Local Fields and p-adic Numbers
Local Fields and p-adic Numbers
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 Qp (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 = pn 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 Qp , 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.
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.
1
Contents III Local Fields
Contents
0 Introduction 3
1 Basic theory 4
1.1 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Topological rings . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 The p-adic numbers . . . . . . . . . . . . . . . . . . . . . . . . . 11
2 Valued fields 15
2.1 Hensel’s lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Extension of norms . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Newton polygons . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
8 Lubin–Tate theory 71
8.1 Motivating example . . . . . . . . . . . . . . . . . . . . . . . . . 71
8.2 Formal groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
8.3 Lubin–Tate extensions . . . . . . . . . . . . . . . . . . . . . . . . 79
Index 88
2
0 Introduction III Local Fields
0 Introduction
What are local fields? Suppose we are interested in some basic number theoretic
problem. Say we have a polynomial f (x1 , · · · , xn ) ∈ Z[x1 , · · · , xn ]. We want to
look for solutions a ∈ Zn , 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 p2 , or p3 , 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.
3
1 Basic theory III Local Fields
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;
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 (
1 x 6= 0
|x| = .
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
4
1 Basic theory III Local Fields
B(x, r) = {y : |x − y| ≤ r}.
So closed balls do not have unique “centers”. Every point can be viewed as
the center.
Proof. Let y ∈ B(z, r). Then
5
1 Basic theory III Local Fields
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:
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 (xn ) be a sequence in K. If xn − xn+1 → 0, then xn is Cauchy.
The converses to all these are of course also true, with the usual proofs.
Proof.
(i) Pick ε > 0 and N such that |xn − xn+1 | < ε for all n ≥ N . Then given
m ≥ n ≥ N , we have
6
1 Basic theory III Local Fields
Definition (Valuation ring). Let K be a valued field. Then the valuation ring
of K is the open subring
OK = {x : |x| ≤ 1}.
We prove that it is actually a ring
Proposition. Let K be a valued field. Then
OK = {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 OK . Moreover, OK = {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 ∈ OR , but 1 + 1 6∈ OR .
Proof. We know that these sets are open since all balls are open.
To see OK is a subring, we have |1| = |−1| = 1. So 1, −1 ∈ OK . If x, y ∈ OK ,
then |x + y| ≤ max(|x|, |y|) ≤ 1. So x + y ∈ OK . Also, |xy| = |x||y| ≤ 1 · 1 = 1.
So xy ∈ OK .
That the other sets are ideals of OK is checked in the same way.
×
To check the units, we have x ∈ OK ⇔ |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 √12 is not.
7
1 Basic theory III Local Fields
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[s1 , · · · , sn ] by a sequence
and each si is integral over R[s1 , · · · , sn−1 ]. 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 s1 .
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, · · · , sdeg f −1 .
In the other direction, let t1 , · · · , td be R-module generators of R[s1 , · · · , sn ].
We show that in fact any element of R[s1 , · · · , sn ] is integral over R. Consider
any element b ∈ R[s1 , · · · , sn ]. Then there is some aij ∈ R such that
d
X
bti = aij tj .
j=1
det(bI − A)tj = 0
P
for all j. Now we know 1 ∈ R. So 1 = cj tj for some cj ∈ R. Then we have
X X
det(bI − A) = det(bI − A) cj t j = cj (det(bI − A)tj ) = 0.
8
1 Basic theory III Local Fields
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((x0 , y0 ), (x1 , y1 )) = max(|x0 − x1 |, |y0 − y1 |).
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 ∈ OK : |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 ⊇ I2 ⊇ I3 ⊇ I4 ⊇ · · ·
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 + In ⊆ 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.
9
1 Basic theory III Local Fields
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
a0 + a1 x + a2 x2 + a3 x3 + · · ·
to “converge”, at least if the ai are “of bounded size”. In general, the ai are
“not too big” if ai xi is genuinely a member of xi 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
(a0 , a0 + a1 x, a0 + a1 x + a2 x2 , · · · )
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 k−1 , then we obtain L mod I k−1 .
In general, suppose we have a sequence
(bn ∈ R/I n )∞
n=1 .
such that bn mod I n−1 = bn−1 . 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 R1 , R2 , , · · · be topological rings, with
continuous homomorphisms fn : Rn+1 → Rn .
R1 f1
R2 f2
R3 f3
R4 ···
Q Q Q
Rn × Rn Rn
10
1 Basic theory III Local Fields
By the definition
Q of the subspace topology, it suffices to show that the correspond-
ing maps on Rn are continuous.Q By the Quniversal property of the product, it
suffices to show that the projects Rn × Rn → Rm is continuous for all m.
But this map can alternatively be obtained by first projecting to Rm , then doing
multiplication in Rm , 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 Rn is the
←−
same as giving a continuous ring homomorphism gn : S → Rn for each n, such
that each of the following diagram commutes:
gn
S Rn
gn−1 fn−1
Rn−1
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 ).
11
1 Basic theory III Local Fields
Note that we must have bd coprime to p, but ab + pm−n 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 pn || x (we say pn || x if
p | x and pn+1 - x).
n
Zp = {x ∈ Qp : |x|p ≤ 1}
Z(p) = {x ∈ Q : |x|p ≤ 1}
So done.
Proposition. The non-zero ideals of Zp are pn Zp for n ≥ 0. Moreover,
Z ∼ Zp
= n .
pn Z p Zp
12
1 Basic theory III Local Fields
{y ∈ Z : |y − x|p ≤ p−n } = x + pn Z.
So done.
Proposition. Zp 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
ν (fn )n
Zp lim Zp /(pn Zp ) lim Z/(pn Z)
←−
zn ∈ lim Zp /pn Zp .
←−
13
1 Basic theory III Local Fields
n = − logp |a|p ∈ Z.
Proof. The second part follows from the first part by multiplying a by p−n .
Example. We have
1
= 1 + p + p2 + p3 + · · · .
1−p
14
2 Valued fields III Local Fields
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) = − logc |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 ∈ Qp , we define
15
2 Valued fields III Local Fields
remaining elements form an ideal (since the field is non-archimedean), and thus
we have a maximal ideal
m = mK = {x ∈ K : |x| < 1}
The quotient
k = kK = OK /mK
is known as the residue field .
Example. Let K = Qp . Then O = Zp , and m = pZp . So
k = O/m = Zp /pZp ∼
= Z/pZ.
max |ai | = 1.
i
f¯(x) = ḡ(x)h̄(x)
f (x) = g(x)h(x)
in O[x] with
ḡ = g, h̄ = h mod m,
with deg g = deg ḡ.
Note that requiring deg g = deg ḡ is the best we can hope for — we cannot
guarantee deg h = deg h̄, since we need not have deg f = deg f¯.
This is one of the most important results in the course.
Proof. Let g0 , h0 be arbitrary lifts of ḡ and h̄ to O[x] with deg ḡ = g0 and
deg h̄ = h0 . Then we have
f = g0 h0 mod m.
16
2 Valued fields III Local Fields
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 = g0 h0 + πr0 , r0 ∈ O[x].
If we are lucky enough that deg r0 b < deg g0 , then we group as we learnt in
secondary school to get
g1 = g0 + πr0 b
h1 = h0 + πr0 a,
If it is not true that deg r0 b ≤ deg g0 , we use the division algorithm to write
r0 b = qg0 + p.
Then we have
f = g0 h0 + π((r0 a + q)g0 + ph0 ),
and then proceed as above.
Given the factorization (∗), we replace r1 by r1 (ag0 + bh0 ), and then repeat
the procedure to get a factorization
f ≡ gk hk mod π k+1
gk ≡ gk−1 mod π k
hk ≡ hk−1 mod π k
deg gk = deg ḡ
Note that we may drop the terms of hk whose coefficient are in π k+1 O, and the
above equations still hold. Moreover, we can then bound deg hk ≤ deg f − deg gk .
It now remains to set
g = lim gk , h = lim hk .
k→∞ k→∞
17
2 Valued fields III Local Fields
18
2 Valued fields III Local Fields
then
kxkmax = max |ai |
i
defines a norm on V .
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 · kmax .
Corollary. V is complete with respect to any norm.
19
2 Valued fields III Local Fields
For n ≥ 2, we let
is also closed. BySconstruction, this does not contain 0. So there is some C > 0
n
such that if x ∈ i=1 xi + Vi , then kxk ≥ C. We claim that
C kxkmax ≤ kxk.
P
Indeed, take x = ai xi ∈ V . Let r be such that
Then
−1
kxkmax kxk =
a−1
r x
a1 ar−1 ar+1 an
=
x1 + · · · +
xr−1 + xr + xr+1 + · · · + xn
ar ar ar ar
≥ C,
20
2 Valued fields III Local Fields
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 q
|α| = n NL/K (α),
L
is a norm.
(i) If |α|L = 0, then NL/K (α) = 0. This is true iff α = 0.
(ii) The multiplicativity
√ of |α| and follows from the multiplicativity of NL/K ,
| · | and n · .
To show the strong triangle inequality, it suffices to show that |α|L ≤ 1 implies
|α + 1|L ≤ 1.
Recall that
We need to show that ai ∈ OK for all i. In other words, |ai | ≤ 1 for all i. This
is easy for a0 , since
NL/K (α) = ±am 0 ,
|ai | ≤ max(|a0 |, 1)
By general properties of the field norm, there is some m ∈ Z≥1 such that
NL/K (α) = ±am0 . So we have
|ai | ≤ max NL/K (α)1/m , 1 = 1.
21
2 Valued fields III Local Fields
for some d ∈ Z≥1 , and each σ(α) is integral over OK , since α is (apply σ to the
minimal polynomial). This implies that NL/K (α) is integral over OK (and lies
in K). So NL/K (α) ∈ OK since OK 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 OL is the integral
closure of OK in L.
t4 + p2 t4 − p3 t2 + pt + p3 .
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:
valuation of coefficient
power of t
0 1 2 3 4
22
2 Valued fields III Local Fields
valuation of coefficient
power of t
0 1 2 3 4
−1
23
2 Valued fields III Local Fields
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 (Q2 , v2 ) with the polynomial
7 9
4t4 + 5t3 + t + .
2 2
valuation of coefficient
power of t
24
2 Valued fields III Local Fields
Note that by lower convexity, there can be at most one line segment for each
slope. So this theorem makes sense.
w(α1 ) = · · · = w(αs1 ) = m1
w(αs1 +1 ) = · · · = w(αs2 ) = m2
..
.
w(αst ) = · · · = w(αn ) = mt+1 ,
where we have
m1 < m2 < · · · < mt+1 .
Then we know
v(an ) = v(1) = 0
X
v(an−1 ) = w αi ≥ min w(αi ) = m1
i
X
v(an−2 ) = w αi αj ≥ min w(αi αj ) = 2m1
i6=j
..
.
X
v(an−s1 ) = w αi1 ...αis = min w(αi1 · · · αis1 ) = s1 m1 .
1
i1 6=...6=is1
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
until we reach
v(αn−s1 −s2 ) = s1 m1 + (s2 − s1 )m2 .
We keep going on.
We draw the Newton polygon.
(n, 0)
(n − s1 , s1 m1 )
···
(n − s1 − s2 , s1 m1 + (s2 − s1 )m1 )
25
2 Valued fields III Local Fields
We don’t know where exactly the other points are, but the inequalities imply
that the (i, v(ai )) are above the lines drawn. So this is the Newton polygon.
Counting from the right, the first line segment has length n − (n − s1 ) = s1
and slope
0 − s1 m1
= −m1 .
n − (n − s1 )
In general, the kth segment has length (n − sk−1 ) − (n − sk ) = sk − sk−1 , and
slope
Pk−2 Pk−1
s1 m1 + i=1 (si+1 − si )mi+1 − s1 m1 + i=1 (si+1 − si )mi+1
sk − sk−1
−(sk − sk−1 )mk
= = −mk .
sk − sk−1
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.
26
3 Discretely valued fields III Local Fields
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
(∞ )
X
n
Ok((T )) = k[[T ]] = an T : an ∈ k .
n=0
Here T is a uniformizer.
27
3 Discretely valued fields III Local Fields
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 ⊆ OK
be a set of coset representatives of Ok /mk = kK containing 0. Then
(i) The non-zero ideals of OK are π n OK for n ≥ 0.
(ii) The ring OK is a PID with unique prime π (up to units), and mK = πOK .
(iii) The topology on OK induced by the absolute value is the π-adic topology.
where an ∈ S, and
|x| = |π|− inf{n:an 6=0} .
Ok ∼ OK̂
π n Ok π n OK̂
is an isomorphism.
Proof. The same as for Qp and Zp , with π instead of p.
28
3 Discretely valued fields III Local Fields
Now the valuation ring OK 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 ∼
= OK 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
(
n x 6= 0
v(x) =
∞ x=0
v(π n u) = n ∈ Z≥0 ⇐⇒ π n u ∈ R.
So we have R = OK .
Now recall our two “standard” examples of valued fields — Fp ((T )) and
Qp . Both of their residue fields are Fp , and in particular has characteristic p.
However, Fp ((T )) itself is also of characteristic p, while Qp 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 kK . Then K has equal characteristic if
char K = char kK .
29
3 Discretely valued fields III Local Fields
x = a0 + a1 p + a2 p2 + · · · ,
[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 (Teichmüller map). The map [−] : R/xR → R is called the Te-
ichm uller map. [x] is called the Teichmüller 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 pth root of a to R, and then take the pth power
of it. We claim that this is a better way of picking a lift. Suppose we have picked
−1
two lifts of ap , say, α1 and α10 . Then α10 = xc + α1 for some c. So we have
where we abuse notation and write O(x2 ) to mean terms that are multiples of
x2 .
We now recall that R/xR has characteristic p, so p ∈ xR. Thus in fact
pxc = O(x2 ). So we have
So while the lift is still arbitrary, any two arbitrary choices can differ by at most
x2 . Alternatively, our lift is now a well-defined element of R/x2 R.
We can, of course, do better. We can lift the p2 th root of a to R, then take
the p2 th power of it. Now any two lifts can differ by at most O(x3 ). More
generally, we can try to lift the pn th root of a, then take the pn th power of
30
3 Discretely valued fields III Local Fields
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 xk (†)
Then we have
αp = β p mod xk+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
p−1
p p
X p
α −β + αp−i β i ∈ xp(k+1) R.
i=1
i
We define n
βn = αnp .
The claim is that
[a] = lim βn
n→∞
31
3 Discretely valued fields III Local Fields
(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 γn ∈ R
−n −n
lifting bp for all n, then αn γn lifts (ab)p . So
n n n n
[ab] = lim αnp γnp = lim αnp lim γnp = [a][b].
−n −n −n
If R has characteristic p, then αn + γn lifts ap + bp = (a + b)p . So
pn n n
[a + b] = lim(αn + γn ) = lim αnp + lim γnp = [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
−n −n
properties. Then we note that φ(ap ) ≡ ap mod x, and is thus a valid choice
of αn . So we have
−n n
[a] = lim φ(ap )p = lim φ(a) = φ(a).
n→∞
32
3 Discretely valued fields III Local Fields
[−] : A/pA → A,
33
3 Discretely valued fields III Local Fields
a0 = a (mod p)
−1
a1 ≡ p (a − [a0 ]) (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
X X
F [an ]pn = [f (an )]pn .
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)].
−n −n
We take αn ∈ A lifting ap ∈ A/pA. Then ψ(αn ) lifts f (a)p . So
−n −n
ψ([a]) = lim ψ(αnp ) = lim ψ(αn )p = [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:
F
A B
.
f
A/pA B/xB
Indeed, the conditions on B are sufficient for Teichmüller 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 vector s of R such that W (R)/pW (R) ∼
= R.
Moreover, for any other perfect ring R, the reduction mod p map gives a
bijection
HomRing (W (R), W (R0 )) ∼ HomRing (R, R0 ) .
Proof sketch. If W (R) and W (R0 ) 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 ᾱ : W (R)/pW (R) → C/pC and its
inverse ᾱ−1 have unique lifts γ : W (R) → C and γ −1 : C → W (R), and these
are inverses by uniqueness of lifts.
34
3 Discretely valued fields III Local Fields
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= [an ]pn ∈ A,
n=0
and
∞
X
px = [an ]pn+1 ∈ J,
n=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) ∼ A
= .
pW (R) J + pA
But we know ( )
X
J + pA = [an ]pn | a0 ∈ I .
n
So we have −∞
W (R) ∼ Fp [xpr | r ∈ R] ∼
= = R.
pW (R) I
So we know that W (R) is a strict p-ring.
35
3 Discretely valued fields III Local Fields
36
3 Discretely valued fields III Local Fields
R = M + p(M + pR) = M + p2 r = · · · = M + pm R
37
4 Some p-adic analysis III Local Fields
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:
∞ ∞
X xn X xn
exp(x) = , log(1 + x) = (−1)n−1
n=0
n! n=1
n
The domain will no longer be all of the field. Instead, we have the following
result:
38
4 Some p-adic analysis III Local Fields
x
is a polynomial in x, and so defines a continuous function Zp → Q p by x →
7 n .
x
When n = 0, weset n = 1 for all x ∈ Zp .
x
x
We know n ∈ Z if x ∈ Z≥0 . So by density of Z≥0 ⊆ Zp , we must have
n ∈ Zp for all x ∈ Zp .
We will eventually want to prove the following result:
Theorem (Mahler’s theorem). Let f : Zp → Qp be any continuous function.
Then there is a unique sequence (an )n≥0 with an ∈ Qp and an → 0 such that
∞
X x
f (x) = an ,
n=0
n
and moreover
sup |f (x)| = max |ak |.
x∈Zp k∈N
kf k = sup |f (x)|p .
x∈Zp
39
4 Some p-adic analysis III Local Fields
So we have
k∆f k ≤ kf k.
In other words, we have
k∆k ≤ 1.
Definition (Mahler coefficient). Let f ∈ C(Zp , Qp ). Then nth-Mahler coeffi-
cient an (f ) ∈ Qp is defined by the formula
n
X n
an (f ) = ∆n (f )(0) = (−1)i f (n − i).
i=0
i
We will eventually show that these are the an ’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(Zp , Qp ). Then there exists some k ≥ 1 such that
k 1
k∆p f k ≤ kf k.
p
Proof. If f = 0, there is nothing to prove. So we will wlog kf k = 1 by scaling
(this is possible since the norm is attained at some x0 , so we can just divide by
f (x0 )). We want to find some k such that
k
∆p f (x) ≡ 0 mod p
for all x. To do so, we use the explicit formula
k
p k
pk
X p i
∆ f (x) = (−1) f (x + pk − i) ≡ f (x + pk ) − f (x) (mod p)
i=0
i
k
because the binomial coefficients pi are divisible by p for i 6= 0, pk . Note that
k
we do have a negative sign in front of f (x) because (−1)p is −1 as long as p is
odd, and 1 = −1 if p = 2.
Now Zp 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 ∈ Zp . 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→ (an (f ))∞
n=0 defines an injective norm-decreasing
linear map C(Zp , Qp ) → c0 .
Proof. First we prove that an (f ) → 0. We know that
kan (f )kp ≤ k∆n f k.
So it suffices to show that k∆n f k → 0. Since k∆k ≤ 1, we know k∆n f k 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 k1 , k2 , · · · such that
≤ 1 kf k.
pk1 +...+kn
∆
pn
40
4 Some p-adic analysis III Local Fields
a0 (f ) = f (0) = 0,
f (n) = ∆k f (0) = an (f ) = 0.
Lemma. We have
x x x+1
+ =
n n−1 n
for all n ∈ Z≥1 and x ∈ Zp .
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.
∞
X x
fa (x) = an .
n=0
n
kfa k ≤ kak.
41
4 Some p-adic analysis III Local Fields
Then we have
Iterating, we have
∆k fa = fa(k) .
So we have
an (fa ) = ∆n fa (0) = fa(n) (0) = an .
with
F (f ) = (an (f ))
G(a) = fa .
F (v − GF v) = F v − F GF v = (F − F )v = 0.
42
5 Ramification theory for local fields III Local Fields
[L : K] = eL/K fL/K .
43
5 Ramification theory for local fields III Local Fields
Increasing r and decreasing s if necessary, we wlog r = |a| and s = |b| for some
a, b ∈ K.
Then we can write
n
M n
M
M= Ok bαi ⊆ OL ⊆ N = OK aαi .
i=1 i=1
[L : K] = eL/K fL/K ,
44
5 Ramification theory for local fields III Local Fields
Proof. We will be lazy and write e = eL/K and f = fL/K . We first note that
kL /kK is separable, so there is some ᾱ ∈ kL such that kL = kK (ᾱ) by the
primitive element theorem. Let
f¯(x) ∈ kK [x]
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 sj 6≡ 0 mod πL ,
vL (a−1
kj sj ) = 0.
So we must have
45
5 Ramification theory for local fields III Local Fields
Now we write
X e−1
X
aij αi π j = sj π j = 0.
j=0
j
6 0, then we have vL (sj π ) = vL (sj ) + j ∈ j + eZ. So no two non-zero
If sj =
Pe−1 Pe−1
terms in j=0 sj π j have the same valuation. This implies that j=0 sj π j 6= 0,
which is a contradiction.
We now want to prove that
M
OL = OK αi π j .
i,j
We let M
M= OK αi π j ,
i,j
and put
f −1
M
N= OL α i .
i=0
Then we have
M = N + πN + π 2 N + · · · + π e−1 N.
We are now going to use the fact that 1, ᾱ, · · · , ᾱf −1 span kL over kK . So we
must have that OL = N + πOL . We iterate this to obtain
OL = N + π(N + OL )
= N + πN + π 2 OL
= ···
= N + πN + π 2 N + · · · + π e−1 N + π n OL
= M + πK OL ,
using the fact that πK and π e have the same valuation, and thus they differ by
a unit in OL . Iterating this again, we have
OL = M + πkn OL
for all n ≥ 1. So M is dense in OL . But M is the closed unit ball in the subspace
M
Kαi π j ⊆ l
i,j
with respect to the maximum norm with respect to the given basis. So it must
be complete, and thus M = OL .
Finally, since αi π j = αi f (α)j is a polynomial in α, we know that OL =
OK [α].
Corollary. If M/L/K is a tower of finite extensions of local fields, then
fM/K = fL/K fM/L
eM/K = eL/K eM/L
Proof. The multiplicativity of fM/K follows from the tower law for the residue
fields, and the multiplicativity of eM/K follows from the tower law for the local
fields and that fM/K eM/K = [M : K].
46
5 Ramification theory for local fields III Local Fields
47
5 Ramification theory for local fields III Local Fields
using the fact that πkn 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 ϕ̄ : kL ∼= kM over kK between unramified extensions.
Then ϕ̄ 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
So L/K is Galois. Moreover, the map AutK (L) → AutkK (kL ) 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 K̄ 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 ᾱ be a primitive element of kK /kL , and f¯ ∈ kK [x] a minimal polyno-
mial of ᾱ, and f ∈ Ok [x] a monic lift of f¯, and α ∈ OL a unique lift of f lifting
ᾱ. Then L = K(α). So LM = M (α).
Let ḡ be the minimal polynomial of ᾱ over kM . Then ḡ | f¯. By Hensel’s
lemma, we can factorize f = gh in OM [x], where g is monic and lifts ḡ. 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
48
5 Ramification theory for local fields III Local Fields
e−1
L/K = w(πL ) = min{w(x) : x ∈ mL },
Proof. We know w and vL differ by a constant. To figure out what this is, we
have
1 = w(πK ) = e−1
L/K vL (πK ).
w(x) = e−1
L/K vL (x).
w(πL ) = e−1 −1
L/K vL (πL ) = eL/K .
The equality
w(πL ) = min{w(x) : x ∈ mL },
is trivially true because the minimum is attained by πL .
Definition (Eisenstein polynomial). A polynomial f (x) ∈ OK [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(πL ) and the minimal polynomial of πL 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], vK be the valuation of K, and w the unique extension
to L. Then
1
[K(πL ) : K]−1 ≤ e−1
K(πL )/K = min w(c) ≤ ,
x∈mK(πL ) n
where the last inequality follows from the fact that πL ∈ mL(πL ) .
But we also know that
[K(πL ) : K] ≤ [L : K].
49
5 Ramification theory for local fields III Local Fields
So we have
n n−1 i
1= w(πL ) = w(a0 + a1 πL + · · · + an−1 πL ) = min vk (ai ) + .
i=0,...,n−1 n
1
e−1
L/K = min w(x) ≤ = [L : K]−1 .
x∈mL n
So [L : K] = eL/K = n. So L/K is totally ramified and α is a uniformizer.
50
6 Further ramification theory III Local Fields
UK /UK ∼
(1) ×
= (kK , · ),
(s) (s+1) ∼
U /U
K = (kK , +).
K
for s ≥ 1.
× ×
Proof. We have a surjective homomorphism OK → kK which is just reduction
(1)
mod πK , and the kernel is just things that are 1 modulo πK , i.e. UK . So this
gives the first part.
(s)
For the second part, we define a surjection UK → kK given by
s
1 + πK x 7→ x mod πk .
This is a group homomorphism because
s s
(1 + πK x)(1 + πK y) = 1 + π S (x + y + π s xy),
and this gets mapped to
x + y + πs x + y ∼
=x+y mod πK .
(s+1)
Then almost by definition, the kernel is UK .
51
6 Further ramification theory III Local Fields
∼
Gal(T /K) Gal(kT /kK ).
52
6 Further ramification theory III Local Fields
f : x 7→ σ −1 ([σ̄(x)]).
σ −1 ([σ̄(x)]) ≡ x mod πL .
given by
σ(πL )
σ 7→
πL
is a well-defined injective group homomorphism, independent of the choice of
πL .
Proof. We define the map
(s)
UL
φ : Gs (L/K) → (s+1)
UL
σ 7→ σ(πL )/πL .
53
6 Further ramification theory III Local Fields
So we trivially have Gs+1 (L/K) ⊆ ker φ. To show the converse, let x ∈ OL and
write
X∞
n
x= [xn ]πL .
n=0
So we know vL (σ(x) − x) ≥ s + 2.
Corollary. Gal(L/K) is solvable.
Proof. Note that \
Gs (L/K) = {id}.
s
|Gs (L/K)|
|Gs+1 (L/K)|
54
6 Further ramification theory III Local Fields
is also a pth power. However, we know that the intersection of all Gs (L/K)
is {id}, and also Gal(L/K) is finite. So for sufficiently large t, we know that
|Gt (L/K)| = 1. So we conclude that
Proposition. G1 (L/K) is always a p-group.
We now use the injection
G0 (L/K) ×
,→ kL ,
G1 (L/K)
×
and the fact that kL has order prime to p. So G1 (L/K) must be the Sylow
p-subgroup of G0 (L/K). Since it is normal, it must be the unique p-subgroup.
Definition (Wild inertia group and tame quotient). G1 (L/K) is called the wild
inertia group, and the quotient G0 (L/K)/G1 (L/K) is the tame quotient.
This is trivial, because the definition uses the valuation vM 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 ηM/L such that
Gs (M/K)
Gt (L/K) ∼
=
Gs (M/L)
55
6 Further ramification theory III Local Fields
This is not very helpful. We now claim that we can compute iL/K using the
following formula:
Proposition. Let L/K be a finite Galois extension of local fields, and pick
α ∈ OL such that OL = OK [α]. Then
Proof. Fix a σ. It is clear that iL/K (σ) ≤ vL (σ(α) − α). Conversely, for any
x ∈ OL , we can find a polynomial g ∈ OK [t] such that
X
x = g(α) = bi α i ,
Here eM/L is just to account for the difference between vL and vM . So the
real content is that the value of iL/K (σ) is the sum of the values of iM/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 OM = OL [α] and OL = OK [β], where α ∈ OM and
β ∈ OL . Then we have
iM/K (τ ) = vM (τ α − α)
56
6 Further ramification theory III Local Fields
We let
b = σ(β) − β = τ (β) − β
and Y
a= (τ g(α) − α).
g∈H
Then we know
n
X
τ (z) − z = zi (τ (β)i − β i )
i=1
is divisible by τ (β) − β = b.
Now let F (x) ∈ OL [x] be the minimal polynomial of α over L. Then explicitly,
we have Y
F (x) = (x − g(α)).
g∈H
Then we have Y
(τ F )(x) = (x − τ g(α)),
g∈H
So b | a.
In other direction, we pick f ∈ OK [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 OL [x]. So we have
f (x) − β = F (x)h(x)
for some h ∈ OL [x].
Then noting that f has coefficients in OK , we have
−b = β − τ β = ±a(τ h)(α).
So a | b.
57
6 Further ramification theory III Local Fields
Now that we understand how the iL/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 ηL/K : [−1, ∞) → [−1, ∞):
!
X
−1
ηL/K (s) = eL/K min(iL/K (σ), s + 1) − 1.
σ∈G
Then we have
Gs (M/K)H
= Gt (L/K).
H
By some isomorphism theorem, and the fact that H ∩ Gs (M/K) = Gs (M/L),
this is equivalent to saying
Gs (M/K)
Gt (L/K) ∼
= .
Gs (M/L)
iM/K (τ ) ≥ iM/K (τ g)
58
6 Further ramification theory III Local Fields
59
6 Further ramification theory III Local Fields
for t ∈ [−1, ∞). The original number is called the lower numbering.
Since ηM/K and ηL/K ◦ ηM/L agree at 0 (they both take value 0), we know that
the functions must agree everywhere. So done.
60
6 Further ramification theory III Local Fields
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 ζpn a primitive pn th root of unity, and K = Qp (ζpn ). The
minimal polynomial of ζpn is the pn th cyclotomic polynomial
n−1 n−1
Φpn (x) = xp (p−1)
+ xp (p−2)
+ · · · + 1.
π = ζpn − 1
is a uniformizer. So we know
OK = Zp [ζpn − 1] = Zp [ζpn ].
σm (ζpn ) = ζpmn .
We have
since ζpn is a unit. If m = 1, then this thing is infinity. If it is not 1, then ζpm−1
n
61
6 Further ramification theory III Local Fields
Thus we have
vK (ζpm−1
n − 1) ≥ pk ⇔ m ≡ 1 mod pk .
It then follows that for
pk ≥ s + 1 ≥ pk−1 + 1,
we have
Gs (K/Qp ) ∼
= {m ∈ (Z/pn )× : m ≡ 1 mod pk }.
Now m ≡ 1 mod pk iff σm (ζpk ) = ζpk . So in fact
Gs (K/Qp ) ∼
= Gal(K/Qp (ζpk )).
Gs (K/Qp ) = 1.
We claim that
ηK/Qp (pk − 1) = k.
So we have
Gk (K/Qp ) = Gal(K/Qp (ζpk )).
This actually looks much nicer!
To actually compute ηK/Qp , we have notice that the function we integrate to
get η looks something like this (not to scale):
1
p−1
1
p(p−1)
1
p2 (p−1)
p−1 p2 − 1 p3 − 1
62
7 Local class field theory III Local Fields
63
7 Local class field theory III Local Fields
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.
fij : Gj → Gi
fii = idGi
and
fik = fij ◦ fjk
whenever i ≤ j ≤ k.
We define the inverse limit on the system (Gi , fij ) to be
( )
Y Y
lim Gi = (gi ) ∈ Gi : fij (gj ) = gi for all i ≤ j ⊆ gi ,
←−
i∈I i∈I i∈I
64
7 Local class field theory III Local Fields
Then we have \
Gal(M/L) = Gal(M/L0 ),
0
L ⊆L
L0 /K finite
and each Gal(M/L0 ) is open, hence closed. So the whole thing is closed.
65
7 Local class field theory III Local Fields
The left hand map is an isomorphism by (infinite) Galois theory, and since all
finite subextensions of kM /kK are of the form kL /kK 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 = TM/K /K.
By general field theory, we know that Gal(kL /kK ) is a cyclic group generated by
FrobL/K : x 7→ xq ,
66
7 Local class field theory III Local Fields
To prove the second part, we again let L/K be a finite subextension. Then
L · TM/K ⊆ TM/L . We then have maps
∼
=
FrobZTM/K /K Gal(TM/K /K) Gal(kM /kK )
∼
=
FrobZTM/L /L Gal(TM/L /L) Gal(kM /kL )
67
7 Local class field theory III Local Fields
68
7 Local class field theory III Local Fields
K× M ∩ K ab
→ Gal .
NM/K (M × ) K
To simplify this, we will write N (L/K) = NL/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]
K× ∼ K×
= Gal(M/K) ∼
= .
N (L/K) N (M/K)
N (L/K) L/K
69
7 Local class field theory III Local Fields
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 = idM . 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) ∼
= ArtK (N (L/K)) ⊆ ArtK (N (M/K)).
So we can find x ∈ M × such that σ = ArtK (NM/K (x)). So σ|M = idM 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.
70
8 Lubin–Tate theory III Local Fields
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 ArtK .
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.
So we have a surjection
K× vK Z
.
N (L/K) fL/K Z
K× ∼ ×
= hπK i × OK .
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
m (n)
hπK i × UK .
71
8 Lubin–Tate theory III Local Fields
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)
∼
= FrobZK × 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 TK ur /K = K ur trivially, and then by
definition W (K ur/K ) ∼
= FrobZK .
Example. We consider the special case of K = Qp and πK = p. We let
Ln = Qp (ζpn ),
Then again this is totally ramified extension, since it is the nested union of
totally ramified extensions.
72
8 Lubin–Tate theory III Local Fields
Then we have
Note that we are a bit sloppy in this deduction. While we know that it is true
that Z× ∼ (Z/pn Z)× , the inverse limit depends not only on the groups
p = lim
←−n
n ×
(Z/p 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
m = a0 + a1 p + · · · ,
∼
= ∼ restriction
hpi × Z×
p W (Qur
p /Qp ) × Gal(Qp (ζp∞ )/Qp )
73
8 Lubin–Tate theory III Local Fields
In fact, we have
Theorem (Local Kronecker-Weber theorem).
[
Qab
p = Qp (ζn ),
n∈Z≥1
[
Qur
p = Qp (ζn ).
n∈Z≥1
(n,p)=1
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, ArtQp (m)|Qp (ζp∞ ) is σm .
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
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
Gs (M/K) ∼
= lim Gs (L/K),
←−
L/K
which corresponds to
hpi × U (0)
s = −1
hpn i × U (m)
n (0)
hp i × U
n (m)
−1 < s ≤ 0
hp i × U
hpn i × U (k)
k−1<s≤k ≤m−1
hpn i × U (m)
1 s>m−1
74
8 Lubin–Tate theory III Local Fields
Gs (Qab k
p /Qp ) = ArtQp (1 + p Zp ) = ArtQp (U
(k)
),
N (L/Qp )(1 + pn Zp )
Gs (L/Qp ) = ArtQp ,
N (L/Qp )
where n − 1 < s ≤ n.
Here ArtQp induces an isomorphism
Q×
p
→ Gal(L/Qp ).
N (L/Qp )
So it follows that L ⊆ Qp (ζpm ) for some n if and only if Gs (L/Qp ) = 1 for all
s > m − 1.
75
8 Lubin–Tate theory III Local Fields
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 OK , where K is a complete valued field,
then F (x, y) converges for all x, y ∈ mK . So mK becomes a (semi)group under
the multiplication
(x, y) 7→ F (x, y) ∈ mk
Example. We can define
Ĝa (X, Y ) = X + Y.
This is called the formal additive group.
Similarly, we can have
Ĝm (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 mK bijects with 1 + mk by sending
x 7→ 1 + x, and the rule sending (x, y) ∈ m2K 7→ x + y + xy ∈ mK is just the
usual multiplication in 1 + mK transported to mK 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 EndR (F ) with addition +F given by
(f +F g)(x) = F (f (x), g(x)).
and multiplication is given by composition.
76
8 Lubin–Tate theory III Local Fields
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 → EndR (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 = |kK |. We let π ∈ OK be a uniformizer.
Definition (Lubin–Tate module). A Lubin–Tate module over OK with respect
to π is a formal OK -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 Ĝm is a Lubin–Tate Zp module with respect to p
given by the following formula: if a ∈ Zp , then we define
∞
X a
[a]Ĝm (X) = (1 + X)a − 1 = X n.
n=1
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) ∈ OK [[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 Eπ .
Now by definition, if F is a Lubin–Tate OK 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(aq−1 X q−1 + · · · + a2 X 2 ) + πX
(1)
with u ∈ UK , and aq−1 , · · · , a2 ∈ OK .
In particular, these are Lubin–Tate series.
77
8 Lubin–Tate theory III Local Fields
and
e1 (F (x1 , · · · , xn )) = F (e2 (x1 ), e2 (x2 ), · · · , e2 (xn )).
For reasons of time, we will not prove this. We just build F by successive
approximation, which is not terribly enlightening.
Corollary. Let e ∈ Eπ be a Lubin–Tate series. Then there are unique power
series Fe (X, Y ) ∈ OK [[X, Y ]] such that
Fe (X, Y ) ≡ X + Y mod (X + Y )2
e(Fe (X, Y )) = Fe (e(X), e(Y ))
a 7→ [a]e .
78
8 Lubin–Tate theory III Local Fields
79
8 Lubin–Tate theory III Local Fields
fn (X) = (e ◦ · · · ◦ e)(X).
| {z }
n times
0
q
fn+1 (x) = (qfn (x)q−1 + π)fn0 (x) = π 1 + fn (x)q−1 fn0 (x).
π
By induction hypothesis, we know fn0 (x) 6= 0, and by assumption |fn (x)| <
1. So the same argument works.
We now prove the lemma. We assume that fn (x) = 0. We want to show that
|fi (x)| < 1 for all i = 0, · · · , n − 1. By induction, we have
n
fn (x) = xq + πgn (x)
for some gn (x) ∈ OK [x]. It follows that if fn (x) = 0, then |x| < 1. So |fi (x)| < 1
for all i. So fn0 (x) 6= 0.
The point of the lemma is to prove the following proposition:
Proposition. F (n) is a free OK /π n OK module of rank 1. In particular, it has
q n elements.
Proof. By definition, we know
π n · F (n) = 0.
OK → F (n)
80
8 Lubin–Tate theory III Local Fields
Ln ⊆ Ln+1 .
The field Ln depends only in π, and not on F . To see this, we let G be another
Lubin–Tate OK -module, and let f : F → G be an isomorphism. Then
Moreover, if F = Fe , where
81
8 Lubin–Tate theory III Local Fields
∼ (n)
Gal(Ln /K) UK /UK
commutes. So the isomorphism
(m)
UK
Gal(Lm /Ln ) ∼
= (n)
UK
follows by looking at the kernels.
82
8 Lubin–Tate theory III Local Fields
∼ ∼ (n) ∼
Gal(L∞ /K) lim Gal(Ln /K) limn UK /UK = UK
←− ←−
σ (σ|Ln )n
83
8 Lubin–Tate theory III Local Fields
Gal(Ln /Lk ) ∼
(k) (n)
= UK /UK
G1 (Ln /K) ∼
(1) (n)
= UK /UK .
So we know that G1 (Ln /K) = Gal(Ln /L1 ). Thus we know that Gs (Ln /K) =
Gal(Ln /K) for 0 < s ≤ 1.
(1) (n)
We now let σ = σu ∈ G1 (Ln /K) and u ∈ UK /UK . We write
u = 1 + επ k
for some ε ∈ UK and some k = k(u) ≥ 1. Since σ is not the identity, we know
k < n. We claim that
iLn /K (σ) = vLn (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 Ln and OLn = OK [λ]. We can compute
σu (λ) = [u]F (λ)
= [1 + επ k ]F (λ)
= F (λ, [επ k ]F (λ))
84
8 Lubin–Tate theory III Local Fields
where we set L0 = K.
Once again, the numbering is a bit more civilized in the upper numbering.
Proof. We have to compute the integral of
1
.
(G0 (Ln /K) : Gx (Ln /K)
We again plot this out
85
8 Lubin–Tate theory III Local Fields
1
q−1
1
q(q−1)
1
q 2 (q−1)
q−1 q2 − 1 q3 − 1
So by the same computation as the ones we did last time, we find that
s
−1 ≤ s ≤ 0
s−(q k−1 −1)
ηLn /K (s) = (k − 1) + q k−1 (q−1)
q k−1 − 1 ≤ s ≤ q k − 1, k = 1, · · · , n − 1
s−(q n−1 −1)
(n − 1) + n−1
− 1.
q n−1 (q−1) s>q
86
8 Lubin–Tate theory III Local Fields
So we have
Gt (K ab /K) = Gt (K ur L∞ /K)
= lim Gt (Km Ln /K)
←−
m,n
= Gal(K ur L∞ /K ur Ldte )
= Gal(K ab /K ur Ldte ),
and
Art−1 ab ur −1
K (Gal(K /K Ldte )) = ArtK lim Gal(Km Ln /Km Ldte )
←−
m,n
n≥dte
= lim Art−1
← − K Gal(Km Ln /Km Ldte )
m,n
n≥dte
(dte)
U
= lim K(n)
←−
m,n UK
n≥dte
= U dte .
Proof. We have
!
(dte)
t Gt (K ab /K)G(K ab /M ) UK N (M/K)
G (M/K) = = Art .
G(K ab /M ) N (M/K)
87
Index III Local Fields
Index
C(Zp , Qp ), 39 DVR, 29
I-adic completion, 11
I-adic topology, 9 Eisenstein criterion, 49
I-adically complete, 11 Eisenstein polynomial, 49
I-adically open, 9 equal characteristic, 29
K ur , 66 equivalence of norm, 19
L/K, 52 equivalent absolute values, 4
N (L/K), 69
(s) formal additive group, 76
Uk , 51
formal group, 75
W (M/K), 66
homomorphism, 76
W (R), 34
module, 77
ηL/K , 58
formal Laurent series, 15
OK , 15
formal module, 77
mK , 16
formal multiplicative group, 76
π n -division points, 79
fundamental theorem of Galois
ψL/K , 60
theory, 65
c0 , 39
iL/K , 55 Galois extension, 63
kK , 16 generalized local Kronecker-Weber
p-adic absolute value, 11 theorem, 83
p-adic integers, 12 geometric Frobenius, 74
p-adic numbers, 12
sth ramification group, 52 Hasse-Arf theorem, 83
x-adic topology, 9 Hensel’s lemma, 16
Herbrand’s theorem, 58
abelian extension, 68 Higher ramification groups, 74
absolute ramification index, 36 higher unit groups, 51
absolute value, 4 homomorphism
archimedean, 5 formal group, 76
non-archimedean, 5
absolute values inertia degree, 43
equivalence, 4 inertia field, 52
adjoint matrix, 7 inertia group, 52
adjugate matrix, 7 integral element, 7
algebraic integer, 7 integrally closed, 8
archimedean absolute value, 5 inverse limit, 10, 64
arithmetic Frobenius, 66 inverse limit topology, 10
inverse system, 64
break points, 23
Krull topology, 63
class field, 70
Coleman operators, 83 length, 24
line segment, 24
directed system, 64 local field, 27
discrete valuation ring, 29 Local Kronecker-Weber theorem, 74
discretely valued field, 27 local Kronecker-Weber theorem, 83
DVF, 27 lower convex hull, 23
88
Index III Local Fields
89