has

an infinite cyclic quotient then it has a normal subgroup of every index. In contrast the

Modular group has only two normal subgroups which are not of index 6k, some k ∈ N ([70]

theorems 8.6 and 8.7). However by (6.4.3), with only finitely many exceptions, every normal

congruence subgroup of SL2(Od,m) is of index 6k, some k ∈ N. We need the following lemma

Lemma 6.4.4. Let K be a Noetherian domain of Krull dimension one. Let q1, q2 K such

that q1 + q2 = K. Then

SL2(K/q1q2) ∼= SL2(K/q1)× SL2(K/q2)

Proof. By the Chinese remainder theorem K/q1q2∼= K/q1 ×K/q2 where the isomorphism

is given by x+ q1q2 7→ (x+ q1, x+ q2) = (x1, x2). Define ϕ : SL2(K/q1q2)→ SL2(K/q1)×

SL2(K/q2) by a b

c d

7→ a1 b1

c1 d1

,

a2 b2

c2 d2

where a = a+q1q2 7→ (a1, a2) etc. It is simple, if tedious to check that ϕ is an monomorphism

(indeed it is true for any commutative ring). To show that ϕ is onto we need Wohlfahrt’s

theorem (5.2.5). By Wohlfahrt SL2(K, q1)SL2(K, q2) = SL2(K, q1 + q2) = SL2(K). Also

SL2(K, q1) ∩ SL2(K, q2) = SL2(K, q1q2). So in the group SL2(K/q1q2) = G we have two

normal subgroups H, K, say such that HK = G and H ∩K = 1. So by [79] theorem 4.1 we

have G ∼= H ×K. Hence result.

Lemma 6.4.5. Let O = Od,m. Let n ∈ N and suppose that 2 | n and 3 - n. Then

120

1. If O has no ideals of index 2 then Nd,m(n) = 0.

2. If O has an ideal of index 2 then Nd,m(n) =

3 if n = 2

1 if n = 4

0 else

Proof. Let N SL2(O) be a congruence subgroup of index n that contains −I. Then by

(6.4.3) SL2(O, 2O) 6 N . There are 3 cases, depending on how many ideals of index two O

has. If O has no ideals of index 2 then SL2(O)/SL2(O, 2O) ∼= SL2(F4) but −I ∈ N and

PSL2(F4) is simple, so Nd,m(n) = 0. Now suppose that O has exactly one ideal of index 2.

Let L denote the local ring O/2O, so SL2(O)/SL2(O, 2O) ∼= SL2(L). Now corresponding

to N is N SL2(L), o(N) = L. So, by (6.3.2), SL2(L)′ 6 N . Now −I ∈ N so, by (5.3.19),

(6.3.5) and (6.3.6), SL2(O)/N is a factor of Z2 × Z2 which, by (3.4.10), has exactly 3 non-

trivial normal subgroups, all of index 2. The result follows in this case. Finally suppose

that O has two ideals of index 2. In this case, by (6.4.4), SL2(O)/SL2(O, 2O) ∼= S3 × S3,

so |SL2(O) : SL2(O, 2O)| = 2232. So N is of index 2 or 4, so SL2(O)/N is abelian. Now

(SL2(O)/SL2(O, 2O))ab ∼= Sab3 ×Sab

3∼= Z2×Z2 and so the answer is as the previous case.

Lemma 6.4.6. Let O = Od,m. Let n ∈ N and suppose that 2 - n and 3 | n. Then

1. If O has no ideals of index 3 then Nd,m(n) = 0.

2. If O has exactly one ideal of index 3 then Nd,m(n) =

1 if n = 3

0 else

3. If O has exactly two ideals of index 3 then Nd,m(n) =

4 if n = 3

1 if n = 9

0 else

Proof. Let N SL2(O) be a congruence subgroup of index n that contains −I. Then

by (6.4.3), SL2(O, 3O) 6 N . First suppose that O has no ideals of index 3. Then

SL2(O)/SL2(O, 3O) ∼= SL2(F9) but −I ∈ N and PSL2(F9) is simple, so Nd,m(n) = 0.

Now suppose that O has exactly one ideal of index three, m, say and let L denote the local

121

ring O/3O. So, by passing to SL2(L) and applying (5.3.15) we see that SL2(O,m) 6 N .

Now SL2(O)/SL2(O,m) = SL2(F3), −I ∈ N and, by [79] exercise 8.11, PSL2(F3) ∼= A4

which has exactly one normal subgroup, which is of index 3. Hence result in this case.

Finally suppose that O has exactly two ideals of index three, mi, i = 1, 2, say. Now, by

(6.4.4),SL2(O)

SL2(O, 3O)∼= SL2(F3)× SL2(F3)

and |SL2(F3)| = 24 (this follows from [79] theorem 8.8), so |SL2(O) : SL2(O, 3O)| = 2632.

So N is of index 3 or 9 and so SL2(O)/N is abelian. Now SL2(F3)ab ∼= Z3 so(

SL2(O)

SL2(O, 3O)

)ab

∼= Z3 × Z3

and, −I2 = I so −I ∈ SL2(O/3O)′. Now, by (3.4.10), Z3 × Z3 has exactly 4 non-trivial

normal subgroups, all of index 3. Hence result in this case.

In [31] Grunewald and Schwermer determine the minimum index of a non-congruence

subgroup of a Bianchi group. Let ncs(d) denote the minimum index of a non-congruence

subgroup of PSL2(Od). Their main result is

Theorem.

ncs(d) =

5 if d = 1

4 if d = 2

22 if d = 3

3 if d = 7

2 else

In the Modular group the least index of a non-congruence subgroup if 7 ([83] theorem

5.4). There is considerable overlap with Grunewald and Schwermer’s theorem and what

follows but our techniques are completely different. First recall the following

Theorem.

ϕ :SL2(Od,m)

U2(Od,m) Fs

where s = r − 1, and r = r(d,m), the rank of the Zimmert Set and U ∈ kerϕ.

122

Theorem 6.4.7. If r = r(d,m) > 2 then SL2(Od,m) has a normal non-congruence subgroup

containing −I of every index.

Proof. Since r > 2 we have

ϕ :SL2(Od,m)

U2(Od,m) Z

and clearly −I ∈ kerϕ. Suppose that N SL2(Od,m) such that kerϕ 6 N , so T, U ∈ N , so

l(N) = Od,m. So, if N is a congruence subgroup then N = SL2(Od,m). Thus SL2(Od,m) has

a normal non-congruence subgroup containing −I of every index.

Corollary 6.4.8. If r = r(d,m) > 2 then PSL2(Od,m) contains a normal non-congruence

subgroup of every index.

Not included in the above theorem are the cases d = 1, 2, 3, 7, 11, 19, and m = 1 and

(d,m) = (1, 2), (3, 2). We are unable to deal with the cases (d,m) = (1, 1), (3, 1) but we take

a closer look at the others later. Recall that it is conjectured that these are the only values

of (d,m) such that PSL(Od,m) does not have a free non-abelian quotient.

All the non-congruence subgroups produced above contain the Zimmert kernel and so

have torsion. We now show that in any Bianchi group the number of normal congruence

subgroups with torsion is finite. Recall

Theorem. (1.1.8) I 6= M ∈ SL2(Od,m) is of finite order if and only if trM = 0, or ±1.

We now show that the order of a subgroup with torsion is severely restricted. Let M ∈

SL2(Od,m) and recall that the order, o(M), of the matrix M =

a b

c d

is the ideal of

Od,m generated by b, c, a− d. The order of any subgroup containing M contains o(M).

Lemma 6.4.9. Let M ∈ SL2(Od,m). Then if trM = 0 then 2 ∈ o(M).

Proof. a + d = 0, so a = −d. Now a − d = 2a ∈ o(M). Also 1 = ad − bc = −a2 − bc, so

1 + a2 ∈ o(M). Thus 2a2 + 2 ∈ o(M) and, because 2a ∈ o(M) we have 2a2 ∈ o(M). Hence

2 ∈ o(M).

Lemma 6.4.10. Let M ∈ SL2(Od,m). Then if trM = ±1 then 3 ∈ o(M).

123

Proof. First suppose that trM = −1, so a + d = −1 and a = −1 − d. Now a − d =

2a + 1 ∈ o(M). Also 1 = ad − bc = −a − a2 − bc, so a2 + a + 1 = −bc ∈ o(M). So

2a2 + 1 = 2(a2 + a+ 1)− (2a+ 1) ∈ o(M). Further 1− a = 2a2 + 1− (2a+ 1)a ∈ o(M). So

3 = 2(1− a) + 2a+ 1 ∈ o(M). Similarly if trM = 1.

Theorem 6.4.11. There are only finitely many normal congruence subgroups with torsion

in SL2(Od,m).

Proof. Let N SL2(Od,m) and suppose that N has torsion and is a congruence subgroup.

So N contains a matrix M , say with trM = 0, or ±1. So 2, or 3 ∈ o(N), so 6 ∈ o(N). Now

N is a normal congruence subgroup so, by (5.4.14), SL2(Od,m, 288Od,m) 6 N and there are

only finitely many such subgroups.

Example 6.4.1. There are infinitely many non-normal congruence subgroups with torsion

in SL2(Od,m). The groups SL2(Od,mn), for n ∈ N, are an example.

Recall that it is conjectured that PSL2(Od,m) has a free non-abelian quotient for all values

of (d,m) except for (1, 1), (2, 1), (3, 1), (7, 1), (11, 1), (19, 1), (1, 2), and (3, 2). The normal

subgroups of PSL2(O1) were studied in [26] where, in particular, the normal subgroups of

index < 60 were classified and it was shown that for a wide collection of n, PSL2(O1) had no

normal subgroups of index n. The normal subgroups PSL2(O3) were studied in [1] where the

normal subgroups of index < 960 were classified and shown to all be congruence subgroups.

We now take a closer look at the others.

6.5 The groups PSL2(Od), d = 2, 7, 11 and PSL2(O3,2) and

PSL2(O1,2)

In this section we deal exclusively with PSL2(Od,m), which we denote by PSL2(Od,m), or

PSL2(Od), if m = 1. We look first at the groups PSL2(Od), for d = 2, 7, 11. In [25]

section 4.5.3, Fine attempts to classify the normal subgroups of PSL2(Od), d = 2, 7, 11

124

but his classification is incomplete. We first of all correct his errors. Recall the following

presentations

Theorem. ([25] theorem 4.3.1)

PSL2(O2) =< a, t, u; a2, (at)3, (u−1aua)2, [t, u] >

PSL2(O7) =< a, t, u; a2, (at)3, (u−1auat)2, [t, u] >

PSL2(O11) =< a, t, u; a2, (at)3, (u−1auat)3, [t, u] >

We now state and prove the correct version of Fine’s results

Theorem 6.5.1. Let NPSL2(O2) be of index n and suppose that 6 - n. Then if (n, 6) = 1

N = N(a, t, un)

if 2 | n and 3 - n then N is one of

N(a, t, un), N(at, un/2), N(at, aun/2)

if 2 - n and 3 | n then N is one of

N(a, t, un), N(a, un/3), N(a, tun/3), N(a, t2un/3)

Proof. Suppose that (n, 6) = 1 so a = at = u−1aua = 1, so a = t = 1, so PSL2(O2)/N is

a factor of Z. SoN = N(a, t, un). Suppose that 2 | n and 3 - n, so at = 1, so PSL2(O2)/N is a

factor of< a, u; a2, au = ua >∼= Z2×Z. So, by (3.4.10), N = N(a, t, un), N(at, un/2), N(at, aun/2).

Now suppose that 2 - n and 3 | n, so a = u−1aua = 1, so PSL2(O2)/N is a factor of <

t, u; t3, tu = ut >∼= Z3×Z. So, by (3.4.10), N = N(a, t, un), N(a, un/3), N(a, tun/3), N(a, t2un/3).

Fine ([25] section 4.5.3) correctly classified the normal subgroups of PSL2(O2) of index

coprime to 6, didn’t deal with the cases 2 | n and 3 - n or 2 - n and 3 | n, and erroneously

claimed to have classified the normal subgroups of PSL2(O2) with torsion of index 6k. As

can be seen above we have dealt with the case of a normal subgroup of index not divisible

125

by 6 but the case of an index divisible by 6 is very complicated and we have been unable

to deal with it. The complications arise because if N PSL2(O2) is of index 6k and has

torsion then PSL2(O2)/N is a factor of Z× PSL2(Z) ([25] theorem 4.5.3).

Lemma. ([25] theorem 4.5.3) If N PSL2(O7) has torsion then PSL2(O7)/N is a factor

of Z2 × Z.

Theorem 6.5.2. Let N PSL2(O7) be of index n. Then if 2 - n

N = N(a, t, un)

if 2 | n and N has torsion then N is one of

N(a, t, un), N(at, un/2), N(at, aun/2)

Proof. Suppose that 2 - n so a, u−1auat ∈ N , so a, t ∈ N , so PSL2(O7)/N is a factor of

< u >∼= Z. So N = N(a, t, u2).

Now suppose that 2 | n and N has torsion, so PSL2(O7)/N is a factor of Z2 × Z so, by

(3.4.10), N is one of

N(a, t, un), N(at, un/2), N(at, aun/2)

Missing from Fine’s classification were the groups N(at, aun/2).

Lemma. ([25] theorem 4.5.3) If N PSL2(O11) has torsion then PSL2(O11)/N is a factor

of Z3 × Z.

Theorem 6.5.3. Let N PSL2(O11) be of index n. Then if 3 - n then

N = N(a, t, un)

if 2 - n, and 3 | n, or 6 | n and N has torsion then N is one of

N(a, t, un), N(a, un/3), N(a, tun/3), N(a, t2un/3)

126

Proof. Suppose that 3 - n, so at = u−1auat = 1, so a = t = 1, so N = N(a, t, un). Suppose

that 2 - n, and 3 | n, so a = 1, so t3 = 1, so PSL2(O11)/N is a factor of < t, u; t3, [t, u] >∼=

Z3 × Z. Similarly, if 6 | n, and N has torsion, then PSL2(O11)/N is a factor of Z3 × Z. So

that, by (3.4.10), N = N(a, t, un), N(a, un/3), N(a, tun/3), N(a, t2un/3).

Missing from Fine’s classification were the groups N(a, tun/3) and N(a, t2un/3).

6.5.1 Normal congruence subgroups of PSL2(O2)

Now consider O2 and let ω = i√

2. By (6.3.3), O2 has one ideal of index 2, namely ωO2

and two ideals of index 3, namely (1 + ω)O2, and (1 − ω)O2. Thus, by (6.4.5) and (6.4.6),

if n ∈ N such that 6 - n then

N2(n) =

3 if n = 2

4 if n = 3

1 if n = 4, 9

0 else

and so from (6.5.1) all the normal subgroups of PSL2(O2) of index 2

N(a, t, u2), N(at, u), N(at, au)

are congruence subgroups, and all the normal subgroups

N(a, t, u3), N(a, u), N(a, tu), N(a, t2u)

of index 3 are congruence subgroups. Exactly one of the normal subgroups

N(a, t, u4), N(at, u2), N(at, au2)

of index 4 is a congruence subgroup and exactly one of the normal subgroups

N(a, t, u9), N(a, u3), N(a, tu3), N(a, t2u3)

of index 9 is a congruence subgroup. All other normal subgroups of PSL2(O2) of index not

divisible by 6 are non-congruence subgroups.

127

Lemma 6.5.4.

PSL2(O2, ωO2) 6 N(at, u)

Proof. PSL2(O2)/PSL2(O2, ωO2) ∼= S3, so PSL2(O2) has a normal congruence subgroup

of index 2. By (6.5.1), the normal subgroups of PSL2(O2) of index 2 are:

N(a, t, u2), N(at, u), N(at, au)

Now u ∈ PSL2(O2, ωO2), so u ∈ N . So, clearly, N 6= N(a, t, u2), N(at, au). Hence N =

N(at, u).

Lemma 6.5.5.

N(a, t, u4), N(at, au2)

are non-congruence subgroups, and

PSL2(O2, 2O2) 6 N(at, u2), N(a, t, u2), N(at, au)

Proof. Suppose N is one of these groups. Then PSL2(O2)/N is abelian. Suppose N is

a congruence subgroup, so because N is of index 2, or 4 we have PSL2(O2, 2O2) 6 N .

Now (PSL2(O2)/PSL2(O2, 2O2))ab ∼= Z2 × Z2. Suppose N = N(a, t, u4), N(at, au2). Then

PSL2(O2)/N ∼= Z4. So N is a non-congruence subgroup. By (6.5.1), the only other normal

subgroup of PSL2(O2) of index 4 is N(at, u2). Thus PSL2(O2, 2O2) 6 N(at, u2).

Now, by (3.4.10), Z2 × Z2 has 3 normal subgroups of index 2. By (6.5.1), the normal

subgroups of PSL2(O2) of index 2 are

N(a, t, u2), N(at, u), N(at, au)

and we have already seen that PSL2(O2, ωO2) 6 N(at, u).

Lemma 6.5.6.

PSL2(O2, (1 + ω)O2) 6 N(a, tu)

128

Proof. PSL2(O2)/PSL2(O2, (1 + ω)O2) ∼= A4, and A4 has a normal subgroup of index 3.

By (6.5.1), the normal subgroups of PSL2(O2) of index 3 are:

N(a, t, u3), N(a, u), N(a, tu), N(a, t2u)

Now tu ∈ PSL2(O2, (1 + ω)O2), so tu ∈ N . So, clearly N 6= N(a, t, u3), N(a, u). Suppose

that N = N(a, t2u). Then t = t2uu−1t−1 ∈ N , so a, t, u ∈ N . Contradiction. Hence

N = N(a, tu).

Lemma 6.5.7.

PSL2(O2, (1− ω)O2) 6 N(a, t2u)

Proof. PSL2(O2)/PSL2(O2, (1 − ω)O2) ∼= A4, so, as above, PSL2(O2) has a normal con-

gruence subgroup, N , of index 3. So, by (6.5.1), N is one of

N(a, t, u3), N(a, u), N(a, t2u)

Now tu−1 ∈ PSL2(O2, (1− ω)O2), so tu−1 ∈ N . So clearly N 6= N(a, t, u3), N(a, u). Hence

N = N(a, t2u).

Lemma 6.5.8.

N(a, t, u9), N(a, tu3), N(a, t2u3)

are non-congruence subgroups, and

PSL2(O2, 3O2) 6 N(a, u3), N(a, t, u3), N(a, u)

Proof. Let N be one of these groups and suppose that N is a congruence subgroup. Then be-

causeN is of index 3, or 9 we have PSL2(O2, 3O2) 6 N . Now ifN = N(a, t, u9), N(a, tu3), N(a, t2u3)

then PSL2(O2)/N ∼= Z9, so as (PSL2(O2)/PSL2(O2, 3O2))ab ∼= Z3 × Z3, N is a non-

congruence subgroup. The only other normal subgroup of index 9 is N(a, u3). Hence

PSL2(O2, 3O2) 6 N(a, u3).

Now, by (3.4.10), Z3 × Z3 has 4 normal subgroups of index 3. By (6.5.1), the normal

subgroups of PSL2(O2) of index 3 are

N(a, t, u3), N(a, u), N(a, tu), N(a, t2u)

129

and we have already seen that PSL2(O2, (1+ω)O2) 6 N(a, tu), and PSL2(O2, (1−ω)O2) 6

N(a, t2u). Hence result.

Hence

Theorem 6.5.9. The normal congruence subgroups of PSL2(O2) of index not divisible by

6 are precisely:

Group Level Index Abelianization

N(at, u) ω 2 Z23 × Z

N(a, t, u2) 2 2 Z32 × Z3 × Z

N(at, au) 2 2 Z3 × Z

N(a, tu) 1 + ω 3 Z32 × Z

N(a, t2u) 1− ω 3 Z22 × Z

N(a, u) 3 3 Z32 × Z

N(a, t, u3) 3 3 Z42 × Z3 × Z

N(at, u2) 2 4 Z23 × Z

N(a, u3) 3 9 Z72 × Z

Proof. Presentations for each of these groups were found using GAP [24]. These presentations

were then abelianized by hand.

Remark. So we have shown that every subgroup of PSL2(O2) of index 2 is a congruence

subgroup and we have an example of a (normal) non-congruence subgroup of PSL2(O2) of

index 4. Now PSL2(O2)/PSL2(O2, ωO2) ∼= S3 and S3 has a non-normal subgroup of index

3. Thus PSL2(O2) has a non-normal congruence subgroup of index 3. It is a simple matter

to use GAP [24] to show that PSL2(O2) has exactly one non-normal subgroup of index 3.

Thus all subgroups of PSL2(O2) of index 3 are congruence subgroups. So the least index

of a non-congruence subgroup of PSL2(O2) is 4. This replicates part of Grunewald and

Schwermer’s theorem ([31] proposition 3.1)

130

6.5.2 Normal congruence subgroups of PSL2(O7), PSL2(O11),

PSL2(O3,2), and PSL2(O1,2)

In exactly the same way as we did for PSL2(O2) we can classify Normal congruence sub-

groups of PSL2(O7), PSL2(O11), PSL2(O3,2), and PSL2(O1,2). However we do have

Lemma 6.5.10. Let N PSL2(O7) be of index n. Suppose that 6 | n and N has torsion.

Then N is a non-congruence subgroup.

Proof. Suppose that N is a congruence subgroup. By (6.5.2) a, or at ∈ N , so by (6.4.1)

there is a normal congruence subgroup of SL2(O7) corresponding to N , of order O7. By

applying (5.4.9), (5.3.19) and then (6.4.1) again we see that PSL2(O7, 2) 6 N . From (6.5.2)

we can see that PSL2(O7)/N is abelian. Now (PSL2(O7)/PSL2(O7, 2))ab ∼= Z2 × Z2, so

6 | 4. Contradiction. Hence N is a non-congruence subgroup.

and in a similar way we can prove

Lemma 6.5.11. Let N PSL2(O11) be of index n. Suppose that 6 | n and N has torsion.

Then N is a non-congruence subgroup.

Lemma 6.5.12. Let N PSL2(O1,2) be of index n. Suppose that N has torsion and 6 | n.

Then N is a non-congruence subgroup.

So that we get:

Theorem 6.5.13. The normal congruence subgroups of PSL2(O7) with torsion are precisely:

Group Level Index Abelianization

N(at, u) ω 2 Z3 × Z

N(at, au) 1− ω 2 Z3 × Z

N(a, t, u2) 2 2 Z2 × Z

N(at, u2) 2 4 Z23 × Z

131

Remark. So we have shown that all subgroups of PSL2(O7) of index 2 are congruence

subgroups and we have given an example of a (normal) non-congruence subgroup of index

3 in PSL2(O7). So 3 is the least index of a non-congruence subgroup of PSL2(O7). This

replicates a part of Grunewald and Schwermer’s theorem ([31] proposition 3.1).

Theorem 6.5.14. The normal congruence subgroups of PSL2(O11) with torsion are pre-

cisely:

Group Level Index Abelianization

N(a, u) ω 3 Z22 × Z

N(a, t2u) 1− ω 3 Z22 × Z

N(a, t, u3) 3 3 Z3 × Z

N(a, tu) 3 3 Z22 × Z

N(a, u3) 3 9 Z62 × Z

Remark. As stated earlier the unique subgroup of PSL2(O11) of index 2 is a non-congruence

subgroup so clearly 2 is the least index of a non-congruence subgroup. Again this replicates

a part of Grunewald and Schwermer’s theorem ([31] proposition 3.1).

Theorem 6.5.15. The normal congruence subgroups of PSL2(O3,2) of index not divisible

by 6 are precisely:

Group Level Index Abelianization

N(at, aw) m2 2 Z23 × Z

N(a, t, w2) 2 2 Z2 × Z3 × Z

N(at, w) 2 2 Z23 × Z

N(a, tw) m3 3 Z2 × Z

N(at, w2) 2 4 Z33 × Z

where m2 = (2, 1 + i√

3) is the ideal of index 2 in O3,2, and m3 = (3, 2 + 2i√

3) is the ideal

of index 3 in O3,2.

132

Remark. So we have shown that every subgroup of PSL2(O3,2) of index 2 is a congruence

subgroup and we have given an example of a (normal) non-congruence subgroup of index 3

in PSL2(O3,2). So 3 is the least index of a non-congruence subgroup of PSL2(O3,2). This is

an extension of Grunewald and Schwermer’s theorem ([31] proposition 3.1).

Theorem 6.5.16. The normal congruence subgroups of PSL2(O1,2) with torsion are pre-

cisely

Group Level Index Abelianization

N(at, tz, w) m2 2 Z3 × Z2

N(a, t, z, w2) 2 2 Z42 × Z

N(at, aw, tz) 2 2 Z2

N(at, tz, w2) 2 4 Z23 × Z

where m2 = (2, 2i) is the ideal of index 2 in O1,2.

Remark. So there are exactly three congruence subgroups of index 2 in PSL2(O1,2). It is

a simple matter to use GAP [24] to show that there are exactly 8 subgroups of PSL2(O1,2)

of index 2. So there are exactly 5 non-congruence subgroup of index 2 in PSL2(O1,2) and

2 is the least index of a non-congruence subgroup of PSL2(O1,2). This is an extension of

Grunewald and Schwermer’s theorem ([31] proposition 3.1).

133

Bibliography

[1] R. G. Alperin. Normal subgroups of PSL2(Z[√−3]). Proc. AMS, 124(10):2935–2941,

1996.

[2] M. F. Atiyah and I. G. MacDonald. Introduction to commutative algebra. Addison-

Wesley Publishing, 1969.

[3] H. Bass. Algebraic K-theory. Benjamin, New York, 1968.

[4] H. Bass, M. Lazard, and J-P. Serre. Sous-groupes d’indice fini dans SL(n,Z). Bull.

AMS, 70:385–392, 1964.

[5] H. Bass, J. Milnor, and J.-P. Serre. Solution of the congruence subgroup problem for

SLn (n > 3) and Sp2n (n > 2). Publ. I.H.E.S., 33:59–137, 1967.

[6] A. F. Beardon. The geometry of discrete groups. Springer-Verlag, 1983.

[7] A. F. Beardon. The geometry of Fuchsian groups. Lecture notes, March 1998. LMS

course on Fuchsian groups at The University of Lancaster.

[8] L. Bianchi. Geometrische darstellung der gruppen linearer substitutionen mit ganzen

complexen coefficienten nebst anwendungen auf die zahlentheorie. Math. Ann., 38:313–

333, 1891.

[9] L. Bianchi. Sui gruppi de sostituzioni lineari con coeficienti appartenenti a corpi quad-

ratici imaginari. Math. Ann., 40:332–412, 1892.

134

[10] J. Britto. On the construction of non-congruence subgroups. Acta. Arith., 33:261–267,

1977.

[11] B. Chandler and W. Magnus. The history of combinatorial group theory: A case study

in the history of ideas. Springer-Verlag, 1982.

[12] D. E. Cohen. Combinatorial Group Theory: a topological approach. London Math.

Soc. Student Text Volume 14, 1989.

[13] H. Cohn. A classical invitation to Algebraic Numbers and Class Fields. Springer-Verlag,

1978.

[14] P. M. Cohn. On the structure of the GL2 of a ring. Publ. I.H.E.S., 30:5–53, 1966.

[15] P. M. Cohn. A presentation of SL2 for euclidean imaginary quadratic number fields.

Mathematika, 15:156–163, 1968.

[16] D. L. Costa. Zero-dimensionality and the GE2 of polynomial rings. J. Pure and Applied

Algebra, 50(3):223–229, 1988.

[17] D. L. Costa and G. E. Keller. On the normal subgroups of SL(2, A). J. Pure and

Applied Alg, 53:201–226, 1988.

[18] D. L. Costa and G. E. Keller. On the normal subgroups of GL(2, A). J. of Alg.,

135:395–406, 1990.

[19] R. K. Dennis. The GE2 property for discrete subrings of C. Proc. AMS, 50:77–82, 1975.

[20] I. M. S. Dey. Schreier systems in free products. Proc. Glasgow Math. Soc., 7:61–79,

1965.

[21] J. Dieudonne. La geometrie des groupes classiques. Springer, 1963.

[22] A. Drillick. The Picard Group. PhD thesis, New York University, 1971.

135

[23] J. Elstrodt, F. Grunewald, and J. Mennicke. Groups acting on Hyperbolic space.

Springer Monographs in Mathematics, 1998.

[24] M. Schonert et al. GAP. RWTH Aachen: Lehrstuhl D fur Mathematik, 1996.

http://www.gap.dcs.st-and.ac.uk/gap.

[25] B. Fine. Algebraic Theory of the Bianchi groups. Marcel Dekker, 1989.

[26] B. Fine and M. Newman. The normal subgroup structure of the Picard group. Trans.

AMS, 302(2):769–786, 1987.

[27] R. Fricke. Uber die substitutionsgruppen welche zu den aus dem Legendreschen Inter-

gralmodul gezogenen Wurzeln gehoren. Math. Annalen, 28:99–118, 1886.

[28] V. N. Gerasimov. The unit group of free products. Mat Sb, 134(1):42–65, 1987.

[29] K. R. Goodearl and P. Menal. Stable range one for rings with many units. J. Pure and

Applied Algebra, 54:261–287, 1988.

[30] F. Grunewald, J Mennicke, and L. Vaserstein. On the groups SL2(Z [x]) and

SL2(k [x, y]). Israel J. Maths, 86:157–193, 1994.

[31] F. Grunewald and J. Schwermer. On the concept of level for the subgroups of SL2 over

arithmetic rings. Preprint, 1998.

[32] F. J. Grunewald and J. Schwermer. Free non-abelian quotients of SL2 over orders of

imaginary quadratic number fields. J. of Alg., 69:298–304, 1981.

[33] S. K. Gupta and M. P. Murthy. Suslin’s work on linear groups over polynomial rings

and Serre problem. Number 8 in ISI Lecture Notes. MacMillan, 1980.

[34] A. J. Hahn and O.T. O’Meara. The Classical Groups and K-Theory. Springer-Verlag,

1989.

[35] M. Hall. Subgroups of finite index in free groups. Canad. J. Math., 1:187–190, 1949.

136

[36] D. L. Johnson. Presentations of Groups. London Mathematical Society Lecture Notes

Series 22, 1976.

[37] G. A. Jones. Congruence and non-congruence subgroups of the modular group: A

survey. In Proceedings of Groups St. Andrews, volume 121 of London Mathematical

Society Lecture Notes Series, pages 223–234. Cambridge University Press, 1985.

[38] S. Katok. Fuchsian groups. University of Chicago Press, 1992.

[39] F. Klein. Zur theorie der elliptischen modulfunctionen. Math. Annalen, 17:62–70, 1880.

[40] F. Klein and R. Fricke. Vorlesungen uber die Theorie der elliptischen Modulfunctionen.

2 vols, Teubner, Leipzig, 1890, 1892. reprinted by Johnson Reprint, New York, 1965.

[41] M. Kline. Mathematics in Western culture. Penguin Books, 1987.

[42] W. Klingenberg. Lineare gruppen uber lokalen ringen. Amer. J. Math., 83:137–153,

1961.

[43] N. H. J. Lacroix. Two dimensional linear groups over local rings. Canad. J. Math.,

21:106–135, 1969.

[44] N. H. J. Lacroix and C. Levesque. Sur les sous-groupes normaux de SL2 sur un anneau

local. Canad. Math. Bull., 26(2):209–219, 1983.

[45] A. Lubotzky. Free quotients and the congruence kernel of SL2. J. of Alg., 77:411–418,

1982.

[46] A. Lubotzky. Counting finite index subgroups. In Groups St Andrews 1993, volume

212 of London Mathematical Society Lecture Notes Series, pages 368–404. Cambridge

University Press, 1995.

[47] A. Lubotzky. Subgroup growth and congruence subgroups. Invent. Maths, 119:267–295,

1995.

137

[48] R. Lyndon and P. Schupp. Combinatorial Group Theory. Springer-Verlag, 1977.

[49] A. M. Macbeath. Groups of homeomorphisms of a simple connected space. Annals of

Mathematics, 79(3):473–488, 1964.

[50] D. MacHale. Comic Sections: The book of Mathematical jokes, humour, wit and wis-

dom. Boole Press, Dublin, 1993.

[51] W. Magnus. Noneuclidean tesselations and their groups. Academic Press, 1974.

[52] Yu. I. Manin and A. A. Panchishkin. Number Theory I, volume 49 of Encyclopaedia of

Mathematical Sciences. Springer, 1995.

[53] A. W. Mason. Free quotients of congruence subgroups of SL2 over a Dedekind ring of

arithmetic type contained in a function field. Math. Proc. Camb. Phil. Soc., 101:421–429,

1987.

[54] A. W. Mason. On the GL2 of a local ring in which 2 is not a unit. Canad. Math. Bull.,

30(2):165–176, 1987.

[55] A. W. Mason. Standard subgroups of GL2(A). Proc. EMS, 30:341–349, 1987.

[56] A. W. Mason. Non-standard normal subgroups and non-normal standard subgroups of

the Modular group. Canad. Math. Bull., 32(1):109–113, 1989.

[57] A. W. Mason. On GL2 of a local ring in which 2 is not a unit II. Comm in Alg,

17:511–551, 1989.

[58] A. W. Mason. The order and level of a subgroup of GL2 over a Dedekind domain of

arithmetic type. Proc. Royal Soc. Edin., 119A:191–212, 1991.

[59] A. W. Mason. Normal subgroups of level zero of the Bianchi groups. Bull. London

Math. Soc., 26:263–267, 1994.

[60] A. W. Mason and R. W. K. Odoni. Non-normal standard subgroups of the Bianchi

groups. Proc. AMS, 124(3):721–726, 1996.

138

[61] A. W. Mason, R. W. K. Odoni, and W. W. Stothers. Almost all Bianchi groups have

free non-cyclic quotients. Math. Proc. Camb. Phil. Soc., 111(1):1–6, 1992.

[62] A. W. Mason and S. J. Pride. Normal subgroups of prescribed order and zero level of

the Modular group and related groups. J. London Math. Soc., 42(2):465–474, 1990.

[63] A. W. Mason and W. W. Stothers. On the subgroups of GL(n,A) which are generated

by commutators. Invent. Math., 23:327–346, 1974.

[64] H. Matsumura. Commutative algebra. W. A. Benjamin, Inc., New York, 1970.

[65] D. L. McQuillan. Classification of the normal congruence subgroups of the Modular

group. Amer. J. Math., 87:285–296, 1965.

[66] J. Mennicke. Finite factor groups of the unimodular group. Ann. Math., 81:31–37, 1965.

[67] J. Milnor. Hyperbolic geometry: The first 150 years. Bull. AMS, 6(1):9–24, 1982.

[68] J. Moody. Groups for Undergraduates. World Scientific Publishing Co., 1994.

[69] P. M. Neumann. The SQ-Universality of some finitely presented groups. J. Austral.

Math. Soc., 16:1–6, 1973.

[70] M. Newman. Integral matrices. Academic Press, 1972.

[71] M. Newman. Asymptotic formulas related to free products of cyclic groups. Math. of

Comp., 30(136):838–846, 1976.

[72] O. T. O’Meara. Lectures on linear groups. In Regional Conference Series in Mathem-

atics, volume 22. American Mathematical Society, 1974.

[73] E. Picard. Sur un groupe de transformations des points de l’espace situe du meme cote

d’un plan. Bull. Soc. Math. France, 12:43–47, 1884.

[74] G. Pick. Uber gewisse ganzzahlige lineare Substitutionen welche sich nicht durch algeb-

raische Congruenzen erklaren lassen. Math. Annalen, 28:119–124, 1886.

139

[75] A. Pilkington. The E2(R)-normalized subgroups of GL2(R). J. of Alg., 172:584–611,

1995.

[76] S. J. Pride. The concept of largeness in group theory. In Word Problems II, volume 95

of Studies in Logic and the Foundations of Mathematics, pages 299–335. North-Holland,

1978.

[77] I. Reiner. Normal subgroups of the unimodular group. Illinois J. Math., 2:142–144,

1958.

[78] R. Riley. Applications of a computer implementation of Poincare’s theorem on funda-

mental polyhedra. Math. of Comp., 40(162):607–632, 1983.

[79] J. J. Rotman. The Theory of Groups. Allyn and Bacon, Inc., 1973.

[80] J.-P. Serre. Le probleme des groupes de congruence pour SL2. Ann. of Math., 92:489–

527, 1970.

[81] J.-P. Serre. Trees. Springer, Berlin, 1980.

[82] S. Stahl. The Poincare half-plane. Jones and Bartlett Publishers, 1993.

[83] W. W. Stothers. Level and index in the Modular group. Proc. Royal Soc. Edin.,

99A:115–126, 1984.

[84] R. G. Swan. Generators and relations for certain special linear groups. Adv. in Math.,

6:1–77, 1971.

[85] L. N. Vaserstein. K1-theory and the congruence subgroup problem. Mat Zametki,

5:233–244, 1969. English Trans: Math Notes 5 (1969) 141-148.

[86] L. N. Vaserstein. On the normal subgroups of GLn over a ring. In Algebraic K-theory,

Evanston 1980, volume 854 of Lecture Notes in Mathematics, pages 456–465. Springer-

Verlag, 1981.

140

[87] L. N. Vaserstein. An answer to the question of M. Newman on matrix completion. Proc.

AMS, 97:189–196, 1986.

[88] L. N. Vaserstein. Normal subgroups of the general linear groups over Banach algebras.

J. Pure and Applied Algebra, 41:99–112, 1986.

[89] L. N. Vaserstein. Normal subgroups of the general linear groups over von Neumann

regular rings. Proc. AMS, 96(2):209–214, 1986.

[90] L. N. Vaserstein. On normal subgroups of GL2 over rings with many units. Comp.

Math., 74:157–164, 1990.

[91] K. Wohlfahrt. An extension of F. Klein’s level concept. Illinois J. Math, 8:529–535,

1964.

[92] K. Wohlfahrt. Uber einen satz von Dey und die modulgruppe. Arch. Math., 29:455–457,

1977.

[93] R. Zimmert. Zur SL2 der ganzen Zahlen eines imaginar-quadratischen Zahlkorpers.

Invent. Math., 19:73–81, 1973.

141