Part II Number Fields
Based on lectures by I. Grojnowski
Notes taken by Dexter Chua
Lent 2016
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Part IB Groups, Rings and Modules is essential and Part II Galois Theory is desirable
Definition of algebraic number fields, their integers and units. Norms, bases and
discriminants. [3]
Ideals, principal and prime ideals, unique factorisation. Norms of ideals. [3]
Minkowski’s theorem on convex bodies. Statement of Dirichlet’s unit theorem. Deter-
mination of units in quadratic fields. [2]
Ideal classes, finiteness of the class group. Calculation of class numbers using statement
of the Minkowski bound. [3]
Dedekind’s theorem on the factorisation of primes. Application to quadratic fields. [2]
Discussion of the cyclotomic field and the Fermat equation or some other topic chosen
by the lecturer. [3]
Contents
0 Introduction
1 Number fields
2 Norm, trace, discriminant, numbers
3 Multiplicative structure of ideals
4 Norms of ideals
5 Structure of prime ideals
6 Minkowski bound and finiteness of class group
7 Dirichlet’s unit theorem
8 L-functions, Dirichlet series*
0 Introduction
Technically, IID Galois Theory is not a prerequisite of this course. However,
many results we have are analogous to what we did in Galois Theory, and we
will not refrain from pointing out the correspondence. If you have not learnt
Galois Theory, then you can ignore them.
1 Number fields
The focus of this course is, unsurprisingly, number fields. Before we define what
number fields are, we look at some motivating examples. Suppose we wanted to
find all numbers of the form
x
2
+
y
2
, where
x, y Z
. For example, if
a, b
can
both be written in this form, does it follow that ab can?
In IB Groups, Rings and Modules, we did the clever thing of working with
Z
[
i
]. The integers of the form
x
2
+
y
2
are exactly the norms of integers in
Z
[
i
],
where the norm of x + iy is
N(x + iy) = |x + iy|
2
= x
2
+ y
2
.
Then the previous result is obvious if
a
=
N
(
z
) and
b
=
N
(
w
), then
ab
=
N(zw). So ab is of the form x
2
+ y
2
.
Similarly, in the IB Groups, Rings and Modules example sheet, we found
all solutions to the equation
x
2
+ 2 =
y
3
by working in
Z
[
2
]. This is a very
general technique working with these rings, and corresponding fields
Q
(
d
)
can tell us a lot about arithmetic we care about.
In this chapter, we will begin by writing down some basic definitions and
proving elementary properties about number fields.
Definition
(Field extension)
.
A field extension is an inclusion of fields
K L
.
We sometimes write this as L/K.
Definition
(Degree of field extension)
.
Let
K L
be fields. Then
L
is a vector
space over K, and the degree of the field extension is
[L : K] = dim
K
(L).
Definition
(Finite extension)
.
A finite field extension is a field extension with
finite degree.
Definition (Number field). A number field is a finite field extension over Q.
A field is the most boring kind of ring the only ideals are the trivial one
and the whole field itself. Thus, if we want to do something interesting with
number fields algebraically, we need to come up with something more interesting.
In the case of
Q
itself, one interesting thing to talk about is the integers
Z
.
It turns out the right generalization to number fields is algebraic integers.
Definition
(Algebraic integer)
.
Let
L
be a number field. An algebraic integer
is an
α L
such that there is some monic
f Z
[
x
] with
f
(
α
) = 0. We write
O
L
for the set of algebraic integers in L.
Example.
It is a fact that if
L
=
Q
(
i
), then
O
L
=
Z
[
i
]. We will prove this in
the next chapter after we have the necessary tools.
These are in fact the main objects of study in this course. Since we say this
is a generalization of Z Q, the following had better be true:
Lemma. O
Q
= Z, i.e. α Q is an algebraic integer if and only if α Z.
Proof. If α Z, then x α Z[x] is a monic polynomial. So α O
Q
.
On the other hand, let
α Q
. Then there is some coprime
r, s Z
such that
α =
r
s
. If it is an algebraic integer, then there is some
f(x) = x
n
+ a
n1
x
n1
+ ··· + a
0
with a
i
Z such that f(α) = 0. Substituting in and multiplying by s
n
, we get
r
n
+ a
n1
r
n1
s + ···+ a
0
s
n
| {z }
divisible by s
= 0,
So
s | r
n
. But if
s 6
= 1, there is a prime
p
such that
p | s
, and hence
p | r
n
. Thus
p | r
. So
p
is a common factor of
s
and
r
. This is a contradiction. So
s
= 1, and
α is an integer.
How else is this a generalization of
Z
? We know
Z
is a ring. So perhaps
O
L
also is.
Theorem. O
L
is a ring, i.e. if α, β O
L
, then so is α ± β and αβ.
Note that in general O
L
is not a field. For example, Z = O
Q
is not a field.
The proof of this theorem is not as straightforward as the previous one.
Recall we have proved a similar theorem in IID Galois Theory before with
“algebraic integer” replaced with “algebraic number”, namely that if
L/K
is a
field extension with
α, β L
algebraic over
K
, then so is
αβ
and
α ± β
, as well
as
1
α
if α 6= 0.
To prove this, we notice that
α K
is algebraic if and only if
K
[
α
] is a finite
extension if α is algebraic, with
f(α) = a
n
α
n
+ ··· + a
0
= 0, a
n
6= 0
then
K
[
α
] has degree at most
n
, since
α
n
(and similarly
α
1
) can be written as
a linear combination of 1
, α, ··· , α
n1
, and thus these generate
K
[
α
]. On the
other hand, if
K
[
α
] is finite, say of degree
k
, then 1
, α, ··· , α
k
are independent,
hence some linear combination of them vanishes, and this gives a polynomial
for which
α
is a root. Moreover, by the same proof, if
K
0
is any finite extension
over K, then any element in K
0
is algebraic.
Thus, to prove the result, notice that if
K
[
α
] is generated by 1
, α, ··· , α
n
and
K
[
β
] is generated by 1
, β, ··· , β
m
, then
K
[
α, β
] is generated by
{α
i
β
j
}
for
1
i n,
1
j m
. Hence
K
[
α, β
] is a finite extension, hence
αβ, α ± β
K[α, β] are algebraic.
We would like to prove this theorem in an analogous way. We will consider
O
L
as a ring extension of
Z
. We will formulate the general notion of “being an
algebraic integer” in general ring extensions:
Definition
(Integrality)
.
Let
R S
be rings. We say
α S
is integral over
R
if there is some monic polynomial f R[x] such that f(α) = 0.
We say S is integral over R if all α S are integral over R.
Definition
(Finitely-generated)
.
We say
S
is finitely-generated over
R
if there
exists elements
α
1
, ··· , α
n
S
such that the function
R
n
S
defined by
(
r
1
, ··· , r
n
)
7→
P
r
i
α
i
is surjective, i.e. every element of
S
can be written as a
R
-
linear combination of elements
α
1
, ··· , α
n
. In other words,
S
is finitely-generated
as an R-module.
This is a refinement of the idea of being algebraic. We allow the use of rings
and restrict to monic polynomials. In Galois theory, we showed that finiteness
and algebraicity “are the same thing”. We will generalize this to integrality of
rings.
Example.
In a number field
Z Q L
,
α L
is an algebraic integer if and
only if α is integral over Z by definition, and O
L
is integral over Z.
Notation.
If
α
1
, ··· , α
r
S
, we write
R
[
α
1
, ··· , α
r
] for the subring of
S
generated by
R, α
1
, ··· , α
r
. In other words, it is the image of the homomorphism
from the polynomial ring R[x
1
, ··· , x
n
] S given by x
i
7→ α
i
.
Proposition.
(i)
Let
R S
be rings. If
S
=
R
[
s
] and
s
is integral over
R
, then
S
is
finitely-generated over R.
(ii)
If
S
=
R
[
s
1
, ··· , s
n
] with
s
i
integral over
R
, then
S
is finitely-generated
over R.
This is the easy direction in identifying integrality with finitely-generated.
Proof.
(i)
We know
S
is spanned by 1
, s, s
2
, ···
over
R
. However, since
s
is integral,
there exists a
0
, ··· , a
n
R such that
s
n
= a
0
+ a
1
s + ··· + a
n1
s
n1
.
So the
R
-submodule generated by 1
, s, ··· , s
n1
is stable under multiplica-
tion by s. So it contains s
n
, s
n+1
, s
n+2
, ···. So it is S.
(ii)
Let
S
i
=
R
[
s
1
, ··· , s
i
]. So
S
i
=
S
i1
[
s
i
]. Since
s
i
is integral over
R
, it is
integral over
S
i1
. By the previous part,
S
i
is finitely-generated over
S
i1
.
To finish, it suffices to show that being finitely-generated is transitive.
More precisely, if
A B C
are rings,
B
is finitely generated over
A
and
C
is finitely generated over
B
, then
C
is finitely generated over
A
. This
is not hard to see, since if
x
1
, ··· , x
n
generate
B
over
A
, and
y
1
, ··· , y
m
generate
C
over
B
, then
C
is generated by
{x
i
y
j
}
1in,1jm
over
A
.
The other direction is harder.
Theorem. If S is finitely-generated over R, then S is integral over R.
The idea of the proof is as follows: if
s S
, we need to find a monic
polynomial which it satisfies. In Galois theory, we have fields and vector spaces,
and the proof is easy. We can just consider 1
, s, s
2
, ···
, and linear dependence
kicks in and gives us a relation. But even if this worked in our case, there is no
way we can make this polynomial monic.
Instead, consider the multiplication-by-
s
map:
m
s
:
S S
by
γ 7→
. If
S
were a finite-dimensional vector space over
R
, then Cayley-Hamilton tells us
m
s
,
and thus
s
, satisfies its characteristic polynomial, which is monic. Even though
S
is not a finite-dimensional vector space, the proof of Cayley-Hamilton will
work.
Proof.
Let
α
1
, ··· , α
n
generate
S
as an
R
-module. wlog take
α
1
= 1
S
. For
any s S, write
i
=
X
b
ij
α
j
for some
b
ij
R
. We write
B
= (
b
ij
). This is the “matrix of multiplication by
S”. By construction, we have
(sI B)
α
1
.
.
.
a
n
= 0. ()
Now recall for any matrix
X
, we have
adj
(
X
)
X
= (
det X
)
I
, where the
i, j
th
entry of
adj
(
X
) is given by the determinant of the matrix obtained by removing
the ith row and jth column of X.
We now multiply () by adj(sI B). So we get
det(sI B)
α
1
.
.
.
α
n
= 0
In particular,
det
(
sI B
)
α
1
= 0. Since we picked
α
1
= 1, we get
det
(
sI B
) = 0.
Hence if f (x) = det(xI B), then f(x) R[x], and f(s) = 0.
Hence we obtain the following:
Corollary. Let L Q be a number field. Then O
L
is a ring.
Proof.
If
α, β O
L
, then
Z
[
α, β
] is finitely-generated by the proposition. But
then
Z
[
α, β
] is integral over
Z
, by the previous theorem. So
α ± β, αβ
Z[α, β].
Note that it is not necessarily true that if
S R
is an integral extension,
then
S
is finitely-generated over
R
. For example, if
S
is the set of all algebraic
integers in
C
, and
R
=
Z
, then by definition
S
is an integral extension of
Z
, but
S is not finitely generated over Z.
Thus the following corollary isn’t as trivial as the case with “integral” replaced
by “finitely generated”:
Corollary.
If
A B C
be ring extensions such that
B
over
A
and
C
over
B
are integral extensions. Then C is integral over A.
The idea of the proof is that while the extensions might not be finitely gener-
ated, only finitely many things are needed to produce the relevant polynomials
witnessing integrality.
Proof. If c C, let
f(x) =
N
X
i=0
b
i
x
i
B[x]
be a monic polynomial such that
f
(
c
) = 0. Let
B
0
=
A
[
b
0
, ··· , b
N
] and let
C
0
=
B
0
[
c
]. Then
B
0
/A
is finitely generated as
b
0
, ··· , b
N
are integral over
A
.
Also,
C
0
is finitely-generated over
B
0
, since
c
is integral over
B
0
. Hence
C
0
is
finitely-generated over
A
. So
c
is integral over
A
. Since
c
was arbitrary, we know
C is integral over A.
Now how do we recognize algebraic integers? If we want to show something
is an algebraic integer, we just have to exhibit a monic polynomial that vanishes
on the number. However, if we want to show that something is not an algebraic
integer, we have to make sure no monic polynomial kills the number. How can
we do so?
It turns out to check if something is an algebraic integer, we don’t have to
check all monic polynomials. We just have to check one. Recall that if
K L
is a field extensions with
α L
, then the minimal polynomial is the monic
polynomial p
α
(x) K[x] of minimal degree such that p
α
(α) = 0.
Note that we can always make the polynomial monic. It’s just that the
coefficients need not lie in Z.
Recall that we had the following lemma about minimal polynomials:
Lemma. If f K[x] with f (α) = 0, then p
α
| f .
Proof. Write f = p
α
h + r, with r K[x] and deg(r) < deg(p
α
). Then we have
0 = f(α) = p(α)h(α) + r(α) = r(α).
So if r 6= 0, this contradicts the minimality of deg p
α
.
In particular, this lemma implies
p
α
is unique. One nice application of this
result is the following:
Proposition.
Let
L
be a number field. Then
α O
L
if and only if the minimal
polynomial p
α
(x) Q[x] for the field extension Q L is in fact in Z[x].
This is a nice proposition. This gives us an necessary and sufficient condition
for whether something is algebraic.
Proof. () is trivial, since this is just the definition of an algebraic integer.
(
) Let
α O
L
and
p
α
Q
[
x
] be the minimal polynomial of
α
, and
h
(
x
)
Z
[
x
] be a monic polynomial which
α
satisfies. The idea is to use
h
to
show that the coefficients of p
α
are algebraic, thus in fact integers.
Now there exists a bigger field M L such that
p
α
(x) = (x α
1
) ···(x α
r
)
factors in
M
[
x
]. But by our lemma,
p
α
| h
. So
h
(
α
i
) = 0 for all
α
i
. So
α
i
O
M
is an algebraic integer. But
O
M
is a ring, i.e. sums and products of the
α
i
’s are
still algebraic integers. So the coefficients of
p
α
are algebraic integers (in
O
M
).
But they are also in Q. Thus the coefficients must be integers.
Alternatively, we can deduce this proposition from the previous lemma plus
Gauss’ lemma.
Another relation between
Z
and
Q
is that
Q
is the fraction field of
Z
. This
is true for general number fields
Lemma. We have
Frac O
L
=
α
β
: α, β O
L
, β 6= 0
= L.
In fact, for any α L, there is some n Z such that O
L
.
Proof.
If
α L
, let
g
(
x
)
Q
[
x
] be its monic minimal polynomial. Then there
exists
n Z
non-zero such that
ng
(
x
)
Z
[
x
] (pick
n
to be the least common
multiple of the denominators of the coefficients of
g
(
x
)). Now the magic is to
put
h(x) = n
deg(g)
g
x
n
.
Then this is a monic polynomial with integral coefficients in effect, we have
just multiplied the coefficient of
x
i
by
n
deg(g)i
! Then
h
(
) = 0. So
is
integral.
2 Norm, trace, discriminant, numbers
Recall that in our motivating example of
Z
[
i
], one important tool was the norm
of an algebraic integer
x
+
iy
, given by
N
(
x
+
iy
) =
x
2
+
y
2
. This can be
generalized to arbitrary number fields, and will prove itself to be a very useful
notion to consider. Apart from the norm, we will also consider a number known
as the trace, which is also useful. We will also study numbers associated with
the number field itself, rather than particular elements of the field, and it turns
out they tell us a lot about how the field behaves.
Norm and trace
Recall the following definition from IID Galois Theory:
Definition
(Norm and trace)
.
Let
L/K
be a field extension, and
α L
. We
write
m
α
:
L L
for the map
7→ α
. Viewing this as a linear map of
L
vector
spaces, we define the norm of α to be
N
L/K
(α) = det m
α
,
and the trace to be
tr
L/K
(α) = tr m
α
.
The following property is immediate:
Proposition.
For a field extension
L/K
and
a, b L
, we have
N
(
ab
) =
N(a)N(b) and tr(a + b) = tr(a) + tr(b).
We can alternatively define the norm and trace as follows:
Proposition.
Let
p
α
K
[
x
] be the minimal polynomial of
α
. Then the
characteristic polynomial of m
α
is
det(xI m
α
) = p
[L:K(α)]
α
Hence if p
α
(x) splits in some field L
0
K(α), say
p
α
(x) = (x α
1
) ···(x α
r
),
then
N
K(α)/K
(α) =
Y
α
i
, tr
K(α)/K
(α) =
X
α
i
,
and hence
N
L/K
(α) =
Y
α
i
[L:K(α)]
, tr
L/K
= [L : K(α)]
X
α
i
.
This was proved in the IID Galois Theory course, and we will just use it
without proving.
Corollary. Let L Q be a number field. Then the following are equivalent:
(i) α O
L
.
(ii) The minimal polynomial p
α
is in Z[x]
(iii) The characteristic polynomial of m
α
is in Z[x].
This in particular implies N
L/Q
(α) Z and tr
L/Q
(α) Z.
Proof.
The equivalence between the first two was already proven. For the
equivalence between (ii) and (iii), if
m
α
Z
[
x
], then
α O
L
since it vanishes
on a monic polynomial in
Z
[
x
]. On the other hand, if
p
α
Z
[
x
], then so is the
characteristic polynomial, since it is just p
N
α
.
The final implication comes from the fact that the norm and trace are just
coefficients of the characteristic polynomial.
It would be nice if the last implication is an if and only if. This is in general
not true, but it occurs, obviously, when the characteristic polynomial is quadratic,
since the norm and trace would be the only coefficients.
Example.
Let
L
=
K
(
d
) =
K
[
z
]
/
(
z
2
d
), where
d
is not a square in
K
. As a
vector space over
K
, we can take 1
,
d
as our basis. So every
α
can be written
as
α = x + y
d.
Hence the matrix of multiplication by α is
m
α
=
x dy
y x
.
So the trace and norm are given by
tr
L/K
(x + y
d) = 2x = (x + y
d) + (x y
d)
N
L/K
(x + y
d) = x
2
dy
2
= (x + y
d)(x y
d)
We can also obtain this by consider the roots of the minimal polynomial of
α = x + y
d, namely (α x)
2
y
2
d = 0, which has roots x ± y
d.
In particular, if
L
=
Q
(
d
), with
d <
0, then the norm of an element is just
the norm of it as a complex number.
Now that we have computed the general trace and norm, we can use the
proposition to find out what the algebraic integers are. It turns out the result is
(slightly) unexpected:
Lemma. Let L = Q(
d), where d Z is not 0, 1 and is square-free. Then
O
L
=
(
Z[
d] d 2 or 3 (mod 4)
Z
h
1
2
(1 +
d)
i
d 1 (mod 4)
Proof.
We know
x
+
y
λ O
L
if and only if 2
x, x
2
dy
2
Z
by the previous
example. These imply 4
dy
2
Z
. So if
y
=
r
s
with
r, s
coprime,
r, s Z
, then we
must have s
2
| 4d. But d is square-free. So s = 1 or 2. So
x =
u
2
, y =
v
2
for some
u, v Z
. Then we know
u
2
dv
2
4
Z
, i.e.
u
2
dv
2
(
mod
4). But we
know the squares mod 4 are always 0 and 1. So if
d 6≡
1 (
mod
4), then
u
2
dv
2
(
mod
4) imply that
u
2
=
v
2
= 0 (
mod
4), and hence
u, v
are even. So
x, y Z
,
giving O
L
= Z[
d].
On the other hand, if
d
1 (
mod
4), then
u, v
have the same parity mod 2,
i.e. we can write x + y
d as a Z-combination of 1 and
1
2
(1 +
d).
As a sanity check, we find that the minimal polynomial of
1
2
(1 +
d
) is
x
2
x +
1
4
(1 d) which is in Z if and only if d 1 (mod 4).
Field embeddings
Recall the following theorem from IID Galois Theory:
Theorem
(Primitive element theorem)
.
Let
K L
be a separable field extension.
Then there exists an α L such that K(α) = L.
For example, Q(
2,
3) = Q(
2 +
3).
Since
Q
has characteristic zero, it follows that all number fields are separable
extensions. So any number field
L/Q
is of the form
L
=
Q
(
α
). This makes it
much easier to study number fields, as the only extra “stuff” we have on top of
Q.
One particular thing we can do is to look at the number of ways we can
embed
L C
. For example, for
Q
(
1
), there are two such embeddings one
sends
1 to i and the other sends
1 to i.
Lemma.
The degree [
L
:
Q
] =
n
of a number field is the number of field
embeddings L C.
Proof.
Let
α
be a primitive element, and
p
α
(
x
)
Q
[
x
] its minimal polynomial.
Then by we have
deg p
α
= [
L
:
Q
] =
n
, as 1
, α, α
2
, ··· , α
n1
is a basis. Moreover,
Q[x]
(p
α
)
=
Q(α) = L.
Since L/Q is separable, we know p
α
has n distinct roots in C. Write
p
α
(x) = (x α
1
) ···(x α
n
).
Now an embedding
Q
[
x
]
/
(
p
α
)
C
is uniquely determined by the image of
x
,
and
x
must be sent to one of the roots of
p
α
. So for each
i
, the map
x 7→ α
i
gives us a field embedding, and these are all. So there are n of them.
Using these field embeddings, we can come up with the following alternative
formula for the norm and trace.
Corollary.
Let
L/Q
be a number field. If
σ
1
, ··· , σ
n
:
L C
are the different
field embeddings and β L, then
tr
L/Q
(β) =
X
σ
i
(β), N
L/Q
(β) =
Y
i
σ
i
(β).
We call σ
1
(β), ··· , σ
n
(β) the conjugates of β in C.
Proof is in the Galois theory course.
Using this characterization, we have the following very concrete test for when
something is a unit.
Lemma. Let x O
L
. Then x is a unit if and only if N
L/Q
(x) = ±1.
Notation. Write O
×
L
= {x O
L
: x
1
O
L
}, the units in O
L
.
Proof.
(
) We know
N
(
ab
) =
N
(
a
)
N
(
b
). So if
x O
×
L
, then there is some
y O
L
such that xy = 1. So N(x)N(y) = 1. So N(x) is a unit in Z, i.e. ±1.
(
) Let
σ
1
, ··· , σ
n
:
L C
be the
n
embeddings of
L
in
C
. For notational
convenience, We suppose that
L
is already subfield of
C
, and
σ
1
is the inclusion
map. Then for each x O
L
, we have
N(x) =
2
(x) ···σ
n
(x).
Now if
N
(
x
) =
±
1, then
x
1
=
±σ
2
(
x
)
···σ
n
(
x
). So we have
x
1
O
L
, since
this is a product of algebraic integers. So x is a unit in O
L
.
Corollary. If x O
L
is such that N(x) is prime, then x is irreducible.
Proof.
If
x
=
ab
, then
N
(
a
)
N
(
b
) =
N
(
x
). Since
N
(
x
) is prime, either
N
(
a
) =
±
1
or N (b) = ±1. So a or b is a unit.
We can consider a more refined notion than just the number of field embed-
dings.
Definition
(
r
and
s
)
.
We write
r
for the number of field embeddings
L R
,
and s the number of pairs of non-real field embeddings L C. Then
n = r + 2s.
Alternatively,
r
is the number of real roots of
p
α
, and
s
is the number of pairs of
complex conjugate roots.
The distinction between real embeddings and complex embeddings will be
important in the second half of the course.
Discriminant
The final invariant we will look at in this chapter is the discriminant. It is based
on the following observation:
Proposition.
Let
L/K
be a separable extension. Then a
K
-bilinear form
L × L K
defined by (
x, y
)
7→ tr
L/K
(
xy
) is non-degenerate. Equivalent, if
α
1
, ··· , α
n
are a
K
-basis for
L
, the Gram matrix (
tr
(
α
i
α
j
))
i,j=1,···,n
has non-zero
determinant.
Recall from Galois theory that if
L/K
is not separable, then
tr
L/K
= 0, and
it is very very degenerate. Also, note that if
K
is of characteristic 0, then there is
a quick and dirty proof of this fact the trace map is non-degenerate, because
for any
x K
, we have
tr
L/K
(
x · x
1
) =
n 6
= 0. This is really the only case
we care about, but in the proof of the general result, we will also find a useful
formula for the discriminant when the basis is 1, θ, θ
2
, . . . , θ
n1
.
We will use the following important notation:
Notation.
∆(α
1
, ··· , α
n
) = det(tr
L/K
(α
i
α
j
)).
Proof.
Let
σ
1
, ··· , σ
n
:
L
¯
K
be the
n
distinct
K
-linear field embeddings
L
¯
K. Put
S = (σ
i
(α
j
))
i,j=1,···,n
=
σ
1
(α
1
) ··· σ
1
(α
n
)
.
.
.
.
.
.
.
.
.
σ
n
(α
1
) ··· σ
n
(α
n
).
Then
S
T
S =
n
X
k=1
σ
k
(α
i
)σ
k
(α
j
)
!
i,j=1,···n
.
We know σ
k
is a field homomorphism. So
n
X
k=1
σ
k
(α
i
)σ
k
(α
j
) =
n
X
k=1
σ
k
(α
i
α
j
) = tr
L/K
(α
i
α
j
).
So
S
T
S = (tr(α
i
α
j
))
i,j=1,···,n
.
So we have
∆(α
1
, ··· , α
n
) = det(S
T
S) = det(S)
2
.
Now we use the theorem of primitive elements to write
L
=
K
(
θ
) such that
1, θ, ··· , θ
n1
is a basis for L over K, with [L : K] = n. Now S is just
S =
1 σ
1
(θ) ··· σ
1
(θ)
n1
.
.
.
.
.
.
.
.
.
.
.
.
1 σ
n
(θ) ··· σ
n
(θ)
n1
.
This is a Vandermonde matrix, and so
∆(1, θ, ··· , θ
n1
) = (det S)
2
=
Y
i<j
(σ
i
(θ) σ
j
(θ))
2
.
Since the field extension is separable, and hence
σ
i
6
=
σ
j
for all
i, j
, this implies
σ
i
(
θ
)
6
=
σ
j
(
θ
), since
θ
generates the field. So the product above is non-zero.
So we have this nice canonical bilinear map. However, this determinant is
not canonical. Recall that if
α
1
, ··· , α
n
is a basis for
L/K
, and
α
0
1
, ··· , α
0
n
is
another basis, then
α
0
i
=
X
a
ij
α
j
for some A = (a
ij
) GL
n
(K). So
∆(α
0
1
, ··· , α
0
n
) = (det A)
2
∆(α
1
, ··· , α
n
).
However, for number fields, we shall see that we can pick a “canonical” basis,
and get a canonical value for ∆. We will call this the discriminant.
Definition
(Integral basis)
.
Let
L/Q
be a number field. Then a basis
α
1
, ··· , α
n
of L is an integral basis if
O
L
=
(
n
X
i=1
m
i
α
i
: m
i
Z
)
=
n
M
1
Zα
i
.
In other words, it is simultaneously a basis for L over Q and O
L
over Z.
Note that integral bases are not unique, just as with usual bases. Given one
basis, you can get any other by acting by GL
n
(Z).
Example.
Consider
Q
(
d
) with
d
square-free,
d 6
= 0
,
1. If
d
=
1 (
mod
4),
we’ve seen that 1
,
1
2
(1 +
λ
) is an integral basis. Otherwise, if
d
=
2
,
3 (
mod
4),
then 1,
d is an integral basis.
The important theorem is that an integral basis always exists.
Theorem.
Let
Q/L
be a number field. Then there exists an integral basis for
O
L
. In particular, O
L
=
Z
n
with n = [L : Q].
Proof.
Let
α
1
, ··· , α
n
be any basis of
L
over
Q
. We have proved that there is
some
n
i
Z
such that
n
i
α
i
O
L
. So wlog
α
1
, ··· , α
n
O
L
, and are an basis of
L
over
Q
. Since
α
i
are integral, so are
α
i
α
j
, and so all these have integer trace,
as we have previously shown. Hence ∆(
α
1
, ··· , α
n
), being the determinant of a
matrix with integer entries, is an integer.
Now choose a
Q
-basis
α
1
, ··· , α
n
O
L
such that ∆(
α
1
, ··· , α
n
)
Z \ {
0
}
has minimal absolute value. We will show that these are an integral basis.
Let x O
L
, and write
x =
X
λ
i
α
i
for some λ
i
Q. These λ
i
are necessarily unique since α
1
, ··· , α
n
is a basis.
Suppose some λ
i
6∈ Z. wlog say λ
1
6∈ Z. We write
λ
1
= n
1
+ ε
1
,
for n
1
Z and 0 < ε
1
< 1. We put
α
0
1
= x n
1
α
1
= ε
1
α
1
+ λ
2
α
2
+ ··· + λ
n
α
n
O
L
.
So α
0
1
, α
2
, ··· , α
n
is still a basis for L/Q, and are still in O
L
. But then
∆(α
0
1
, ··· , α
n
) = ε
2
1
· ∆(α
1
, ··· , α
n
) < ∆(α
1
, ··· , α
n
).
This contradicts minimality. So we must have
λ
i
Z
for all
Z
. So this is a basis
for O
L
.
Now if
α
0
1
, ··· , α
0
n
is another integral basis of
L
over
Q
, then there is some
g GL
n
(
Z
) such that
gα
i
=
α
0
i
. Since
det
(
g
) is invertible in
Z
, it must be 1 or
1, and hence
det ∆(α
0
1
, ··· , α
0
n
) = det(g)
2
∆(α
1
, ··· , α
n
) = ∆(α
1
, ··· , α
n
)
and is independent of the choice of integral basis.
Definition
(Discriminant)
.
The discriminant
D
L
of a number field
L
is defined
as
D
L
= ∆(α
1
, ··· , α
n
)
for any integral basis α
1
, ··· , α
n
.
Example.
Let
L
=
Q
(
d
), where
d 6
= 0
,
1 and
d
is square-free. If
d
=
2
,
3
(mod 4), then it has an integral basis 1,
d. So
D
L
= det
1
d
1
d
2
= 4d.
Otherwise, if d
=
1 (mod 4), then
D
L
= det
1
1
2
(1 +
d)
1
1
2
(1
d)
2
= d.
Recall that we have seen the word discriminant before, and let’s make sure
these concepts are more-or-less consistent. Recall that the discriminant of a
polynomial f(x) =
Q
(x α
i
) is defined as
disc(f) =
Y
i<j
(α
i
α
j
)
2
= (1)
n(n1)/2
Y
i6=j
(α
i
α
j
).
If
p
θ
(
x
)
K
[
x
] is the minimal polynomial of
θ
(where
L
=
K
[
θ
]), then the roots
of p
θ
are σ
i
(θ). Hence we get
disc(p
θ
) =
Y
i<j
(σ
i
(θ) σ
j
(θ))
2
.
In other words,
disc(p
θ
) = ∆(1, θ, ··· , θ
n1
).
So this makes sense.
3 Multiplicative structure of ideals
Again, let
L/Q
be a number field. It turns out that in general, the integral
ring
O
L
is not too well-behaved as a ring. In particular, it fails to be a UFD in
general.
Example. Let L = Q(
5). Then O
L
= Z[
5]. Then we find
3 · 7 = (1 + 2
5)(1 2
5).
These have norms 9, 49, 21, 21. So none of 3, 7, 1 + 2
5 are associates.
Moreover, 3
,
7
,
1
±
2
5
are all irreducibles. The proof is just a straightfor-
ward check on the norms.
For example, to show that 3 is irreducible, if 3 =
αβ
, then 9 =
N
(3) =
N
(
α
)
N
(
β
). Since none of the terms on the right are
±
1, we must have
N
(
α
) =
±
3.
But there are no solutions to
x
2
+ 5y
2
= ±3
where x, y are integers. So there is no α = x + y
5 such that N(α) = ±3.
So unique factorization fails.
Note that it is still possible to factor any element into irreducibles, just not
uniquely we induct on
|N
(
α
)
|
. If
|N
(
α
)
|
= 1, then
α
is a unit. Otherwise,
α
is either irreducible, or
α
=
βγ
. Since
N
(
β
)
N
(
γ
) =
N
(
α
), and none of them are
±1, we must have |N(β)|, |N(γ)| < |N(α)|. So done by induction.
To fix the lack of unique factorization, we instead look at ideals in
O
L
.
This has a natural multiplicative structure the product of two ideals
a, b
is
generated by products
ab
, with
a a, b b
. The big theorem is that every ideal
can be written uniquely as a product of prime ideals.
Definition
(Ideal multiplication)
.
Let
a, b C O
L
be ideals. Then we define the
product ab as
ab =
X
i,j
α
i
β
j
: α
i
a, β
j
b
.
We write a | b if there is some ideal c such that ac = b, and say a divides b.
The proof of unique factorization is the same as the proof that
Z
is a UFD.
Usually, when we want to prove factorization is unique, we write an object as
a = x
1
x
2
···x
m
= y
1
y
2
···y
n
.
We then use primality to argue that
x
1
must be equal to some of the
y
i
, and
then cancel them from both sides. We can usually do this because we are working
with an integral domain. However, we don’t have this luxury when working with
ideals.
Thus, what we are going to do is to find inverses for our ideals. Of course,
given any ideal
a
, there is no ideal
a
1
such that
aa
1
=
O
L
, as for any
b
,
we know
ab
is contained in
a
. Thus we are going to consider more general
objects known as fractional ideals, and then this will allow us to prove unique
factorization.
Even better, we will show that
a | b
is equivalent to
b a
. This is a very
useful result, since it is often very easy to show that
b a
, but it is usually very
hard to actually find the quotient a
1
b.
We first look at some examples of multiplication and factorization of ideals
to get a feel of what these things look like.
Example. We have
hx
1
, ··· , x
n
ihy
1
, ··· , y
m
i = hx
i
y
j
: 1 i n, 1 j mi.
In particular,
hxihyi = hxyi.
It is also an easy exercise to check (ab)c = a(bc).
Example. In Z[
5], we claim that we have
h3i = h3, 1 +
5ih3, 1
5i.
So h3i is not irreducible.
Indeed, we can compute
h3, 1 +
5ih3, 1
5i = h9, 3(1 + 2
5), 3(1 2
5), 21i.
But we know
gcd
(9
,
21) = 3. So
h
9
,
21
i
=
h
3
i
by Euclid’s algorithm. So this is in
fact equal to h3i.
Notice that when we worked with elements, the number 3 was irreducible, as
there is no element of norm 3. Thus, scenarios such as 2
·
3 = (1+
5
)(1
5
)
could appear and mess up unique factorization. By passing on to ideals, we can
further factorize
h
3
i
into a product of smaller ideals. Of course, these cannot be
principal ideals, or else we would have obtained a factorization of 3 itself. So
we can think of these ideals as “generalized elements” that allow us to further
break elements down.
Indeed, given any element in
α O
L
, we obtain an ideal
hαi
corresponding
to α. This map is not injective if two elements differ by a unit, i.e. they are
associates, then they would give us the same ideal. However, this is fine, as we
usually think of associates as being “the same”.
We recall the following definition:
Definition
(Prime ideal)
.
Let
R
be a ring. An ideal
p R
is prime if
R/p
is
an integral domain. Alternatively, for all
x, y R
,
xy p
implies
x p
or
y p
.
In this course, we take the convention that a prime ideal is non-zero. This is
not standard, but it saves us from saying “non-zero” all the time.
It turns out that the ring of integers
O
L
is a very special kind of rings, known
as Dedekind domains:
Definition (Dedekind domain). A ring R is a Dedekind domain if
(i) R is an integral domain.
(ii) R is a Noetherian ring.
(iii) R
is integrally closed in
Frac R
, i.e. if
x Frac R
is integral over
R
, then
x R.
(iv) Every proper prime ideal is maximal.
This is a rather specific list of properties
O
L
happens to satisfy, and it turns
out most interesting properties of
O
L
can be extended to arbitrary Dedekind
domains. However, we will not do the general theory, and just study number
fields in particular.
The important result is, of course:
Proposition.
Let
L/Q
be a number field, and
O
L
be its ring of integers. Then
O
L
is a Dedekind domain.
The first three parts of the definition are just bookkeeping and not too
interesting. The last one is what we really want. This says that
O
L
is “one
dimensional”, if you know enough algebraic geometry.
Proof of (i) to (iii).
(i) Obvious, since O
L
L.
(ii)
We showed that as an abelian group,
O
L
=
Z
n
. So if
a O
L
is an ideal,
then
a Z
n
as a subgroup. So it is finitely generated as an abelian group,
and hence finitely generated as an ideal.
(iii)
Note that
Frac O
L
=
L
. If
x L
is integral over
O
L
, as
O
L
is integral
over Z, x is also integral over Z. So x O
L
, by definition of O
L
.
To prove the last part, we need the following lemma, which is also very
important on its own right.
Lemma.
Let
a C O
L
be a non-zero ideal. Then
a Z 6
=
{
0
}
and
O
L
/a
is finite.
Proof. Let α a and α 6= 0. Let
p
α
= x
m
+ a
m1
x
m1
+ ··· + a
0
be its minimal polynomial. Then
p
α
Z
[
x
]. We know
a
0
6
= 0 as
p
α
is irreducible.
Since p
α
(α) = 0, we know
a
0
= α(α
m1
+ a
m1
α
m2
+ ··· + a
2
α + a
1
).
We know
α a
by assumption, and the mess in the brackets is in
O
L
. So the
whole thing is in a. But a
0
Z. So a
0
Z a.
Thus, we know ha
0
i a. Thus we get a surjection
O
L
ha
0
i
O
L
a
.
Hence it suffices to show that O
L
/ha
0
i is finite. But for every d Z, we know
O
L
hdi
=
Z
n
dZ
n
=
Z
dZ
n
,
which is finite.
Finally, recall that a finite integral domain must be a field let
x R
with
x 6
= 0. Then
m
x
:
y 7→ xy
is injective, as
R
is an integral domain. So it is a
bijection, as R is finite. So there is some y R such that xy = 1.
This allows us to prove the last part
Proof of (iv).
Let
p
be a prime ideal. Then
O
L
/p
is an integral domain. Since
the lemma says O
L
/p is finite, we know O
L
/p is a field. So p is maximal.
We now continue on to prove a few more technical results.
Lemma.
Let
p
be a prime ideal in a ring
R
. Then for
a, b C R
ideals, then
ab p implies a p or b p.
Once we’ve shown that inclusion of ideals is equivalent to divisibility, this in
effect says “prime ideals are primes”.
Proof.
If not, then there is some
a a \ p
and
b b \ p
. Then
ab ab p
. But
then a p or b p. Contradiction.
Eventually, we will prove that every ideal is a product of prime ideals.
However, we cannot prove that just yet. Instead, we will prove the following
“weaker” version of that result:
Lemma.
Let 0
6
=
a C O
L
a non-zero ideal. Then there is a subset of
a
that is a
product of prime ideals.
The proof is some unenlightening abstract nonsense.
Proof.
We are going to use the fact that
O
L
is Noetherian. If this does not hold,
then there must exist a maximal ideal
a
not containing a product of prime ideals
(by which we mean any ideal greater than
a
contains a product of prime ideals,
not that
a
is itself a maximal ideal). In particular,
a
is not prime. So there are
some x, y O
L
such that x, y 6∈ a but xy a.
Consider
a
+
hxi
. This is an ideal, strictly bigger than
a
. So there exists
prime ideals p
1
, ··· , p
r
such that p
1
···p
r
a + hxi, by definition.
Similarly, there exists q
1
, ···q
s
such that q
1
···q
s
a + hyi.
But then
p
1
···p
r
q
1
···q
s
(a + hxi)(a + hyi) a + hxyi = a
So a contains a product of prime ideals. Contradiction.
Recall that for integers, we can multiply, but not divide. To make life easier,
we would like to formally add inverses to the elements. If we do so, we obtain
things like
1
3
, and obtain the rationals.
Now we have ideals. What can we do? We can formally add some inverse
and impose some nonsense rules to make sure it is consistent, but it is helpful
to actually construct something explicitly that acts as an inverse. We can then
understand what significance these inverses have in terms of the rings.
Proposition.
(i) Let 0 6= a C O
L
be an ideal. If x L has xa a, then x O
L
.
(ii) Let 0 6= a C O
L
be a proper ideal. Then
{y L : ya O
L
}
contains elements that are not in O
L
. In other words,
{y L : ya O
L
}
O
L
6= 0.
We will see that the object
{y L
:
ya O
L
}
is in some sense an inverse to
a.
Before we prove this, it is helpful to see what this means in a concrete setting.
Example.
Consider
O
L
=
Z
and
a
= 3
Z
. Then the first part says if
a
b
·
3
Z
3
Z
,
then
a
b
Z. The second says
n
a
b
:
a
b
· 3 Z
o
contains something not in Z, say
1
3
. These are both “obviously true”.
Proof.
(i)
Let
a O
L
. Then since
O
L
is Noetherian, we know
a
is finitely generated,
say by
α
1
, ··· , α
m
. We consider the multiplication-by-
x
map
m
x
:
a a
,
i.e. write
i
=
X
a
ij
α
j
,
where A = (a
ij
) is a matrix in O
L
. So we know
(xI A)
α
1
.
.
.
α
n
= 0.
By multiplying by the adjugate matrix, this implies
det
(
xI A
) = 0. So
x
satisfies a monic polynomial with coefficients in
O
L
, i.e.
x
is integral over
O
L
. Since O
L
is integrally closed, x O
L
.
(ii)
It is clear that if the result is true for
a
, then it is true for all
a
0
a
. So
it is enough to prove this for
a
=
p
, a maximal, and in particular prime,
ideal.
Let
α p
be non-zero. By the previous lemma, there exists prime ideals
p
1
, ··· , p
r
such that
p
1
···p
r
hαi
. We also have that
hαi p
by definition.
Assume
r
is minimal with this property. Since
p
is prime, there is some
i
such that
p
i
p
. wlog, we may as well assume
i
= 1, i.e.
p
1
p
. But
p
1
is
a prime ideal, and hence maximal. So p
1
= p.
Also, since r is minimal, we know p
2
···p
r
6⊆ hai.
Pick β p
2
···p
r
\ hai. Then
βp = βp
1
p
1
p
2
···p
r
hαi.
Dividing by
α
, we get
β
α
p O
L
. But
β 6∈ hαi
. So we know
β
α
6∈ O
L
. So
done.
What is this
{x L
:
xa O
L
}
? This is not an ideal, but it almost is. The
only way in which it fails to be an ideal is that it is not contained inside
O
L
. By
this we mean it is closed under addition and multiplication by elements in
O
L
.
So it is an
O
L
module, which is finitely generated (we will see this in a second),
and a subset of L. We call this a “fractional ideal”.
Definition
(Fractional ideal)
.
A fractional ideal of
O
L
is a subset of
L
that is
also an O
L
module and is finitely generated.
Definition
(Integral/honest ideal)
.
If we want to emphasize that
a C O
L
is an
ideal, we say it is an integral or honest ideal. But we never use “ideal” to mean
fractional ideal.
Note that the definition of fractional ideal makes sense only because
O
L
is
Noetherian. Otherwise, the non-finitely-generated honest ideals would not qualify
as fractional ideals, which is bad. Rather, in the general case, the following
characterization is more helpful:
Lemma.
An
O
L
module
q L
is a fractional ideal if and only if there is some
c L
×
such that
cq
is an ideal in
O
L
. Moreover, we can pick
c
such that
c Z
.
In other words, each fractional ideal is of the form
1
c
a
for some honest ideal
a and integer c.
Proof.
()
We have to prove that
q
is finitely generated. If
q L
×
,
c L
non-zero,
then
cq
=
q
as an
O
L
module. Since
O
L
is Noetherian, every ideal is
finitely-generated. So cq, and hence q is finitely generated.
()
Suppose
x
1
, ··· , x
n
generate
q
as an
O
L
-module. Write
x
i
=
y
i
n
i
, with
y
i
O
L
and n
i
Z, n
i
6= 0, which we have previously shown is possible.
We let
c
=
lcm
(
n
1
, ··· , n
k
). Then
cq O
L
, and is an
O
L
-submodule of
O
L
, i.e. an ideal.
Corollary.
Let
q
be a fractional ideal. Then as an abelian group,
q
=
Z
n
, where
n = [L : Q].
Proof.
There is some
c L
×
such that
cqC O
L
as an ideal, and
cq
=
q
as abelian
groups. So it suffices to show that any non-zero ideal
q O
L
is isomorphic to
Z
n
. Since
q O
L
=
Z
n
as abelian groups, we know
q
=
Z
m
for some
m
. But
also there is some
a
0
Z q
, and
Z
n
=
ha
0
i q
. So we must have
n
=
m
, and
q
=
Z
n
.
Corollary.
Let
a O
L
be a proper ideal. Then
{x L
:
xa O
L
}
is a
fractional ideal.
Proof.
Pick
a a
. Then
a · {x L
:
xa O
L
} O
L
and is an ideal in
O
L
.
Finally, we can state the proposition we want to prove, after all that nonsense
work.
Definition
(Invertible fractional ideal)
.
A fractional ideal
q
is invertible if there
exists a fractional ideal r such that qr = O
L
= h1i.
Notice we can multiply fractional ideals using the same definition as for
integral ideals.
Proposition. Every non-zero fractional ideal is invertible. The inverse of q is
{x L : xq O
L
}.
This is good.
Note that if q =
1
n
a and r =
1
m
b, and a, b C O
L
are integral ideals, then
qr =
1
mn
ab = O
L
if and only if
ab
=
hmni
. So the proposition is equivalent to the statement that
for every a C O
L
, there exists an ideal b C O
L
such that ab is principal.
Proof.
Note that for any
n O
L
non-zero, we know
q
is invertible if and only if
nq
is invertible. So if the proposition is false, there is an integral ideal
a C O
L
which is not invertible. Moreover, as
O
L
is Noetherian, we can assume
a
is
maximal with this property, i.e. if a < a
0
< O
L
, then a
0
is invertible.
Let
b
=
{x L
:
xa O
L
}
, a fractional ideal. We clearly have
O
L
b
, and
by our previous proposition, we know this inclusion is strict.
As
O
L
b
, we know
a ab
. Again, this inclusion is strict if
ab
=
a
, then
for all
x b
, we have
xa a
, and we have shown that this implies
x O
L
, but
we cannot have b O
L
.
So
a ( ab
. By assumption, we also have
ab O
L
, and since
a
is not invertible,
this is strict. But then by definition of
a
, we know
ab
is invertible, which implies
a
is invertible (if
c
is an inverse of
ab
, then
bc
is an inverse of
a
). This is a
contradiction. So all fractional ideals must be invertible.
Finally, we have to show that the formula for the inverse holds. We write
c = {x L : xq O
L
}.
Then by definition, we know q
1
c. So
O
L
= qq
1
qc O
L
.
Hence we must have qc = O
L
, i.e. c = q
1
.
We’re now done with the annoying commutative algebra, and can finally
prove something interesting.
Corollary. Let a, b, c C O
L
be ideals, c 6= 0. Then
(i) b a if and only if bc ac
(ii) a | b if and only if ac | bc
(iii) a | b if and only if b a.
Proof.
(i) () is clear, and () is obtained by multiplying with c
1
.
(ii) () is clear, and () is obtained by multiplying with c
1
.
(iii)
(
) is clear. For the other direction, we notice that the result is easy if
a
=
hαi
is principal. Indeed, if
b
=
hβ
1
, ··· , β
r
i
, then
b hαi
means there
are some β
0
a
, ··· , β
0
r
O
L
such that β
i
= β
0
i
α. But this says
hβ
a
, ··· , β
r
i = hβ
0
1
, ··· , β
0
r
ihαi,
So hαi | b.
In general, suppose we have
b a
. By the proposition, there exists an
ideal c C O
L
such that ac = hαi is principal with α O
L
, α 6= 0. Then
b a if and only if bc hαi by (i); and
a | b if and only if hαi | bc by (ii).
So the result follows.
Finally, we can prove the unique factorization of prime ideals:
Theorem.
Let
a C O
L
be an ideal,
a 6
= 0. Then
a
can be written uniquely as a
product of prime ideals.
Proof.
To show existence, if
a
is prime, then there is nothing to do. Otherwise,
if
a
is not prime, then it is not maximal. So there is some
b ) a
with
b C O
L
.
Hence
b | a
, i.e. there is some
c C O
L
with
a
=
bc
, and
c a
. We can continue
factoring this way, and it must stop eventually, or else we have an infinite chain
of strictly ascending ideals.
We prove uniqueness the usual way. We have shown
p | ab
implies
p | a
or
p | b
. So if
p
1
···p
r
=
q
1
···q
s
, with
p
i
, q
j
prime, then we know
p
1
| q
1
···q
s
,
which implies
p
1
| q
i
for some
i
, and wlog
i
= 1. So
q
1
p
1
. But
q
1
is prime
and hence maximal. So p
1
= q
1
.
Multiply the equation
p
1
···p
r
=
q
1
···q
s
by
p
1
1
, and we get
p
2
···p
r
=
q
2
···q
s
. Repeat, and we get
r
=
s
and
p
i
=
q
i
for all
i
(after renumbering).
Corollary.
The non-zero fractional ideals form a group under multiplication.
We denote this
I
L
. This is a free abelian group generated by the prime ideals,
i.e. any fractional ideal
q
can be written uniquely as
p
a
1
1
···p
q
r
r
, with
p
i
distinct
prime ideals and a
i
Z.
Moreover, if q is an integral ideal, i.e. q C O
L
, then a
i
, ··· , a
r
0.
Proof.
We already have unique factorization of honest ideals. Now take any
fractional ideal, and write it as
q
=
ab
1
, with
a, b O
L
(e.g. take
b
=
hni
for
some n), and the result follows.
Unimportant side note: we have shown that there are two ways we can
partially order the ideals of
O
L
by inclusion and by division. We have shown
that these two orders are actually the same. Thus, it follows that the “least
common multiple” of two ideals
a, b
is their intersection
a b
, and the “greatest
common divisor” of two ideals is their sum
a + b = {a + b : a a, b b}.
Example. Again let [L : Q] = 2, i.e. L = Q(
d) with d 6= 0, 1 and square-free.
While we proved that every ideal can be factorized into prime ideals, we have
completely no idea what prime ideals look like. We just used their very abstract
properties like being prime and maximal. So we would like to play with some
actual ideals.
Recall we had the example
h3, 1 + 2
5ih3, 1 2
5i = h3i.
This is an example where we multiply two ideals together to get a principal ideal,
and the key to this working is that 1 + 2
5
is conjugate to 1
2
5
. We will
use this idea to prove the previous result for number fields of this form.
Let
a C O
L
be a non-zero ideal. We want to find some
b C O
L
such that
ab
is principal.
We know
O
L
=
Z
2
, and
a O
L
as a subgroup. Moreover, we have shown
that we must have
a
=
Z
2
as abelian groups. So it is generated by 2 elements as
subgroups of
Z
2
. Since
Z
is a subring of
O
L
, we know
a
must be generated by
at most 2 elements as
O
L
-modules, i.e. as ideals of
O
L
. If it is generated by one
element, then it is already principal. Otherwise, suppose
a
=
hα, βi
for some
α, β O
L
.
Further, we claim that we can pick
α, β
such that
β Z
. We write
α
=
a
+
b
d
and
β
=
a
0
+
b
0
d
. Then let
=
gcd
(
b, b
0
) =
mb
+
m
0
b
0
, with
m, m
0
Z
(by
Euclid’s algorithm). We set
β
0
= ( + m
0
β) ·
b
0
+ β
= (ma + m
0
a
0
+
d)
b
0
+ a
0
+ b
0
d
= (ma + m
0
a
0
)
b
0
+ a
0
Z
using the fact that that
b
0
`
Z. Then hα, β
0
i = hα, βi.
So suppose a = hb, αi, with b Z and α O
L
. We now claim
hb, αihb, ¯αi
is principal (where
α
=
x
+
y
d
,
¯α
=
x y
d
). In particular, if
a C O
L
, then
a
¯
a is principal, so the proposition is proved by hand.
To show this, we can manually check
hb, αihb, ¯αi = hb
2
, bα, b¯α, α¯αi
= hb
2
, bα, b tr(α), N(α)i,
using the fact that
tr
(
α
) =
α
+
¯α
and
N
(
α
) =
α¯α
. Now note that
b
2
, b tr
(
α
) and
N
(
α
) are all integers. So we can take the gcd
c
=
gcd
(
b
2
, b tr
(
α
)
, N
(
α
)). Then
this ideal is equal to
= hc, i.
Finally, we claim that hci.
Write
=
cx
, with
x L
. Then
tr x
=
b
c
tr α Z
since
b
c
Z
by definition,
and
N(x) = N
c
=
b
2
N(α)
c
2
=
b
2
c
N(α)
c
Z.
So x O
L
. So c | in O
L
. So hc, i = hci.
Finally, after all these results, we can get to the important definition of the
course.
Definition
(Class group)
.
. The class group or ideal class group of a number
field L is
cl
L
= I
L
/P
L
,
where
I
L
is the group of fractional ideals, and
P
L
is the subgroup of principal
fractional ideals.
If
a I
L
, we write [
a
] for its equivalence class in
cl
L
. So [
a
] = [
b
] if and only
if there is some γ L
×
such that γa = b.
The significance is that cl
L
measures the failure of unique factorization:
Theorem. The following are equivalent:
(i) O
L
is a principal ideal domain
(ii) O
L
is a unique factorization domain
(iii) cl
L
is trivial.
Proof.
(i) and (iii) are equivalent by definition, while (i) implies (ii) is well-known
from IB Groups, Rings and Modules. So the real content is (ii) to (i), which is
specific to Dedekind domains.
If
p C O
L
is prime, and
x p \ {
0
}
, we factor
x
=
α
1
···α
k
such that
α
i
is
irreducible in
O
L
. As
p
is prime, there is some
α
i
p
. But then
hα
i
i p
, and
hα
i
i
is prime as
O
L
is a UFD. So we must have
hα
i
i
=
p
as prime ideals are
maximal. So p is principal.
In the next few chapters, we will come up with methods to explicitly compute
the class group of any number field.
4 Norms of ideals
In the previous chapter, we defined the class group, and we know it is generated
by prime ideals of
O
L
. So we now want to figure out what the prime ideals are.
In the case of finding irreducible elements, one very handy tool was the norm
we know an element of
O
L
is a unit iff it has norm 1. So if
x O
L
is not
irreducible, then there must be some element whose norm strictly divides
N
(
x
).
Similarly, we would want to come up with the notion of the norm of an ideal,
which turns out to be an incredibly useful notion.
Definition (Norm of ideal). Let a C O
L
be an ideal. We define
|N(a)| = |O
L
/a| N.
Recall that we’ve already proved that
|O
L
/a|
is finite. So this definition
makes sense. It is also clear that N(a) = 1 iff a = O
L
(i.e. a is a “unit”).
Example. Let d Z. Then since O
L
=
Z
n
, we have dO
L
=
dZ
n
. So we have
N(hdi) = |Z
n
/(dZ)
n
| = |Z/dZ|
n
= d
n
.
We start with a simple observation:
Proposition. For any ideal a, we have N(a) a Z.
Proof.
It suffices to show that
N
(
a
)
a
. Viewing
O
L
/a
as an additive group,
the order of 1 is a factor of
N
(
a
). So
N
(
a
) =
N
(
a
)
·
1 = 0
O
L
/a
. Hence
N(a) a.
The most important property of the norm is the following:
Proposition. Let a, b C O
L
be ideals. Then N(ab) = N(a)N(b).
We will provide two proofs of the result.
Proof.
By the factorization into prime ideals, it suffices to prove this for
b
=
p
prime, i.e.
N(ap) = N(a)N(p).
In other words, we need to show that
O
L
a
=
O
L
ap
.
O
L
p
.
By the third isomorphism theorem, we already know that
O
L
a
=
O
L
ap
a
ap
.
So it suffices to show that O
L
/p
=
a/ap as abelian groups.
In the case of the integers, we could have, say,
p
= 7
Z
,
a
= 12
Z
. We would
then simply define
Z
7Z
12Z
7 · 12Z
x 12x
However, in general, we do not know that
a
is principal, but it turns out it
doesn’t really matter. We can just pick an arbitrary element to multiply with.
By unique factorization, we know a 6= ap. So we can find some α a \ ap.
We now claim that the homomorphism of abelian groups
O
L
p
a
ap
x + p αx + ap
is an isomorphism. We first check this is well-defined if
p p
, then
αp ap
since
α a
. So the image of
x
+
p
and (
x
+
p
)+
p
are equal. So this is well-defined.
To prove our claim, we have to show injectivity and surjectivity. To show
injectivity, since
hαi a
, we have
a | hai
, i.e. there is an ideal
c O
L
such that
ac = hαi. If x O
L
is in the kernel of the map, then αx ap. So
xac ap.
So
xc p.
As
p
is prime, either
c p
or
x p
. But
c p
implies
hαi
=
ac ap
,
contradicting the choice of α. So we must have x p, and the map is injective.
To show this is surjective, we notice that surjectivity means
hαi/ap
=
a/ap
,
or equivalently ap + hαi = a.
Using our knowledge of fractional ideals, this is equivalent to saying (
ap
+
hαi)a
1
= O
L
. But we know
ap < ap + hαi a.
We now multiply by a
1
to obtain
p < (ap + hαi)a
1
= p + c O
L
.
Since
p
is a prime, and hence maximal ideal, the last inclusion must be an
equality. So ap + hαi = a, and we are done.
Now we provide the sketch of a proof that makes sense. The details are left
as an exercise in the second example sheet.
Proof.
It is enough to show that
N
(
p
a
1
1
···p
a
r
r
) =
N
(
p
1
)
a
1
···N
(
p
r
)
a
r
by unique
factorization.
By the Chinese remainder theorem, we have
O
L
p
a
1
1
···p
a
r
r
=
O
L
p
a
1
1
× ··· ×
O
L
p
a
r
r
where p
1
, ··· , p
r
are distinct prime ideals.
Next, we show by hand that
O
L
p
r
=
O
L
p
×
p
p
2
× ··· ×
p
r1
p
r
=
O
L
p
r
,
by showing that
p
k
/p
k+1
is a 1-dimensional vector space over the field
O
L
/p
.
Then the result follows.
This is actually the same proof, but written in a much saner form. This is
better because we are combining a general statement (the Chinese remainder
theorem), with a special property of the integral rings. In the first proof, what
we really did was simultaneously proving two parts using algebraic magic.
We’ve taken an obvious invariant of an ideal, the size, and found it is
multiplicative. How does this relate to the other invariants?
Recall that
∆(α
1
, ··· , α
n
) = det(tr
L/Q
(α
i
α
j
)) = det(σ
i
(α
j
))
2
.
Proposition. Let a C O
L
be an ideal, n = [L : Q]. Then
(i) There exists α
1
, ··· , α
n
a such that
a =
n
X
r
i
α
i
: r
i
Z
o
=
n
M
1
α
i
Z,
and
α
1
, ··· , α
n
are a basis of
L
over
Q
. In particular,
a
is a free
Z
-module
of n generators.
(ii) For any such α
1
, ··· , α
n
,
∆(α
1
, ··· , α
n
) = N(a)
2
D
L
.
The prove this, we recall the following lemma from IB Groups, Rings and
Modules:
Lemma.
Let
M
be a
Z
-module (i.e. abelian group), and suppose
M Z
n
.
Then M
=
Z
r
for some 0 r n.
Moreover, if
r
=
n
, then we can choose a basis
v
1
, ··· , v
n
of
M
such that the
change of basis matrix A = (a
ij
) M
n×n
(Z) is upper triangular, where
v
j
=
X
a
ij
e
i
,
where e
1
, ··· , e
n
is the standard basis of Z
n
.
In particular,
|Z
n
/M | = |a
11
a
12
···a
nn
| = |det A|.
Proof of proposition.
Let
d a Z
, say
d
=
N
(
α
). Then
dO
L
a O
L
. As
abelian groups, after picking an integral basis α
0
1
, ··· , α
0
n
of O
L
, we have
Z
n
=
dZ
n
a Z
n
.
So
a
=
Z
n
. Then the lemma gives us a basis
α
1
, ··· , α
n
of
a
as a
Z
-module. As
a
Q
-module, since the
α
i
are obtained from linear combinations of
α
0
i
, by basic
linear algebra, α
1
, ··· , α
n
is also a basis of L over Q.
Moreover, we know that we have
∆(α
1
, ··· , α
n
) = det(A)
2
∆(α
0
1
, ··· , α
0
n
).
Since
det
(
A
)
2
=
|O
L
/a|
2
=
N
(
a
) and
D
L
= ∆(
α
0
1
, ··· , α
0
n
) by definition, the
second part follows.
This result is very useful for the following reason:
Corollary.
Suppose
a C O
L
has basis
α
1
, ··· , α
n
, and ∆(
α
1
, ··· , α
n
) is square-
free. Then a = O
L
(and D
L
is square-free).
This is a nice trick, since it allows us to determine immediately whether a
particular basis is an integral basis.
Proof. Immediate, since this forces N(a)
2
= 1.
Note that nothing above required
a
to be an actual ideal. It merely had
to be a subgroup of
O
L
that is isomorphic to
Z
n
, since the quotient
O
L
/a
is
well-defined as long as
a
is a subgroup. With this, we can have the following
useful result:
Example.
Let
α
be an algebraic integer and
L
=
Q
(
α
). Let
n
= [
Q
(
α
) :
Q
].
Then a = Z[α] C O
L
. We have
disc(p
α
) = ∆(1, α, α
2
, ··· , α
n1
) = discriminant of minimal polynomial of α.
Thus if disc(p
α
) is square-free, then Z[α] = O
L
.
Even if
disc
(
p
α
) is not square-free, it still says something: let
d
2
| disc
(
p
α
)
be such that disc(p
α
)/d
2
is square-free. Then |N(Z[α])| divides d.
Let
x O
L
. Then the order of
x
+
Z
[
α
]
O
L
/Z
[
α
] divides
N
(
Z
[
α
]), hence
d. So d · x Z[α]. So x
1
d
Z[α]. Hence we have
Z[α] O
L
1
d
Z[α].
For example, if
α
=
a
for some square-free
a
, then
disc
(
a
) is the discriminant
of x
2
a, which is 4a. So the d above is 2, and we have
Z[α] O
Q(
d)
1
2
Z[α],
as we have previously seen.
We shall prove one more lemma, and start factoring things. Recall that we
had two different notions of norm. Given
α O
L
, we can take the norm
N
(
hαi
),
or
N
L/Q
(
α
). It would be great if they are related, like if they are equal. However,
that cannot possibly be true, since
N
(
hαi
) is always positive, but
N
L/Q
(
α
) can
be negative. So we take the absolute value.
Lemma. If α O
L
, then
N(hαi) = |N
L/Q
(α)|.
Proof.
Let
α
1
, ··· , α
n
be an integral basis of
O
L
. Then
αα
1
, .., αα
n
is an integral
basis of hαi. So by the previous lemma,
∆(αα
1
, ··· , αα
n
) = N(hαi)
2
D
L
.
But
∆(αα
1
, ··· , αα
n
) = det(σ
i
(αα
j
)
ij
)
2
= det(σ
i
(α)σ
i
(α
j
))
2
=
n
Y
i=1
σ
i
(α)
!
2
∆(α
1
, ··· , α
n
)
= N
L/Q
(α)
2
D
L
.
So
N
L/Q
(α)
2
= N(hαi)
2
.
But N (hαi) is positive. So the result follows.
5 Structure of prime ideals
We can now move on to find all prime ideals of
O
L
. We know that every ideal
factors as a product of prime ideals, but we don’t know what the prime ideals
are. The only obvious way we’ve had to obtain prime ideals is to take a usual
prime, take its principal ideal and factor it in
O
L
, and get the resultant prime
ideals.
It turns out this gives us all prime ideals.
Lemma.
Let
p C O
L
be a prime ideal. Then there exists a unique
p Z
,
p
prime, with p | hpi. Moreover, N (p) = p
f
for some 1 f n.
This is not really too exciting, as soon as we realize that
p | hpi
is the same
as saying hpi p, and we already know p Z is non-empty.
Proof.
Well
p Z
is an ideal in
Z
, and hence principal. So
p Z
=
pZ
for some
p Z.
We now claim
p
is a prime integer. If
p
=
ab
with
ab Z
. Then since
p p
,
either a p or b p. So a p Z = pZ or b p Z = pZ. So p | a or p | b.
Since
hpi p
, we know
hpi
=
pa
for some ideal
a
by factorization. Taking
norms, we get
p
n
= N(hpi) = N(p)N(a).
So the result follows.
This is all good. So all we have to do is to figure out how principal ideals
hpi
factor into prime ideals.
We write
hpi = p
e
1
1
···p
e
m
m
for some distinct prime ideals
p
i
, with
N
(
p
i
) =
p
f
i
for some positive integers
e
i
.
Taking norms, we get
p
n
=
Y
p
f
i
e
i
.
So
n =
X
e
i
f
i
.
We start by giving some names to the possible scenarios.
Definition
(Ramification indices)
.
Let
hpi
=
p
e
1
1
···p
e
m
m
be the factorization
into prime ideals. Then e
1
, ··· , e
m
are the ramification indices.
Definition (Ramified prime). We say p is ramified if some e
i
> 1.
Definition
(Inert prime)
.
We say
p
is inert if
m
= 1 and
e
m
= 1, i.e.
hpi
remains prime.
Definition
(Splitting prime)
.
We say
p
splits completely if
e
1
=
···
=
e
m
= 1 =
f
1
= ··· = f
m
. So m = n.
Note that this does not exhaust all possibilities. The importance of these
terms, especially ramification, will become clear later.
So how do we actually compute
p
i
and
e
i
? In other words, how can we factor
the ideal
hpi C O
L
into prime ideals? The answer is given very concretely by
Dedekind’s criterion.
Theorem
(Dedekind’s criterion)
.
Let
α O
L
and
g
(
x
)
Z
[
x
] be its minimal
polynomial. Suppose
Z
[
α
]
O
L
has finite index, coprime to
p
(i.e.
p - |O
L
/Z
[
α
]
|
).
We write
¯g(x) = g(x) (mod p),
so ¯g(x) F
p
[x]. We factor
¯g(x) = ϕ
e
1
1
···ϕ
e
m
m
into distinct irreducibles in F
p
[x]. We define the ideal
p
i
= hp, ˜ϕ
i
(α)i C O
L
,
generated by
p
and
˜ϕ
i
, where
˜ϕ
i
is any polynomial in
Z
[
x
] such that
˜ϕ
i
mod p
=
ϕ
i
. Notice that if
˜ϕ
0
is another such polynomial, then
p |
(
˜ϕ
i
ϕ
i
), so
hp, ˜ϕ
0
(
α
)
i
=
hp, ˜ϕ(α)i.
Then the p
i
are prime, and
hpi = p
e
1
1
···p
e
m
m
.
Moreover, f
i
= deg ϕ
i
, so N (a) = p
deg ϕ
i
.
If we are lucky, we might just find an
α
such that
Z
[
α
] =
O
L
. If not, we
can find something close, and as long as
p
is not involved, we are fine. After
finding
α
, we get its minimal polynomial, factor it, and immediately get the
prime factorization of hpi.
Example.
Consider
L
=
Q
(
11
). We want to factor
h
5
i
in
O
L
. We consider
Z
[
11
]
O
L
. This has index 2, and (hopefully) 5
-
2. So this is good
enough. The minimal polynomial is
x
2
+ 11. Taking mod 5, this reduces to
x
2
4 = (x 2)(x + 2). So Dedekind says
h5i = h5,
11 + 2ih5,
11 2i.
In general, consider
L
=
Q
(
d
),
d 6
= 0
,
1 and square-free, and
p
an odd
prime. Then
Z
[
d
]
O
L
has index 1 or 2, both of which are coprime to
p
. So
Dedekind says factor x
2
d mod p. What are the possibilities?
(i)
There are two distinct roots mod
p
, i.e.
d
is a square mod
p
, i.e.
d
p
= 1.
Then
x
2
d = (x + r)(x r) (mod p)
for some r. So Dedekind says
hpi = p
1
p
2
,
where
p
1
= hp,
d ri, p
2
= hp,
d + ri,
and N (p
1
) = N(p
2
) = p. So p splits.
(ii) x
2
d
is irreducible, i.e.
d
is not a square mod
p
, i.e.
d
p
=
1. Then
Dedekind says hpi = p is prime in O
L
. So p is inert.
(iii) x
2
d
has a repeated root mod
p
, i.e.
p | d
, or alternatively
d
p
= 0. Then
by Dedekind, we know
hpi = p
2
,
where
p = hp,
di.
So p ramifies.
So in fact, we see that the Legendre symbol encodes the ramification behaviour
of the primes.
What about the case where p = 2, for L = Q(
d)? How do we factor h2i?
Lemma. In L = Q(
d),
(i) 2 splits in L if and only if d 1 (mod 8);
(ii) 2 is inert in L if and only if d 5 (mod 8);
(iii) 2 ramifies in L if d 2, 3 (mod 4).
Proof.
If
d
1 (
mod
4), then then
O
L
=
Z
[
α
], where
α
=
1
2
(1 +
d
). This has
minimal polynomial
x
2
x +
1
4
(1 d).
We reduce this mod 2.
If d 1 (mod 8), we get x(x + 1). So 2 splits.
If
d
5 (
mod
8), then we get
x
2
+
x
+ 1, which is irreducible. So
h
2
i
is prime, hence 2 is inert.
If
d
2
,
3 (
mod
4), then
O
L
=
Z
[
d
], and
x
2
d
is the minimal polynomial.
Taking mod 2, we get
x
2
or
x
2
+ 1 = (
x
+ 1)
2
. In both cases, 2 ramifies.
Note how important
p - |O
L
/Z
[
α
]
|
is. If we used
Z
[
d
] when
d
1 (
mod
4),
we would have gotten the wrong answer.
Recall
D
L
=
(
4d d 2, 3 (mod 4)
d d 1 (mod 4)
The above computations show that
p | D
L
if and only if
p
ramifies in
L
. This
happens to be true in general. This starts to hint how important these invariants
like D
L
are.
Now we get to prove Dedekind’s theorem.
Proof of Dedekind’s criterion. The key claim is that
Claim. We have
O
L
p
i
=
F
p
[x]
hϕ
i
i
.
Suppose this is true. Then since
ϕ
i
is irreducible, we know
F
p
[x]
hϕ
i
i
is a field.
So p
i
is maximal, hence prime.
Next notice that
p
e
1
1
= hp, ˜ϕ
i
(α)i
e
i
hp, ˜ϕ
i
(α)
e
i
i.
So we have
p
e
1
1
···p
e
m
m
hp, ˜ϕ
1
(α)
e
1
··· ˜ϕ
m
(α)
e
m
i = hp, g(α)i = hpi,
using the fact that g(α) = 0.
So to prove equality, we notice that if we put
f
i
=
deg ϕ
i
, then
N
(
p
i
) =
p
f
i
,
and
N(p
e
1
1
···p
e
m
m
) = N(p
1
)
e
1
···N(p
m
)
e
m
= p
P
e
i
f
i
= p
deg g
.
Since
N
(
hpi
) =
p
n
, it suffices to show that
deg g
=
n
. Since
Z
[
α
]
O
L
has finite
index, we know Z[α]
=
Z
n
. So 1, α, ··· , α
n1
are independent over Z, hence Q.
So deg g = [Q(α) : Q] = n = [L : Q], and we are done.
So it remains to prove that
O
L
p
i
=
Z[α]
p
i
Z[α]
=
F
p
[x]
hϕ
i
i
.
The second isomorphism is clear, since
Z[α]
hp, ˜ϕ
i
(α)i
=
Z[x]
hp, ˜ϕ
i
(x), g(x)i
=
F
p
[x]
h˜ϕ
i
(x), g(x)i
=
F
p
[x]
hϕ
i
(x), ¯g(x)i
=
F
p
[x]
hϕ
i
i
.
To prove the first isomorphism, it suffices to show that the following map is an
isomorphism:
Z[α]
pZ[α]
O
L
pO
L
()
x + pZ[α] 7→ x + pO
L
If this is true, then quotienting further by ˜ϕ
i
gives the desired isomorphism.
To prove the claim, we consider a slightly different map. We notice
p -
|O
L
/Z[α]| means the “multiplication by p map
O
L
Z[α]
O
L
Z[α]
p
()
is injective. But
O
L
/Z
[
α
] is a finite abelian group. So the map is an isomorphism.
By injectivity of (
), we have
Z
[
α
]
pO
L
=
pZ
[
α
]. By surjectivity, we have
Z
[
α
] +
pO
L
=
O
L
. It thus follows that (
) is injective and surjective respectively.
So it is an isomorphism. We have basically applied the snake lemma to the
diagram
0 Z[α] O
L
O
L
Z[α]
0
0 Z[α] O
L
O
L
Z[α]
0
p
p
p
Corollary.
If
p
is prime and
p < n
= [
L
:
Q
], and
Z
[
α
]
O
L
has finite index
coprime to p, then p does not split completely in O
L
.
Proof.
By Dedekind’s theorem, if
g
(
x
) is the minimal polynomial of
α
, then the
factorization of
¯g
(
x
) =
g
(
x
)
mod p
determines the factorization of
hpi
into prime
ideals. In particular,
p
splits completely if and only if
¯g
factors into distinct
linear factors, i.e.
¯g(x) = (x α
1
) ···(x α
n
),
where
α
i
F
p
and
α
i
are distinct. But if
p < n
, then there aren’t
n
distinct
elements of F
p
!
Example.
Let
L
=
Q
(
α
), where
α
has minimal polynomial
x
3
x
2
2
x
8.
This is the case where
n
= 3
>
2 =
p
. On example sheet 2, you will see that 2
splits completely, i.e.
O
L
/
2
O
L
=
F
2
× F
2
× F
2
. But then this corollary shows
that for all
β O
L
,
Z
[
β
]
O
L
has even index, i.e. there does not exist an
β O
L
with |O
L
/Z[β]| odd.
As we previously alluded to, the following is true:
Theorem. p | D
L
if and only if p ramifies in O
L
.
We will not prove this.
6 Minkowski bound and finiteness of class group
Dedekind’s criterion allowed us to find all prime factors of
hpi
, but if we want
to figure out if, say, the class group of a number field is trivial, or even finite,
we still have no idea how to do so, because we cannot go and check every single
prime p and see what happens.
What we are now going to do is the following we are going to use purely
geometric arguments to reason about ideals, and figure that each element of
the class group
cl
L
=
I
L
/P
L
has a representative whose norm is bounded by
some number
c
L
, which we will find rather explicitly. After finding the
c
L
, to
understand the class group, we just need to factor all prime numbers less than
c
L
and see what they look like.
We are first going to do the case of quadratic extensions explicitly, since
2-dimensional pictures are easier to draw. We will then do the full general case
afterwards.
Quadratic extensions
Consider again the case L = Q(
d), where d < 0. Then O
L
= Z[α], where
α =
(
d d 2, 3 (mod 4)
1
2
(1 +
d) d 1 (mod 4)
We can embed this as a subfield
L C
. We can then plot the points on the
complex plane. For example, if
d
2
,
3 (
mod
4), then the points look like this:
0
d
1
1 +
d
2
d
Then an ideal of
O
L
, say
a
=
h
2
,
di
, would then be the sub-lattice given by
the blue crosses.
×
×
×
×
×
×
×
×
×
×
×
×
×
×
×
We always get this picture, since any ideal of
O
L
is isomorphic to
Z
2
as an
abelian group.
If we are in the case where d 1 (mod 4), then the lattice is hexagonal:
0
d
1
1
2
(1 +
d)
The key result is the following purely geometric lemma:
Lemma
(Minskowski’s lemma)
.
Let Λ =
Zv
1
+
Zv
2
R
2
be a lattice, with
v
1
, v
2
linearly independent in
R
(i.e.
Rv
1
+
Rv
2
=
R
2
). We write
v
i
=
a
i
e
1
+
b
i
e
2
.
Then let
A(Λ) = area of fundamental parallelogram =
det
a
1
a
2
b
1
b
2
,
where the fundamental parallelogram is the following:
v
1
v
2
v
1
+ v
2
Then a closed disc S around 0 contains a non-zero point of Λ if
area(S) 4A(Λ).
In particular, there exists an α Λ with α 6= 0, such that
0 < |α|
2
4A(Λ)
π
.
This is just an easy piece of geometry. What is remarkable is that the radius
of the disc needed depends only on the area of the fundamental parallelogram,
and not its shape.
Proof. We will prove a general result in any dimensions later.
We now apply this to ideals
a O
L
, regarded as a subset of
C
=
R
2
via
some embedding
L C
. The following proposition gives us the areas of the
relevant lattices:
Proposition.
(i) If α = a + b
λ, then as a complex number,
|α|
2
= (a + b
λ)(a b
λ) = N(α).
(ii) For O
L
, we have
A(O
L
) =
1
2
p
|D
L
|.
(iii) In general, we have
A(a) =
1
2
p
|∆(α
1
, α
2
)|,
where α
1
, α
2
are the integral basis of a.
(iv) We have
A(a) = N(a)A(O
L
).
Proof.
(i) This is clear.
(ii) We know O
L
has basis 1, α, where again
α =
(
d d 2, 3 (mod 4)
1
2
(1 +
d) d 1 (mod 4)
.
So we can just look at the picture of the lattice, and compute to get
A(O
L
) =
(
p
|d| d 2, 3 (mod 4)
1
2
p
|d| d 1 (mod 4)
=
1
2
p
|D
L
|.
(iii)
If
α
1
, α
2
are the integral basis of
a
, then the lattice of
a
is in fact spanned
by the vectors α
1
= a + bi, α
2
= a
0
+ b
0
i. This has area
A(a) = det
a b
a
0
b
0
,
whereas we have
∆(α
1
, α
2
) = det
α
1
¯α
1
α
2
¯α
2
2
= (α
1
¯α
2
α
2
¯α
1
)
2
= Im(2α
1
¯α
2
)
2
= 4(a
0
b ab
0
)
2
= 4A(a)
2
.
(iv) This follows from (ii) and (iii), as
∆(α
1
, ··· , α
n
) = N(a)
2
D
L
in general.
Now what does Minkowski’s lemma tell us? We know there is an
α a
such
that
N(α)
4A(a)
π
= N(a)c
L
,
where
c
L
=
2
p
|D
L
|
π
.
But α a implies hαi a, which implies hαi = ab for some ideal b. So
|N(α)| = N (hαi) = N(a)N(b).
So this implies
N(b) c
L
=
2
p
|D
L
|
π
.
Recall that the class group is
cl
L
=
I
L
/P
L
, the fractional ideals quotiented by
principal ideals, and we write [
a
] for the class of
a
in
cl
L
. Then if
ab
=
hαi
, then
we have
[b] = [a
1
]
in cl
L
. So we have just shown,
Proposition
(Minkowski bound)
.
For all [
a
]
cl
L
, there is a representative
b
of [a] (i.e. an ideal b O
L
such that [b] = [a]) such that
N(b) c
L
=
2
p
|D
L
|
π
.
Proof. Find the b such that [b] = [(a
1
)
1
] and N (b) c
L
.
Combining this with the following easy lemma, we know that the class group
is finite!
Lemma.
For every
n Z
, there are only finitely many ideals
a O
L
with
N(a) = m.
Proof.
If
N
(
a
) =
m
, then by definition
|O
L
/a|
=
m
. So
m a
by Lagrange’s
theorem. So
hmi a
, i.e.
a | hmi
. Hence
a
is a factor of
hmi
. By unique
factorization of prime ideals, there are only finitely many such ideals.
Another proof is as follows:
Proof.
Each ideal bijects with an ideal in
O
L
/mO
L
= (
Z/m
)
n
. So there are
only finitely many.
Thus, we have proved
Theorem.
The class group
cl
L
is a finite group, and the divisors of ideals of
the form hpi for p Z, p a prime, and 0 < p < c
L
, collectively generate cl
L
.
Proof.
(i)
Each element is represented by an ideal of norm less than 2
p
|D
L
|
, and
there are only finitely many ideals of each norm.
(ii)
Given any element of
cl
L
, we pick a representative
a
such that
N
(
a
)
< c
L
.
We factorize
a = p
e
1
1
···p
e
r
r
.
Then
N(p
i
) N (a) < c
L
.
Suppose
p
i
| hpi
. Then
N
(
p
) is a power of
p
, and is thus at least
p
. So
p < c
L
.
We now try to work with some explicit examples, utilizing Dedekind’s criterion
and the Minkowski bound.
Example. Consider d = 7. So Q(
7) = L, and D
L
= 7. Then we have
1 < c
L
=
2
7
π
< 2.
So cl
L
= {1}, since there are no primes p < c
L
. So O
L
is a UFD.
Similarly, if d = 1, 2, 3, then O
L
is a UFD.
Example. Let d = 5. Then D
L
= 20. We have
2 < c
L
=
4
5
π
< 3.
So cl
L
is generated by primes dividing h2i.
Recall that Dirichlet’s theorem implies
h2i = h2, 1 +
5i
2
= p
2
.
Also,
p
=
h
2
,
1 +
5i
is not principal. If it were, then
p
=
hβi
, with
β
=
x
+
y
5
, and
N
(
β
) = 2. But there are no solutions in
Z
of
x
2
+ 5
y
2
= 2. So
cl
L
= hpi = Z/2.
Example.
Consider
d
=
17
3 (
mod
4). So
c
L
5
.
3. So
cl
L
is generated by
primes dividing by h2i, h3i, h5i. We factor
x
2
+ 17 x
2
+ 1 (x + 1)
2
(mod 2).
So
h2i = p
2
= h2, 1 +
di
2
.
Doing this mod 3, we have
x
2
+ 17 x
2
1 (x 1)(x + 1) (mod 3).
So we have
h3i = q
¯
q = h3, 1 +
dih3, 1
di.
Finally, mod 5, we have
x
2
+ 17 x
2
+ 2 (mod 5).
So 5 is inert, and [h5i] = 1 in cl
L
. So
cl
L
= h[p], [q]i,
and we need to compute what this is. We can just compute powers
q
2
, q
3
, ···
,
pq, pq
2
, ···, and see what happens.
But a faster way is to look for principal ideals with small norms that are
multiples of 2 and 3. For example,
N(h1 +
di) = 18 = 2 · 3
2
.
But we have
1 +
d p, q.
So
p, q | h
1 +
di
. Thus we know
pq | h
1 +
di
. We have
N
(
pq
) = 2
·
3 = 6. So
there is another factor of 3 to account for. In fact, we have
h1 +
di = pq
2
,
which we can show by either thinking hard or expanding it out. So we must have
[p] = [q]
2
in
cl
L
. So we have
cl
L
=
h
[
q
]
i
. Also, [
q
]
2
= [
p
]
6
= 1 in
cl
L
, as if it did, then
p
is
principal, i.e.
p
=
hx
+
y
di
, but 2 =
N
(
p
) =
x
2
+ 7
y
2
has no solution in the
integers. Also, we know [p]
2
= [1]. So we know
cl
L
= Z/4Z.
In fact, we have
Theorem. Let L = Q(
d) with d < 0. Then O
L
is a UFD if
d {1, 2, 3, 7, 11, 19, 43, 67, 163}.
Moreover, this is actually an “if and only if”.
The first part is a straightforward generalization of what we have been doing,
but the proof that no other d’s work is hard.
General case
Now we want to extend these ideas to higher dimensions. We are really just
doing the same thing, but we need a bit harder geometry and proper definitions.
Definition
(Discrete subset)
.
A subset
X R
n
is discrete if for every
x X
,
there is some
ε >
0 such that
B
ε
(
x
)
X
=
{x}
. This is true if and only if for
every compact K R
n
, K X is finite.
We have the following very useful characterization of discrete subgroups of
R
n
:
Proposition.
Suppose Λ
R
n
is a subgroup. Then Λ is a discrete subgroup of
(R
n
, +) if and only if
Λ =
(
m
X
1
n
i
x
i
: n
i
Z
)
for some x
1
, ··· , x
m
linearly independent over R.
Note that linear independence is important. For example,
Z
2
+
Z
3 R
is not discrete. On the other hand, if Λ =
a C O
L
is an ideal, where
L
=
Q
(
d
)
and d < 0, then this is discrete.
Proof.
Suppose Λ is generated by
x
1
, ··· , x
m
. By linear independence, there
is some
g GL
n
(
R
) such that
gx
i
=
e
i
for all 1
i m
, where
e
1
, ··· , e
n
is
the standard basis. We know acting by
g
preserves discreteness, since it is a
homeomorphism, and
g
Λ =
Z
m
R
m
× R
nm
is clearly discrete (take
ε
=
1
2
).
So this direction follows.
For the other direction, suppose Λ is discrete. We pick
y
1
, ··· , y
m
Λ
which are linearly independent over
R
, with
m
maximal (so
m n
). Then by
maximality, we know
(
m
X
i=1
λ
i
y
i
: λ
i
R
)
=
(
m
X
1
λ
i
z
i
: λ
i
R, z
i
Λ
)
,
and this is the smallest vector subspace of R
n
containing Λ. We now let
X =
(
m
X
i=1
λ
i
y
i
: λ
i
[0, 1]
)
=
[0, 1]
m
.
This is closed and bounded, and hence compact. So X Λ is finite.
Also, we know
M
Zy
i
= Z
m
Λ,
and if γ is any element of Λ, we can write it as γ = γ
0
+ γ
1
, where γ
0
X and
γ
1
Z
m
. So
Λ
Z
m
|X Λ| < .
So let d = |Λ/Z
m
|. Then dΛ Z
m
, i.e. Λ
1
d
Z
m
. So
Z
m
Λ
1
d
Z
m
.
So Λ is a free abelian group of rank
m
. So there exists
x
1
, ··· , x
m
1
d
Z
m
which
is an integral basis of Λ and are linearly independent over R.
Definition (Lattice). If rank Λ = n = dim R
n
, then Λ is a lattice in R
n
.
Definition
(Covolume and fundamental domain)
.
Let Λ
R
n
be a lattice, and
x
1
, ··· , x
n
be a basis of Λ, then let
P =
(
n
X
i=1
λ
i
x
i
: λ
i
[0, 1]
)
,
and define the covolume of Λ to be
covol(Λ) = vol(P ) = |det A|,
where A is the matrix such that x
i
=
P
a
ij
e
j
.
We say P is a fundamental domain for the action of Λ on R
n
, i.e.
R
n
=
[
γΛ
(γ + P ),
and
(γ + P ) (µ + P ) (γ + P ).
In particular, the intersection has zero volume.
This is called the covolume since if we consider the space
R
n
/
Λ, which is an
n-dimensional torus, then this has volume covol(Λ).
Observe now that if
x
0
1
, ··· , x
0
n
is a different basis of Λ, then the transition
matrix
x
0
i
=
P
b
ij
x
j
has
B GL
n
(
Z
). So we have
det B
=
±
1, and
covol
(Λ) is
independent of the basis choice.
With these notations, we can now state Minkowski’s theorem.
Theorem
(Minkowski’s theorem)
.
Let Λ
R
n
be a lattice, and
P
be a funda-
mental domain. We let
S R
n
be a measurable set, i.e. one for which
vol
(
S
) is
defined.
(i)
Suppose
vol
(
S
)
> covol
(Λ). Then there exists distinct
x, y S
such that
x y Λ.
(ii)
Suppose
0 S
, and
S
is symmetric around 0, i.e.
s S
if and only if
s S, and S is convex, i.e. for all x, y S and λ [0, 1], then
λx + (1 λ)y S.
Then suppose either
(a) vol(S) > 2
n
covol(Λ); or
(b) vol(S) 2
n
covol(Λ) and S is closed.
Then S contains a γ Λ with γ 6= 0.
Note that for n = 2, this is what we used for quadratic fields.
By considering Λ =
Z
n
R
n
and
S
= [
1
,
1]
n
, we know the bounds are
sharp.
Proof.
(i)
Suppose
vol
(
S
)
> covol
(Λ) =
vol
(
P
). Since
P R
n
is a fundamental
domain, we have
vol(S) = vol(S R
n
) = vol
S
X
γΛ
(P + γ)
=
X
γΛ
vol(S (P + γ)).
Also, we know
vol(S (P + γ)) = vol((S γ) P ),
as volume is translation invariant. We now claim the sets (
S γ
)
P
for
γ Λ are not pairwise disjoint. If they were, then
vol(P )
X
γΛ
vol((S γ) P ) =
X
γΛ
vol(S (P + γ)) = vol(S),
contradicting our assumption.
Then in particular, there are some distinct
γ
and
µ
such that (
S γ
) and
(
S µ
) are not disjoint. In other words, there are
x, y S
such that
x γ = y µ, i.e. x y = γ µ Λ 6= 0.
(ii) We now let
S
0
=
1
2
S =
1
2
s : s S
.
So we have
vol(S
0
) = 2
n
vol(S) > covol(Λ),
by assumption.
(a)
So there exists some distinct
y, z S
0
such that
y z
Λ
\ {
0
}
. We
now write
y z =
1
2
(2y + (2z)),
Since 2
z S
implies
2
z S
by symmetry around
0
, so we know
y z S by convexity.
(b)
We apply the previous part to
S
m
=
1 +
1
m
S
for all
m N
,
m >
0.
So we get a non-zero γ
m
S
m
Λ.
By convexity, we know
S
m
S
1
= 2
S
for all
m
. So
γ
1
, γ
2
, ··· S
1
Λ.
But
S
1
is compact set. So
S
1
Λ is finite. So there exists
γ
such that
γ
m
is γ infinitely often. So
γ
\
m0
S
m
= S.
So γ S.
We are now going to use this to mimic our previous proof that the class
group of an imaginary quadratic field is finite.
To begin with, we need to produce lattices from ideals of
O
L
. Let
L
be a
number field, and [
L
:
Q
] =
n
. We let
σ
1
, ··· , σ
r
:
L R
be the real embeddings,
and
σ
r+1
, ··· , σ
r+s
, ¯σ
r+1
, ··· , ¯σ
r+s
:
L C
be the complex embeddings (note
that which embedding is σ
r+i
and which is ¯σ
r+i
is an arbitrary choice).
Then this defines an embedding
σ = (σ
1
, σ
2
, ··· , σ
r
, σ
r+1
, ··· , σ
r+s
) : L R
r
× C
s
=
R
r
× R
2s
= R
r+2s
= R
n
,
under the isomorphism C R
2
by x + iy 7→ (x, y).
Just as we did for quadratic fields, we can relate the norm of ideals to their
covolume.
Lemma.
(i) σ(O
L
) is a lattice in R
n
of covolume 2
s
|D
L
|
1
2
.
(ii)
More generally, if
aC O
L
is an ideal, then
σ
(
a
) is a lattice and the covolume
covol(σ(a)) = 2
s
|D
L
|
1
2
N(a).
Proof.
Obviously (ii) implies (i). So we just prove (ii). Recall that
a
has an
integral basis γ
1
, ··· , γ
n
. Then a is the integer span of the vectors
(σ
1
(γ
i
), σ
2
(γ
i
), ··· , σ
r+s
(γ
i
))
for
i
= 1
, ··· , n
, and they are independent as we will soon see when we compute
the determinant. So it is a lattice.
We also know that
∆(γ
1
, ··· , γ
n
) = det(σ
i
(γ
j
))
2
= N(a)
2
D
L
,
where the σ
i
run over all σ
1
, ··· , σ
r
, σ
r+1
, ··· , σ
r+s
, ¯σ
r+1
, ··· ¯σ
r+s
.
So we know
|det(σ
i
(γ
j
))| = N (a)|D
L
|
1
2
.
So what we have to do is to relate
det
(
σ
i
(
γ
j
)) to the covolume of
σ
(
a
). But
these two expressions are very similar.
In the σ
i
(γ
j
) matrix, we have columns that look like
σ
r+i
(γ
j
) ¯σ
r+i
(γ
j
)
=
z ¯z
.
On the other hand, the matrix of σ(γ) has corresponding entries
Re(z) Im(z)
=
1
2
(z + ¯z)
i
2
(¯z z)
=
1
2
1 1
i i
z
¯z
We call the last matrix A =
1
2
1 1
i i
. We can compute the determinant as
|det A| =
det
1
2
1 1
i i
=
1
2
.
Hence the change of basis matrix from (
σ
i
(
γ
j
)) to
σ
(
γ
) is
s
diagonal copies of
A
,
so has determinant 2
s
. So this proves the lemma.
Proposition.
Let
a C O
L
be an ideal. Then there exists an
α a
with
α 6
= 0
such that
|N(α)| c
L
N(a),
where
c
L
=
4
π
s
n!
n
n
|D
L
|
1
2
.
This is the Minkowski bound.
Proof. Let
B
r,s
(t) =
n
(y
1
, ··· , y
r
, z
1
, ··· , z
s
) R
r
× C
s
:
X
|y
i
| + 2
X
|z
i
| t
o
.
This
(i) is closed and bounded;
(ii) is measurable (it is defined by polynomial inequalities);
(iii) has volume
vol(B
r,s
(t)) = 2
r
π
2
s
t
n
n!
;
(iv) is convex and symmetric about 0.
Only (iii) requires proof, and it is on the second example sheet, i.e. we are not
doing it here. It is just doing the integral.
We now choose t so that
vol B
r,s
(t) = 2
n
covol(σ(a)).
Explicitly, we let
t
n
=
4
π
s
n!|D
L
|
1/2
N(a).
Then by Minkowski’s lemma, there is some
α a
non-zero such that
σ
(
α
)
B
r,s
(t). We write
σ(α) = (y
1
, ··· , y
r
, z
1
, ··· , z
s
).
Then we observe
N(α) = y
1
···y
r
z
1
¯z
1
z
2
¯z
2
···z
s
¯z
s
=
Y
y
i
Y
|z
j
|
2
.
By the AM-GM inequality, we know
|N(α)|
1/n
1
n
X
y
i
+ 2
X
|z
j
|
t
n
,
as we know σ(a) B
r,s
(t). So we get
|N(α)|
t
n
n
n
= c
L
N(a).
Corollary. Every [a] cl
L
has a representative a O
L
with N (a) c
L
.
Theorem
(Dirichlet)
.
The class group
cl
L
is finite, and is generated by prime
ideals of norm c
L
.
Proof. Just as the case for imaginary quadratic fields.
7 Dirichlet’s unit theorem
We have previously characterized the units on
O
L
as the elements with unit
norm, i.e.
α O
L
is a unit if and only if
|N
(
α
)
|
= 1. However, this doesn’t tell
us much about how many units there are, and how they are distributed. The
answer to this question is given by Dirichlet’s unit theorem.
Theorem (Dirichlet unit theorem). We have the isomorphism
O
×
L
=
µ
L
× Z
r+s1
,
where
µ
L
= {α L : α
N
= 1 for some N > 0}
is the group of roots of unity in L, and is a finite cyclic group.
Just as in the finiteness of the class group, we do it for an example first, or
else it will be utterly incomprehensible.
We do the example of real quadratic fields,
L
=
Q
(
d
), where
d >
1 is
square-free. So r = 2, s = 0, and L R implies µ
L
= 1}. So
O
×
L
=
1} × Z.
Also, we know that
N(x + y
d) = (x + y
d)(x y
d) = x
2
dy
2
.
So Dirichlet’s theorem is saying that there are infinitely many solutions of
x
2
dy
2
= ±1, and are all (plus or minus) the powers of one single element.
Theorem
(Pell’s equation)
.
There are infinitely many
x
+
y
d O
L
such that
x
2
dy
2
= ±1.
You might have seen this in IIC Number Theory, where we proved it directly
by continued fractions. We will provide a totally unconstructive proof here, since
this is more easily generalized to arbitrary number fields.
This is actually just half of Dirichlet’s theorem. The remaining part is to
show that they are all powers of the same element.
Proof. Recall that σ : O
L
R
2
sends
α = x + y
d 7→ (σ
1
(α), σ
2
(α)) = (x + y
d, x y
d).
(in the domain,
d
is a formal symbol, while in the codomain, it is a real number,
namely the positive square root of d)
Also, we know
covol(σ(O
L
)) = |D
L
|
1
2
.
(1, 1)
Z[X]
N(α) = 1
Consider
s
t
=
(y
1
, y
2
) R
2
: |y
1
| t, |y
2
|
|D
L
|
1/2
t
.
So
vol(s
t
) = 4|D
L
|
1
2
= 2
n
covol(O
L
),
as
n
= [
L
:
Q
] = 2. Now Minkowski implies there is an
α O
L
non-zero such
that σ(α) s
t
. Also, if we write
σ(α) = (y
1
, y
2
),
then
N(α) = y
1
y
2
.
So such an α will satisfy
1 |N (α)| |D
L
|
1/2
.
This is not quite what we want, since we need
|N
(
α
)
|
= 1 exactly. Nevertheless,
this is a good start. So let’s try to find infinitely such elements.
First notice that no points on the lattice (apart from the origin) hits the
x
or
y
axis, since any such point must satisfy
x ± y
d
= 0, but
d
is not rational.
Also,
s
t
is compact. So
s
t
σ
(
O
L
) contains finitely many points. So we can
find a
t
2
such that for each (
y
1
, y
2
)
s
t
O
L
, we have
|y
1
| > t
2
. In particular,
s
t
2
does not contain any point in
s
t
σ
(
O
L
). So we get a new set of points
α s
t
2
O
L
such that 1 |N(α)| |D
L
|
1/2
.
s
1
s
2
We can do the same thing for
s
t
2
and get a new
t
3
. In general, given
t
1
> ··· > t
n
,
pick t
n+1
be such that
0 < t
n+1
< min
(
|y
1
| : (y
1
, y
2
)
n
[
i=1
s
t
i
σ(O
L
)
)
,
and the minimum is finite since
s
t
is compact and hence contains finitely many
lattice points on σ(O
L
).
Then we get an infinite sequence of
t
i
such that
s
t
i
σ
(
O
L
) are disjoint for
different
i
. Since each must contain at least one point, we have got infinitely
many points in O
L
satisfying 1 |N (α)| |D
L
|
1/2
.
Since there are only finitely many integers between 1 and
|D
L
|
1/2
, we can
apply the pigeonhole principle, and get that there is some integer satisfying
1
|m| |D
L
|
1/2
such that there exists infinitely many
α O
L
with
N
(
α
) =
m
.
This is not quite good enough. We consider
O
L
/mO
L
=
(Z/mZ)
[L:Q]
,
another finite set. We notice that each
α O
L
must fall into one of finitely
many the cosets of
mO
L
in
O
L
. In particular, each
α
such that
N
(
α
) =
m
must
belong to one of these cosets.
So again by the pigeonhole principle, there exists a
β O
L
with
N
(
β
) =
m
,
and infinitely many α O
L
with N (α) = m and α = β (mod mO
L
).
Now of course
α
and
β
are not necessarily units, if
m 6
= 1. However, we will
show that
α/β
is. The hard part is of course showing that it is in
O
L
itself,
because it is clear that
α/β
has norm 1 (alternatively, by symmetry,
β
is in
O
L
, so an inverse exists).
Hence all it remains is to prove the general fact that if
α = β + ,
where α, β, γ O
L
and N (α) = N(β) = m, then α/β O
L
.
To show this, we just have to compute
α
β
= 1 +
m
β
γ = 1 +
N(β)
β
γ = 1 +
¯
βγ O
L
,
since N (β) = β
¯
β. So done.
We have thus constructed infinitely many units.
We now prove the remaining part
Theorem
(Dirchlet’s unit theorem for real quadratic fields)
.
Let
L
=
Q
(
d
).
Then there is some ε
0
O
×
L
such that
O
×
L
= ε
n
0
: n Z}.
We call such an ε
0
a fundamental unit (which is not unique). So
O
×
L
=
1} × Z.
Proof.
We have just proved the really powerful theorem that there are infinitely
many
ε
with
N
(
ε
) = 1. We are not going to need the full theorem. All we need
is that there are three in particular, something that is not ±1.
We pick some
ε O
×
L
with
ε 6
=
±
1. This exists by what we just proved.
Then we know
|σ
1
(ε)| 6= 1,
as
|σ
1
(
ε
)
|
= 1 if and only if
ε
=
±
1. Replacing by
ε
1
if necessary, we wlog
E = |σ
1
(ε)| > 1. Now consider
{α O
L
: N(α) = ±1, 1 |σ
1
(α)| E}.
This is again finite, since it is specified by a compact subset of the
O
L
-lattice.
So we pick
ε
0
in this set with
ε
0
6
=
±
1 and
|σ
1
(
ε
0
)
|
minimal (
>
1). Replacing
ε
0
by ε
0
if necessary, we can assume σ
1
(ε) > 1.
Finally, we claim that if
ε O
×
L
and
σ
1
(
ε
)
>
0, then
ε
=
ε
N
0
for some
N Z
.
This is obvious if we have addition instead of multiplication. So we take logs.
Suppose
log ε
log ε
0
= N + γ,
where N Z and 0 γ < 1. Then we know
εε
N
0
= ε
γ
0
O
×
L
,
but
|ε
γ
0
|
=
|ε
0
|
γ
< |ε
0
|
, as
|ε
0
| >
1. By our choice of
ε
0
, we must have
γ
= 0. So
done.
Now we get to prove the Dirichlet unit theorem in its full glory.
Theorem (Dirichlet unit theorem). We have the isomorphism
O
×
L
=
µ
L
× Z
r+s1
,
where
µ
L
= {α L : α
N
= 1 for some N > 0}
is the group of roots of unity in L, and is a finite cyclic group.
Proof.
We do the proof in the opposite order. We throw in the logarithm at the
very beginning. We define
: O
×
L
R
r+s
by
x 7→ (log |σ
1
(x)|, ··· , log |σ
r
(x)|, 2 log |σ
r+1
(x)|, ··· , 2 log |σ
r+s
(x)|).
Note that
|σ
r+i
(
x
)
|
=
|σ
r+`
(x)|
. So this is independent of the choice of one of
σ
r+i
, ¯σ
r+i
.
Claim.
We now claim that
im
is a discrete group in
R
r+s
and
ker
=
µ
L
is a
finite cyclic group.
We note that
log |ab| = log |a| + log |b|.
So this is a group homomorphism, and the image is a subgroup. To prove the
first part, it suffices to show that
im
[
A, A
]
r+s
is finite for all
A >
0. We
notice factors as
O
×
L
O
L
R
r
× C
s
R
r+s
.
σ
j
where σ maps α 7→ (σ
1
(α), ··· , σ
r+s
(α)), and
j : (y
1
, ··· , y
r
, z
1
, ··· , z
s
) 7→ (log |y
1
|, ··· , log |y
r
|, 2 log |z
1
|, ··· , 2 log |z
2
|).
We see
j
1
([A, A]
r+s
) = {(y
i
, z
j
) : e
A
|y
i
| e
A
, e
A
2|z
j
| e
A
}
is a compact set, and
σ
(
O
L
) is a lattice, in particular discrete. So
σ
(
O
L
)
j
1
([
A, A
]
r+s
) is finite. This also shows the kernel is finite, since the kernel is
the inverse image of a compact set.
Now as
ker
is finite, all elements are of finite order. So
ker µ
L
. Con-
versely, it is clear that
µ
L
ker
. So it remains to show that
µ
L
is cyclic. Since
L
embeds in
C
, we know
µ
L
is contained in the roots of unity in
C
. Since
µ
L
is
finite, we know
L
is generated by a root of unity with the smallest argument
(from, say, IA Groups).
Claim. We claim that
im
n
(y
1
, ··· , y
r+s
) :
X
y
i
= 0
o
=
R
r+s1
.
To show this, note that if α O
×
L
, then
N(α) =
n
Y
i=1
σ
i
(α)
s
Y
`=1
σ
r+`
(α)¯σ
r+`
= ±1.
Taking the log of the absolute values, we get
0 =
X
log |σ
i
(α)| + 2
X
log |σ
r+i
(α)|.
So we know
im R
r+s1
as a discrete subgroup. So it is isomorphic to
Z
a
for some
a r
+
s
1. Then what we want to show is that
im R
r+s1
is a
lattice, i.e. it is congruent to Z
r+s1
.
Note that so far what we have done is the second part of what we did for the
real quadratic fields. We took the logarithm to show that these form a discrete
subgroup. Next, we want to find
r
+
s
1 independent elements to show it is a
lattice.
Claim.
Fix a
k
such that 1
k r
+
s
and
α O
L
with
α 6
= 0. Then there
exists a β O
L
such that
|N(β)|
2
π
s
|D
L
|
1/2
,
and moreover if we write
(α) = (a
1
, ··· , a
r+s
)
(β) = (b
1
, ··· , b
r+s
),
then we have b
i
< a
i
for all i 6= k.
We can apply Minkowski to the region
S = {(y
1
, ··· , y
r
, z
1
, ··· , z
s
) R
r
× C
s
: |y
i
| c
i
, |z
j
| c
r+j
}
(we will decide what values of c
i
to take later). Then this has volume
vol(S) = 2
r
π
s
c
1
···c
r+s
.
We notice
S
is convex and symmetric around 0. So if we choose 0
< c
i
< e
a
i
for
i 6= k, and choose
c
k
=
2
π
s
|D
L
|
1/2
1
c
1
···ˆc
k
···c
r+s
.
Then Minkowski gives β σ(O
L
) S, satisfying the two conditions above.
Claim.
For any
k
= 1
, ··· , r
+
s
, there is a unit
u
k
O
×
L
such that if
(
u
k
) =
(y
1
, ··· , y
r+s
, then y
i
< 0 for all i 6= k (and hence y
k
> 0 since
P
y
i
= 0).
This is just as in the proof for the real quadratic case. We can repeatedly
apply the previous claim to get a sequence
α
1
, α
2
, ··· O
L
such that
N
(
α
t
)
is bounded for all
t
, and for all
i 6
=
k
, the
i
th coordinate of
(
α
1
)
,
(
α
2
)
, ···
is strictly decreasing. But then as with real quadratic fields, the pigeonhole
principle implies we can find t, t
0
such that
N(α
t
) = N(α
t
0
) = m,
say, and
α
t
α
t
0
(mod mO
L
),
i.e.
α
t
=
α
t
0
in
O
L
/mO
L
. Hence for each
k
, we get a unit
u
k
=
α
t
t
0
such that
(u
k
) = (α
t
) (α
0
t
) = (y
1
, ··· , y
r+s
)
has
y
i
<
0 if
i 6
=
k
(and hence
y
k
>
0, since
P
y
i
= 0). We need a final trick to
show the following:
Claim.
The units
u
1
, ··· , u
r+s1
are linearly independent in
R
r+s1
. Hence
the rank of (O
×
L
) = r + s 1, and Dirichlet’s theorem is proved.
We let
A
be the (
r
+
s
)
×
(
r
+
s
) matrix whose
j
th row is
(
u
j
), and apply
the following lemma:
Claim.
Let
A Mat
m
(
R
) be such that
a
ii
>
0 for all
i
and
a
ij
<
0 for all
i 6
=
j
,
and
P
j
a
ij
0 for each i. Then rank(A) m 1.
To show this, we let
v
i
be the
i
th column of
A
. We show that
v
1
, ··· , v
m1
are linearly independent. If not, there exists a sequence t
i
R such that
m1
X
i=1
t
i
v
i
= 0, ()
with not all of the
t
i
non-zero. We choose
k
so that
|t
k
|
is maximal among the
t
1
, ··· , t
m1
’s. We divide the whole equation by
t
k
. So we can wlog assume
t
k
= 1, t
i
1 for all i.
Now consider the kth row of (). We get
0 =
m1
X
i=1
t
i
a
ki
m1
X
i=1
a
ki
,
as
a <
0 and
t
1 implies
at a
. Moreover, we know
a
mi
>
0 strictly. So we
get
0 >
m
X
i=1
a
ki
0.
This is a contradiction. So done.
You should not expect this to be examinable.
We make a quick definition that we will need later.
Definition (Regulator). The regulator of a number field L is
R
L
= covol((O
×
L
) R
r+s1
).
More concretely, we pick fundamental units ε
1
, ··· , ε
r+s1
O
×
L
so that
O
×
L
= µ
L
× {ε
n
1
1
···ε
n
r+s1
r+s1
: n
i
Z}.
We take any (
r
+
s
1)(
r
+
s
1) subminor of the matrix
(ε
1
) ··· (ε
r+s
)
.
Their determinants all have the same absolute value, and
|det(subminor)| = R
L
.
This is a definition we will need later.
We quickly look at some examples with quadratic fields. Consider
L
=
Q
(
d
),
where d 6= 0, 1 square-free.
Example.
If
d <
0, then
r
= 0 and
s
= 1. So
r
+
s
1 = 0. So
O
×
L
=
µ
L
is a
finite group. So R
L
= 1.
Lemma.
(i) If d = 1, then Z[i]
×
= 1, ±i} = Z/4Z.
(ii)
If
d
=
3, then let
ω
=
1
2
(1 +
d
), and we have
ω
6
= 1. So
Z
[
ω
]
×
=
{1, ω, ··· , ω
5
}
=
Z/6Z.
(iii) For any other d < 0, we have O
×
L
= 1}.
Proof. This is just a direct check.
If
d
2
,
3 (
mod
4), then by looking at the solution of
x
2
dy
2
=
±
1 in the
integers, we get (i) and (iii).
If
d
1 (
mod
4), then by looking at the solutions to
x +
y
2
2
d
4
y
2
=
±
1
in the integers, we get (ii) and (iii).
Now if
d >
0, then
R
L
=
|log |ε||
, where
ε
is a fundamental unit. So how do
we find a fundamental unit? In general, there is no good algorithm for finding the
fundamental unit of a fundamental field. The best algorithm takes exponential
time. We do have a good algorithm for quadratic fields using continued fractions,
but we are not allowed to use that.
Instead, we could just guess a solution we find a unit by guessing, and
then show there is no smaller one by direct check.
Example.
Consider the field
Q
(
2
). We can try
ε
= 1 +
2
. We have
N
(
ε
) = 1
2 =
1. So this is a unit. We claim this is fundamental. If not,
there there exists
u
=
a
+
b
2
, where
a, b Z
and 1
< u < ε
(as real numbers).
Then we have
¯u = a b
2
has u¯u = ±1. Since u > 1, we know |¯u| < 1. Then we must have u ± ¯u > 0. So
we need a, b > 0. We know can only be finitely many possibilities for
1 < a + b
2 < 1 +
2,
where a, b are positive integers. But there actually are none. So done.
Example.
Consider
Q
(
11
). We guess
ε
= 10
3
11
is a unit. We can
compute N (ε) = 100 99 = 1. Note that ε < 1 and ε
1
> 1.
Suppose this is not fundamental. Then we have some u such that
1 < u = a + b
11 < 10 + 3
11 = ε
1
< 20. ()
We can check all the cases, but there is a faster way.
We must have
N
(
u
) =
±
1. If
N
(
u
) =
1, then
a
2
11
b
2
=
1. But
1 is
not a square mod 11.
So there we must have
N
(
u
) = 1. Then
u
1
=
¯u
. We get 0
< ε < u
1
=
¯u <
1
also. So
1 < a + b
11 < 0.
Adding this to (), we get
0 < 2b
11 < 10 + 3
11 < 7
11.
So b = 1, 2 or 3, but 11b
2
+ 1 is not a square for each of these. So done.
8 L-functions, Dirichlet series*
This section is non-examinable.
We start by proving the exciting fact that there are infinitely many primes.
Theorem (Euclid). There are infinitely many primes.
Proof. Consider the function
Y
p primes
1
1
p
1
=
Y
p prime
1 +
1
p
+
1
p
2
+ ···
=
X
n>0
1
n
.
This is since every
n
=
p
e
1
1
···p
e
r
r
factors uniquely as a product of primes, and
each such product appears exactly once in this. If there were finitely many
primes, as
P
1
p
n
converges to
1
1
p
1
, the sum
X
n1
1
n
=
Y
p prime
1
1
p
must be finite. But the harmonic series diverges. This is a contradiction.
We all knew that. What we now want to prove is something more interesting.
Theorem
(Dirichlet’s theorem)
.
Let
a, q Z
be coprime. Then there exists
infinitely many primes in the sequence
a, a + q, a + 2q, ··· ,
i.e. there are infinitely many primes in any such arithmetic progression.
We want to imitate the Euler proof, but then that would amount to showing
that
Y
pa mod q
p prime
1
1
p
1
is divergent, and there is no nice expression for this. So it will be a lot more
work.
To begin with, we define the Riemann zeta function.
Definition
(Riemann zeta function)
.
The Riemann zeta function is defined as
ζ(s) =
X
n1
n
s
for s C.
There are some properties we will show (or assert):
Proposition.
(i) The Riemann zeta function ζ(s) converges for Re(s) > 1.
(ii) The function
ζ(s)
1
s 1
extends to a holomorphic function when Re(s) > 0.
In other words,
ζ
(
s
) extends to a meromorphic function on
Re
(
s
)
>
0 with
a simple pole at 1 with residue 1.
(iii) We have the expression
ζ(s) =
Y
p prime
1
1
p
s
1
for
Re
(
s
)
>
1, and the product is absolutely convergent. This is the Euler
product.
The first part follows from the following general fact about Dirichlet series.
Definition
(Dirichlet series)
.
A Dirichlet series is a series of the form
P
a
n
n
s
,
where a
1
, a
2
, ··· C.
Lemma. If there is a real number r R such that
a
1
+ ··· + a
N
= O(N
r
),
then
X
a
n
n
s
converges for Re(s) > r, and is a holomorphic function there.
Then (i) is immediate by picking
r
= 1, since in the Riemann zeta function,
a
1
= a
2
= ··· = 1.
Recall that x
s
= e
s log x
has
|x
s
| = |x
Re(s)
|
if x R, x > 0.
Proof. This is just IA Analysis. Suppose Re(s) > r. Then we can write
N
X
n=1
a
n
n
s
= a
1
(1
s
2
s
) + (a
1
+ a
2
)(2
s
3
s
) + ···
+ (a
1
+ ··· + a
N1
)((N 1)
s
N
s
) + R
N
,
where
R
N
=
a
1
+ ··· + a
N
N
s
.
This is getting annoying, so let’s write
T (N) = a
1
+ ··· + a
N
.
We know
T (N)
N
s
=
T (N)
N
r
1
N
Re(s)r
0
as N , by assumption. Thus we have
X
n1
a
n
n
s
=
X
n1
T (n)(n
s
(n + 1)
s
)
if
Re
(
s
)
> r
. But again by assumption,
T
(
n
)
B · n
r
for some constant
B
and
all n. So it is enough to show that
X
n
n
r
(n
s
(n + 1)
s
)
converges. But
n
s
(n + 1)
s
=
Z
n+1
n
s
x
s+1
dx,
and if x [n, n + 1], then n
r
x
r
. So we have
n
r
(n
s
(n + 1)
s
)
Z
n+1
n
x
r
s
x
s+1
dx = s
Z
n+1
n
dx
x
s+1r
.
It thus suffices to show that
Z
n
1
dx
x
s+1r
converges, which it does (to
s
sr
).
We omit the proof of (ii). The idea is to write
1
s 1
=
X
n=1
Z
n+1
n
dx
x
s
,
and show that
P
φ
n
is uniformly convergent when Re(s) > 0, where
φ
n
= n
s
Z
n+1
n
dx
x
s
.
For (iii), consider the first r primes p
1
, ··· , p
r
, and
r
Y
i=1
(1 p
s
i
)
1
=
X
n
s
,
where the sum is over the positive integers
n
whose prime divisors are among
p
1
, ··· , p
r
. Notice that 1, ··· , r are certainly in the set.
So
ζ(s)
r
Y
i=1
(1 p
s
i
)
1
X
nr
|n
s
| =
X
nr
n
Re(s)
.
But
P
nr
n
Re(s)
0 as
r
, proving the result, if we also show that it
converges absolutely. We omit this proof, but it follows from the fact that
X
p prime
p
s
X
n
n
s
.
and the latter converges absolutely, plus the fact that
Q
(1
a
n
) converges if
and only if
P
a
n
converges, by IA Analysis I.
This is good, but not what we want. Let’s mimic this definition for an
arbitrary number field!
Definition
(Zeta function)
.
Let
L Q
be a number field, and [
L
:
Q
] =
n
. We
define the zeta function of L by
ζ
L
(s) =
X
aCO
L
N(a)
s
.
It is clear that if
L
=
Q
and
O
L
=
Z
, then this is just the Riemann zeta
function.
Theorem.
(i) ζ
L
(s) converges to a holomorphic function if Re(s) > 1.
(ii)
Analytic class number formula:
ζ
L
(
s
) is a meromorphic function if
Re
(
s
)
>
1
1
n
and has a simple pole at s = 1 with residue
|cl
L
|2
r
(2π)
s
R
L
|D
L
|
1/2
|µ
L
|
,
where
cl
L
is the class group,
r
and
s
are the number of real and complex
embeddings, you know what
π
is,
R
L
is the regulator,
D
L
is the discriminant
and µ
L
is the roots of unity in L.
(iii)
ζ
L
(s) =
Y
pCO
L
prime ideal
(1 N(p)
s
)
1
.
This is again known as the Euler product.
We will not prove this, but the proof does not actually require any new ideas.
Note that
X
aCO
L
N(a)
s
=
Y
pCO
L
,p prime
(1 N(p)
s
)
1
holds “formally”, as in the terms match up when you expand, as an immediate
consequence of the unique factorization of ideals into a product of prime ideals.
The issue is to study convergence of
P
N
(
a
)
s
, and this comes down to estimating
the number of ideals of fixed norm geometrically, and that is where all the factors
in the pole come it.
Example.
We try to compute
ζ
L
(
s
), where
L
=
Q
(
d
). This has discriminant
D, which may be d or 4d. We first look at the prime ideals.
If
p
is a prime ideal in
O
L
, then
p | hpi
for a unique
p
. So let’s enumerate
the factors of η
L
controlled by p Z.
Now if
p | |D
L
|
, then
hpi
=
p
2
ramifies, and
N
(
p
) =
p
. So this contributes a
factor of (1 p
s
)
1
.
Now if p remains prime, then we have N (hpi) = p
2
. So we get a factor of
(1 p
2s
)
1
= (1 p
s
)
1
(1 + p
s
)
1
.
If p splits completely, then
hpi = p
1
p
2
.
So
N(p
i
) = p,
and so we get a factor of
(1 p
s
)
1
(1 p
s
)
1
.
So we find that
ζ
L
(s) = ζ(s)L(χ
D
, s),
where we define
Definition (L-function). We define the L-function by
L(χ, s) =
Y
p prime
(1 χ(p)p
s
)
1
.
In our case, χ is given by
χ
D
(p) =
0 p | D
1 p remains prime
1 p splits
=
(
D
p
p is odd
depends on d mod 8 p = 2
.
Example. If L = Q(
1), then we know
4
p
=
1
p
= (1)
p1
2
if p 6= 2,
and χ
D
(2) = 0 as 2 ramifies. We then have
L(χ
D
, s) =
Y
p>2 prime
(1 (1)
p1
2
p
s
)
1
= 1
1
3
s
+
1
5
s
1
7
s
+ ··· .
Note that
χ
D
was defined for primes only, but we can extend it to a function
χ
D
: Z C by imposing
χ
D
(nm) = χ
D
(n)χ
D
(m),
i.e. we define
χ
D
(p
e
1
1
···p
e
r
r
) = χ
D
(p
1
)
e
1
···χ
D
(p
r
)
e
r
.
Example. Let L = Q(
1). Then
χ
4
(m) =
(
(1)
m1
2
m odd
0 m even.
It is an exercise to show that this is really the extension, i.e.
χ
4
(mn) = χ
4
(m)χ
4
(n).
Notice that this has the property that
χ
4
(m 4) = χ
4
(m).
We give these some special names
Definition
(Dirichlet character)
.
A function
χ
:
Z C
is a Dirichlet character
of modulus D if there exists a group homomorphism
w :
Z
DZ
×
C
×
such that
χ(m) =
(
w(m mod D) gcd(m, D) = 1
0 otherwise
.
We say χ is non-trivial if ω is non-trivial.
Example. χ
4
is a Dirichlet character of modulus 4.
Note that
χ(mn) = χ(m)χ(n)
for such Dirichlet characters, and so
L(χ, s) =
Y
p prime
(1 χ(p)p
s
)
1
=
X
n1
χ(n)
n
s
for such χ.
Proposition. χ
D
, as defined for
L
=
Q
(
d
) is a Dirichlet character of modulus
D.
Note that this is a very special Dirichlet character, as it only takes values
0, ±1. We call this a quadratic Dirichlet character.
Proof. We must show that
χ
D
(p + Da) = χ
D
(p)
for all p, a.
(i) If d 3 (mod 4), then D = 4d. Then
χ
D
(2) = 0,
as (2) ramifies. So χ
D
(even) = 0. For p > 2, we have
χ
D
(p) =
D
p
=
d
p
=
p
d
(1)
p1
2
as
d1
2
1 (mod 2), by quadratic reciprocity. So
χ
D
(p + Da) =
p + Da
d
(1)
p1
2
(1)
4da/2
= χ
D
(p).
(ii) If d 1, 2 (mod 4), see example sheet.
Lemma.
Let
χ
be any non-trivial Dirichlet character. Then
L
(
χ, s
) is holomor-
phic for Re(s) > 0.
Proof. By our lemma on convergence of Dirichlet series, we have to show that
N
X
i=1
χ(i) = O(1),
i.e. it is bounded. Recall from Representation Theory that distinct irreducible
characters of a finite group G are orthogonal, i.e.
1
|G|
X
gG
χ
1
(g)χ
2
(g) =
(
1 χ
1
= χ
2
0 otherwise
.
We apply this to
G
= (
Z/DZ
)
×
, where
χ
1
is trivial and
χ
2
=
χ
. So orthogonality
gives
X
aD<i(a+1)D
χ(i) =
X
i(Z/DZ)
×
χ(i) = 0,
using that χ(i) = 0 if i is not coprime to D. So we are done.
Corollary. For quadratic characters χ
D
, we have
L(χ
D
, 1) 6= 0.
For example, if D < 0, then
L(χ
D
, 1) =
2π|cl
Q(
d)
|
|D|
1/2
|µ
Q(
d)
|
.
Proof. We have shown that
ζ
Q(
d)
(s) = ζ
Q
(s)L(χ
D
, s).
Note that
ζ
Q(
d)
(
s
) and
ζ
Q
(
s
) have simple poles at
s
= 1, while
L
(
χ
D
, s
) is
holomorphic at s = 1.
Since the residue of
ζ
Q
(
s
) at
s
= 1 is 1, while the residue of
ζ
Q(
D)
at
s
= 1
is non-zero by the analytic class number formula. So
L
(
χ
D
,
1) is non-zero, and
given by the analytic class number formula.
Example. If L = Q(
1), then
1
1
3
+
1
5
1
7
+ ··· =
2π · 1
2 · 4
=
π
4
.
In general, for any field whose class number we know, we can get a series
expansion for π. And it converges incredibly slow.
Note that this corollary required two things the analytic input for the
analytic class number formula, and quadratic reciprocity (to show that
χ
D
is a
Dirichlet character).
More ambitiously, we now compute the zeta function of a cyclotomic field,
L
=
Q
(
ω
q
), where
ω
q
is the primitive
q
th root of unity and
q N
. We need to
know the following facts about cyclotomic extensions:
Proposition.
(i) We have [L : Q] = ϕ(q), where
ϕ(q) = |(Z/qZ)
×
|.
(ii) L Q is a Galois extension, with
Gal(L/Q) = (Z/qZ)
×
,
where if
r
(
Z/qZ
)
×
, then
r
acts on
Q
(
w
Q
) by sending
ω
q
7→ ω
r
q
. This is
what plays the role of quadratic reciprocity for cyclotomic fields.
(iii) The ring of integers is
O
L
= Z[ω
q
] = Z[x]/Φ
q
(x),
where
Φ
q
(x) =
x
q
1
Q
d|q,d6=q
Φ
d
(x)
is the qth cyclotomic polynomial.
(iv)
Let
p
be a prime. Then
p
ramifies in
O
L
if and only if
p | D
L
, if and only
if
p | q
. So while
D
might be messy, the prime factors of
D
are the prime
factors of q.
(v)
Let
p
be a prime and
p - q
. Then
hpi
factors as a product of
ϕ
(
q
)
/f
distinct
prime ideals, each of norm p
f
, where f is the order of p in (Z/qZ)
×
.
Proof.
(i) In the Galois theory course.
(ii) In the Galois theory course.
(iii) In the example sheet.
(iv) In the example sheet.
(v) Requires proof, but is easy Galois theory, and is omitted.
Example. Let q = 8. Then
Φ
8
=
x
8
1
(x + 1)(x 1)(x
2
+ 1)
=
x
8
1
x
4
1
= x
4
+ 1.
So given a prime p (that is not 2), we need to understand
O
L
/p =
F
p
[x]
Φ
8
,
i.e. how Φ
8
factors factors mod p (Dedekind’s criterion). We have
(Z/8)
×
= {1, 3, 5, 7} = {1, 3, 3, 1} = Z/2 × Z/2.
Then (v) says if p = 17, then x
4
factors into 4 linear factors, which it does.
If p = 3, then (v) says x
4
factors into 2 quadratic factors. Indeed, we have
(x
2
x 1)(x
2
+ x 1) = (x
2
1)
2
x
2
= x
4
+ 1.
Given all of these, let’s compute the zeta function! Recall that
ζ
Q(ω
q
)
(s) =
Y
p
(1 N(p)
s
)
1
.
We consider the prime ideals
p
dividing
hpi
, where
p
is a fixed integer prime
number. If p - q, then (v) says this contributes a factor of
(1 p
fs
)
ϕ(q)/f
,
to the zeta function, where
f
is the order of
p
in (
Z/qZ
)
×
. We observe that this
thing factors, since
1 t
f
=
Y
γµ
f
(1 γt),
with
µ
f
= {γ C : γ
f
= 1},
and we can put t = p
s
.
We let
ω
1
, ··· , ω
ϕ(q)
: (Z/qZ)
×
C
×
be the distinct irreducible (one-dimensional) representations of (
Z/qZ
)
×
, with
ω
1
being the trivial representation, i.e. ω
1
(a) = 1 for all a (Z/qZ)
×
.
The claim is that
ω
1
(
p
)
, ··· , ω
ϕ(q)
(
p
) are
f
th roots of 1, each repeated
ϕ
(
q
)
/f
times. We either say this is obvious, or we can use some representation theory.
We know
p
generates a cyclic subgroup
hpi
of (
Z/qZ
)
×
of order
f
, by definition
of
f
. So this is equivalent to saying the restrictions of
ω
1
, ··· , ω
ϕ(q)
to
p
are the
f distinct irreducible characters of hpi
=
Z/f, each repeated ϕ(q)/f times.
Equivalently, note that
Res
(Z/qZ)
×
hpi
(ω
1
··· ω
ϕ(q)
) = Res
(Z/qZ)
×
hpi
(regular representation of (Z/qZ)
×
).
So this claims that
Res
(Z/qZ)
×
hpi
(regular rep. of (Z/qZ)
×
) =
ϕ(q)
f
(regular rep. of Z/f).
But this is true for any group, since
Res
G
H
CG = |G/H|CH,
as the character of both sides is |G|δ
e
.
So we have
(1 p
fs
)
ϕ(q)/f
=
ϕ(q)
Y
i=1
(1 ω
i
(p)p
s
)
1
.
So we let
χ
i
(n) =
(
w
i
(n mod q) gcd(n, q) = 1
0 otherwise
be the corresponding Dirichlet characters. So we have just shown that
Proposition. We have
ζ
Q(ω
q
)
(s) =
ϕ(q)
Y
i=1
L(χ
i
, s) · (corr. factor) = ζ
Q
(s)
ϕ(q)
Y
i=2
L(χ
i
, s) · (corr. factor)
where the correction factor is a finite product coming from the primes that divide
q.
By defining the
L
functions in a slightly more clever way, we can hide the
correction factors into the
L
(
χ, s
), and then the
ζ
function is just the product of
these L-functions.
Proof.
Our analysis covered all primes
p - q
, and the correction factor is just to
include the terms with p | q. The second part is just saying that
ζ
Q
(s) = L(χ
1
, s)
Y
p|q
(1 p
s
)
1
.
This allows us to improve our result on the non-vanishing of
L
(
χ,
1) to all
Dirichlet characters, and not just quadratic Dirichlet characters.
Corollary. If χ is any non-trivial Dirichlet character, then L(χ, 1) 6= 0.
Proof.
By definition, Dirichlet characters come from representations of some
(
Z/qZ
)
×
, so they appear in the formula of the
ζ
function of some cyclotomic
extension.
Consider the formula
ζ
Q(ω
q
)
(s) = ζ
Q
(s)
ϕ(q)
Y
i=2
L(χ
i
, s) · (corr. factor)
at
s
= 1. We know that the
L
(
χ
i
, s
) are all holomorphic at
s
= 1. Moreover,
both
ζ
Q(ω
q
)
and
ζ
Q
have a simple pole at 0. Since the correction terms are finite,
it must be the case that all L(χ
i
, s) are non-zero.
Theorem
(Dirichlet, 1839)
.
Let
a, q N
be coprime, i.e.
gcd
(
a, q
) = 1. Then
there are infinitely many primes in the arithmetic progression
a, a + q, a + 2q, a + 3q, ··· .
Proof. As before, let
ω
1
, ··· , ω
ϕ(q)
: (Z/qZ)
×
C
×
be the irreducible characters, and let
χ
1
, ··· , χ
ϕ(q)
: Z C
be the corresponding Dirichlet character, with ω
1
the trivial one.
Recall the orthogonality of columns of the character table, which says that if
gcd(p, q) = 1, then
1
ϕ(q)
X
i
ω
i
(a)ω
i
(p) =
(
1 a p (mod q)
0 otherwise
.
Hence we know
1
ϕ(q)
X
i
χ
i
(a)χ
i
(p) =
(
1 a p (mod q)
0 otherwise
,
even if gcd(p, q) 6= 1, as then χ
i
(p) = 0. So
X
pa mod q
p prime
p
s
=
1
ϕ(q)
X
i
χ
i
(a)
X
all primes p
χ
i
(p)p
s
. ()
We want to show this has a pole at s = 1, as in Euclid’s proof.
To do so, we show that
P
p
χ
i
(
p
)
p
s
is “essentially”
log L
(
χ
i
, s
), up to some
bounded terms. We Taylor expand
log L(χ, s) =
X
log(1 χ(p)p
s
) =
X
n1
p prime
χ(p)
n
np
ns
=
X
n1
p prime
χ(p
n
)
np
ns
.
What we care about is the n = 1 term. So we claim that
X
n2,p prime
χ(p
n
)
np
ns
converges at s = 1. This follows from the geometric sum
X
p
X
n2
χ(p
n
)
np
ns
X
p
X
n2
p
ns
=
X
p prime
1
p
s
(p
s
1)
X
n2
1
n
s
(n
s
1)
< .
Hence we know
log L(χ, s) =
X
p
χ
i
(p)p
s
+ bounded stuff
near s = 1.
So at s = 1, we have
()
1
ϕ(q)
X
i
χ
i
(a) log L(χ
i
, s).
and we have to show that the right hand side has a pole at s = 1.
We know that for
i 6
= 1, i.e.
χ
i
non-trivial,
L
(
χ
i
, s
) is holomorphic and
non-zero at
s
= 1. So we just have to show that
log L
(
χ
1
, s
) has a pole. Note
that L(χ
1
, s) is essentially ζ
Q
(s). Precisely, we have
L(χ
1
, s) = ζ
Q
(s)
Y
p|q
(1 p
s
).
Moreover, we already know that ζ
Q
(s) blows up at s = 1. We have
ζ
Q
(s) =
1
s 1
+ holomorphic function
=
1
s 1
(1 + (s 1)(holomorphic function)).
So we know
log L(χ
1
, s) log ζ
Q
(s) log
1
s 1
,
and this does blow up at s = 1.
So far, we have been working with abelian extensions over
Q
, i.e. extensions
L/Q
whose Galois group is abelian. By the Kronecker–Weber theorem, every
abelian extension of
Q
is contained within some cyclotomic extension. So in
some sense, we have considered the “most general” abelian extension.
Can we move on to consider more complicated number fields? In general,
suppose
L/Q
is Galois, and
G
=
Gal
(
L/Q
). We can still make sense of the
ζ
functions, and it turns out it always factors as
ζ
L
(s) =
Y
ρ
L(ρ, s)
dim ρ
,
where
ρ
ranges over all the irreducible representations of
G
, and
L
(
ρ, s
) is the
Artin
L
-function. It takes some effort to define the Artin
L
-function, and we
shall not do so here. However, it is worth noting that
L
(1
, s
) is just
ζ
Q
(
s
), and
for ρ 6= 1, we still have a factorization of the form
L(ρ, s) =
Y
p prime
L
p
(ρ, s).
This L
p
(ρ, s) is known as the Euler factor .
One can show that
L
(
ρ, s
) is always a meromorphic function of
s
, and is
conjectured to be holomorphic for all s (if ρ 6= 1, of course).
If
ρ
is one-dimensional, then
L
(
ρ, s
) is a Dirichlet series
L
(
χ, s
) for some
χ
.
Recall that to establish this fact for quadratic fields, we had to use quadratic
reciprocity. In general given a
ρ
, finding
χ
is a higher version of “quadratic
reciprocity”. This area is known as class field theory. If dim ρ > 1, then this is
“non-abelian class field theory”, known as Langlands programme.