Part III Local Fields
Based on lectures by H. C. Johansson
Notes taken by Dexter Chua
Michaelmas 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.
The
p
-adic numbers
Q
p
(where
p
is any prime) were invented by Hensel in the late 19th
century, with a view to introduce function-theoretic methods into number theory. They
are formed by completing
Q
with respect to the
p
-adic absolute value
| |
p
, defined
for non-zero
x Q
by
|x|
p
=
p
n
, where
x
=
p
n
a/b
with
a, b, n Z
and
a
and
b
are
coprime to
p
. The
p
-adic absolute value allows one to study congruences modulo all
powers of
p
simultaneously, using analytic methods. The concept of a local field is an
abstraction of the field
Q
p
, and the theory involves an interesting blend of algebra and
analysis. Local fields provide a natural tool to attack many number-theoretic problems,
and they are ubiquitous in modern algebraic number theory and arithmetic geometry.
Topics likely to be covered include:
The p-adic numbers. Local fields and their structure.
Finite extensions, Galois theory and basic ramification theory.
Polynomial equations; Hensel’s Lemma, Newton polygons.
Continuous functions on the p-adic integers, Mahler’s Theorem.
Local class field theory (time permitting).
Pre-requisites
Basic algebra, including Galois theory, and basic concepts from point set topology
and metric spaces. Some prior exposure to number fields might be useful, but is not
essential.
Contents
0 Introduction
1 Basic theory
1.1 Fields
1.2 Rings
1.3 Topological rings
1.4 The p-adic numbers
2 Valued fields
2.1 Hensel’s lemma
2.2 Extension of norms
2.3 Newton polygons
3 Discretely valued fields
3.1 Teichm¨uller lifts
3.2 Witt vectors*
4 Some p-adic analysis
5 Ramification theory for local fields
5.1 Ramification index and inertia degree
5.2 Unramified extensions
5.3 Totally ramified extensions
6 Further ramification theory
6.1 Some filtrations
6.2 Multiple extensions
7 Local class field theory
7.1 Infinite Galois theory
7.2 Unramified extensions and Weil group
7.3 Main theorems of local class field theory
8 Lubin–Tate theory
8.1 Motivating example
8.2 Formal groups
8.3 Lubin–Tate extensions
0 Introduction
What are local fields? Suppose we are interested in some basic number theoretic
problem. Say we have a polynomial
f
(
x
1
, ··· , x
n
)
Z
[
x
1
, ··· , x
n
]. We want to
look for solutions
a Z
n
, or show that there are no solutions at all. We might
try to view this polynomial as a real polynomial, look at its roots, and see if
they are integers. In lucky cases, we might be able to show that there are no
real solutions at all, and conclude that there cannot be any solutions at all.
On the other hand, we can try to look at it modulo some prime
p
. If there
are no solutions mod
p
, then there cannot be any solution. But sometimes
p
is
not enough. We might want to look at it mod
p
2
, or
p
3
, or . . . . One important
application of local fields is that we can package all these information together.
In this course, we are not going to study the number theoretic problems, but
just look at the properties of the local fields for their own sake.
Throughout this course, all rings will be commutative with unity, unless
otherwise specified.
1 Basic theory
We are going to start by making loads of definitions, which you may or may not
have seen before.
1.1 Fields
Definition
(Absolute value)
.
Let
K
be a field. An absolute value on
K
is a
function | · | : K R
0
such that
(i) |x| = 0 iff x = 0;
(ii) |xy| = |x||y| for all x, y K;
(iii) |x + y| |x| + |y|.
Definition (Valued field). A valued field is a field with an absolute value.
Example.
The rationals, reals and complex numbers with the usual absolute
values are absolute values.
Example
(Trivial absolute value)
.
The trivial absolute value on a field
K
is the
absolute value given by
|x| =
(
1 x 6= 0
0 x = 0
.
The only reason we mention the trivial absolute value here is that from
now on, we will assume that the absolute values are not trivial, because trivial
absolute values are boring and break things.
There are some familiar basic properties of the absolute value such as
Proposition. ||x| |y|| |x y|
. Here the outer absolute value on the left
hand side is the usual absolute value of
R
, while the others are the absolute
values of the relevant field.
An absolute value defines a metric d(x, y) = |x y| on K.
Definition
(Equivalence of absolute values)
.
Let
K
be a field, and let
| · |, | · |
0
be absolute values. We say they are equivalent if they induce the same topology.
Proposition.
Let
K
be a field, and
| · |, | · |
0
be absolute values on
K
. Then
the following are equivalent.
(i) | · | and | · |
0
are equivalent
(ii) |x| < 1 implies |x|
0
< 1 for all x K
(iii) There is some s R
>0
such that |x|
s
= |x|
0
for all x K.
Proof.
(i)
(ii) and (iii)
(i) are easy exercises. Assume (ii), and we shall
prove (iii). First observe that since
|x
1
|
=
|x|
1
, we know
|x| >
1 implies
|x|
0
>
1, and hence
|x|
= 1 implies
|x|
0
= 1. To show (iii), we have to show that
the ratio
log |x|
log |x
0
|
is independent of x.
Suppose not. We may assume
log |x|
log |x|
0
<
log |y|
log |y|
0
,
and moreover the logarithms are positive. Then there are
m, n Z
>0
such that
log |x|
log |y|
<
m
n
<
log |x|
0
log |y|
0
.
Then rearranging implies
x
n
y
m
< 1 <
x
n
y
m
0
,
a contradiction.
Exercise.
Let
K
be a valued field. Then equivalent absolute values induce the
same the completion
ˆ
K
of
K
, and
ˆ
K
is a valued field with an absolute value
extending | · |.
In this course, we are not going to be interested in the usual absolute values.
Instead, we are going to consider some really weird ones, namely non-archimedean
ones.
Definition
(Non-archimedean absolute value)
.
An absolute value
| · |
on a field
K
is called non-archimedean if
|x
+
y| max
(
|x|, |y|
). This condition is called
the strong triangle inequality.
An absolute value which isn’t non-archimedean is called archimedean.
Metrics satisfying
d
(
x, z
)
max
(
d
(
x, y
)
, d
(
y, z
)) are often known as ultra-
metrics.
Example. Q, R and C under the usual absolute values are archimedean.
In this course, we will only consider non-archimedean absolute values. Thus,
from now on, unless otherwise mentioned, an absolute value is assumed to be
non-archimedean. The metric is weird!
We start by proving some absurd properties of non-archimedean absolute
values.
Recall that the closed balls are defined by
B(x, r) = {y : |x y| r}.
Proposition.
Let (
K, | · |
) be a non-archimedean valued field, and let
x K
and r R
>0
. Let z B(x, r). Then
B(x, r) = B(z, r).
So closed balls do not have unique “centers”. Every point can be viewed as
the center.
Proof. Let y B(z, r). Then
|x y| = |(x z) + (z y)| max(|x z|, |z y|) r.
So y B(x, r). By symmetry, y B(x, r) implies y B(z, r).
Corollary. Closed balls are open.
Proof. To show that B(x, r) is open, we let z B(x, r). Then we have
{y : |y z| < r} B(z, r) = B(x, r).
So we know the open ball of radius
r
around
z
is contained in
B
(
x, r
). So
B
(
x, r
)
is open.
Norms in non-archimedean valued fields are easy to compute:
Proposition.
Let
K
be a non-archimedean valued field, and
x, y K
. If
|x| > |y|, then |x + y| = |x|.
More generally, if
x
=
P
c=0
x
i
and the non-zero
|x
i
|
are distinct, then
|x| = max |x
i
|.
Proof.
On the one hand, we have
|x
+
y| max{|x|, |y|}
. On the other hand,
we have
|x| = |(x + y) y| max(|x + y|, |y|) = |x + y|,
since we know that we cannot have
|x| |y|
. So we must have
|x|
=
|x
+
y|
.
Convergence is also easy for valued fields.
Proposition. Let K be a valued field.
(i) Let (x
n
) be a sequence in K. If x
n
x
n+1
0, then x
n
is Cauchy.
If we assume further that K is complete, then
(ii)
Let (
x
n
) be a sequence in
K
. If
x
n
x
n+1
0, then a sequence (
x
n
) in
K converges.
(iii) Let
P
n=0
y
n
be a series in K. If y
n
0, then
P
n=0
y
n
converges.
The converses to all these are of course also true, with the usual proofs.
Proof.
(i)
Pick
ε >
0 and
N
such that
|x
n
x
n+1
| < ε
for all
n N
. Then given
m n N , we have
|x
m
x
n
| = |x
m
x
m1
+ x
m1
x
m2
+ ··· x
n
|
max(|x
m
x
m1
|, ··· , |x
n+1
x
n
|)
< ε.
So the sequence is Cauchy.
(ii) Follows from (1) and the definition of completeness.
(iii) Follows from the definition of convergence of a series and (2).
The reason why we care about these weird non-archimedean fields is that
they have very rich algebraic structure. In particular, there is this notion of the
valuation ring.
Definition
(Valuation ring)
.
Let
K
be a valued field. Then the valuation ring
of K is the open subring
O
K
= {x : |x| 1}.
We prove that it is actually a ring
Proposition. Let K be a valued field. Then
O
K
= {x : |x| 1}
is an open subring of
K
. Moreover, for each
r
(0
,
1], the subsets
{x
:
|x| < r}
and {x : |x| r} are open ideals of O
K
. Moreover, O
×
K
= {x : |x| = 1}.
Note that this is very false for usual absolute values. For example, if we take
R with the usual absolute value, we have 1 O
R
, but 1 + 1 6∈ O
R
.
Proof. We know that these sets are open since all balls are open.
To see
O
K
is a subring, we have
|
1
|
=
|
1
|
= 1. So 1
,
1
O
K
. If
x, y O
K
,
then
|x
+
y| max
(
|x|, |y|
)
1. So
x
+
y O
K
. Also,
|xy|
=
|x||y|
1
·
1 = 1.
So xy O
K
.
That the other sets are ideals of O
K
is checked in the same way.
To check the units, we have
x O
×
K
|x|, |x
1
|
1
|x|
=
|x|
1
= 1.
1.2 Rings
Definition
(Integral element)
.
Let
R S
be rings and
s S
. We say
s
is
integral over R if there is some monic f R[x] such that f(s) = 0.
Example. Any r R is integral (take f(x) = x r).
Example.
Take
Z C
. Then
z C
is integral over
Z
if it is an algebraic
integer (by definition of algebraic integer). For example,
2
is an algebraic
integer, but
1
2
is not.
We would like to prove the following characterization of integral elements:
Theorem.
Let
R S
be rings. Then
s
1
, ··· , s
n
S
are all integral iff
R[s
1
, ··· , s
n
] S is a finitely-generated R-module.
Note that
R
[
s
1
, ··· , s
n
] is by definition a finitely-generated
R
-algebra, but
requiring it to be finitely-generated as a module is stronger.
Here one direction is easy. It is not hard to show that if
s
1
, ··· , s
n
are all
integral, then
R
[
s
1
, ··· , s
n
] is finitely-generated. However to show the other
direction, we need to find some clever trick to produce a monic polynomial that
kills the s
i
.
The trick we need is the adjugate matrix we know and love from IA Vectors
and Matrices.
Definition
(Adjoint/Adjugate matrix)
.
Let
A
= (
a
ij
) be an
n × n
matrix with
coefficients in a ring
R
. The adjugate matrix or adjoint matrix
A
= (
a
ij
) of
A
is defined by
a
ij
= (1)
i+j
det(A
ij
),
where
A
ij
is an (
n
1)
×
(
n
1) matrix obtained from
A
by deleting the
i
th
column and the jth row.
As we know from IA, the following property holds for the adjugate matrix:
Proposition.
For any
A
, we have
A
A
=
AA
=
det
(
A
)
I
, where
I
is the
identity matrix.
With this, we can prove our claim:
Proof of theorem. Note that we can construct R[s
1
, ··· , s
n
] by a sequence
R R[s
1
] R[s
1
, s
2
] ··· R[s
1
, ··· , s
n
] S,
and each
s
i
is integral over
R
[
s
1
, ··· , s
n1
]. Since the finite extension of a finite
extension is still finite, it suffices to prove it for the case
n
= 1, and we write
s
for s
1
.
Suppose
f
(
x
)
R
[
x
] is monic such that
f
(
s
) = 0. If
g
(
x
)
R
[
x
], then there
is some
q, r R
[
x
] such that
g
(
x
) =
f
(
x
)
q
(
x
) +
r
(
x
) with
deg r < deg f
. Then
g
(
s
) =
r
(
s
). So any polynomial expression in
s
can be written as a polynomial
expression with degree less than
deg f
. So
R
[
s
] is generated by 1
, s, ··· , s
deg f 1
.
In the other direction, let
t
1
, ··· , t
d
be
R
-module generators of
R
[
s
1
, ··· , s
n
].
We show that in fact any element of
R
[
s
1
, ··· , s
n
] is integral over
R
. Consider
any element b R[s
1
, ··· , s
n
]. Then there is some a
ij
R such that
bt
i
=
d
X
j=1
a
ij
t
j
.
In matrix form, this says
(bI A)t = 0.
We now multiply by (bI A)
to obtain
det(bI A)t
j
= 0
for all j. Now we know 1 R. So 1 =
P
c
j
t
j
for some c
j
R. Then we have
det(bI A) = det(bI A)
X
c
j
t
j
=
X
c
j
(det(bI A)t
j
) = 0.
Since det(bI A) is a monic polynomial in b, it follows that b is integral.
Using this characterization, the following result is obvious:
Corollary.
Let
R S
be rings. If
s
1
, s
2
S
are integral over
R
, then
s
1
+
s
2
and
s
1
s
2
are integral over
R
. In particular, the set
˜
R S
of all elements in
S
integral over R is a ring, known as the integral closure of R in S.
Proof.
If
s
1
, s
2
are integral, then
R
[
s
1
, s
2
] is a finite extension over
R
. Since
s
1
+ s
2
and s
1
s
2
are elements of R[s
1
, s
2
], they are also integral over R.
Definition
(Integrally closed)
.
Given a ring extension
R S
, we say
R
is
integrally closed in S if
˜
R = R.
1.3 Topological rings
Recall that we previously constructed the valuation ring
O
K
. Since the valued
field
K
itself has a topology, the valuation ring inherits a subspace topology.
This is in fact a ring topology.
Definition
(Topological ring)
.
Let
R
be a ring. A topology on
R
is called a
ring topology if addition and multiplication are continuous maps
R × R R
. A
ring with a ring topology is a topological ring.
Example. R
and
C
with the usual topologies and usual ring structures are
topological rings.
Exercise.
Let
K
be a valued field. Then
K
is a topological ring. We can see
this from the fact that the product topology on
K ×K
is induced by the metric
d((x
0
, y
0
), (x
1
, y
1
)) = max(|x
0
x
1
|, |y
0
y
1
|).
Now if we are just randomly given a ring, there is a general way of constructing
a ring topology. The idea is that we pick an ideal
I
and declare its elements to
be small. For example, in a valued ring, we can pick
I
=
{x O
K
:
|x| <
1
}
.
Now if you are not only in
I
, but
I
2
, then you are even smaller. So we have a
hierarchy of small sets
I I
2
I
3
I
4
···
Now to make this a topology on
R
, we say that a subset
U R
is open if every
x U
is contained in some translation of
I
n
(for some
n
). In other words, we
need some y R such that
x y + I
n
U.
But since
I
n
is additively closed, this is equivalent to saying
x
+
I
n
U
. So we
make the following definition:
Definition
(
I
-adically open)
.
Let
R
be a ring and
I R
an ideal. A subset
U R
is called
I
-adically open if for all
x U
, there is some
n
1 such that
x + I
n
U.
Proposition.
The set of all
I
-adically open sets form a topology on
R
, called
the I-adic topology.
Note that the
I
-adic topology isn’t really the kind of topology we are used
to thinking about, just like the topology on a valued field is also very weird.
Instead, it is a “filter” for telling us how small things are.
Proof.
By definition, we have
and
R
are open, and arbitrary unions are clearly
open. If
U, V
are
I
-adically open, and
x U V
, then there are
n, m
such that
x + I
n
U and x + I
m
V . Then x + I
max(m,n)
U V .
Exercise. Check that the I-adic topology is a ring topology.
In the special case where
I
=
xR
, we often call the
I
-adic topology the
x
-adic
topology.
Now we want to tackle the notion of completeness. We will consider the case
of I = xR for motivation, but the actual definition will be completely general.
If we pick the
x
-adic topology, then we are essentially declaring that we take
x to be small. So intuitively, we would expect power series like
a
0
+ a
1
x + a
2
x
2
+ a
3
x
3
+ ···
to “converge”, at least if the
a
i
are “of bounded size”. In general, the
a
i
are
“not too big” if
a
i
x
i
is genuinely a member of
x
i
R
, as opposed to some silly thing
like x
i
.
As in the case of analysis, we would like to think of these infinite series as a
sequence of partial sums
(a
0
, a
0
+ a
1
x, a
0
+ a
1
x + a
2
x
2
, ···)
Now if we denote the limit as
L
, then we can think of this sequence alternatively
as
(L mod I, L mod I
2
, L mod I
3
, ···).
The key property of this sequence is that if we take
L mod I
k
and reduce it mod
I
k1
, then we obtain L mod I
k1
.
In general, suppose we have a sequence
(b
n
R/I
n
)
n=1
.
such that
b
n
mod I
n1
=
b
n1
. Then we want to say that the ring is
I
-adically
complete if every sequence satisfying this property is actually of the form
(L mod I, L mod I
2
, L mod I
3
, ···)
for some
L
. Alternatively, we can take the
I
-adic completion to be the collection
of all such sequences, and then a space is
I
-adically complete it is isomorphic to
its I-adic completion.
To do this, we need to build up some technical machinery. The kind of
sequences we’ve just mentioned is a special case of an inverse limit.
Definition
(Inverse/projective limit)
.
Let
R
1
, R
2
, , ···
be topological rings, with
continuous homomorphisms f
n
: R
n+1
R
n
.
R
1
R
2
R
3
R
4
···
f
1
f
2
f
3
The inverse limit or projective limit of the R
i
is the ring
lim
R
n
=
(
(x
n
)
Y
n
R
n
: f
n
(x
n+1
) = x
n
)
,
with coordinate-wise addition and multiplication, together with the subspace
topology coming from the product topology of
Q
R
n
. This topology is known as
the inverse limit topology.
Proposition. The inverse limit topology is a ring topology.
Proof sketch. We can fit the addition and multiplication maps into diagrams
lim
R
n
× lim
R
n
lim
R
n
Q
R
n
×
Q
R
n
Q
R
n
By the definition of the subspace topology, it suffices to show that the cor-
responding maps on
Q
R
n
are continuous. By the universal property of the
product, it suffices to show that the projects
Q
R
n
×
Q
R
n
R
m
is continuous
for all
m
. But this map can alternatively be obtained by first projecting to
R
m
,
then doing multiplication in
R
m
, and projection is continuous. So the result
follows.
It is easy to see the following universal property of the inverse limit topology:
Proposition.
Giving a continuous ring homomorphism
g
:
S lim
R
n
is the
same as giving a continuous ring homomorphism
g
n
:
S R
n
for each
n
, such
that each of the following diagram commutes:
S R
n
R
n1
g
n
g
n1
f
n1
Definition
(
I
-adic completion)
.
Let
R
be a ring and
I
be an ideal. The
I
-adic
completion of R is the topological ring
lim
R/I
n
,
where
R/I
n
has the discrete topology, and
R/I
n+1
R/I
n
is the quotient map.
There is an evident map
ν : R lim
R/I
n
r 7→ (r mod I
n
)
.
This map is a continuous ring homomorphism if
R
is given the
I
-adic topology.
Definition
(
I
-adically complete)
.
We say that
R
is
I
-adically complete if
ν
is a
bijection.
Exercise. If ν is a bijection, then ν is in fact a homeomorphism.
1.4 The p-adic numbers
For the rest of this course, p is going to be a prime number.
We consider a particular case of valued fields, namely the
p
-adic numbers,
and study some of its basic properties.
Let
x Q
be non-zero. Then by uniqueness of factorization, we can write
x
uniquely as
x = p
n
a
b
,
where a, b, n Z, b > 0 and a, b, p are pairwise coprime.
Definition
(
p
-adic absolute value)
.
The
p
-adic absolute value on
Q
is the
function | · |
p
: Q R
0
given by
|x|
p
=
(
0 x = 0
p
n
x = p
n
a
b
as above
.
Proposition. The p-adic absolute value is an absolute value.
Proof. It is clear that |x|
p
= 0 iff x = 0.
Suppose we have
x = p
n
a
b
, y = p
m
c
d
.
We wlog m n. Then we have
|xy|
p
=
p
n+m
ac
bd
= p
mn
= |x|
p
|y|
p
.
So this is multiplicative. Finally, we have
|x + y|
p
=
p
n
ab + p
mn
cb
bd
p
n
= max(|x|
p
, |y|
p
).
Note that we must have
bd
coprime to
p
, but
ab
+
p
mn
cb
need not be. However,
any extra powers of
p
could only decrease the absolute value, hence the above
result.
Note that if x Z is an integer, then |x|
p
= p
n
iff p
n
|| x (we say p
n
|| x if
p
n
| x and p
n+1
- x).
Definition
(
p
-adic numbers)
.
The
p
-adic numbers
Q
p
is the completion of
Q
with respect to | · |
p
.
Definition (p-adic integers). The valuation ring
Z
p
= {x Q
p
: |x|
p
1}
is the p-adic integers.
Proposition. Z
p
is the closure of Z inside Q
p
.
Proof. If x Z is non-zero, then x = p
n
a with n 0. So |x|
p
1. So Z Z
p
.
We now want to show that Z is dense in Z
p
. We know the set
Z
(p)
= {x Q : |x|
p
1}
is dense inside
Z
p
, essentially by definition. So it suffices to show that
Z
is dense
in Z
(p)
. We let x Z
(p)
\ {0}, say
x = p
n
a
b
, n 0.
It suffices to find x
i
Z such that x
i
1
b
. Then we have p
n
ax
i
x.
Since (
b, p
) = 1, we can find
x
i
, y
i
Z
such that
bx
i
+
p
i
y
i
= 1 for all
i
1.
So
x
i
1
b
p
=
1
b
p
|bx
i
1|
p
= |p
i
y
i
|
p
p
i
0.
So done.
Proposition. The non-zero ideals of Z
p
are p
n
Z
p
for n 0. Moreover,
Z
p
n
Z
=
Z
p
p
n
Z
p
.
Proof.
Let 0
6
=
I Z
p
be an ideal, and pick
x I
such that
|x|
p
is maximal.
This supremum exists and is attained because the possible values of the absolute
values are discrete and bounded above. If
y I
, then by maximality, we have
|y|
p
|x|
p
. So we have
|yx
1
|
p
1. So
yx
1
Z
p
, and this implies that
y
= (
yx
1
)
x xZ
p
. So
I xZ
p
, and we obviously have
xZ
p
I
. So we have
I = xZ
p
.
Now if
x
=
p
n
a
b
, then since
a
b
is invertible in
Z
p
, we have
xZ
p
=
p
n
Z
p
. So
I = p
n
Z
p
.
To show the second part, consider the map
f
n
: Z
Z
p
p
n
Z
p
given by the inclusion map followed by quotienting. Now
p
n
Z
p
=
{x
:
|x|
p
p
n
.
So we have
ker f
n
= {x Z : |x|
p
p
n
} = p
n
Z.
Now since
Z
is dense in
Z
p
, we know the image of
f
n
is dense in
Z
p
/p
n
Z
p
.
But
Z
p
/p
n
Z
p
has the discrete topology. So
f
n
is surjective. So
f
n
induces an
isomorphism Z/p
n
Z
=
Z
p
/p
n
Z
p
.
Corollary. Z
p
is a PID with a unique prime element p (up to units).
This is pretty much the point of the
p
-adic numbers there are a lot of
primes in Z, and by passing on to Z
p
, we are left with just one of them.
Proposition.
The topology on
Z
induced by
| · |
p
is the
p
-adic topology (i.e.
the pZ-adic topology).
Proof.
Let
U Z
. By definition,
U
is open wrt
| · |
p
iff for all
x U
, there is
an n N such that
{y Z : |y x|
p
p
n
} U.
On the other hand,
U
is open in the
p
-adic topology iff for all
x U
, there is
some n 0 such that x + p
n
Z U. But we have
{y Z : |y x|
p
p
n
} = x + p
n
Z.
So done.
Proposition. Z
p
is p-adically complete and is (isomorphic to) the p-adic com-
pletion of Z.
Proof. The second part follows from the first as follows: we have the maps
Z
p
lim
Z
p
/(p
n
Z
p
) lim Z/(p
n
Z)
ν
(f
n
)
n
We know the map induced by (
f
n
)
n
is an isomorphism. So we just have to show
that ν is an isomorphism
To prove the first part, we have
x ker ν
iff
x p
n
Z
p
for all
n
iff
|x|
p
p
n
for all n iff |x|
p
= 0 iff x = 0. So the map is injective.
To show surjectivity, we let
z
n
lim
Z
p
/p
n
Z
p
.
We define a
i
{0, 1, ··· , p 1} recursively such that
x
n
=
n1
X
i=0
a
i
p
i
is the unique representative of
z
n
in the set of integers
{
0
,
1
, ··· , p
n
1
}
. Then
x =
X
i=0
a
i
p
i
exists in
Z
p
and maps to
x x
n
z
n
(
mod p
n
) for all
n
0. So
ν
(
x
) = (
z
n
).
So the map is surjective. So ν is bijective.
Corollary. Every a Z
p
has a unique expansion
a =
X
i=0
a
i
p
i
.
with a
i
{0, ··· , p 1}.
More generally, for any a Q
×
, there is a unique expansion
a =
X
i=n
a
i
p
i
for a
i
{0, ··· , p 1}, a
n
6= 0 and
n = log
p
|a|
p
Z.
Proof.
The second part follows from the first part by multiplying
a
by
p
n
.
Example. We have
1
1 p
= 1 + p + p
2
+ p
3
+ ··· .
2 Valued fields
2.1 Hensel’s lemma
We return to the discussion of general valued fields. We are now going to introduce
an alternative to the absolute value that contains the same information, but is
presented differently.
Definition
(Valuation)
.
Let
K
be a field. A valuation on
K
is a function
v : K R {∞} such that
(i) v(x) = 0 iff x = 0
(ii) v(xy) = v(x) + v(y)
(iii) v(x + y) min{v(x), v(y)}.
Here we use the conventions that r + = and r for all r .
In some sense, this definition is sort-of pointless, since if
v
is a valuation,
then the function
|x| = c
v(x)
for any
c >
1 is a (non-archimedean) absolute value. Conversely, if
| · |
is a
valuation, then
v(x) = log
c
|x|
is a valuation.
Despite this, sometimes people prefer to talk about the valuation rather than
the absolute value, and often this is more natural. As we will later see, in certain
cases, there is a canonical normalization of
v
, but there is no canonical choice
for the absolute value.
Example. For x Q
p
, we define
v
p
(x) = log
p
|x|
p
.
This is a valuation, and if x Z
p
, then v
p
(x) = n iff p
n
|| x.
Example. Let K be a field, and define
k((T )) =
(
X
i=n
a
i
T
i
: a
i
k, n Z
)
.
This is the field of formal Laurent series over k. We define
v
X
a
i
T
i
= min{i : a
i
6= 0}.
Then v is a valuation of k((T)).
Recall that for a valued field K, the valuation ring is given by
O
K
= {x K : |x| 1} = {x K : v(x) 0}.
Since this is a subring of a field, and the absolute value is multiplicative, we
notice that the units in
O
are exactly the elements of absolute value 1. The
remaining elements form an ideal (since the field is non-archimedean), and thus
we have a maximal ideal
m = m
K
= {x K : |x| < 1}
The quotient
k = k
K
= O
K
/m
K
is known as the residue field.
Example. Let K = Q
p
. Then O = Z
p
, and m = pZ
p
. So
k = O/m = Z
p
/pZ
p
=
Z/pZ.
Definition
(Primitive polynomial)
.
If
K
is a valued field and
f
(
x
) =
a
0
+
a
1
x
+
··· + a
n
x
n
K[x] is a polynomial, we say that f is primitive if
max
i
|a
i
| = 1.
In particular, we have f O[x].
The point of a primitive polynomial is that such a polynomial is naturally,
and non-trivially, an element of
k
[
x
]. Moreover, focusing on such polynomials is
not that much of a restriction, since any polynomial is a constant multiple of a
primitive polynomial.
Theorem
(Hensel’s lemma)
.
Let
K
be a complete valued field, and let
f K
[
x
]
be primitive. Put
¯
f = f mod m k[x]. If there is a factorization
¯
f(x) = ¯g(x)
¯
h(x)
with (¯g,
¯
h) = 1, then there is a factorization
f(x) = g(x)h(x)
in O[x] with
¯g = g,
¯
h = h mod m,
with deg g = deg ¯g.
Note that requiring
deg g
=
deg ¯g
is the best we can hope for we cannot
guarantee deg h = deg
¯
h, since we need not have deg f = deg
¯
f.
This is one of the most important results in the course.
Proof.
Let
g
0
, h
0
be arbitrary lifts of
¯g
and
¯
h
to
O
[
x
] with
deg ¯g
=
g
0
and
deg
¯
h = h
0
. Then we have
f = g
0
h
0
mod m.
The idea is to construct a “Taylor expansion” of the desired
g
and
h
term by
term, starting from
g
0
and
h
0
, and using completeness to guarantee convergence.
To proceed, we use our assumption that
¯g,
¯
h
are coprime to find some
a, b O
[
x
]
such that
ag
0
+ bh
0
1 mod m. ()
It is easier to work modulo some element
π
instead of modulo the ideal
m
, since
we are used to doing Taylor expansion that way. Fortunately, since the equations
above involve only finitely many coefficients, we can pick an
π m
with absolute
value large enough (i.e. close enough to 1) such that the above equations hold
with m replaced with π. Thus, we can write
f = g
0
h
0
+ πr
0
, r
0
O[x].
Plugging in (), we get
f = g
0
h
0
+ πr
0
(ag
0
+ bh
0
) + π
2
(something).
If we are lucky enough that
deg r
0
b < deg g
0
, then we group as we learnt in
secondary school to get
f = (g
0
+ πr
0
b)(h
0
+ πr
0
a) + π
2
(something).
We can then set
g
1
= g
0
+ πr
0
b
h
1
= h
0
+ πr
0
a,
and then we can write
f = g
1
h
1
+ π
2
r
1
, r
1
O[x], deg g
1
= deg ¯g. ()
If it is not true that deg r
0
b deg g
0
, we use the division algorithm to write
r
0
b = qg
0
+ p.
Then we have
f = g
0
h
0
+ π((r
0
a + q)g
0
+ ph
0
),
and then proceed as above.
Given the factorization (
), we replace
r
1
by
r
1
(
ag
0
+
bh
0
), and then repeat
the procedure to get a factorization
f g
2
h
2
mod π
3
, deg g
2
= deg ¯g.
Inductively, we constrict g
k
, h
k
such that
f g
k
h
k
mod π
k+1
g
k
g
k1
mod π
k
h
k
h
k1
mod π
k
deg g
k
= deg ¯g
Note that we may drop the terms of
h
k
whose coefficient are in
π
k+1
O
, and the
above equations still hold. Moreover, we can then bound
deg h
k
deg f deg g
k
.
It now remains to set
g = lim
k→∞
g
k
, h = lim
k→∞
h
k
.
Corollary.
Let
f
(
x
) =
a
0
+
a
1
x
+
···
+
a
n
x
n
K
[
x
] where
K
is complete and
a
0
, a
n
6= 0. If f is irreducible, then
|a
| max(|a
0
|, |a
n
|)
for all .
Proof.
By scaling, we can wlog
f
is primitive. We then have to prove that
max
(
|a
0
|, |a
n
|
) = 1. If not, let
r
be minimal such that
|a
r
|
= 1. Then 0
< r < n
.
Moreover, we can write
f(x) x
r
(a
r
+ a
r+1
x + ··· + a
n
x
nr
) mod m.
But then Hensel’s lemma says this lifts to a factorization of
f
, a contradiction.
Corollary (of Hensel’s lemma). Let f O[x] be monic, and K complete. If f
mod m
has a simple root
¯α k
, then
f
has a (unique) simple root
α O
lifting
¯α.
Example.
Consider
x
p1
1
Z
p
[
x
]. We know
x
p1
splits into distinct linear
factors over
F
p
[
x
]. So all roots lift to
Z
p
. So
x
p1
1 splits completely in
Z
p
.
So Z
p
contains all p roots of unity.
Example. Since 2 is a quadratic residue mod 7, we know
2 Q
7
.
2.2 Extension of norms
The main goal of this section is to prove the following theorem:
Theorem.
Let
K
be a complete valued field, and let
L/K
be a finite extension.
Then the absolute value on
K
has a unique extension to an absolute value on
L
,
given by
|α|
L
=
n
q
|N
L/K
(α)|,
where
n
= [
L
:
K
] and
N
L/K
is the field norm. Moreover,
L
is complete with
respect to this absolute value.
Corollary.
Let
K
be complete and
M/K
be an algebraic extension of
K
. Then
| · | extends uniquely to an absolute value on M.
This is since any algebraic extension is the union of finite extensions, and
uniqueness means we can patch the absolute values together.
Corollary.
Let
K
be a complete valued field and
L/K
a finite extension. If
σ Aut(L/K), then |σ(α)|
L
= |α|
L
.
Proof.
We check that
α 7→ |σ
(
α
)
|
L
is also an absolute value on
L
extending the
absolute value on K. So the result follows from uniqueness.
Before we can prove the theorem, we need some preliminaries. Given a finite
extension
L/K
, we would like to consider something more general than a field
norm on
L
. Instead, we will look at norms of
L
as a
K
-vector space. There
are less axioms to check, so naturally there will be more choices for the norm.
However, just as in the case of
R
-vector spaces, we can show that all choices of
norms are equivalent. So to prove things about the extended field norm, often
we can just pick a convenient vector space norm, prove things about it, then
apply equivalence.
Definition
(Norm on vector space)
.
Let
K
be a valued field and
V
a vector
space over K. A norm on V is a function k·k : V R
0
such that
(i) kxk = 0 iff x = 0.
(ii) kλk = |λ|kxk for all λ K and x V .
(iii) kx + yk max{kxk, kyk}.
Note that our norms are also non-Archimedean.
Definition
(Equivalence of norms)
.
Let
k·k
and
k·k
0
be norms on
V
. Then
two norms are equivalent if they induce the same topology on
V
, i.e. there are
C, D > 0 such that
C kxk kxk
0
D kxk
for all x V .
One of the most convenient norms we will work with is the max norm:
Example
(Max norm)
.
Let
K
be a complete valued field, and
V
a finite-
dimensional K-vector space. Let x
1
, ··· , x
n
be a basis of V . Then if
x =
X
a
i
x
i
,
then
kxk
max
= max
i
|a
i
|
defines a norm on V .
Proposition.
Let
K
be a complete valued field, and
V
a finite-dimensional
K-vector space. Then V is complete under the max norm.
Proof.
Given a Cauchy sequence in
V
under the max norm, take the limit of each
coordinate to get the limit of the sequence, using the fact that
K
is complete.
That was remarkably easy. We can now immediately transfer this to all other
norms we can think of by showing all norms are equivalent.
Proposition.
Let
K
be a complete valued field, and
V
a finite-dimensional
K-vector space. Then any norm k·k on V is equivalent to k·k
max
.
Corollary. V is complete with respect to any norm.
Proof. Let k·k be a norm. We need to find C, D > 0 such that
C kxk
max
kxk D kxk
max
.
We set D = max
i
(kx
i
k). Then we have
kxk =
X
a
i
x
i
max (|a
i
|kx
i
k) (max |a
i
|)D = kxk
max
D.
We find
C
by induction on
n
. If
n
= 1, then
kxk
=
ka
1
x
1
k
=
|a
1
|kxk
=
kxk
max
kx
1
k. So C = kx
1
k works.
For n 2, we let
V
i
= Kx
1
··· Kx
i1
Kx
i+1
··· Kx
n
= span{x
1
, ··· , x
i1
, x
i+1
, ··· , x
n
}.
By the induction hypothesis, each
V
i
is complete with respect to (the restriction
of) k·k. So in particular V
i
is closed in V . So we know that the union
n
[
i=1
x
i
+ V
i
is also closed. By construction, this does not contain 0. So there is some
C >
0
such that if x
S
n
i=1
x
i
+ V
i
, then kxk C. We claim that
C kxk
max
kxk.
Indeed, take x =
P
a
i
x
i
V . Let r be such that
|a
r
| = max
i
(|a
i
|) = kxk
max
.
Then
kxk
1
max
kxk =
a
1
r
x
=
a
1
a
r
x
1
+ ··· +
a
r1
a
r
x
r1
+ x
r
+
a
r+1
a
r
x
r+1
+ ··· +
a
n
a
r
x
n
C,
since the last vector is an element of x
r
+ V
r
.
Before we can prove our theorem, we note the following two easy lemmas:
Lemma.
Let
K
be a valued field. Then the valuation ring
O
K
is integrally
closed in K.
Proof.
Let
x K
and
|x| >
1. Suppose we have
a
n1
, ··· , a
0
O
K
. Then we
have
|x
n
| > |a
0
+ a
1
x + ··· + a
n1
x
n1
|.
So we know
x
n
+ a
n1
x
n1
+ ··· + a
1
x + a
0
has non-zero norm, and in particular is non-zero. So
x
is not integral over
O
K
.
So O
K
is integrally closed.
Lemma.
Let
L
be a field and
| · |
a function that satisfies all axioms of an
absolute value but the strong triangle inequality. Then
| · |
is an absolute value
iff |α| 1 implies |α + 1| 1.
Proof.
It is clear that if
| · |
is an absolute value, then
|α|
1 implies
|α
+ 1
|
1.
Conversely, if this holds, and
|x| |y|
, then
|x/y|
1. So
|x/y
+ 1
|
1. So
|x + y| |y|. So |x + y| max{|x|, |y|}.
Finally, we get to prove our theorem.
Theorem.
Let
K
be a complete valued field, and let
L/K
be a finite extension.
Then the absolute value on
K
has a unique extension to an absolute value on
L
,
given by
|α|
L
=
n
q
N
L/K
(α)
,
where
n
= [
L
:
K
] and
N
L/K
is the field norm. Moreover,
L
is complete with
respect to this absolute value.
Proof.
For uniqueness and completeness, if
|·|
L
is an absolute value on
L
, then
it is in particular a
K
-norm on
L
as a finite-dimensional vector space. So we
know L is complete with respect to |·|
L
.
If
|·|
0
L
is another absolute value extending
|·|
, then we know
|·|
L
and
|·|
0
L
are equivalent in the sense of inducing the same topology. But then from one of
the early exercises, when field norms are equivalent, then we can find some
s >
0
such that
|·|
s
L
=
|·|
0
L
. But the two norms agree on
K
, and they are non-trivial.
So we must have s = 1. So the norms are equal.
To show existence, we have to prove that
|α|
L
=
n
q
N
L/K
(α)
is a norm.
(i) If |α|
L
= 0, then N
L/K
(α) = 0. This is true iff α = 0.
(ii)
The multiplicativity of
|α|
and follows from the multiplicativity of
N
L/K
,
|·| and
n
·.
To show the strong triangle inequality, it suffices to show that
|α|
L
1 implies
|α + 1|
L
1.
Recall that
O
L
= {α L : |α|
L
1} = {α L : N
L/K
(α) O
K
}.
We claim that
O
L
is the integral closure of
O
K
in
L
. This implies what we
want, since the integral closure is closed under addition (and 1 is in the integral
closure).
Let
α O
L
. We may assume
α 6
= 0, since that case is trivial. Let the
minimal polynomial of α over K be
f(x) = a
0
+ a
1
x + ··· + a
n1
x
n1
+ x
n
K[x].
We need to show that
a
i
O
K
for all
i
. In other words,
|a
i
|
1 for all
i
. This
is easy for a
0
, since
N
L/K
(α) = ±a
m
0
,
and hence |a
0
| 1.
By the corollary of Hensel’s lemma, for each i, we have
|a
i
| max(|a
0
|, 1)
By general properties of the field norm, there is some
m Z
1
such that
N
L/K
(α) = ±a
m
0
. So we have
|a
i
| max
N
L/K
(α)
1/m
, 1
= 1.
So f O
K
[x]. So α is integral over O
K
.
On the other hand, suppose
α
is integral over
O
K
. Let
¯
K/K
be an algebraic
closure of K. Note that
N
L/K
(α) =
Y
σ:L
¯
K
σ(α)
!
d
,
for some
d Z
1
, and each
σ
(
α
) is integral over
O
K
, since
α
is (apply
σ
to the
minimal polynomial). This implies that
N
L/K
(
α
) is integral over
O
K
(and lies
in K). So N
L/K
(α) O
K
since O
K
is integrally closed in K.
Corollary
(of the proof)
.
Let
K
be a complete valued field, and
L/K
a finite
extension. We equip
L
with
| · |
L
extending
| · |
on
K
. Then
O
L
is the integral
closure of O
K
in L.
2.3 Newton polygons
We are going to have a small digression to Newton polygons. We will not make
use of them in this course, but it is a cute visual devices that tell us about roots
of polynomials. It is very annoying to write down a formal definition, so we first
look at some examples. We will work with valuations rather than the absolute
value.
Example. Consider the valued field (Q
p
, v
p
), and the polynomial
t
4
+ p
2
t
4
p
3
t
2
+ pt + p
3
.
We then plot the coefficients for each power of
t
, and then draw a “convex
polygon” so that all points lie on or above it:
power of t
valuation of coefficient
1 2 3 40
1
2
3
Example. Consider (Q
2
, v
2
) with the polynomial
4t
4
+ 5t
3
+
7
2
t +
9
2
.
Here there is no t
2
term, so we simply don’t draw anything.
power of t
valuation of coefficient
1 2 3 40
1
1
2
We now go to come up with a formal definition.
Definition (Lower convex set). We say a set S R
2
is lower convex if
(i) Whenever (x, y) S, then (x, z) S for all z y.
(ii) S is convex.
Definition
(Lower convex hull)
.
Given any set of points
T R
2
, there is a
minimal lower convex set
S T
(by the intersection of all lower convex sets
containing
T
this is a non-empty definition because
R
2
satisfies the property).
This is known as the lower convex hull of the points.
Example.
The lower convex hull of the points (0
,
3)
,
(1
,
1)
,
(2
,
3)
,
(3
,
2)
,
(4
,
0) is
given by the region denoted below:
Definition
(Newton polygon)
.
Let
f
(
x
) =
a
0
+
a
1
x
+
···
+
a
n
x
n
K
[
x
], where
(
K, v
) is a valued field. Then the Newton polygon of
f
is the lower convex hull
of {(i, v(a
i
)) : i = 0, ··· , n, a
i
6= 0}.
This is the formal definition, so in our first example, the Newton polygon
really should be the shaded area shown above, but most of the time, we only
care about the lower line.
Definition
(Break points)
.
Given a polynomial, the points (
i, v
(
a
i
)) lying on
the boundary of the Newton polygon are known as the break points.
Definition
(Line segment)
.
Given a polynomial, the line segment between two
adjacent break points is a line segment.
Definition
(Multiplicity/length)
.
The length or multiplicity of a line segment
is the horizontal length.
Definition (Slope). The slope of a line segment is its slope.
Example. Consider again (Q
2
, v
2
) with the polynomial
4t
4
+ 5t
3
+
7
2
t +
9
2
.
power of t
valuation of coefficient
The middle segment has length 2 and slope 1/2.
Example. In the following Newton polygon:
The second line segment has length 3 and slope
1
3
.
It turns out the Newton polygon tells us something about the roots of the
polynomial.
Theorem. Let K be complete valued field, and v the valuation on K. We let
f(x) = a
0
+ a
1
x + ··· + a
n
x
n
K[x].
Let
L
be the splitting field of
f
over
K
, equipped with the unique extension
w
of v.
If (
r, v
(
a
r
))
(
s, v
(
a
s
)) is a line segment of the Newton polygon of
f
with
slope m R, then f has precisely s r roots of valuation m.
Note that by lower convexity, there can be at most one line segment for each
slope. So this theorem makes sense.
Proof.
Dividing by
a
n
only shifts the polygon vertically, so we may wlog
a
n
= 1.
We number the roots of f such that
w(α
1
) = ··· = w(α
s
1
) = m
1
w(α
s
1
+1
) = ··· = w(α
s
2
) = m
2
.
.
.
w(α
s
t
) = ··· = w(α
n
) = m
t+1
,
where we have
m
1
< m
2
< ··· < m
t+1
.
Then we know
v(a
n
) = v(1) = 0
v(a
n1
) = w
X
α
i
min
i
w(α
i
) = m
1
v(a
n2
) = w
X
α
i
α
j
min
i6=j
w(α
i
α
j
) = 2m
1
.
.
.
v(a
ns
1
) = w
X
i
1
6=...6=i
s
1
α
i
1
...α
i
s
1
= min w(α
i
1
···α
i
s
1
) = s
1
m
1
.
It is important that in the last one, we have equality, not an inequality, because
there is one term in the sum whose valuation is less than all the others.
We can then continue to get
v(α
ns
1
1
) min w(α
i
1
···α
i
s
1
+1
) = s
1
m
1
+ m
2
,
until we reach
v(α
ns
1
s
2
) = s
1
m
1
+ (s
2
s
1
)m
2
.
We keep going on.
We draw the Newton polygon.
(n, 0)
(n s
1
, s
1
m
1
)
(n s
1
s
2
, s
1
m
1
+ (s
2
s
1
)m
1
)
···
We don’t know where exactly the other points are, but the inequalities imply
that the (i, v(a
i
)) are above the lines drawn. So this is the Newton polygon.
Counting from the right, the first line segment has length
n
(
n s
1
) =
s
1
and slope
0 s
1
m
1
n (n s
1
)
= m
1
.
In general, the
k
th segment has length (
n s
k1
)
(
n s
k
) =
s
k
s
k1
, and
slope
s
1
m
1
+
P
k2
i=1
(s
i+1
s
i
)m
i+1
s
1
m
1
+
P
k1
i=1
(s
i+1
s
i
)m
i+1
s
k
s
k1
=
(s
k
s
k1
)m
k
s
k
s
k1
= m
k
.
and the others follow similarly.
Corollary.
If
f
is irreducible, then the Newton polygon has a single line segment.
Proof.
We need to show that all roots have the same valuation. Let
α, β
be in
the splitting field
L
. Then there is some
σ Aut
(
L/K
) such that
σ
(
α
) =
β
.
Then w(α) = w(σ(α)) = β. So done.
Note that Eisenstein’s criterion is a (partial) converse to this!
3 Discretely valued fields
We are now going to further specialize. While a valued field already has some
nice properties, we can’t really say much more about them without knowing
much about their valuations.
Recall our previous two examples of valued fields:
Q
p
and
F
p
((
T
)). The
valuations had the special property that they take values in
Z
. Such fields are
known as discretely valued fields.
Definition (Discretely valued field). Let K be a valued field with valuation v.
We say
K
is a discretely valued field (DVF) if
v
(
k
×
)
R
is a discrete subgroup
of R, i.e. v(k
×
) is infinite cyclic.
Note that we do not require the image to be exactly
Z R
. So we allow
scaled versions of the valuation. This is useful because the property of mapping
into
Z
is not preserved under field extensions in general, as we will later see. We
will call those that do land in Z normalized valuations.
Definition
(Normalized valuation)
.
Let
K
be a DVF. The normalized valuation
V
K
on
K
is the unique valuation on
K
in the given equivalence class of valuations
whose image is Z.
Note that the normalized valuation does not give us a preferred choice of
absolute value, since to obtain an absolute value, we still have to arbitrarily pick
the base c > 1 to define |x| = c
v(x)
.
Definition
(Uniformizer)
.
Let
K
be a discrete valued field. We say
π K
is
uniformizer if
v
(
π
)
>
0 and
v
(
π
) generates
v
(
k
×
) (iff
v
(
π
) has minimal positive
valuation).
So with a normalized valuation, we have v
K
(π) = 1.
Example.
The usual valuation on
Q
p
is normalized, and so is the usual valuation
on k((T )). p is a uniformizer for Q
p
and T is a uniformizer for k((T )).
The kinds of fields we will be interested are local fields. The definition we
have here might seem rather ad hoc. This is just one of the many equivalent
characterizations of a local field, and the one we pick here is the easiest to state.
Definition
(Local field)
.
A local field is a complete discretely valued field with
a finite residue field.
Example. Q
and
Q
p
with
v
p
are both discretely valued fields, and
Q
p
is a local
field. p is a uniformizer.
Example. The Laurent series field k((T)) with valuation
v
X
a
n
T
n
= inf{n : a
n
6= 0}
is a discrete valued field, and is a local field if and only if
k
is finite field, as the
residue field is exactly k. We have
O
k((T ))
= k[[T ]] =
(
X
n=0
a
n
T
n
: a
n
k
)
.
Here T is a uniformizer.
These discretely valued field are pretty much like the p-adic numbers.
Proposition.
Let
K
be a discretely valued field with uniformizer
π
. Let
S O
K
be a set of coset representatives of O
k
/m
k
= k
K
containing 0. Then
(i) The non-zero ideals of O
K
are π
n
O
K
for n 0.
(ii)
The ring
O
K
is a PID with unique prime
π
(up to units), and
m
K
=
πO
K
.
(iii)
The topology on
O
K
induced by the absolute value is the
π
-adic topology.
(iv) If K is complete, then O
K
is π-adically complete.
(v) If K is complete, then any x K can be written uniquely as
x =
X
n−∞
a
n
π
n
,
where a
n
S, and
|x| = |π|
inf{n:a
n
6=0}
.
(vi)
The completion
ˆ
K
is also discretely valued and
π
is a uniformizer, and
moreover the natural map
O
k
π
n
O
k
O
ˆ
K
π
n
O
ˆ
K
is an isomorphism.
Proof. The same as for Q
p
and Z
p
, with π instead of p.
Proposition.
Let
K
be a discretely valued field. Then
K
is a local field iff
O
K
is compact.
Proof.
If
O
K
is compact, then
π
n
O
K
is compact for all
n
0 (where
π
is the
uniformizer), and in particular complete. So
K =
[
n0
π
n
O
K
is complete, as this is an increasing union, and Cauchy sequences are bounded.
Also, we know the quotient map
O
K
k
K
is continuous when
k
K
is given the
discrete topology, by definition of the
π
-adic topology. So
k
K
is compact and
discrete, hence finite.
In the other direction, if
K
is local, then we know
O
K
n
O
K
is finite for
all
n
0 (by induction and finiteness of
k
K
). We let (
x
i
) be a sequence in
O
K
.
Then by finiteness of
O
K
O
K
, there is a subsequence (
x
1,i
) which is constant
modulo
π
. We keep going, choosing a subsequence (
x
n+1,i
) of (
x
n
i
) such that
(
x
n+1,i
) is constant modulo
π
n+1
. Then (
x
i,i
)
i=1
converges, since it is Cauchy as
|x
ii
x
jj
| |π|
j
for j i. So O
K
is sequentially compact, hence compact.
Now the valuation ring
O
K
inherits a valuation from
K
, and this gives it a
structure of a discrete valuation ring. We will define a discrete valuation ring in
a funny way, but there are many equivalent definitions that we will not list.
Definition
(Discrete valuation ring)
.
A ring
R
is called a discrete valuation
ring (DVR) if it is a PID with a unique prime element up to units.
Proposition. R is a DVR iff R
=
O
K
for some DVF K.
Proof.
We have already seen that valuation rings of discrete valuation fields are
DVRs. In the other direction, let
R
be a DVR, and
π
a prime. Let
x R \ {
0
}
.
Then we can find a unique unit
u R
×
and
n Z
0
such that
x
=
π
n
u
(say,
by unique factorization of PIDs). We define
v(x) =
(
n x 6= 0
x = 0
This is then a discrete valuation of
R
. This extends uniquely to the field of
fractions K. It remains to show that R = O
K
. First note that
K = R
1
π
.
This is since any non-zero element in
R
1
π
looks like
π
n
u, u R
×
, n Z
, and
is already invertible. So it must be the field of fractions. Then we have
v(π
n
u) = n Z
0
π
n
u R.
So we have R = O
K
.
Now recall our two “standard” examples of valued fields
F
p
((
T
)) and
Q
p
. Both of their residue fields are
F
p
, and in particular has characteristic
p
.
However,
F
p
((
T
)) itself is also of characteristic
p
, while
Q
p
has characteristic 0.
It would thus be helpful to split these into two different cases:
Definition
(Equal and mixed characteristic)
.
Let
K
be a valued field with
residue field k
K
. Then K has equal characteristic if
char K = char k
K
.
Otherwise, we have K has mixed characteristic.
If
K
has mixed characteristic, then necessarily
char K
= 0, and
char k
K
>
0.
Example. Q
p
has mixed characteristic, since
char Q
p
= 0 but
char k
Q
p
=
Z/pZ = p.
We will also need the following definition:
Definition
(Perfect ring)
.
Let
R
be a ring of characteristic
p
. We say
R
is
perfect if the Frobenius map
x 7→ x
p
is an automorphism of
R
, i.e. every element
of R has a pth root.
Fact.
Let
F
be a field of characteristic
p
. Then
F
is perfect if and only if every
finite extension of F is separable.
Example. F
q
is perfect for every q = p
n
.
3.1 Teichm¨uller lifts
Take our favorite discretely valued ring
Z
p
. This is
p
-adically complete, so we
can write each element as
x = a
0
+ a
1
p + a
2
p
2
+ ··· ,
where each
a
i
is in
{
0
,
1
, ··· , p
1
}
. The reason this works is that 0
,
1
, ··· , p
1 are coset representatives of the ring
Z
p
/pZ
p
=
Z/pZ
. While these coset
representatives might feel like a “natural” thing to do in this context, this is
because we have implicitly identified with
Z
p
/pZ
p
=
Z/pZ
as a particular subset
of
Z Z
p
. However, this identification respects effectively no algebraic structure
at all. For example, we cannot multiplying the cosets simply by multiplying the
representatives as elements of
Z
p
, because, say, (
p
1)
2
=
p
2
2
p
+ 1, which is
not 1. So this is actually quite bad, at least theoretically.
It turns out that we can actually construct “natural” lifts in a very general
scenario.
Theorem.
Let
R
be a ring, and let
x R
. Assume that
R
is
x
-adically
complete and that
R/xR
is perfect of characteristic
p
. Then there is a unique
map [] : R/xR R such that
[a] a mod x
and
[ab] = [a][b].
for all
a, b R/xR
. Moreover, if
R
has characteristic
p
, then [
] is a ring
homomorphism.
Definition
(Teichm¨uller map)
.
The map [
] :
R/xR R
is called the Te-
ichm uller map. [x] is called the Teichm¨uller lift or representative of x.
The idea of the proof is as follows: suppose we have an
a R/xR
. If we
randomly picked a lift
α
, then chances are it would be a pretty “bad” choice,
since any two such choices can differ by a multiple of x.
Suppose we instead lifted a
p
th root of
a
to
R
, and then take the
p
th power
of it. We claim that this is a better way of picking a lift. Suppose we have picked
two lifts of a
p
1
, say, α
1
and α
0
1
. Then α
0
1
= xc + α
1
for some c. So we have
(α
0
1
)
p
α
p
1
= α
p
1
+ pxc + O(x
2
) α
p
1
= pxc + O(x
2
),
where we abuse notation and write
O
(
x
2
) to mean terms that are multiples of
x
2
.
We now recall that
R/xR
has characteristic
p
, so
p xR
. Thus in fact
pxc = O(x
2
). So we have
(α
0
1
)
p
α
p
1
= O(x
2
).
So while the lift is still arbitrary, any two arbitrary choices can differ by at most
x
2
. Alternatively, our lift is now a well-defined element of R/x
2
R.
We can, of course, do better. We can lift the p
2
th root of a to R, then take
the
p
2
th power of it. Now any two lifts can differ by at most
O
(
x
3
). More
generally, we can try to lift the
p
n
th root of
a
, then take the
p
n
th power of
it. We keep picking a higher and higher
n
, take the limit, and hopefully get
something useful out!
To prove this result, we will need the following messy lemma:
Lemma.
Let
R
be a ring with
x R
such that
R/xR
has characteristic
p
. Let
α, β R be such that
α = β mod x
k
()
Then we have
α
p
= β
p
mod x
k+1
.
Proof.
It is left as an exercise to modify the proof to work for
p
= 2 (it is actually
easier). So suppose p is odd. We take the pth power of () to obtain
α
p
β
p
+
p1
X
i=1
p
i
α
pi
β
i
x
p(k+1)
R.
We can now write
p1
X
i=1
(1)
i
p
i
α
pi
β
i
=
p1
2
X
i=1
(1)
i
p
i
(αβ)
i
α
p2i
β
p2i
= p(α β)(something).
Now since
R/xR
has characteristic
p
, we know
p xR
. By assumption, we know
α β x
k+1
R. So this whole mess is in x
k+2
R, and we are done.
Proof of theorem.
Let
a R/xR
. For each
n
, there is a unique
a
p
n
R/xR
.
We lift this arbitrarily to some α
n
R such that
α
n
a
p
n
mod x.
We define
β
n
= α
p
n
n
.
The claim is that
[a] = lim
n→∞
β
n
exists and is independent of the choices.
Note that if the limit exists no matter how we choose the
α
n
, then it
must be independent of the choices. Indeed, if we had choices
β
n
and
β
0
n
,
then
β
1
, β
0
2
, β
3
, β
0
4
, β
5
, β
0
6
, ···
is also a respectable choice of lifts, and thus must
converge. So β
n
and β
0
n
must have the same limit.
Since the ring is
x
-adically complete and is discretely valued, to show the
limit exists, it suffices to show that β
n+1
β
n
0 x-adically. Indeed, we have
β
n+1
β
n
= (α
p
n+1
)
p
n
α
p
n
n
.
We now notice that
α
p
n+1
(a
p
n1
)
p
= a
p
n
α
n
mod x.
So by applying the previous the lemma many times, we obtain
(α
p
n+1
)
p
n
α
p
n
n
mod x
n+1
.
So β
n+1
β
n
x
n+1
R. So lim β
n
exists.
To see [a] = a mod x, we just have to note that
lim
n→∞
α
p
n
n
lim
n→∞
(a
p
n
)
p
n
= lim a = a mod x.
(here we are using the fact that the map
R R/xR
is continuous when
R
is
given the x-adic topology and R/xR is given the discrete topology)
The remaining properties then follow trivially from the uniqueness of the
above limit.
For multiplicativity, if we have another element
b R/xR
, with
γ
n
R
lifting b
p
n
for all n, then α
n
γ
n
lifts (ab)
p
n
. So
[ab] = lim α
p
n
n
γ
p
n
n
= lim α
p
n
n
lim γ
p
n
n
= [a][b].
If R has characteristic p, then α
n
+ γ
n
lifts a
p
n
+ b
p
n
= (a + b)
p
n
. So
[a + b] = lim(α
n
+ γ
n
)
p
n
= lim α
p
n
n
+ lim γ
p
n
n
= [a] + [b].
Since 1 is a lift of 1 and 0 is a lift of 0, it follows that this is a ring homomorphism.
Finally, to show uniqueness, suppose
φ
:
R/xR R
is a map with these
properties. Then we note that
φ
(
a
p
n
)
a
p
n
mod x
, and is thus a valid choice
of α
n
. So we have
[a] = lim
n→∞
φ(a
p
n
)
p
n
= lim φ(a) = φ(a).
Example. Let R = Z
p
and x = p. Then [] : F
p
Z
p
satisfies
[x]
p1
= [x
p1
] = [1] = 1.
So the image of [
x
] must be the unique
p
1th root of unity lifting
x
(recall we
proved their existence via Hensel’s lemma).
When proving theorems about these rings, the Teichm¨uller lifts would be
very handy and natural things to use. However, when we want to do actual
computations, there is absolutely no reason why these would be easier!
As an application, we can prove the following characterization of equal
characteristic complete DVF’s.
Theorem.
Let
K
be a complete discretely valued field of equal characteristic
p
,
and assume that k
K
is perfect. Then K
=
k
K
((T )).
Proof.
Let
K
be a complete DVF. Since every DVF the field of fractions of
its valuation ring, it suffices to prove that
O
K
=
k
K
[[
T
]]. We know
O
K
has
characteristic
p
. So [
] :
k
K
O
K
is an injective ring homomorphism. We
choose a uniformizer π O
K
, and define
k
K
[[T ]] O
K
by
X
n=0
a
n
T
n
7→
X
n=0
[a
n
]π
n
.
Then this is a ring homomorphism since [
] is. The bijectivity follows from
property (v) in our list of properties of complete DVF’s.
Corollary.
Let
K
be a local field of equal characteristic
p
. Then
k
K
=
F
q
for
some q a power of p, and K
=
F
q
((T )).
3.2 Witt vectors*
We are now going to look at the mixed characteristic analogue of this result. We
want something that allows us to go from characteristic
p
to characteristic 0.
This is known as Witt vectors, which is non-examinable.
We start with the notion of a strict
p
-ring. Roughly this is a ring that satisfies
all the good properties whose name has the word p in it.
Definition
(Strict
p
-ring)
.
Let
A
be a ring. A is called a strict
p
-ring if it is
p-torsion free, p-adically complete, and A/pA is a perfect ring.
Note that a strict
p
-ring in particular satisfies the conditions for the Te-
ichm¨uller lift to exist, for x = p.
Example. Z
p
is a strict p-ring.
The next example we are going to construct is more complicated. This is in
some sense a generalization of the usual polynomial rings
Z
[
x
1
, ··· , x
n
], or more
generally,
Z[x
i
| i I],
for
I
possibly infinite. To construct the “free” strict
p
-ring, after adding all these
variables
x
i
, to make it a strict
p
-ring, we also need to add their
p
th roots, and
the p
2
th roots etc, and then take the p-adic completion, and hope for the best.
Example. Let X = {x
i
: i I} be a set. Let
B = Z[x
p
−∞
i
| i I] =
[
n=0
Z[x
p
n
i
| i I].
Here the union on the right is taken by treating
Z[x
i
| i I] Z[x
p
1
i
| i I] ···
in the natural way.
We let
A
be the
p
-adic completion of
B
. We claim that
A
is a strict
p
-ring
and A/pA
=
F
p
[x
p
−∞
i
| i I].
Indeed, we see that
B
is
p
-torsion free. By Exercise 13 on Sheet 1, we know
A is p-adically complete and torsion free. Moreover,
A/pA
=
B/pB
=
F
p
[x
p
−∞
i
| i I],
which is perfect since every element has a p-th root.
If A is a strict p-ring, then we know that we have a Teichm¨uller map
[] : A/pA A,
Lemma.
Let
A
be a strict
p
-ring. Then any element of
A
can be written
uniquely as
a =
X
n=0
[a
n
]p
n
,
for a unique a
n
A/pA.
Proof. We recursively construct the a
n
by
a
0
= a (mod p)
a
1
p
1
(a [a
0
]) (mod p)
.
.
.
Lemma.
Let
A
and
B
be strict
p
-rings and let
f
:
A/pA B/pB
be a ring
homomorphism. Then there is a unique homomorphism
F
:
A B
such that
f = F mod p, given by
F
X
[a
n
]p
n
=
X
[f(a
n
)]p
n
.
Proof sketch.
We define
F
by the given formula and check that it works. First of
all, by the formula,
F
is
p
-adically continuous, and the key thing is to check that
it is additive (which is slightly messy). Multiplicativity then follows formally
from the continuity and additivity.
To show uniqueness, suppose that we have some
ψ
lifting
f
. Then
ψ
(
p
) =
p
.
So ψ is p-adically continuous. So it suffices to show that ψ([a]) = [ψ(a)].
We take α
n
A lifting a
p
n
A/pA. Then ψ(α
n
) lifts f(a)
p
n
. So
ψ([a]) = lim ψ(α
p
n
n
) = lim ψ(α
n
)
p
n
= [f(a)].
So done.
There is a generalization of this result:
Proposition.
Let
A
be a strict
p
-ring and
B
be a ring with an element
x
such that
B
is
x
-adically complete and
B/xB
is perfect of characteristic
p
. If
f
:
A/pA B/xB
is a ring homomorphism. Then there exists a unique ring
homomorphism
F
:
A B
with
f
=
F mod x
, i.e. the following diagram
commutes:
A B
A/pA B/xB
F
f
.
Indeed, the conditions on
B
are sufficient for Teichm¨uller lifts to exist, and
we can at least write down the previous formula, then painfully check it works.
We can now state the main theorem about strict p-rings.
Theorem.
Let
R
be a perfect ring. Then there is a unique (up to isomorphism)
strict p-ring W (B) called the Witt vectors of R such that W (R)/pW (R)
=
R.
Moreover, for any other perfect ring
R
, the reduction mod
p
map gives a
bijection
Hom
Ring
(W (R), W (R
0
)) Hom
Ring
(R, R
0
)
.
Proof sketch.
If
W
(
R
) and
W
(
R
0
) are such strict
p
-rings, then the second part
follows from the previous lemma. Indeed, if
C
is a strict
p
-ring with
C/pC
=
R
=
W
(
R
)
/pW
(
R
), then the isomorphism
¯α
:
W
(
R
)
/pW
(
R
)
C/pC
and its
inverse
¯α
1
have unique lifts
γ
:
W
(
R
)
C
and
γ
1
:
C W
(
R
), and these
are inverses by uniqueness of lifts.
To show existence, let R be a perfect ring. We form
F
p
[x
p
−∞
r
| r R] R
x
r
7→ r
Then we know that the
p
-adic completion of
Z
[
x
p
−∞
r
| r R
], written
A
, is a
strict p-ring with
A/pA
=
F
p
[x
p
−∞
r
| r R].
We write
I = ker(F
p
[x
p
−∞
r
| r R] R).
Then define
J =
(
X
n=0
[a
k
]p
n
A : a
n
I for all n
)
.
This turns out to be an ideal.
J A R
0 I A/pA R 0
We put
W
(
R
) =
A/J
. We can then painfully check that this has all the required
properties. For example, if
x =
X
n=0
[a
n
]p
n
A,
and
px =
X
n=0
[a
n
]p
n+1
J,
then by definition of
J
, we know [
a
n
]
I
. So
x J
. So
W
(
R
)
/J
is
p
-torsion
free. By a similar calculation, one checks that
\
n=0
p
n
W (R) = {0}.
This implies that
W
(
R
) injects to its
p
-adic completion. Using that
A
is
p
-adically
complete, one checks the surjectivity by hand.
Also, we have
W (R)
pW (R)
=
A
J + pA
.
But we know
J + pA =
(
X
n
[a
n
]p
n
| a
0
I
)
.
So we have
W (R)
pW (R)
=
F
p
[x
p
−∞
r
| r R]
I
=
R.
So we know that W (R) is a strict p-ring.
Example. W
(
F
p
) =
Z
p
, since
Z
p
satisfies all the properties
W
(
F
p
) is supposed
to satisfy.
Proposition.
A complete DVR
A
of mixed characteristic with perfect residue
field and such that p is a uniformizer is the same as a strict p-ring A such that
A/pA is a field.
Proof.
Let
A
be a complete DVR such that
p
is a uniformizer and
A/pA
is
perfect. Then
A
is
p
-torsion free, as
A
is an integral domain of characteristic 0.
Since it is also p-adically complete, it is a strict p-ring.
Conversely, if
A
is a strict
p
-ring, and
A/pA
is a field, then we have
A
×
A \ pA, and we claim that A
×
= A \ pA. Let
x =
X
n=0
[x
n
]p
n
with
x
0
6
= 0, i.e.
x 6∈ pA
. We want to show that
x
is a unit. Since
A/pA
is a
field, we can multiply by [
x
1
0
], so we may wlog
x
0
= 1. Then
x
= 1
py
for
some y A. So we can invert this with a geometric series
x
1
=
X
n=0
p
n
y
n
.
So
x
is a unit. Now, looking at Teichm¨uller expansions and factoring out multiple
of
p
, any non-zero element
z
can be written as
p
n
u
for a unique
n Z
0
and
u A
×
. Then we have
v(z) =
(
n z 6= 0
z = 0
is a discrete valuation on A.
Definition
(Absolute ramification index)
.
Let
R
be a DVR with mixed charac-
teristic
p
with normalized valuation
v
R
. The integer
v
R
(
p
) is called the absolute
ramification index of R.
Corollary.
Let
R
be a complete DVR of mixed characteristic with absolute
ramification index 1 and perfect residue field k. Then R
=
W (k).
Proof.
Having absolute ramification index 1 is the same as saying
p
is a uni-
formizer. So
R
is a strict
p
-ring with
R/pR
=
k
. By uniqueness of the Witt
vector, we know R
=
W (k).
Theorem.
Let
R
be a complete DVR of mixed characteristic
p
with a perfect
residue field k and uniformizer π. Then R is finite over W (k).
Proof.
We need to first exhibit
W
(
k
) as a subring of
R
. We know that
id
:
k k
lifts to a homomorphism
W
(
k
)
R
. The kernel is a prime ideal because
R
is
an integral domain. So it is either 0 or
pW
(
k
). But
R
has characteristic 0. So it
can’t be pW(k). So this must be an injection.
Let e be the absolute ramification index of R. We want to prove that
R =
e1
M
i=0
π
i
W (k).
Looking at valuations, one sees that 1
, π, π, ··· , π
e1
are linearly independent
over W (k). So we can form
M =
e1
M
i=0
π
i
W (k) R.
We consider R/pR. Looking at Teichm¨uller expansions
X
n=0
[x
n
]π
n
e1
X
n=0
[x
n
]π
n
mod pR,
we see that 1
, π, ··· , π
e1
generate
R/pR
as
W
(
k
)-modules (all the Teichm¨uller
lifts live in W (k)). Therefore R = M + pR. We iterate to get
R = M + p(M + pR) = M + p
2
r = ··· = M + p
m
R
for all
m
1. So
M
is dense in
R
. But
M
is also
p
-adically complete, hence
closed in R. So M = R.
The important statement to take away is
Corollary.
Let
K
be a mixed characteristic local field. Then
K
is a finite
extension of Q
p
.
Proof.
Let
F
q
be the residue field of
K
. Then
O
K
is finite over
W
(
F
q
) by the
previous theorem. So it suffices to show that
W
(
F
q
) is finite over
W
(
F
p
) =
Z
p
.
Again the inclusion
F
p
F
q
gives an injection
W
(
F
p
)
W
(
F
q
). Write
q
=
p
d
,
and let x
1
, ··· , x
d
W (F
q
) be lifts of an F
p
-bases of F
q
.. Then we have
W (F
q
) =
d
M
i=1
x
d
Z
p
+ pW (F
q
),
and then argue as in the end of the previous theorem to get
W (F
q
) =
d
M
i=1
x
d
Z
p
.
4 Some p-adic analysis
We are now going to do some fun things that is not really related to the course.
In “normal” analysis, the applied mathematicians hold the belief that every
function can be written as a power series
f(x) =
X
n=0
a
n
x
n
.
When we move on to
p
-adic numbers, we do not get such a power series expansion.
However, we obtain an analogous result using binomial coefficients.
Before that, we have a quick look at our familiar functions
exp
and
log
, which
we shall continue to define as a power series:
exp(x) =
X
n=0
x
n
n!
, log(1 + x) =
X
n=1
(1)
n1
x
n
n
The domain will no longer be all of the field. Instead, we have the following
result:
Proposition.
Let
K
be a complete valued field with an absolute value
| · |
and
assume that
K Q
p
and
| · |
restricts to the usual
p
-adic norm on
Q
p
. Then
exp
(
x
) converges for
|x| < p
1/(p1)
and
log
(1 +
x
) converges for
|x| <
1, and
then define continuous maps
exp : {x K : |x| < p
1/(p1)
} O
K
log : {1 + x K : |x| < 1} K.
Proof.
We let
v
=
log
p
| · |
be a valuation extending
v
p
. Then we have the
dumb estimate
v(n) log
p
n.
Then we have
v
x
n
n
n · v(x) log
p
n
if v(x) > 0. So log converges.
For exp, we have
v(n!) =
n s
p
(n)
p 1
,
where s
p
(n) is the sum of the p-adic digits of n. Then we have
v
x
n
n!
n · v(x)
n
p 1
= n ·
v(x)
1
p 1
if v(x) > 1/(p 1). Since v
x
n
n!
0, this lands in O
K
.
For the continuity, we just use uniform convergence as in the real case.
What we really want to talk about is binomial coefficients. Let
n
1. Then
we know that
x
n
=
x(x 1) ···(x n + 1)
n!
is a polynomial in
x
, and so defines a continuous function
Z
p
Q
p
by
x 7→
x
n
.
When n = 0, we set
x
n
= 1 for all x Z
p
.
We know
x
n
Z
if
x Z
0
. So by density of
Z
0
Z
p
, we must have
x
n
Z
p
for all x Z
p
.
We will eventually want to prove the following result:
Theorem
(Mahler’s theorem)
.
Let
f
:
Z
p
Q
p
be any continuous function.
Then there is a unique sequence (a
n
)
n0
with a
n
Q
p
and a
n
0 such that
f(x) =
X
n=0
a
n
x
n
,
and moreover
sup
xZ
p
|f(x)| = max
kN
|a
k
|.
We write
C
(
Z
p
, Q
p
) for the set of continuous functions
Z
p
Q
p
as usual.
This is a Q
p
vector space as usual, with
(λf + µg)(x) = λf(x) + µg(x)
for all λ, µ Q
p
and f, g C(Z
p
, Q
p
) and x Z
p
.
If f C(Z
p
, Q
p
), we set
kfk = sup
xZ
p
|f(x)|
p
.
Since
Z
p
is compact, we know that
f
is bounded. So the supremum exists and
is attained.
Proposition.
The norm
k · k
defined above is in fact a (non-archimedean)
norm, and that C(Z
p
, Q
p
) is complete under this norm.
Let
c
0
denote the set of sequences (
a
n
)
n=0
in
Q
p
such that
a
n
0. This is
a Q
p
-vector space with a norm
k(a
n
)k = max
nN
|a
n
|
p
,
and
c
0
is complete. So what Mahler’s theorem gives us is an isometric isomor-
phism between c
0
and C(Z
p
, Q
p
).
We define
∆ : C(Z
p
, Q
p
) C(Z
p
, Q
p
)
by
f(x) = f (x + 1) f(x).
By induction, we have
n
f(x) =
n
X
i=0
(1)
i
n
i
f(x + n i).
Note that is a linear operator on C(Z
p
, Q
p
), and moreover
|f(x)|
p
= |f(x + 1) f(x)|
p
kfk.
So we have
kfk kfk.
In other words, we have
kk 1.
Definition
(Mahler coefficient)
.
Let
f C
(
Z
p
, Q
p
). Then
n
th-Mahler coeffi-
cient a
n
(f) Q
p
is defined by the formula
a
n
(f) = ∆
n
(f)(0) =
n
X
i=0
(1)
i
n
i
f(n i).
We will eventually show that these are the
a
n
’s that appear in Mahler’s
theorem. The first thing to prove is that these coefficients do tend to 0. We
already know that they don’t go up, so we just have to show that they always
eventually go down.
Lemma. Let f C(Z
p
, Q
p
). Then there exists some k 1 such that
k
p
k
fk
1
p
kfk.
Proof.
If
f
= 0, there is nothing to prove. So we will wlog
kfk
= 1 by scaling
(this is possible since the norm is attained at some
x
0
, so we can just divide by
f(x
0
)). We want to find some k such that
p
k
f(x) 0 mod p
for all x. To do so, we use the explicit formula
p
k
f(x) =
p
k
X
i=0
(1)
i
p
k
i
f(x + p
k
i) f(x + p
k
) f(x) (mod p)
because the binomial coefficients
p
k
i
are divisible by
p
for
i 6
= 0
, p
k
. Note that
we do have a negative sign in front of
f
(
x
) because (
1)
p
k
is
1 as long as
p
is
odd, and 1 = 1 if p = 2.
Now
Z
p
is compact. So
f
is uniformly continuous. So there is some
k
such
that
|x y|
p
p
k
implies
|f
(
x
)
f
(
y
)
|
p
p
1
for all
x, y Z
p
. So take this
k, and we’re done.
We can now prove that the Mahler’s coefficients tend to 0.
Proposition.
The map
f 7→
(
a
n
(
f
))
n=0
defines an injective norm-decreasing
linear map C(Z
p
, Q
p
) c
0
.
Proof. First we prove that a
n
(f) 0. We know that
ka
n
(f)k
p
k
n
fk.
So it suffices to show that
k
n
fk
0. Since
k
k
1, we know
k
n
fk
is
monotonically decreasing. So it suffices to find a subsequence that tends to 0.
To do so, we simply apply the lemma repeatedly to get k
1
, k
2
, ··· such that
p
k
1
+...+k
n
1
p
n
kfk.
This gives the desired sequence.
Note that
|a
n
(f)|
p
k
n
k kfk.
So we know
k(a
n
(f))
n
k = max |a
n
(f)|
p
kfk.
So the map is norm-decreasing. Linearity follows from linearity of ∆. To finish,
we have to prove injectivity.
Suppose a
n
(f) = 0 for all n 0. Then
a
0
(f) = f (0) = 0,
and by induction,we have that
f(n) = ∆
k
f(0) = a
n
(f) = 0.
for all
n
0. So
f
is constantly zero on
Z
0
. By continuity, it must be zero
everywhere on Z
p
.
We are almost at Mahler’s theorem. We have found some coefficients already,
and we want to see that it works. We start by proving a small, familiar, lemma.
Lemma. We have
x
n
+
x
n 1
=
x + 1
n
for all n Z
1
and x Z
p
.
Proof.
It is well known that this is true when
x Z
n
. Since the expressions
are polynomials in
x
, them agreeing on infinitely many values implies that they
are indeed the same.
Proposition. Let a = (a
n
)
n=0
c
0
. We define f
a
: Z
p
Q
p
by
f
a
(x) =
X
n=0
a
n
x
n
.
This defines a norm-decreasing linear map
c
0
C
(
Z
p
, Q
p
). Moreover
a
n
(
f
a
) =
a
n
for all n 0.
Proof. Linearity is clear. Norm-decreasing follows from
|f
a
(x)| =
X
a
n
x
n
sup
n
|a
n
|
p
x
n
p
sup
n
|a
n
|
p
= ka
n
k,
where we used the fact that
x
n
Z
p
, hence
x
n
p
1.
Taking the supremum, we know that
kf
a
k kak.
For the last statement, for all k Z
0
, we define
a
(k)
= (a
k
, a
k+1
, a
k+1
, ···).
Then we have
f
a
(x) = f
a
(x + 1) f
a
(x)
=
X
n=1
a
n

x + 1
n
x
n

=
X
n=1
a
n
x
n 1
=
X
n=0
a
n+1
x
n
= f
a
(
1)
(x)
Iterating, we have
k
f
a
= f
a
(k)
.
So we have
a
n
(f
a
) = ∆
n
f
a
(0) = f
a
(n)
(0) = a
n
.
Summing up, we now have maps
C(Z
p
, Q
p
) c
0
F
G
with
F (f) = (a
n
(f))
G(a) = f
a
.
We now that
F
is injective and norm-decreasing, and
G
is norm-decreasing
and
F G
=
id
. It then follows formally that
GF
=
id
and the maps are norm-
preserving.
Lemma.
Suppose
V, W
are normed spaces, and
F
:
V W
,
G
:
W V
are
maps such that
F
is injective and norm-decreasing, and
G
is norm-decreasing
and F G = id
W
. Then GF = id
V
and F and G are norm-preserving.
Proof. Let v V . Then
F (v GF v) = F v F GF v = (F F )v = 0.
Since F is injective, we have
v = GF v.
Also, we have
kvk kF vk kGF vk = kvk.
So we have equality throughout. Similarly, we have kvk = kGvk.
This finishes the proof Mahler’s theorem, and also finishes this section on
p-adic analysis.
5 Ramification theory for local fields
From now on, the characteristic of the residue field of any local field will be
denoted p, unless stated otherwise.
5.1 Ramification index and inertia degree
Suppose we have an extension
L/K
of local fields. Then since
m
K
m
L
, and
O
L
O
L
, we obtain an injection
k
K
=
O
K
m
k
O
L
m
L
= k
L
.
So we also get an extension of residue fields
k
L
/k
K
. The question we want to ask
is how much of the extension is “due to” the extension of residue fields
k
L
/k
K
,
and how much is “due to” other things happening.
It turns out these are characterized by the following two numbers:
Definition
(Inertia degree)
.
Let
L/K
be a finite extension of local fields. The
inertia degree of L/K is
f
L/K
= [k
L
: k
K
].
Definition
(Ramification index)
.
Let
L/K
be a finite extension of local fields,
and let
v
L
be the normalized valuation of
L
and
π
K
a uniformizer of
K
. The
integer
e
L/K
= v
L
(π
K
)
is the ramification index of L/K.
The goal of the section is to show the following result:
Theorem. Let L/K be a finite extension. Then
[L : K] = e
L/K
f
L/K
.
We then have two extreme cases of ramification:
Definition
(Unramified extension)
.
Let
L/K
be a finite extension of local fields.
We say L/K is unramified if e
L/K
= 1, i.e. f
L/K
= [L : K].
Definition
(Totally ramified extension)
.
Let
L/K
be a finite extension of local
fields. We say L/K is totally ramified if f
L/K
= 1, i.e. e
L/K
= [L : K].
In the next section we will, amongst many things, show that every extension
of local fields can be written as an unramified extension followed by a totally
ramified extension.
Recall the following: let
R
be a PID and
M
a finitely-generated
R
-module.
Assume that
M
is torsion-free. Then there is a unique integer
n
0 such that
M
=
R
n
. We say
n
has rank
n
. Moreover, if
N M
is a submodule, then
N
is
finitely-generated, so N
=
R
m
for some m n.
Proposition.
Let
K
be a local field, and
L/K
a finite extension of degree
n
.
Then
O
L
is a finitely-generated and free
O
K
module of rank
n
, and
k
L
/k
K
is
an extension of degree n.
Moreover, L is also a local field.
Proof.
Choose a
K
-basis
α
1
, ··· , α
n
of
L
. Let
k · k
denote the maximum norm
on L.
n
X
i=1
x
i
α
i
= max
i=1,...,n
|x
i
|
as before. Again, we know that
k · k
is equivalent to the extended norm
| · |
on
L as K-norms. So we can find r > s > 0 such that
M = {x L : kxk s} O
L
N = {x L : kxk r}.
Increasing r and decreasing s if necessary, we wlog r = |a| and s = |b| for some
a, b K.
Then we can write
M =
n
M
i=1
O
k
i
O
L
N =
n
M
i=1
O
K
i
.
We know that
N
is finitely generated and free of rank
n
over
O
K
, and so is
M
.
So O
L
must be finitely generated and free of rank n over O
K
.
Since m
k
= m
k
O
K
, we have a natural injection
O
K
m
k
O
L
m
L
= k
L
.
Since
O
L
is generated over
O
K
by
n
elements, we know that
k
K
is generated by
n elements over k
K
, so it has rank at most n.
To see that
L
is a local field, we know that
k
L
/k
K
is finite and
k
K
is finite,
so
k
L
is finite. It is complete under the norm because it is a finite-dimensional
vector space over a complete field.
Finally, to see that the valuation is discrete, suppose we have a normalized
valuation on K, and w the unique extension of v
K
to L. Then we have
w(α) =
1
n
v
K
(N
L/K
(α)).
So we have
w(L
×
)
1
n
v(K
×
) =
1
n
Z.
So it is discrete.
Note that we cannot just pick an arbitrary basis of
L/K
and scale it to give
a basis of
O
L
/O
K
. For example,
Q
2
(
2
)
/Q
2
has basis 1
,
2
, but
|
2|
=
1
2
and cannot be scaled to 1 by an element in Q
2
.
Even if such a scaled basis exists, it doesn’t necessarily give a basis of the
integral rings. For example,
Q
3
(
1
)
/Q
3
has a
Q
3
-basis 1
,
1 + 3
1
and
|1 + 3
1| = 1, but
1 6∈ Z
3
+ Z
3
(1 + 3
1).
So this is not a basis of O
Q
3
(
1)
over Z
3
.
Theorem. Let L/K be a finite extension. Then
[L : K] = e
L/K
f
L/K
,
and there is some α O
L
such that O
L
= O
K
[α].
Proof.
We will be lazy and write
e
=
e
L/K
and
f
=
f
L/K
. We first note that
k
L
/k
K
is separable, so there is some
¯α k
L
such that
k
L
=
k
K
(
¯α
) by the
primitive element theorem. Let
¯
f(x) k
K
[x]
be the minimal polynomial of
¯α
over
k
K
and let
f O
L
[
x
] be a monic lift of
¯
f
with deg f = deg
¯
f.
We first claim that there is some
α O
L
lifting
¯α
such that
v
L
(
f
(
α
)) = 1
(note that it is always
1). To see this, we just take any lift
β
. If
v
L
(
f
(
β
)) = 1,
then we are happy and set
α
=
β
. If it doesn’t work, we set
α
=
β
+
π
L
, where
π
L
is the uniformizer of L.
Then we have
f(α) = f (β + π
L
) = f(β) + f
0
(β)π
L
+
2
L
for some
b O
L
, by Taylor expansion around
β
. Since
v
L
(
f
(
β
))
2 and
v
L
(
f
0
(
β
)) = 0 (since
¯
f
is separable, we know
f
0
(
β
) does not vanish when we
reduce mod m), we know v
L
(f(α)) = 1. So f(α) is a uniformizer of L.
We now claim that the elements
α
i
π
j
for
i
= 0
, ··· , f
1 and
j
= 0
, ··· , e
1
are an O
K
-basis of O
L
. Suppose we have
X
i,j
a
ij
α
i
π
j
= 0
for some a
ij
K not all 0. We put
s
j
=
f1
X
i=0
a
ij
α
i
.
We know that 1
, α, ··· , α
f1
are linearly independent over
K
since their re-
ductions are linearly independent over
k
K
. So there must be some
j
such that
s
j
6= 0.
The next claim is that if
s
j
6
= 0, then
e | v
L
(
s
j
). We let
k
be an index for
which |a
kj
| is maximal. Then we have
a
1
kj
s
j
=
f1
X
i=0
a
1
kj
a
ij
α
i
.
Now note that by assumption, the coefficients on the right have absolute value
1, and is 1 when i = k. So we know that
a
1
kj
s
j
6≡ 0 mod π
L
,
because 1, ¯α, ··· , ¯α
f1
are linearly independent. So we have
v
L
(a
1
kj
s
j
) = 0.
So we must have
v
L
(s
j
) = v
L
(a
kj
) + v
L
(a
1
kj
s
j
) v
L
(K
×
) = ev
L
(L
×
) = eZ.
Now we write
X
a
ij
α
i
π
j
=
e1
X
j=0
s
j
π
j
= 0.
If
s
j
6
= 0, then we have
v
L
(
s
j
π
j
) =
v
L
(
s
j
) +
j j
+
eZ
. So no two non-zero
terms in
P
e1
j=0
s
j
π
j
have the same valuation. This implies that
P
e1
j=0
s
j
π
j
6
= 0,
which is a contradiction.
We now want to prove that
O
L
=
M
i,j
O
K
α
i
π
j
.
We let
M =
M
i,j
O
K
α
i
π
j
,
and put
N =
f1
M
i=0
O
L
α
i
.
Then we have
M = N + πN + π
2
N + ··· + π
e1
N.
We are now going to use the fact that 1
, ¯α, ··· , ¯α
f1
span
k
L
over
k
K
. So we
must have that O
L
= N + πO
L
. We iterate this to obtain
O
L
= N + π(N + O
L
)
= N + πN + π
2
O
L
= ···
= N + πN + π
2
N + ··· + π
e1
N + π
n
O
L
= M + π
K
O
L
,
using the fact that
π
K
and
π
e
have the same valuation, and thus they differ by
a unit in O
L
. Iterating this again, we have
O
L
= M + π
n
k
O
L
for all
n
1. So
M
is dense in
O
L
. But
M
is the closed unit ball in the subspace
M
i,j
Kα
i
π
j
l
with respect to the maximum norm with respect to the given basis. So it must
be complete, and thus M = O
L
.
Finally, since
α
i
π
j
=
α
i
f
(
α
)
j
is a polynomial in
α
, we know that
O
L
=
O
K
[α].
Corollary. If M/L/K is a tower of finite extensions of local fields, then
f
M/K
= f
L/K
f
M/L
e
M/K
= e
L/K
e
M/L
Proof.
The multiplicativity of
f
M/K
follows from the tower law for the residue
fields, and the multiplicativity of
e
M/K
follows from the tower law for the local
fields and that f
M/K
e
M/K
= [M : K].
5.2 Unramified extensions
Unramified extensions are easy to classify, since they just correspond to extensions
of the residue field.
Theorem.
Let
K
be a local field. For every finite extension
/k
K
, there is a
unique (up to isomorphism) finite unramified extension
L/K
with
k
L
=
over
k
K
. Moreover, L/K is Galois with
Gal(L/K)
=
Gal(/k
K
).
Proof.
We start with existence. Let
¯α
be a primitive element of
/k
K
with
minimal polynomial
¯
f k
K
[
x
]. Take a monic lift
f O
K
[
x
] of
¯
f
such that
deg f
=
deg
¯
f
. Note that since
¯
f
is irreducible, we know
f
is irreducible. So we
can take L = K(α), where α is a root of f (i.e. L = K[x]/f). Then we have
[L : K] = deg f = deg(
¯
f) = [ : k
K
].
Moreover,
k
L
contains a root of
¯
f
, namely the reduction
α
. So there is an
embedding k
L
, sending ¯α to the reduction of α. So we have
[k
L
: k
K
] [ : k
L
] = [L : K].
So L/K must be unramified and k
L
=
over k
K
.
Uniqueness and the Galois property follow from the following lemma:
Lemma.
Let
L/K
be a finite unramified extension of local fields and let
M/K
be a finite extension. Then there is a natural bijection
Hom
K - Alg
(L, M) Hom
k
K
- Alg
(k
L
, k
M
)
given in one direction by restriction followed by reduction.
Proof.
By the uniqueness of extended absolute values, any
K
-algebra homomor-
phism
ϕ
:
L M
is an isometry for the extended absolute values. In particular,
we have
ϕ
(
O
L
)
O
M
and
ϕ
(
m
L
)
m
M
. So we get an induced
k
K
-algebra
homomorphism ¯ϕ : k
L
k
M
.
So we obtain a map
Hom
K-Alg
(L, M) Hom
k
K
-Alg
(k
L
, k
M
)
To see this is bijective, we take a primitive element
¯α k
L
over
k
K
, and take a
minimal polynomial
¯
f k
K
[
x
]. We take a monic lift of
¯
f
to
O
k
[
x
], and
α O
L
the unique root of
f
which lifts
¯α
, which exists by Hensel’s lemma. Then by
counting dimensions, the fact that the extension is unramified tells us that
k
L
= k
K
(¯α), L = K(α).
So we can construct the following diagram:
ϕ Hom
K-Alg
(L, M) Hom
k
K
-Alg
(k
L
, k
M
) ¯ϕ
ϕ(α) {x M : f (x) = 0} {¯x k
M
:
¯
f(¯x) = 0} ¯ϕ(¯α)
=
reduction
=
reduction
But the bottom map is a bijection by Hensel’s lemma. So done.
Alternatively, given a map
¯ϕ
:
k
L
k
M
, we can lift it to the map
ϕ
:
L M
given by
ϕ
X
[a
n
]π
n
k
=
X
[ ¯ϕ(a
n
)]π
n
k
,
using the fact that
π
n
k
is a uniformizer in
L
since the extension is unramified.
So we get an explicit inverse.
Proof of theorem (continued).
To finish off the proof of the theorem, we just
note that an isomorphism
¯ϕ
:
k
L
=
k
M
over
k
K
between unramified extensions.
Then
¯ϕ
lifts to a
K
-embedding
ϕ
:
L M
and [
L
:
K
] = [
M
:
K
] implies that
ϕ is an isomorphism.
To see that the extension is Galois, we just notice that
|Aut
K
(L)| = |Aut
k
K
(k
L
)| = [k
L
: k
K
] = [L : K].
So
L/K
is Galois. Moreover, the map
Aut
K
(
L
)
Aut
k
K
(
k
L
) is really a
homomorphism, hence an isomorphism.
Proposition.
Let
K
be a local field, and
L/K
a finite unramified extension,
and
M/K
finite. Say
L, M
are subfields of some fixed algebraic closure
¯
K
of
K
.
Then
LM/M
is unramified. Moreover, any subextension of
L/K
is unramified
over K. If M/K is unramified as well, then LM/K is unramified.
Proof.
Let
¯α
be a primitive element of
k
K
/k
L
, and
¯
f k
K
[
x
] a minimal polyno-
mial of
¯α
, and
f O
k
[
x
] a monic lift of
¯
f
, and
α O
L
a unique lift of
f
lifting
¯α. Then L = K(α). So LM = M(α).
Let
¯g
be the minimal polynomial of
¯α
over
k
M
. Then
¯g |
¯
f
. By Hensel’s
lemma, we can factorize
f
=
gh
in
O
M
[
x
], where
g
is monic and lifts
¯g
. Then
g
(
α
) = 0 and
g
is irreducible in
M
[
x
]. So
g
is the minimal polynomial of
α
over
M. So we know that
[LM : M] = deg g = deg ¯g [k
LM
: k
M
] [LM : M ].
So we have equality throughout and LM/M is unramified.
The second assertion follows from the multiplicativity of
f
L/K
, as does the
third.
Corollary.
Let
K
be a local field, and
L/K
finite. Then there is a unique
maximal subfield
K T L
such that
T/K
is unramified. Moreover, [
T
:
K
] =
f
L/K
.
Proof.
Let
T/K
be the unique unramified extension with residue field extension
k
L
/k
K
. Then
id
:
k
T
=
k
L
k
L
lifts to a
K
-embedding
T L
. Identifying
T
with its image, we know
[T : K] = f
L/K
.
Now if
T
0
is any other unramified extension, then
T
0
T
is an unramified extension
over K, so
[T : K] [T T
0
: K] f
L/K
= [T : K].
So we have equality throughout, and T
0
T . So this is maximal.
5.3 Totally ramified extensions
We now quickly look at totally ramified extensions. Recall the following irre-
ducibility criterion:
Theorem
(Eisenstein criterion)
.
Let
K
be a local field, and
f
(
x
) =
x
n
+
a
n1
x
n1
+
···
+
a
0
O
K
[
x
]. Let
π
K
be the uniformizer of
K
. If
π
K
|
a
n1
, ··· , a
0
and π
2
K
- a
0
, then f is irreducible.
Proof.
Left as an exercise. You’ve probably seen this already in a much more
general context, but in this case there is a neat proof using Newton polygons.
We will need to use the following characterization of the ramification index:
Proposition.
Let
L/K
be an extension of local fields, and
v
K
be the normalized
valuation. Let
w
be the unique extension of
v
K
to
L
. Then the ramification
index e
L/K
is given by
e
1
L/K
= w(π
L
) = min{w(x) : x m
L
},
Proof.
We know
w
and
v
L
differ by a constant. To figure out what this is, we
have
1 = w(π
K
) = e
1
L/K
v
L
(π
K
).
So for any x L, we have
w(x) = e
1
L/K
v
L
(x).
In particular, putting x = π
L
, we have
w(π
L
) = e
1
L/K
v
L
(π
L
) = e
1
L/K
.
The equality
w(π
L
) = min{w(x) : x m
L
},
is trivially true because the minimum is attained by π
L
.
Definition
(Eisenstein polynomial)
.
A polynomial
f
(
x
)
O
K
[
x
] satisfying the
assumptions of Eisenstein’s criterion is called an Eisenstein polynomial.
We can now state the proposition:
Proposition.
Let
L/K
be a totally ramified extension of local fields. Then
L = K(π
L
) and the minimal polynomial of π
L
over K is Eisenstein.
Conversely, if
L
=
K
(
α
) and the minimal polynomial of
α
over
K
is Eisenstein,
then L/K is totally ramified and α is a uniformizer of L.
Proof.
Let
n
= [
L
:
K
],
v
K
be the valuation of
K
, and
w
the unique extension
to L. Then
[K(π
L
) : K]
1
e
1
K(π
L
)/K
= min
xm
K(π
L
)
w(c)
1
n
,
where the last inequality follows from the fact that π
L
m
L(π
L
)
.
But we also know that
[K(π
L
) : K] [L : K].
So we know that L = K(π
L
).
Now let
f
(
x
) =
x
n
+
a
n1
x
n1
+
···
+
a
0
O
K
[
x
] be the minimal polynomial
of π
L
/K. Then we have
π
n
L
= (a
0
+ a
1
π
L
+ ··· + a
n1
π
n1
L
).
So we have
1 = w(π
n
L
) = w(a
0
+ a
1
π
L
+ ··· + a
n1
π
n1
L
) = min
i=0,...,n1
v
k
(a
i
) +
i
n
.
This implies that v
K
(a
i
) 1 for all i, and v
K
(x
0
) = 1. So it is Eisenstein.
For the converse, if K = K(α) and n = [L : K], take
g(x) = x
n
+ b
n1
x
n1
+ .. + b
0
O
K
[x]
be the minimal polynomial of
α
. So all roots have the same valuation. So we
have
1 = w(b
0
) = n · w(α).
So we have w(α) =
1
n
. So we have
e
1
L/K
= min
xm
L
w(x)
1
n
= [L : K]
1
.
So [L : K] = e
L/K
= n. So L/K is totally ramified and α is a uniformizer.
In fact, more is true. We have
O
L
=
O
K
[
π
L
], since every element in
O
L
can
be written as
X
i0
a
i
π
i
L
,
where a
i
is a lift of an element in k
L
= k
K
, which can be chosen to be in O
K
.
6 Further ramification theory
6.1 Some filtrations
If we have a field
K
, then we have a unit group
U
K
=
O
×
K
. We would like to
come up with a filtration of subgroups of the unit group, namely a sequence
··· U
(2)
K
U
(1)
K
U
(0)
K
= U
K
of subgroups that tells us how close a unit is to being 1. The further down we
are in the chain, the closer we are to being 1.
Similarly, given a field extension
L/K
, we want a filtration on the Galois
group (the indexing is conventional)
··· G
2
(L/K) G
1
(L/K) G
0
(L/K) G
1
(L/K) = Gal(L/K).
This time, the filtration tells us how close the automorphisms are to being the
identity map.
The key thing about these filtrations is that we can figure out information
about the quotients
U
(s)
K
/U
(s+1)
K
and
G
s
(
L/K
)
/G
s+1
(
L/K
), which is often easier.
Later, we might be able to patch these up to get more useful information about
U
K
and Gal(L/K).
We start with the filtration of the unit group.
Definition (Higher unit groups). We define the higher unit groups to be
U
(s)
K
= U
(s)
= 1 + π
s
K
O
K
.
We also put
U
K
= U
(0)
K
= U
(0)
= O
×
K
.
The quotients of these units groups are surprisingly simple:
Proposition. We have
U
K
/U
(1)
K
=
(k
×
K
, ·),
U
(s)
K
/U
(s+1)
K
=
(k
K
, +).
for s 1.
Proof.
We have a surjective homomorphism
O
×
K
k
×
K
which is just reduction
mod
π
K
, and the kernel is just things that are 1 modulo
π
K
, i.e.
U
(1)
K
. So this
gives the first part.
For the second part, we define a surjection U
(s)
K
k
K
given by
1 + π
s
K
x 7→ x mod π
k
.
This is a group homomorphism because
(1 + π
s
K
x)(1 + π
s
K
y) = 1 + π
S
(x + y + π
s
xy),
and this gets mapped to
x + y + π
s
x + y
=
x + y mod π
K
.
Then almost by definition, the kernel is U
(s+1)
K
.
The next thing to consider is a filtration of the Galois group.
Definition (Higher ramification group). Let L/K be a finite Galois extension
of local fields, and v
L
the normalized valuation of L.
Let s R
≥−1
. We define the sth ramification group by
G
s
(L/K) = {σ Gal(L/K) : v
L
(σ(x) x) s + 1 for all x O
L
}.
So if you belong to
G
s
for a large
s
, then you move things less. Note that we
could have defined these only for
s Z
≥−1
, but allowing fractional indices will
be helpful in the future.
Now since σ(x) x O
L
for all x O
L
, we know
G
1
(L/K) = Gal(L/K).
We next consider the case of G
0
(L/K). This is, by definition
G
0
(L/K) = {σ Gal(L/K) : v
L
(σ(x) x) 1 for all x O
L
}
= {σ Gal(L/K) : σ(x) x mod m for all x O
L
}.
In other words, these are all the automorphisms that reduce to the identity when
we reduce it to Gal(k
L
/k
K
).
Definition
(Inertia group)
.
Let
L/K
be a finite Galois extension of local fields.
Then the inertia group of L/K is the kernel of the natural homomorphism
Gal(L/K) Gal(k
L
/k
K
)
given by reduction. We write this as
I(L/K) = G
0
(L/K).
Proposition.
Let
L/K
be a finite Galois extension of local fields. Then the
homomorphism
Gal(L/K) Gal(k
L
/k
K
)
given by reduction is surjective.
Proof.
Let
T/K
be maximal unramified subextension. Then by Galois theory,
the map
Gal
(
L/K
)
Gal
(
T/K
) is a surjection. Moreover, we know that
k
T
= k
L
. So we have a commutative diagram
Gal(L/K) Gal(k
L
/k
K
)
Gal(T/K) Gal(k
T
/k
K
).
So the map Gal(L/K) Gal(k
L
/k
K
) is surjective.
Then the inertia group is trivial iff
L/K
is unramified. The field
T
is
sometimes called the inertia field.
Lemma.
Let
L/K
be a finite Galois extension of local fields, and let
σ I
(
L/K
).
Then σ([x]) = [x] for all x.
More generally, let
x k
L
and
σ Gal
(
L/K
) with image
¯σ Gal
(
k
L
/k
K
).
Then we have
[¯σ(x)] = σ([x]).
Proof. Consider the map k
L
O
L
given by
f : x 7→ σ
1
([¯σ(x)]).
This is multiplicative, because every term is multiplicative, and
σ
1
([¯σ(x)]) x mod π
L
.
So this map f has to be the Teichm¨uller lift by uniqueness.
That’s all we’re going to say about the inertia group. We now consider the
general properties of this filtration.
Proposition.
Let
L/K
be a finite Galois extension of local fields, and
v
L
the
normalized valuation of
L
. Let
π
L
be the uniformizer of
L
. Then
G
s+1
(
L/K
) is
a normal subgroup of G
s
(L/K) for s Z
0
, and the map
G
s
(L/K)
G
s+1
(L/K)
U
(s)
L
U
(s+1)
L
given by
σ 7→
σ(π
L
)
π
L
is a well-defined injective group homomorphism, independent of the choice of
π
L
.
Proof. We define the map
φ : G
s
(L/K)
U
(s)
L
U
(s+1)
L
σ 7→ σ(π
L
)
L
.
We want to show that this has kernel G
s+1
(L/K).
First we show it is well-defined. If σ G
s
(L/K), we know
σ(π
L
) = π
L
+ π
s+1
L
x
for some x O
L
. So we know
σ(π
L
)
π
L
= 1 + π
s
L
x U
(s)
L
.
So it has the right image. To see this is independent of the choice of
π
L
, we let
u O
×
L
. Then σ(u) = u + π
s+1
L
y for some y O
L
.
Since any other uniformizer must be of the form π
L
u, we can compute
σ(π
L
u)
π
L
u
=
(π
L
+ π
s+1
L
)(u + π
s+1
L
y)
π
L
u
= (1 + π
s
L
x)(1 + π
s+1
L
u
1
y)
1π
s
L
x (mod U
s+1
L
).
So they represent the same element in in U
(s)
L
/U
(s+1)
L
.
To see this is a group homomorphism, we know
φ(στ) =
σ(τ(π
L
))
π
L
=
σ(τ(π
L
))
τ(π
L
)
τ(π
L
)
π
L
= φ(σ)φ(t),
using the fact that τ(π
L
) is also a uniformizer.
Finally, we have to show that ker φ = G
s+1
(L/K). We write down
ker φ = {σ G
s
(L/K) : v
L
(σ(π
L
) π
L
) s + 2}.
On the other hand, we have
G
s+1
(L/K) = {σ G
s
(L/K) : v
L
(σ(z) z) s + 2 for all z O
L
}.
So we trivially have
G
s+1
(
L/K
)
ker φ
. To show the converse, let
x O
L
and
write
x =
X
n=0
[x
n
]π
n
L
.
Take σ ker φ G
s
(L/K) I(L/K). Then we have
σ(π
L
) = π
L
+ π
s+2
L
y, y O
L
.
Then by the previous lemma, we know
σ(x) x =
X
n=1
[x
n
] ((σ(π
L
))
n
π
n
L
)
=
X
n=1
[x
n
]
(π
L
+ π
s+2
L
y)
n
π
n
L
= π
s+2
L
(things).
So we know v
L
(σ(x) x) s + 2.
Corollary. Gal(L/K) is solvable.
Proof. Note that
\
s
G
s
(L/K) = {id}.
So (
G
s
(
L/K
))
sZ
≥−1
is a subnormal series of
Gal
(
L/K
), and all quotients are
abelian, because they embed into
U
(s)
L
U
(s+1)
L
=
(
k
K
,
+) (and
s
=
1 can be checked
separately).
Thus if
L/K
is a finite extension of local fields, then we have, for
s
1,
injections
G
s
(L/K)
G
s+1
(L/K)
k
L
.
Since k
L
is a p-group, it follows that
|G
s
(L/K)|
|G
s+1
(L/K)|
is a pth power. So it follows that for any t, the quotient
|G
1
(L/K)|
|G
t
(L/K)|
is also a
p
th power. However, we know that the intersection of all
G
s
(
L/K
)
is
{id}
, and also
Gal
(
L/K
) is finite. So for sufficiently large
t
, we know that
|G
t
(L/K)| = 1. So we conclude that
Proposition. G
1
(L/K) is always a p-group.
We now use the injection
G
0
(L/K)
G
1
(L/K)
k
×
L
,
and the fact that
k
×
L
has order prime to
p
. So
G
1
(
L/K
) must be the Sylow
p-subgroup of G
0
(L/K). Since it is normal, it must be the unique p-subgroup.
Definition
(Wild inertia group and tame quotient)
. G
1
(
L/K
) is called the wild
inertia group, and the quotient G
0
(L/K)/G
1
(L/K) is the tame quotient.
6.2 Multiple extensions
Suppose we have tower
M/L/K
of finite extensions of local fields. How do the
ramification groups of the different extensions relate? We first do the easy case.
Proposition.
Let
M/L/K
be finite extensions of local fields, and
M/K
Galois.
Then
G
s
(M/K) Gal(M/L) = G
s
(M/L).
Proof. We have
G
s
(M/K) = {σ Gal(M/L) : v
M
(σx x) s + 1} = G
s
(M/K) Gal(M/L).
This is trivial, because the definition uses the valuation
v
M
of the bigger field
all the time. What’s more difficult and interesting is quotients, namely going
from M/K to L/K.
We want to prove the following theorem:
Theorem
(Herbrand’s theorem)
.
Let
M/L/K
be finite extensions of local fields
with M/K and L/K Galois. Then there is some function η
M/L
such that
G
t
(L/K)
=
G
s
(M/K)
G
s
(M/L)
for all s, where t = η
M/L
(s).
To better understand the situation, it helps to provide an alternative charac-
terization of the Galois group.
Definition (i
L/K
). We define
i
L/K
(σ) = min
x∈O
L
v
L
(σ(x) x).
It is then immediate that
G
s
(L/K) = {σ Gal(L/K) : i
L/K
(σ) s + 1}.
This is not very helpful. We now claim that we can compute
i
L/K
using the
following formula:
Proposition.
Let
L/K
be a finite Galois extension of local fields, and pick
α O
L
such that O
L
= O
K
[α]. Then
i
L/K
(σ) = v
L
(σ(α) α).
Proof.
Fix a
σ
. It is clear that
i
L/K
(
σ
)
v
L
(
σ
(
α
)
α
). Conversely, for any
x O
L
, we can find a polynomial g O
K
[t] such that
x = g(α) =
X
b
i
α
i
,
where b
i
O
K
. In particular, b
i
is fixed by σ.
Then we have
v
L
(σ(x) x) = v
L
(σg(α) g(α))
= v
L
n
X
i=1
b
i
(σ(α)
i
α
i
)
!
v
L
(σ(α) α),
using the fact that σ(α) α | σ(α)
i
α
i
for all i. So done.
Now if
M/L/K
are finite Galois extensions of local fields, then
O
M
=
O
K
[
α
]
implies O
M
= O
L
[α]. So for σ Gal(M/L), we have
i
M/L
(σ) = i
M/K
(σ).
Going in the other direction is more complicated.
Proposition.
Let
M/L/K
be a finite extension of local fields, such that
M/K
and L/K are Galois. Then for σ Gal(L/K), we have
i
L/K
(σ) = e
1
M/L
X
τGal(M/K)
τ|
L
=σ
i
M/K
(τ).
Here
e
M/L
is just to account for the difference between
v
L
and
v
M
. So the
real content is that the value of
i
L/K
(
σ
) is the sum of the values of
i
M/K
(
τ
) for
all τ that restrict to σ.
Proof.
If
σ
= 1, then both sides are infinite by convention, and equality holds.
So we assume
σ 6
= 1. Let
O
M
=
O
L
[
α
] and
O
L
=
O
K
[
β
], where
α O
M
and
β O
L
. Then we have
e
M/L
i
L/K
(σ) = e
M/L
v
L
(σβ β) = v
M
(σβ β).
Now if τ Gal(M/K), then
i
M/K
(τ) = v
M
(τα α)
Now fix a τ such that τ|
L
= σ. We set H = Gal(M/L). Then we have
X
τ
0
Gal(M/K)
0
|
L
=σ
i
M/K
(τ
0
) =
X
gH
v
M
(τg(α) α) = v
M
Y
gH
(τg(α) α)
.
We let
b = σ(β) β = τ(β) β
and
a =
Y
gH
(τg(α) α).
We want to prove that v
M
(b) = v
M
(a). We will prove that a | b and b | a.
We start with a general observation about elements in
O
L
. Given
z O
L
,
we can write
z =
n
X
i=1
z
i
β
i
, z
i
O
K
.
Then we know
τ(z) z =
n
X
i=1
z
i
(τ(β)
i
β
i
)
is divisible by τ(β) β = b.
Now let
F
(
x
)
O
L
[
x
] be the minimal polynomial of
α
over
L
. Then explicitly,
we have
F (x) =
Y
gH
(x g(α)).
Then we have
(τF )(x) =
Y
gH
(x τg(α)),
where
τF
is obtained from
F
by applying
τ
to all coefficients of
F
. Then all
coefficients of
τF F
are of the form
τ
(
z
)
z
for some
z O
L
. So it is divisible
by b. So b divides every value of this polynomial, and in particular
b | (τF F )(α) =
Y
gH
(α g(α)) = ±a,
So b | a.
In other direction, we pick
f O
K
[
x
] such that
f
(
α
) =
β
. Then
f
(
α
)
β
= 0.
This implies that the polynomial
f
(
x
)
β
divides the minimal polynomial of
α
in O
L
[x]. So we have
f(x) β = F (x)h(x)
for some h O
L
[x].
Then noting that f has coefficients in O
K
, we have
(f τβ)(x) = (τf τb)(x) = (τF )(x)(τh)(x).
Finally, set x = α. Then
b = β τβ = ±a(τh)(α).
So a | b.
Now that we understand how the
i
L/K
behave when we take field extensions,
we should be able to understand how the ramification groups behave!
We now write down the right choice of η
L/K
: [1, ) [1, ):
η
L/K
(s) =
e
1
L/K
X
σG
min(i
L/K
(σ), s + 1)
!
1.
Theorem
(Herbrand’s theorem)
.
Let
M/L/K
be a finite extension of local
fields with M/K and L/K Galois. We set
H = Gal(M/L), t = η
M/L
(s).
Then we have
G
s
(M/K)H
H
= G
t
(L/K).
By some isomorphism theorem, and the fact that
H G
s
(
M/K
) =
G
s
(
M/L
),
this is equivalent to saying
G
t
(L/K)
=
G
s
(M/K)
G
s
(M/L)
.
Proof.
Let
G
=
Gal
(
M/K
). Fix a
σ Gal
(
L/K
). We let
τ Gal
(
M/K
) be an
extension of σ to M that maximizes i
M/K
, i.e.
i
M/K
(τ) i
M/K
(τg)
for all g H. This is possible since H is finite.
We claim that
i
L/K
(σ) 1 = η
M/L
(i
M/K
(τ) 1).
If this were true, then we would have
σ
G
s
(M/K)H
H
τ G
s
(M/K)
i
M/K
(τ) 1 s
Since η
M/L
is strictly increasing, we have
η
M/L
(i
M/K
(τ) 1) η
M/L
(s) = t
i
L/K
(σ) 1 t
σ G
t
(L/K),
and we are done.
To prove the claim, we now use our known expressions for
i
L/K
(
σ
) and
η
M/L
(i
M/K
(τ) 1) to rewrite it as
e
1
M/L
X
gH
i
M/K
(τg) = e
1
M/L
X
gH
min(i
M/L
(g), i
M/K
(τ)).
We then make the stronger claim
i
M/K
(τg) = min(i
M/L
(g), i
M/K
(τ)).
We first note that
i
M/K
(τg) = v
M
(τg(α) α)
= v
M
(τg(α) g(α) + g(α) α)
min(v
M
(τg(α) g(α)), v
M
(g(α) α))
= min(i
M/K
(τ), i
M/K
(g))
We cannot conclude our (stronger) claim yet, since we have a
in the middle.
We now have to split into two cases.
(i)
If
i
M/K
(
g
)
i
M/K
(
τ
), then the above shows that
i
M/K
(
τg
)
i
M/K
(
τ
).
But we also know that it is bounded above by
m
. So
i
M/K
(
τg
) =
i
M/K
(
τ
).
So our claim holds.
(ii)
If
i
M/K
(
g
)
< i
M/K
(
τ
), then the above inequality is in fact an equality as
the two terms have different valuations. So our claim also holds.
So done.
We now prove an alternative characterization of the function
η
L/K
, using a
funny integral.
Proposition. Write G = Gal(L/K). Then
η
L/K
(s) =
Z
s
0
dx
(G
0
(L/K) : G
x
(L/K))
.
When 1 x < 0, our convention is that
1
(G
0
(L/K) : G
x
(L/K))
= (G
x
(L/K) : G
0
(L/K)),
which is just equal to 1 when 1 < x < 0. So
η
L/K
(s) = s if 1 s 0.
Proof.
We denote the RHS by
θ
(
s
). It is clear that both
η
L/K
(
s
) and
θ
(
s
) are
piecewise linear and the break points are integers (since
i
L/K
(
σ
) is always an
integer). So to see they are the same, we see that they agree at a point, and
that they have equal derivatives. We have
η
L/K
(0) =
|{σ G : i
L/K
(σ) 1}|
e
L/K
1 = 0 = θ(0),
since the numerator is the size of the inertia group.
If s [1, ) \ Z, then
η
0
L/K
(s) = e
1
L/K
(|{σ G : i
L/K
(σ) s + 1}|)
=
|G
s
(L/K)|
|G
0
(L/K)|
=
1
(G
0
(L/K) : G
s
(L/K))
= θ
0
(s).
So done.
We now tidy up the proof by inventing a different numbering of the ramifica-
tion groups. Recall that
η
L/K
: [1, ) [1, )
is continuous, strictly increasing, and
η
L/K
(1) = 1, η
L/K
(s) as s .
So this is invertible. We set
Notation.
ψ
L/K
= η
1
L/K
.
Definition
(Upper numbering)
.
Let
L/K
be a Galois extension of local fields.
Then the upper numbering of the ramification groups of L/K is defined by
G
t
(L/K) = G
ψ
L/K
(t)
(L/K)
for t [1, ). The original number is called the lower numbering.
To rephrase our previous theorem using the upper numbering, we need a
little lemma:
Lemma.
Let
M/L/K
be a finite extension of local fields, and
M/K
and
L/K
be Galois. Then
η
M/K
= η
L/K
η
M/L
.
Hence
ψ
M/K
= ψ
M/L
ψ
L/K
.
Proof.
Let
s
[
1
,
), and let
t
=
η
M/L
(
s
), and
H
=
Gal
(
M/L
). By Her-
brand’s theorem, we know
G
t
(L/K)
=
G
s
(M/K)H
H
=
G
s
(M/K)
H G
s
(M/K)
=
G
s
(M/K)
G
s
(M/L)
.
Thus by multiplicativity of the inertia degree, we have
|G
s
(M/K)|
e
M/K
=
|G
t
(L/K)|
e
L/K
|G
s
(M/L)|
e
M/L
.
By the fundamental theorem of calculus, we know that whenever the derivatives
make sense, we have
η
0
M/K
(s) =
|G
s
(M/K)|
e
M/K
.
So putting this in, we know
η
0
M/K
(s) = η
0
L/K
(t)η
0
M/L
(s) = (η
L/K
η
M/L
)
0
(s).
Since
η
M/K
and
η
L/K
η
M/L
agree at 0 (they both take value 0), we know that
the functions must agree everywhere. So done.
Corollary.
Let
M/L/K
be finite Galois extensions of local fields, and
H
=
Gal(M/L). Let t [1, ). Then
G
t
(M/K)H
H
= G
t
(L/K).
Proof. Put s = η
L/K
(t). Then by Herbrand’s theorem, we have
G
t
(M/K)H
H
=
G
ψ
M/K
(t)
(M/K)H
H
=
G
η
M/L
(ψ
M/K
(t))
(L/K)
= G
s
(L/K)
= G
t
(L/K).
This upper numbering might seem like an unwieldy beast that was invented
just so that our theorem looks nice. However, it turns out that often the upper
numberings are rather natural, as we could see in the example below:
Example.
Consider
ζ
p
n
a primitive
p
n
th root of unity, and
K
=
Q
p
(
ζ
p
n
). The
minimal polynomial of ζ
p
n
is the p
n
th cyclotomic polynomial
Φ
p
n
(x) = x
p
n1
(p1)
+ x
p
n1
(p2)
+ ··· + 1.
It is an exercise on the example sheet to show that this is indeed irreducible.
So
K/Q
p
is a Galois extension of degree
p
n1
(
p
1). Moreover, it is totally
ramified by question 6 on example sheet 2, with uniformizer
π = ζ
p
n
1
is a uniformizer. So we know
O
K
= Z
p
[ζ
p
n
1] = Z
p
[ζ
p
n
].
We then have an isomorphism
Z
p
n
Z
×
Gal(L/Q
p
)
obtained by sending m σ
m
, where
σ
m
(ζ
p
n
) = ζ
m
p
n
.
We have
i
K/Q
p
(σ
m
) = v
K
(σ
m
(ζ
p
n
) ζ
p
n
)
= v
K
(ζ
m
p
n
ζ
p
n
)
= v
K
(ζ
m1
p
n
1)
since
ζ
p
n
is a unit. If
m
= 1, then this thing is infinity. If it is not 1, then
ζ
m1
p
n
is a primitive
p
nk
th root of unity for the maximal
k
such that
p
k
| m
1. So
by Q6 on example sheet 2, we have
v
K
(ζ
m1
p
n
1) =
p
n1
(p 1)
p
nk1
(p 1)
= p
k
.
Thus we have
v
K
(ζ
m1
p
n
1) p
k
m 1 mod p
k
.
It then follows that for
p
k
s + 1 p
k1
+ 1,
we have
G
s
(K/Q
p
)
=
{m (Z/p
n
)
×
: m 1 mod p
k
}.
Now m 1 mod p
k
iff σ
m
(ζ
p
k
) = ζ
p
k
. So in fact
G
s
(K/Q
p
)
=
Gal(K/Q
p
(ζ
p
k
)).
Finally, when s p
n
1, we have
G
s
(K/Q
p
) = 1.
We claim that
η
K/Q
p
(p
k
1) = k.
So we have
G
k
(K/Q
p
) = Gal(K/Q
p
(ζ
p
k
)).
This actually looks much nicer!
To actually compute
η
K/Q
p
, we have notice that the function we integrate to
get η looks something like this (not to scale):
1
p1
1
p(p1)
1
p
2
(p1)
p 1
p
2
1 p
3
1
The jumps in the lower numbering are at
p
k
1 for
k
= 1
, ··· , n
1. So we have
η
K/Q
p
(p
k
1) = (p 1)
1
p 1
+ ((p
2
1) (p 1))
1
p(p 1)
+ ··· + ((p
k
1) (p
k1
1))
1
p
k1
(p 1)
= k.
7 Local class field theory
Local class field theory is the study of abelian extensions of local fields, i.e. a
Galois extension whose Galois group is abelian.
7.1 Infinite Galois theory
It turns out that the best way of formulating this theory is to not only use
finite extensions, but infinite extensions as well. So we need to begin with some
infinite Galois theory. We will mostly just state the relevant results instead of
proving them, because this is not a course on Galois theory.
In this section, we will work with any field.
Definition
(Separable and normal extensions)
.
Let
L/K
be an algebraic exten-
sion of fields. We say that
L/K
is separable if, for every
α L
, the minimal
polynomial
f
α
K
[
α
] is separable. We say
L/K
is normal if
f
α
splits in
L
for
every α L.
Definition
(Galois extension)
.
Let
L/K
be an algebraic extension of fields.
Then it is Galois if it is normal and separable. If so, we write
Gal(L/K) = Aut
K
(L).
These are all the same definitions as in the finite case.
In finite Galois theory, the subgroups of
Gal
(
L/K
) match up with the
intermediate extensions, but this is no longer true in the infinite case. The
Galois group has too many subgroups. To fix this, we need to give
Gal
(
L/K
) a
topology, and talk about closed subgroups.
Definition
(Krull topology)
.
Let
M/K
be a Galois extension. We define the
Krull topology on M/K by the basis
{Gal(M/L) : L/K is finite}.
More explicitly, we say that
U Gal
(
M/K
) is open if for every
σ U
, we can
find a finite subextension L/K of M/K such that σ Gal(M/L) U .
Note that any open subgroup of a topological group is automatically closed,
but the converse does not hold.
Note that when
M/K
is finite, then the Krull topology is discrete, since we
can just take the finite subextension to be M itself.
Proposition.
Let
M/K
be a Galois extension. Then
Gal
(
M/K
) is compact
and Hausdorff, and if
U Gal
(
M/K
) is an open subset such that 1
U
, then
there is an open normal subgroup N Gal(M/K) such that N U.
Groups with properties in this proposition are known as profinite groups.
Proof. We will not prove the first part.
For the last part, note that by definition, there is a finite subextension of
M/K
such that
Gal
(
M/L
)
U
. We then let
L
0
be the Galois closure of
L
over
K
. Then
Gal
(
M/L
0
)
Gal
(
M/L
)
U
, and
Gal
(
M/L
0
) is open and normal.
Recall that we previously defined the inverse limit of a sequence rings. More
generally, we can define such an inverse limit for any sufficiently nice poset
of things. Here we are going to do it for topological groups (for those doing
Category Theory, this is the filtered limit of topological groups).
Definition
(Directed system)
.
Let
I
be a set with a partial order. We say that
I
is a directed system if for all
i, j I
, there is some
k I
such that
i k
and
j k.
Example. Any total order is a directed system.
Example. N with divisibility | as the partial order is a directed system.
Definition
(Inverse limit)
.
Let
I
be a directed system. An inverse system (of
topological groups) indexed by
I
is a collection of topological groups
G
i
for each
i I and continuous homomorphisms
f
ij
: G
j
G
i
for all i, j I such that i j, such that
f
ii
= id
G
i
and
f
ik
= f
ij
f
jk
whenever i j k.
We define the inverse limit on the system (G
i
, f
ij
) to be
lim
iI
G
i
=
(
(g
i
)
Y
iI
G
i
: f
ij
(g
j
) = g
i
for all i j
)
Y
iI
g
i
,
which is a group under coordinate-wise multiplication and a topological space
under the subspace topology of the product topology on
Q
iI
G
i
. This makes
lim
iI
G
i
into a topological group.
Proposition.
Let
M/K
be a Galois extension. The set
I
of finite Galois
subextensions
L/K
is a directed system under inclusion. If
L, L
0
I
and
L L
0
,
then we have a restriction map
·|
L
0
L
: Gal(L
0
/K) Gal(L/K).
Then (Gal(L/K), ·|
L
0
L
) is an inverse system, and the map
Gal(M/K) lim
iI
Gal(L/K)
σ 7→ (σ|
L
)
iI
is an isomorphism of topological groups.
We now state the main theorem of Galois theory.
Theorem
(Fundamental theorem of Galois theory)
.
Let
M/K
be a Galois exten-
sion. Then the map
L 7→ Gal
(
M/L
) defines a bijection between subextensions
L/K
of
M/K
and closed subgroups of
Gal
(
M/K
), with inverse given by sending
H 7→ M
H
, the fixed field of H.
Moreover,
L/K
is finite if and only if
Gal
(
M/L
) is open, and
L/K
is Galois
iff Gal(M/L) is normal, and then
Gal(L/K)
Gal(M/L)
Gal(L/K)
is an isomorphism of topological groups.
Proof.
This follows easily from the fundamental theorem for finite field extensions.
We will only show that
Gal
(
M/L
) is closed and leave the rest as an exercise. We
can write
L =
[
L
0
L
L
0
/K finite
L
0
.
Then we have
Gal(M/L) =
\
L
0
L
L
0
/K finite
Gal(M/L
0
),
and each Gal(M/L
0
) is open, hence closed. So the whole thing is closed.
7.2 Unramified extensions and Weil group
We first define what it means for an infinite extension to be unramified or totally
ramified. To do so, we unexcitingly patch up the definitions for finite cases.
Definition
(Unramified extension)
.
Let
K
be a local field, and
M/K
be alge-
braic. Then
M/K
is unramified if
L/K
is unramified for every finite subextension
L/K of M/K.
Note that since the extension is not necessarily finite, in general
M
will not
be a local field, since chances are its residue field would be infinite.
Definition
(Totally ramified extension)
.
Let
K
be a local field, and
M/K
be
algebraic. Then
M/K
is totally ramified if
L/K
is totally ramified for every
finite subextension L/K of M/K.
Proposition.
Let
M/K
be an unramified extension of local fields. Then
M/K
is Galois, and
Gal(M/K)
=
Gal(k
M
/k
K
)
via the reduction map.
Proof.
Every finite subextension of
M/K
is unramified, so in particular is Galois.
So
M/K
is Galois (because normality and separability is checked for each
element). Then we have a commutative diagram
Gal(M/K) Gal(k
M
/k
K
)
lim
L/K
Gal(L/K) lim
L/K
Gal(k
L
/k
K
)
reduction
reduction
The left hand map is an isomorphism by (infinite) Galois theory, and since all
finite subextensions of
k
M
/k
K
are of the form
k
L
/k
K
by our finite theory, we
know the right-hand map is an isomorphism. The bottom map is an isomorphism
since it is an isomorphism in each component. So the top map must be an
isomorphism.
Since the compositor of unramified extensions is unramified, it follows that
any algebraic extension M/K has a maximal unramified subextension
T = T
M/K
/K.
In particular, every field K has a maximal unramified extension K
ur
.
We now try to understand unramified extensions. For a finite unramified
extension L/K, we have an isomorphism
Gal(L/K) Gal(k
L
/k
K
)
,
By general field theory, we know that
Gal
(
k
L
/k
K
) is a cyclic group generated by
Frob
L/K
: x 7→ x
q
,
where
q
=
|k
K
|
is the size of
k
K
. So by the isomorphism, we obtain a generator
of Gal(L/K).
Definition
(Arithmetic Frobenius)
.
Let
L/K
be a finite unramified extension
of local fields, the (arithmetic) Frobenius of
L/K
is the lift of
Frob
L/K
Gal(k
L
/k
K
) under the isomorphism Gal(L/K)
=
Gal(k
L
/k
K
).
There is also a geometric Frobenius, which is its inverse, but we will not use
that in this course.
We know
Frob
is compatible in towers, i.e. if
M/L/K
is a tower of finite
unramified extension of local fields, then
Frob
M/K
|
L
=
Frob
L/K
, since they both
reduce to the map
x 7→ x
|k
K
|
in
Gal
(
k
L
/k
K
), and the map between
Gal
(
k
L
/k
K
)
and Gal(L/K) is a bijection.
So if M/K is an arbitrary unramified extension, then we have an element
(Frob
L/K
) lim
L/K
Gal(L/K)
=
Gal(M/K).
So we get an element
Frob
M/K
Gal
(
M/K
). By tracing through the proof of
Gal(M/K)
=
Gal(k
M
/k
K
), we see that this is the unique lift of x 7→ x
|k
K
|
.
Note that while for finite unramified extensions
M/K
, the Galois group is
generated by the Frobenius, this is not necessarily the case when the extension
is infinite. However, powers of the Frobenius are the only things we want to
think about, so we make the following definition:
Definition
(Weil group)
.
Let
K
be a local field and
M/K
be Galois. Let
T
=
T
M/K
be the maximal unramified subextension of
M/K
. The Weil group
of M/K is
W (M/K) = {σ Gal(M/K) : σ|
T
= Frob
n
T/K
for some n Z}.
We define a topology on
W
(
M/K
) by saying that
U
is open iff there is a finite
extension L/T such that σ Gal(L/T ) U.
In particular, if M/K is unramified, then W(M/K) = Frob
Z
T/K
.
It is helpful to put these groups into a diagram of topological groups to see
what is going on.
Gal(M/T ) W (M/K) Frob
Z
T/K
Gal(M/T ) Gal(M/K) Gal(T /K)
Here we put the discrete topology on the subgroup generated by the Frobenius.
The topology of
W
(
M/K
) is then chosen so that all these maps are continuous
homomorphisms of groups.
In many ways, the Weil group works rather like the Galois group.
Proposition.
Let
K
be a local field, and
M/K
Galois. Then
W
(
M/K
) is dense
in
Gal
(
M/K
). Equivalently, for any finite Galois subextension
L/K
of
M/K
,
the restriction map W(M/K) Gal(L/K) is surjective.
If L/K is a finite subextension of M/K, then
W (M/L) = W (M/K) Gal(M/L).
If L/K is also Galois, then
W (M/K)
W (M/L)
=
Gal(L/K)
via restriction.
Proof.
We first prove density. To see that density is equivalent to
W
(
M/K
)
Gal
(
L/K
) being surjective for all finite subextension
L/K
, note that by the
topology on
Gal
(
M/K
), we know density is equivalent to saying that
W
(
M/K
)
hits every coset of
Gal
(
M/L
), which means that
W
(
M/K
)
Gal
(
L/K
) is
surjective.
Let
L/K
be a subextension. We let
T
=
T
M/K
. Then
T
L/K
=
T L
. Then
we have a diagram
Gal(M/T ) W (M/K) Frob
Z
T/K
Gal(L/T L) Gal(L/K) Gal(T L/K)
Here the surjectivity of the left vertical arrow comes from field theory, and the
right hand vertical map is surjective because
T L/K
is finite and hence the
Galois group is generated by the Frobenius. Since the top and bottom rows are
short exact sequences (top by definition, bottom by Galois theory), by diagram
chasing (half of the five lemma), we get surjectivity in the middle.
To prove the second part, we again let
L/K
be a finite subextension. Then
L · T
M/K
T
M/L
. We then have maps
Frob
Z
T
M/K
/K
Gal(T
M/K
/K) Gal(k
M
/k
K
)
Frob
Z
T
M/L
/L
Gal(T
M/L
/L) Gal(k
M
/k
L
)
=
=
So the left hand vertical map is an inclusion. So we know
Frob
Z
T
M/L
/L
= Frob
Z
T
M/K
/K
Gal(T
M/L
/L).
Now if σ Gal(M/L), then we have
σ W (M/L) σ|
T
M/L
/L
Frob
Z
T
M/L
/L
σ|
T
M/K
/K
Frob
Z
T
M/K
/K
σ W (M/K).
So this gives the second part.
Now
L/K
is Galois as well. Then
Gal
(
M/L
) is normal in
Gal
(
M/K
). So
W (M/L) is normal in W(M/K) by the second part. Then we can compute
W (M/K)
W (M/L)
=
W (M/K)
W (M/K) Gal(M/L)
=
W (M/K) Gal(M/L)
Gal(M/L)
=
Gal(M/K)
Gal(M/L)
=
Gal(L/K).
The only non-trivial part in this chain is the assertion that
W
(
M/K
)
Gal
(
M/L
) =
Gal
(
M/K
), i.e. that
W
(
M/K
) hits every coset of
Gal
(
M/L
), which is what
density tells us.
7.3 Main theorems of local class field theory
We now come to the main theorems of local class field theory.
Definition
(Abelian extension)
.
Let
K
be a local field. A Galois extension
L/K is abelian if Gal(L/K) is abelian.
We will fix an algebraic closure
¯
K
of
K
, and all algebraic extensions we will
consider will be taken to be subextensions of
¯
K/K
. We let
K
sep
be the separable
closure of K inside
¯
K.
If
M/K
and
M/K
are Galois extensions, then
LM/K
is Galois, and the map
given by restriction
Gal(LM/K) Gal(L/K) × Gal(M/K).
is an injection. In particular, if
L/K
and
M/K
are both abelian, then so is
LM/K. This implies that there is a maximal abelian extension K
ab
.
Finally, note that we know an example of an abelian extension, namely the
maximal unramified extension
K
ur
=
T
K
sep
/K
K
ab
, and we put
Frob
K
=
Frob
K
ur
/K
.
Theorem
(Local Artin reciprocity)
.
There exists a unique topological isomor-
phism
Art
K
: K
×
W (K
ab
/K)
characterized by the properties
(i) Art
K
(π
K
)|
K
ur
= Frob
K
, where π
K
is any uniformizer.
(ii) We have
Art
K
(N
L/K
(x))|
L
= id
L
for all L/K finite abelian and x L
×
.
Moreover, if
M/K
is finite, then for all
x M
×
, we know
Art
M
(
x
) is an
automorphism of
M
ab
/M
, and restricts to an automorphisms of
K
ab
/K
. Then
we have
Art
M
(x)|
K
ab
K
= Art
K
(N
M/K
(x)).
Moreover, Art
K
induces an isomorphism
K
×
N
M/K
(M
×
)
Gal
M K
ab
K
.
To simplify this, we will write
N
(
L/K
) =
N
L/K
(
L
×
) for
L/K
finite. From
this theorem, we can deduce a lot of more precise statements.
Corollary. Let L/K be finite. Then N(L/K) = N((L K
ab
)/K), and
(K
×
: N(L/K)) [L : K]
with equality iff L/K is abelian.
Proof.
To see this, we let
M
=
L K
ab
. Applying the isomorphism twice gives
K
×
N(L/K)
=
Gal(M/K)
=
K
×
N(M/K)
.
Since
N
(
L/K
)
N
(
M/K
), and [
L
:
K
]
[
M
:
K
] =
|Gal
(
M/K
)
|
, we are
done.
The theorem tells us if we have a finite abelian extension
M/K
, then we
obtain an open finite-index subgroup
N
M/K
(
M
×
)
K
×
. Conversely, if we are
given an open finite index subgroup of
K
×
, we might ask if there is an abelian
extension of
K
whose norm group is corresponds to this subgroup. The following
theorem tells us this is the case:
Theorem. Let K be a local field. Then there is an isomorphism of posets
open finite index
subgroups of K
×
finite abelian
extensions of L/K
H (K
ab
)
Art
K
(H)
N(L/K) L/K
.
In particular, for L/K and M/K finite abelian extensions, we have
N(LM/K) = N(L/K) N(M/K),
N(L M/K) = N(L/K)N (M/K).
While proving this requires quite a bit of work, a small part of it follows from
local Artin reciprocity:
Theorem.
Let
L/K
be a finite extension, and
M/K
abelian. Then
N
(
L/K
)
N(M/K) iff M L.
Proof.
By the previous theorem, we may wlog
L/K
abelian by replacing with
L K
ab
. The direction is clear by the last part of Artin reciprocity.
For the other direction, we assume that we have
N
(
L/K
)
N
(
M/K
), and
let
σ Gal
(
K
ab
/L
). We want to show that
σ|
M
=
id
M
. This would then imply
that M is a subfield of L by Galois theory.
We know
W
(
K
ab
/L
) is dense in
Gal
(
K
ab
/L
). So it suffices to show this for
σ W (K
ab
/L). Then we have
W (K
ab
/L)
=
Art
K
(N(L/K)) Art
K
(N(M/K)).
So we can find
x M
×
such that
σ
=
Art
K
(
N
M/K
(
x
)). So
σ|
M
=
id
M
by local
Artin reciprocity.
Side note: Why is this called “class field theory”? Usually, we call the field
corresponding to the subgroup
H
the class field of
H
. Historically, the first type
of theorems like this are proved for number fields. The groups that appear on
the left would be different, but in some cases, they are the class group of the
number field.
8 Lubin–Tate theory
For the rest of the course, we will indicate how one can explicitly construct the
field K
ab
and the map Art
K
.
There are many ways we can approach local class field theory. The approach
we use, using Lubin–Tate theory, is the most accessible one. Another possible
approach is via Galois cohomology. This, however, relies on more advanced
machinery, namely Galois cohomology.
8.1 Motivating example
We will work out the details of local Artin reciprocity in the case of
Q
p
as a
motivating example for the proof we are going to come up with later. Here we
will need the results of local class field theory to justify our claims, but this is
not circular since this is not really part of the proof.
Lemma. Let L/K be a finite abelian extension. Then we have
e
L/K
= (O
×
K
: N
L/K
(O
×
L
)).
Proof.
Pick
x L
×
, and
w
the valuation on
L
extending
v
K
, and
n
= [
L
:
K
].
Then by construction of w, we know
v
K
(N
L/K
(x)) = nw(x) = f
L/K
v
L
(x).
So we have a surjection
K
×
N(L/K)
Z
f
L/K
Z
v
K
.
The kernel of this map is equal to
O
×
K
N(L/K)
N(L/K)
=
O
×
K
O
×
K
N (L/K)
=
O
×
K
N
L/K
(O
×
L
)
.
So by local class field theory, we know
n = (K
×
: N(L/K)) = f
L/K
(O
×
K
: N
L/K
(O
×
L
)),
and this implies what we want.
Corollary.
Let
L/K
be a finite abelian extension. Then
L/K
is unramified if
and only if N
L/K
(O
×
L
) = O
×
K
.
Now we fix a uniformizer
π
K
. Then we have a topological group isomorphism
K
×
=
hπ
K
i × O
×
K
.
Since we know that the finite abelian extensions correspond exactly to finite
index subgroups of
K
×
by taking the norm groups, we want to understand
subgroups of K
×
. Now consider the subgroups of K
×
of the form
hπ
m
K
i × U
(n)
K
.
We know these form a basis of the topology of
K
×
, so it follows that finite-index
open subgroups must contain one of these guys. So we can find the maximal
abelian extension as the union of all fields corresponding to these guys.
Since we know that
N
(
LM/K
) =
N
(
L/K
)
N
(
M/K
), it suffices to further
specialize to the cases
hπ
K
i × U
(n)
K
and
hπ
m
K
i × O
K
separately. The second case is easy, because this corresponds to an unramified
extension by the above corollary, and unramified extensions are completely
characterized by the extension of the residue field. Note that the norm group
and the extension are both independent of the choice of uniformizer. The
extensions corresponding to the first case are much more difficult to construct,
and they depend on the choice of
π
K
. We will get them from Lubin–Tate theory.
Lemma.
Let
K
be a local field, and let
L
m
/K
be the extension corresponding
to hπ
m
K
i × O
K
. Let
L =
[
m
L
m
.
Then we have
K
ab
= K
ur
L,
Lemma. We have isomorphisms
W (K
ab
/K)
=
W (K
ur
L/K)
=
W (K
ur
/K) × Gal(L/K)
=
Frob
Z
K
× Gal(L/K)
Proof.
The first isomorphism follows from the previous lemma. The second
follows from the fact that
K
ur
L
=
K
as
L
is totally ramified. The last
isomorphism follows from the fact that
T
K
ur
/K
=
K
ur
trivially, and then by
definition W (K
ur/K
)
=
Frob
Z
K
.
Example. We consider the special case of K = Q
p
and π
K
= p. We let
L
n
= Q
p
(ζ
p
n
),
where
ζ
p
n
is the primitive
p
n
th root of unity. Then by question 6 on example
sheet 2, we know this is a field with norm group
N(Q
p
(ζ
p
n
)/Q
p
) = hpi × (1 + p
n
Z
p
) = hpi × U
(n)
Q
p
,
and thus this is a totally ramified extension of Q
p
.
We put
Q
p
(ζ
p
) =
[
n=1
Q
p
(ζ
p
n
).
Then again this is totally ramified extension, since it is the nested union of
totally ramified extensions.
Then we have
Gal(Q
p
(ζ
p
)/Q
p
)
=
lim
n
Gal(Q
p
(ζ
p
n
)/Q
p
)
= lim
n
(Z/p
n
Z)
×
= Z
×
p
.
Note that we are a bit sloppy in this deduction. While we know that it is true
that
Z
×
p
=
lim
n
(
Z/p
n
Z
)
×
, the inverse limit depends not only on the groups
(
Z/p
n
Z
)
×
themselves, but also on the maps we use to connect the groups together.
Fortunately, from the discussion below, we will see that the maps
Gal(Q
p
(ζ
p
n
)/Q
p
) Gal(Q
p
(ζ
p
n1
)/Q
p
)
indeed correspond to the usual restriction maps
(Z/p
n
Z)
×
(Z/p
n1
Z)
×
.
It is a fact that this is the inverse of the Artin map of
Q
p
restricted to
Z
×
p
.
Note that we have
W
(
Q
p
(
ζ
p
)
/Q
p
) =
Gal
(
Q
p
(
ζ
p
)
/Q
p
) because its maximal
unramified subextension is trivial.
We can trace through the above chains of isomorphisms to figure out what
the Artin map does. Let m = Z
×
p
. Then we can write
m = a
0
+ a
1
p + ··· ,
where a
i
{0, ··· , p 1} and a
0
6= 0. Now for each n, we know
m a
0
+ a
1
p + ··· + a
n1
p
n1
mod p
n
.
By the usual isomorphism Gal(Q
p
(ζ
p
n
)/Q
p
)
=
Z/p
n
Z, we know m acts as
ζ
p
n
7→ ζ
a
0
+a
1
p+...+a
n1
p
n1
p
n
“=” ζ
m
p
n
on
Q
p
(
ζ
p
n
), where we abuse notation because taking
ζ
p
n
to powers of
p
greater
than
n
gives 1. It can also be interpreted as (1 +
λ
p
n
)
m
, where
λ
p
n
=
ζ
p
n
1 is
a uniformizer, which makes sense using binomial expansion.
So the above isomorphisms tells us that
Art
Q
p
restricted to
Z
×
p
acts on
Q
p
(ζ
p
) as
Art
Q
p
(m)(ζ
p
n
) σ
m
1
(ζ
p
n
) = ζ
m
1
p
n
.
The full Artin map can then be read off from the following diagram:
Q
×
p
W (Q
ab
p
/Q
p
)
hpi × Z
×
p
W (Q
ur
p
/Q
p
) × Gal(Q
p
(ζ
p
)/Q
p
)
=
Art
Q
p
restriction
where the bottom map sends
hp
n
, mi 7→ (Frob
n
Q
p
, σ
m
1
).
In fact, we have
Theorem (Local Kronecker-Weber theorem).
Q
ab
p
=
[
nZ
1
Q
p
(ζ
n
),
Q
ur
p
=
[
nZ
1
(n,p)=1
Q
p
(ζ
n
).
Not a proof.
We will comment on the proof of the generalized version later.
Remark.
There is another normalization of the Artin map which sends a
uniformizer to the geometric Frobenius, defined to be the inverse of the arithmetic
Frobenius. With this convention, Art
Q
p
(m)|
Q
p
(ζ
p
)
is σ
m
.
We can define higher ramification groups for general Galois extensions.
Definition
(Higher ramification groups)
.
Let
K
be a local field and
L/K
Galois.
We define, for s R
≥−1
G
s
(M/K) = {σ Gal(M/K) : σ|
L
G
s
(L/K) for all finite
Galois subextension M/K}.
This definition makes sense, because the upper number behaves well when
we take quotients. This is one of the advantages of upper numbering. Note that
we can write the ramification group as the inverse limit
G
s
(M/K)
=
lim
L/K
G
s
(L/K),
as in the case of the Galois group.
Example.
Going back to the case of
K
=
Q
p
. We write
Q
p
n
for the unramified
extension of degree n of Q
p
. By question 11 of example sheet 3, we know that
G
s
(Q
p
n
(ζ
p
m
)/Q
p
) =
Gal(Q
p
n
(ζ
p
m
)/Q
p
) s = 1
Gal(Q
p
n
(ζ
p
m
)/Q
p
n
) 1 < s 0
Gal(Q
p
n
(ζ
p
m
)
p
k
) k 1 < s k m 1
1 s > m 1
,
which corresponds to
hpi × U
(0)
hp
n
i × U
(m)
s = 1
hp
n
i × U
(0)
hp
n
i × U
(m)
1 < s 0
hp
n
i × U
(k)
hp
n
i × U
(m)
k 1 < s k m 1
1 s > m 1
under the Artin map.
By taking the limit as n, m , we get
Theorem. We have
G
s
(Q
ab
p
/Q
p
) = Art
Q
p
(1 + p
k
Z
p
) = Art
Q
p
(U
(k)
),
where k is chosen such that k 1 < s k, k Z
0
.
Corollary. If L/Q
p
is a finite abelian extension, then
G
s
(L/Q
p
) = Art
Q
p
N(L/Q
p
)(1 + p
n
Z
p
)
N(L/Q
p
)
,
where n 1 < s n.
Here Art
Q
p
induces an isomorphism
Q
×
p
N(L/Q
p
)
Gal(L/Q
p
).
So it follows that
L Q
p
(
ζ
p
m
) for some
n
if and only if
G
s
(
L/Q
p
) = 1 for all
s > m 1.
8.2 Formal groups
The proof of local Artin reciprocity will be done by constructing the analogous
versions of
L
n
for an arbitrary local field, and then proving that it works. To
do so, we will need the notion of a formal group. The idea of a formal group is
that a formal group is a rule that specifies how we should multiply two elements
via a power series over a ring
R
. Then if we have a complete
R
-module, then
the formal group will turn the
R
-module into an actual group. There is then a
natural notion of a formal module, which is a formal group
F
with an
R
-action.
At the end, we will pick
R
=
O
K
. The idea is then that we can fix an
algebraic closure
¯
K
, and then a formal
O
K
-module will turn
m
¯
K
into an actual
O
K
-module. Then if we adjoin the right elements of
m
¯
K
to
K
, then we obtain
an extension of
K
with a natural
O
K
action, and we can hope that this restricts
to field automorphisms when we restrict to the unit group.
Notation. Let R be a ring. We write
R[[x
1
, ··· , x
n
]] =
X
k
1
,...,k
n
Z
0
a
k
1
,...,k
n
x
k
1
1
···x
k
n
n
: a
k
1
,...,k
n
R
for the ring of formal power series in n variables over R.
Definition
(Formal group)
.
A (one-dimensional, commutative) formal group
over R is a power series F (X, Y ) R[X, Y ] such that
(i) F (X, Y ) X + Y mod (X
2
, XY, Y
2
)
(ii) Commutativity: F (X, Y ) = F (Y, X)
(iii) Associativity: F (X, F (Y, Z)) = F (F (X, Y ), Z).
This is most naturally understood from the point of view of algebraic geometry,
as a generalization of the Lie algebra over a Lie group. Instead of talking about
the tangent space of a group (the “first-order neighbourhood”), we talk about its
infinitesimal (formal) neighbourhood, which contains all higher-order information.
A lot of the seemingly-arbitrary compatibility conditions we later impose have
such geometric motivation that we unfortunately cannot go into.
Example.
If
F
is a formal group over
O
K
, where
K
is a complete valued field,
then
F
(
x, y
) converges for all
x, y m
K
. So
m
K
becomes a (semi)group under
the multiplication
(x, y) 7→ F (x, y) m
k
Example. We can define
ˆ
G
a
(X, Y ) = X + Y.
This is called the formal additive group.
Similarly, we can have
ˆ
G
m
(X, Y ) = X + Y + XY.
This is called the formal multiplicative group. Note that
X + Y + XY = (1 + X)(1 + Y ) 1.
So if
K
is a complete valued field, then
m
K
bijects with 1 +
m
k
by sending
x 7→
1 +
x
, and the rule sending (
x, y
)
m
2
K
7→ x
+
y
+
xy m
K
is just the
usual multiplication in 1 + m
K
transported to m
K
via the bijection above.
We can think of this as looking at the group in a neighbourhood of the
identity 1.
Note that we called this a formal group, rather than a formal semi-group. It
turns out that the existence of identity and inverses is automatic.
Lemma. Let R be a ring and F a formal group over R. Then
F (X, 0) = X.
Also, there exists a power series i(X) X · R[[X]] such that
F (X, i(X)) = 0.
Proof. See example sheet 4.
The next thing to do is to define homomorphisms of formal groups.
Definition
(Homomorphism of formal groups)
.
Let
R
be a ring, and
F, G
be
formal groups over
R
. A homomorphism
f
:
F G
is an element
f R
[[
X
]]
such that f (X) 0 mod X and
f(F (X, Y )) = G(f(X), f(Y )).
The endomorphisms
f
:
F F
form a ring
End
R
(
F
) with addition +
F
given by
(f +
F
g)(x) = F (f(x), g(x)).
and multiplication is given by composition.
We can now define a formal module in the usual way, plus some compatibility
conditions.
Definition
(Formal module)
.
Let
R
be a ring. A formal
R
-module is a formal
group
F
over
R
with a ring homomorphism
R End
R
(
F
), written,
a 7→
[
a
]
F
,
such that
[a]
F
(X) = aX mod X
2
.
Those were all general definitions. We now restrict to the case we really care
about. Let K be a local field, and q = |k
K
|. We let π O
K
be a uniformizer.
Definition
(Lubin–Tate module)
.
A Lubin–Tate module over
O
K
with respect
to π is a formal O
K
-module F such that
[π]
F
(X) X
q
mod π.
We can think of this condition of saying “uniformizer corresponds to the
Frobenius”.
Example.
The formal group
ˆ
G
m
is a Lubin–Tate
Z
p
module with respect to
p
given by the following formula: if a Z
p
, then we define
[a]
ˆ
G
m
(X) = (1 + X)
a
1 =
X
n=1
a
n
X
n
.
The conditions
(1 + X)
a
1 aX mod X
2
and
(1 + X)
p
1 X
p
mod p
are clear.
We also have to check that
a 7→
[
a
]
F
is a ring homomorphism. This follows
from the identities
((1 + X)
a
)
b
= (1 + X)
ab
, (1 + X)
a
(1 + X)
b
= (1 + X)
ab
,
which are on the second example sheet.
The objective of the remainder of the section is to show that all Lubin–Tate
modules are isomorphic.
Definition
(Lubin–Tate series)
.
A Lubin–Tate series for
π
is a power series
e(X) O
K
[[X]] such that
e(X) πX mod X
2
, e(X) X
q
mod π.
We denote the set of Lubin–Tate series for π by E
π
.
Now by definition, if
F
is a Lubin–Tate
O
K
module for
π
, then [
π
]
F
is a
Lubin–Tate series for π.
Definition
(Lubin–Tate polynomial)
.
A Lubin–Tate polynomial is a polynomial
of the form
uX
q
+ π(a
q1
X
q1
+ ··· + a
2
X
2
) + πX
with u U
(1)
K
, and a
q1
, ··· , a
2
O
K
.
In particular, these are Lubin–Tate series.
Example. X
q
+ πX is a Lubin–Tate polynomial.
Example.
If
K
=
Q
p
and
π
=
p
, then (1 +
X
)
p
1 is a Lubin–Tate polynomial.
The result that allows us to prove that all Lubin–Tate modules are isomorphic
is the following general result:
Lemma. Let e
1
, e
2
E
π
and take a linear form
L(x
1
, ··· , x
n
) =
n
X
i=1
a
i
X
i
, a
i
O
K
.
Then there is a unique power series
F
(
x
1
, ··· , x
n
)
O
K
[[
x
1
, ··· , x
n
]] such that
F (x
1
, ··· , x
n
) L(x
1
, ··· , x
n
) mod (x
1
, ··· , x
n
)
2
,
and
e
1
(F (x
1
, ··· , x
n
)) = F (e
2
(x
1
), e
2
(x
2
), ··· , e
2
(x
n
)).
For reasons of time, we will not prove this. We just build
F
by successive
approximation, which is not terribly enlightening.
Corollary.
Let
e E
π
be a Lubin–Tate series. Then there are unique power
series F
e
(X, Y ) O
K
[[X, Y ]] such that
F
e
(X, Y ) X + Y mod (X + Y )
2
e(F
e
(X, Y )) = F
e
(e(X), e(Y ))
Corollary.
Let
e
1
, e
2
E
π
be Lubin–Tate series and
a O
K
. Then there exists
a unique power series [a]
e
1
,e
2
O
K
[[X]] such that
[a]
e
1
,e
2
(X) aX mod X
2
e
1
([a]
e
1
,e
2
(X)) = [a]
e
1
,e
2
(e
2
(X)).
To simplify notation, if e
1
= e
2
= e, we just write [a]
e
= [a]
e,e
.
We now state the theorem that classifies all Lubin–Tate modules in terms of
Lubin–Tate series.
Theorem.
The Lubin–Tate
O
K
modules for
π
are precisely the series
F
e
for
e E
π
with formal O
K
-module structure given by
a 7→ [a]
e
.
Moreover, if
e
1
, e
2
E
π
and
a O
K
, then [
a
]
e
1
,e
2
is a homomorphism from
F
e
2
F
e
1
.
If a O
×
K
, then it is an isomorphism with inverse [a
1
]
e
2
,e
1
.
So in some sense, there is only one Lubin–Tate module.
Proof sketch.
If
F
is a Lubin–Tate
O
K
-module for
π
, then
e
= [
π
]
F
E
π
by
definition, and
F
satisfies the properties that characterize the series
F
e
. So
F = F
e
by uniqueness.
For the remaining parts, one has to verify the following for all
e, e
1
, e
2
, e
3
E
π
and a, b O
K
.
(i) F
e
(X, Y ) = F
e
(Y, X).
(ii) F
e
(X, F
e
(Y, Z)) = F
e
(F
e
(X, Y ), Z).
(iii) [a]
e
1
,e
2
(F
e
(X, Y )) = F
e
1
([a]
e
1
,e
2
(X), [a]
e
1
,e
2
(Y )).
(iv) [ab]
e
1
,e
3
(X) = [a]
e
1
,e
2
([b]
e
2
,e
3
(X)).
(v) [a + b]
e
1
,e
2
(X) = [a]
e
1
,e
2
(X) + [b]
e
1
,e
2
(X).
(vi) [π]
e
(X) = e(X).
The proof is just repeating the word “uniqueness” ten times.
8.3 Lubin–Tate extensions
We now use the Lubin–Tate modules to do things. As before, we fixed an
algebraic closure
¯
K of K. We let
¯
m = m
¯
K
be the maximal ideal in O
¯
K
.
Proposition.
If
F
is a formal
O
K
-module, then
¯
m
becomes a (genuine)
O
K
module under the operations +
F
and ·
x +
F
y = F (x, y)
a · x = [a]
F
(x)
for all x, y
¯
m and a O
K
.
We denote this
¯
m
F
.
This isn’t exactly immediate, because
¯
K
need not be complete. However,
this is not a problem as each multiplication given by
F
only involves finitely
many things (namely two of them).
Proof.
If
x, y
¯
m
, then
F
(
x, y
) is a series in
K
(
x, y
)
¯
K
. Since
K
(
x, y
) is
a finite extension, we know it is complete. Since the terms in the sum have
absolute value
<
1 and
0, we know it converges to an element in
m
K(x,y)
¯
m
.
The rest then essentially follows from definition.
To prove local class field theory, we want to find elements with an
U
K
/U
(n)
K
action for each
n
, or equivalently elements with an
O
K
/O
(n)
K
action. Note that
the first quotient is a quotient of groups, while the second quotient is a quotient
of a ring by an ideal. So it is natural to consider the following elements:
Definition
(
π
n
-division points)
.
Let
F
be a Lubin–Tate
O
K
-module for
π
. Let
n 1. The group F (n) of π
n
-division points of F is defined to be
F (n) = {x
¯
m
F
| [π
n
]
F
x = 0} = ker([π
n
]
F
).
This is a group under the operation given by F , and is indeed an O
K
module.
Example. Let F =
ˆ
G
m
, K = Q
p
and π = p. Then for x
¯
m
ˆ
G
m
, we have
p
n
· x = (1 + x)
p
n
1.
So we know
ˆ
G
m
(n) = {ζ
i
p
n
1 | i = 0, 1, ··· , p
n
1},
where ζ
p
n
¯
Q
p
is the primitive p
n
th root of unity.
So
ˆ
G
m
(n) generates Q
p
(ζ
p
n
).
To prove this does what we want, we need the following lemma:
Lemma. Let e(X) = X
q
+ πX. We let
f
n
(X) = (e ··· e)
| {z }
n times
(X).
Then f
n
has no repeated roots. Here we take f
0
to be the identity function.
Proof.
Let
x
¯
K
. We claim that if
|f
i
(
x
)
| <
1 for
i
= 0
, ··· , n
1, then
f
0
n
(X) 6= 0.
We proceed by induction on n.
(i) When n = 1, we assume |x| < 1. Then
f
0
1
(x) = e
0
(x) = qx
q1
+ π = π
1 +
q
π
x
q1
6= 0,
since we know
q
π
has absolute value
1 (
q
vanishes in
k
K
, so
q
lives in
O
K
), and x
q1
has absolute value < 1.
(ii) in the induction step, we have
f
0
n+1
(x) = (qf
n
(x)
q1
+ π)f
0
n
(x) = π
1 +
q
π
f
n
(x)
q1
f
0
n
(x).
By induction hypothesis, we know
f
0
n
(
x
)
6
= 0, and by assumption
|f
n
(
x
)
| <
1. So the same argument works.
We now prove the lemma. We assume that
f
n
(
x
) = 0. We want to show that
|f
i
(x)| < 1 for all i = 0, ··· , n 1. By induction, we have
f
n
(x) = x
q
n
+ πg
n
(x)
for some
g
n
(
x
)
O
K
[
x
]. It follows that if
f
n
(
x
) = 0, then
|x| <
1. So
|f
i
(
x
)
| <
1
for all i. So f
0
n
(x) 6= 0.
The point of the lemma is to prove the following proposition:
Proposition. F
(
n
) is a free
O
K
n
O
K
module of rank 1. In particular, it has
q
n
elements.
Proof. By definition, we know
π
n
· F (n) = 0.
So F (n) is indeed an O
K
n
O
K
-module.
To prove that it is free of rank 1, we note that all Lubin–Tate modules
for
π
are isomorphic. This implies that all the honest
O
K
modules
F
(
n
) are
isomorphic. We choose
F
=
F
e
, where
e
=
X
q
+
πX
. Then
F
(
n
) consists
of the roots of the polynomial
f
n
=
e
n
(
X
), which is of degree
q
n
and has no
repeated roots. So
|F
(
n
)
|
=
q
n
. To show that it is actually the right thing, if
λ
n
F (n) \ F (n 1), then we have a homomorphism
O
K
F (n)
given by
A 7→ a · λ
n
. Its kernel is
π
n
O
K
by our choice of
λ
n
. By counting, we
get an O
K
-module isomorphism
O
K
π
n
O
K
F (n)
as desired.
Corollary. We have isomorphisms
O
K
π
n
O
K
=
End
O
K
(F (n))
U
K
U
(n)
K
=
Aut
O
K
(F (n)).
Given a Lubin–Tate O
K
-module F for π, we consider
L
n,π
= L
n
= K(F(n)),
which is the field of
π
n
division points of
F
. From the inclusions
F
(
n
)
F
(
n
+1)
for all n, we obtain a corresponding inclusion of fields
L
n
L
n+1
.
The field
L
n
depends only in
π
, and not on
F
. To see this, we let
G
be another
Lubin–Tate O
K
-module, and let f : F G be an isomorphism. Then
G(n) = f(F (n)) K(F(n))
since the coefficients of f lie in K. So we know
K(G(n)) K(F (n)).
By symmetry, we must have equality.
Theorem. L
n
/K
is a totally ramified abelian extension of degree
q
n1
(
q
1)
with Galois group
Gal(L
n
/K)
=
Aut
O
K
(F (n))
=
U
K
U
(n)
K
.
Explicitly, for any σ Gal(L
n
/K), there is a unique u U
K
/U
(n)
K
such that
σ(λ) = [u]
F
(λ)
for all λ F (n). Under this isomorphism, for m n, we have
Gal(L
m
/L
n
)
=
U
(n)
K
U
(m)
K
.
Moreover, if F = F
e
, where
e(X) = X
q
+ π(a
q1
π
q1
+ ··· + a
2
X
2
) + πX,
and λ
n
F (n) \ F (n 1), then λ
n
is a uniformizer of L
n
and
φ
n
(X) =
e
n
(X)
e
n1
(X)
= X
q
n1
(q1)
+ ··· + π
is the minimal polynomial of λ
n
. In particular,
N
L
n
/K
(λ
n
) = π.
Proof. Consider a Lubin–Tate polynomial
e(X) = x
q
+ π(a
q1
X
q1
+ ··· + a
2
X
2
) + πX.
We set F = F
e
. Then
φ
n
(X) =
e
n
(X)
e
n1
(X)
= (e
n1
(X))
q1
+ π(a
q
1
e
n1
(X)
q2
+ ··· + a
2
e
n1
(X)) + π
is an Eisenstein polynomial of degree
q
n1
(
q
1) by starting at it long enough.
So if
λ
n
F
(
n
)
\ F
(
n
1), then
λ
n
is a root of
φ
n
(
x
), so
K
(
λ
n
)
/K
is totally
ramified of degree q
n1
(q 1), and λ
n
is a uniformizer, and
N
K(λ
n
)/K
(λ
n
) = π
as the norm is just the constant coefficient of the minimal polynomial.
Now let
σ Gal
(
L
n
/K
). Then
σ
induces a permutation of
F
(
n
), as these
are the roots of e
n
(X), which is in fact O
K
-linear, i.e.
σ(x) +
F
σ(y) = F(σ(x), σ(y)) = σ(F (x, y)) = σ(x +
F
y)
σ(a · x) = σ([a]
F
(x)) = [a]
F
(σ(x)) = a · σ(x)
for all x, y m
L
n
and a O
K
.
So we have an injection of groups
Gal(L
n
/K) Aut
O
K
(F (n)) =
U
K
U
(n)
K
But we know
U
K
U
(n)
K
= q
n1
(q 1) = [K(λ
n
) : K] [L
n
: K] = |Gal(L
n
/K)|.
So we must have equality throughout, the above map is an isomorphism, and
K(λ
n
) = L
n
.
It is clear from the construction of the isomorphism that for
m n
, the
diagram
Gal(L
m
/K) U
K
/U
(m)
K
Gal(L
n
/K) U
K
/U
(n)
K
restriction
quotient
commutes. So the isomorphism
Gal(L
m
/L
n
)
=
U
(m)
K
U
(n)
K
follows by looking at the kernels.
Example. In the case where K = Q
p
and π = p, recall that
ˆ
G
m
(n) = {ζ
i
p
n
1 | i = 0, ··· , p
n1
1},
where ζ
p
n
is the principal p
n
th root of unity. The theorem then gives
Gal(Q
p
(ζ
p
n
)/Q
p
)
=
(Z/p
n
)
×
given by if a Z
0
and (a, p) = 1, then
σ
a
(ζ
i
p
n
1) = [a]
ˆ
G
m
(n)
(ζ
i
p
n
1) = (1 + (ζ
i
p
n
1))
a
1 = ζ
ai
p
n
1.
This agrees with the isomorphism we previously constructed.
Back to the general situation, setting
L
=
[
n=1
L
n
,
we know L
/K is Galois, and we have isomorphisms
Gal(L
/K) lim
Gal(L
n
/K) lim
n
U
K
/U
(n)
K
=
U
K
σ (σ|
L
n
)
n
This map will be the inverse of the Artin map restricted to L
.
To complete the proof of Artin reciprocity, we need to use the following
theorem without proof:
Theorem (Generalized local Kronecker-Weber theorem). We have
K
ab
= K
ur
L
(for any π).
Comments on the proof.
One can prove this from the Hasse-Arf theorem, which
states that in an abelian extension, the jumps in the upper ramification groups
occur only at integer values. This, together with the calculation of ramification
groups done later, easily implies the theorem. Essentially,
L
maxed out all
possible jumps of the upper ramification groups. However, the Hasse-Arf theorem
is difficult to prove.
Another approach is to prove the existence of the Artin map using other
techniques (e.g. Galois cohomology). Consideration of the norm group (cf. the
next theorem) then implies the theorem. The content of this section then
becomes an explicit construction of a certain family of abelian extensions.
We can characterize the norm group by
Theorem. We have
N(L
n
/K) = hπi × U
(n)
k
.
Comments on the proof.
This can be done by defining Coleman operators, which
are power series representations of the norm. Alternatively, assuming the
description of the local Artin map given below and local Artin reciprocity,
U
(n)
k
is in the kernel of
Art|
L
n
, so
hπi × U
(n)
k
N
(
L
n
/K
). The result follows by
comparing order.
We can then construct the Artin map as follows:
Theorem.
Let
K
be a local field. Then we have an isomorphism
Art
:
K
×
W (K
ab
/K) given by the composition
K
×
W (K
ab
/K)
hπi× U
K
Frob
Z
K
× Gal(L
/K)
Art
where the bottom map is given by (π
m
, u) 7→ (Frob
m
K
, σ
u
1
), where
σ
u
(λ) = [u]
F
(λ)
for all λ
S
n=1
F (n).
The inverse shows up in the proof to make sure the map defined above is
independent of the choice of uniformizer. We will not prove this, nor that the
map obtained has the desired properties. Instead, we will end the course by
computing the higher ramification groups of these extensions.
Theorem. We have
G
s
(L
n
/K) =
Gal(L
n
/K) 1 s 0
Gal(L
n
/L
k
) q
k1
1 < s q
k
1, 1 k n 1
1 s > q
n1
Proof. The case for 1 s 0 is clear.
For 0 s 1 (which we may wlog is actually 1), we know that
Gal(L
n
/L
k
)
=
U
(k)
K
/U
(n)
K
under the isomorphism
Gal
(
L
n
/K
)
=
U
K
/U
(n)
K
. On the other hand, we know
G
1
(L
n
/K) is the Sylow p-subgroup of Gal(L
n
/K). So we must have
G
1
(L
n
/K)
=
U
(1)
K
/U
(n)
K
.
So we know that
G
1
(
L
n
/K
) =
Gal
(
L
n
/L
1
). Thus we know that
G
s
(
L
n
/K
) =
Gal(L
n
/K) for 0 < s 1.
We now let σ = σ
u
G
1
(L
n
/K) and u U
(1)
K
/U
(n)
K
. We write
u = 1 + επ
k
for some
ε U
K
and some
k
=
k
(
u
)
1. Since
σ
is not the identity, we know
k < n. We claim that
i
L
n
/K
(σ) = v
L
n
(σ(λ) λ) = q
k
.
Indeed, we let
λ F
(
n
)
\ F
(
n
1), where
F
is a choice of Lubin–Tate module
for π. Then λ is a uniformizer of L
n
and O
L
n
= O
K
[λ]. We can compute
σ
u
(λ) = [u]
F
(λ)
= [1 + επ
k
]
F
(λ)
= F (λ, [επ
k
]
F
(λ))
Now we can write
[επ
k
]
F
(λ) = [ε]
F
([π
k
]
F
(λ)) F (n k) \ F (n k 1),
since [
ε
]
F
is invertible, and applying [
π
nk
]
F
to [
π
k
]
F
(
λ
) kills it, but applying
[π
nk1
]
F
gives [π
n1
]
F
, which does not kill.
So we know [
επ
k
]
F
(
λ
) is a uniformizer of
L
nk
. Since
L
n
/L
nk
is totally
ramified of degree q
k
, we can find ε
0
O
×
L
n
such that
[επ
k
]
F
(λ) = ε
0
λ
q
k
Recall that F (X, 0) = X and F (0, Y ) = Y . So we can write
F (X, Y ) = X + Y + XY G(X, Y ),
where G(X, Y ) O
K
[[X, Y ]]. So we have
σ(λ) λ = F (λ, [επ
k
]
F
(λ)) λ
= F (λ, ε
0
λ
q
k
) λ
= λ + ε
0
λ
q
k
+ ε
0
λ
q
k
+1
G(λ, ε
0
λ
q
k
) λ
= ε
0
λ
q
k
+ ε
0
λ
q
k
+1
G(λ, ε
0
λ
q
k
).
In terms of valuation, the first term is the dominating term, and
i
L
n
/K
(σ) = v
L
n
(σ(λ) λ) = q
k
So we know
i
L
n
/K
(σ
k
) s + 1 q
k(u)
1 s.
So we know
G
s
(L
n
/K) = {σ
K
G
1
(L
n
/K) : q
k(u)
1 s} = Gal(L
n
/L
k
),
where q
k1
1 < s q
k
1 for k = 1, ··· , n 1, and 1 if s > q
n1
= 1.
Corollary. We have
G
t
(L
n
/K) =
Gal(L
n
/K) 1 t 0
Gal(L
n
/L
k
) k 1 < t k, k = 1, ··· , n 1
1 t > n 1
In other words, we have
G
t
(L
n
/K) =
(
Gal(L
n
/L
dte
) 1 t n 1
1 t > n 1
,
where we set L
0
= K.
Once again, the numbering is a bit more civilized in the upper numbering.
Proof. We have to compute the integral of
1
(G
0
(L
n
/K) : G
x
(L
n
/K)
.
We again plot this out
1
q1
1
q(q1)
1
q
2
(q1)
q 1
q
2
1 q
3
1
So by the same computation as the ones we did last time, we find that
η
L
n
/K
(s) =
s 1 s 0
(k 1) +
s(q
k1
1)
q
k1
(q1)
q
k1
1 s q
k
1, k = 1, ··· , n 1
(n 1) +
s(q
n1
1)
q
n1
(q1)
s > q
n1
1.
Inverting this, we find that
ψ
L
n
/K
=
t 1 t 0
q
dte−1
(q 1)(t (dte 1)) + q
dte−1
1 1 < t n 1
q
n1
(q 1)(t (n 1)) + q
n1
1 t > n 1
.
Then we have
G
t
(L
n
/K) = G
ψ(L
n
/K)(t)
(L
n
/K),
which gives the desired by the previous theorem.
So we know that
Art
1
K
(G
t
(L
n
/K)) =
(
U
dte
K
/U
(n)
K
1 t n
1 t n
.
Corollary. When t > 1, we have
G
t
(K
ab/K
) = Gal(K
ab
/K
ur
L
dte
),
and
Art
1
K
(G
t
(K
ab
/K)) = U
(dte)
.
Proof.
Recall the following fact from the examples class: If
L/K
is finite un-
ramified and
M/K
is finite totally ramified, then
LM/L
is totally ramified, and
Gal(LM/L)
=
Gal(M/K) by restriction, and
G
t
(LM/K)
=
G
t
(M/K).
via this isomorphism (for t > 1).
Now let
K
m
/K
be the unramified extension of degree
m
. By the lemma and
the previous corollary, we have
G
t
(K
m
L
n
/K)
=
G
t
(L
n
/K) =
(
Gal(L
n
/L
dte
) 1 < t n
1 t n
=
(
Gal(K
m
L
n
/K
m
L
dte
) 1 < t n
1 t n
So we have
G
t
(K
ab
/K) = G
t
(K
ur
L
/K)
= lim
m,n
G
t
(K
m
L
n
/K)
= lim
m,n
n≥dte
Gal(K
m
L
n
/K
m
L
dte
)
= Gal(K
ur
L
/K
ur
L
dte
)
= Gal(K
ab
/K
ur
L
dte
),
and
Art
1
K
(Gal(K
ab
/K
ur
L
dte
)) = Art
1
K
lim
m,n
n≥dte
Gal(K
m
L
n
/K
m
L
dte
)
= lim
m,n
n≥dte
Art
1
K
Gal(K
m
L
n
/K
m
L
dte
)
= lim
m,n
n≥dte
U
(dte)
K
U
(n)
K
= U
dte
.
Corollary.
Let
M/K
be a finite abelian extension. Then we have an isomor-
phism
Art
K
:
K
×
N(M/K)
=
Gal(M/K).
Moreover, for t > 1, we have
G
t
(M/K) = Art
K
U
(dte)
K
N(M/K)
N(M/K)
!
.
Proof. We have
G
t
(M/K) =
G
t
(K
ab
/K)G(K
ab
/M )
G(K
ab
/M )
= Art
U
(dte)
K
N(M/K)
N(M/K)
!
.