Part II Galois Theory
Based on lectures by C. Birkar
Notes taken by Dexter Chua
Michaelmas 2015
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.
Groups, Rings and Modules is essential
Field extensions, tower law, algebraic extensions; irreducible polynomials and relation
with simple algebraic extensions. Finite multiplicative subgroups of a field are cyclic.
Existence and uniqueness of splitting fields. [6]
Existence and uniqueness of algebraic closure. [1]
Separability. Theorem of primitive element. Trace and norm. [3]
Normal and Galois extensions, automorphic groups. Fundamental theorem of Galois
theory. [3]
Galois theory of finite fields. Reduction mod p. [2]
Cyclotomic polynomials, Kummer theory, cyclic extensions. Symmetric functions.
Galois theory of cubics and quartics. [4]
Solubility by radicals. Insolubility of general quintic equations and other classical
problems. [3]
Artin’s theorem on the subfield fixed by a finite group of automorphisms. Polynomial
invariants of a finite group; examples. [2]
Contents
0 Introduction
1 Solving equations
2 Field extensions
2.1 Field extensions
2.2 Ruler and compass constructions
2.3 K-homomorphisms and the Galois Group
2.4 Splitting fields
2.5 Algebraic closures
2.6 Separable extensions
2.7 Normal extensions
2.8 The fundamental theorem of Galois theory
2.9 Finite fields
3 Solutions to polynomial equations
3.1 Cyclotomic extensions
3.2 Kummer extensions
3.3 Radical extensions
3.4 Solubility of groups, extensions and polynomials
3.5 Insolubility of general equations of degree 5 or more
4 Computational techniques
4.1 Reduction mod p
4.2 Trace, norm and discriminant
0 Introduction
The most famous result of Galois theory is that there is no general solution
to polynomial equations of degree 5 or above in terms of radicals. However,
this result was, in fact, proven before Galois theory existed, and goes under the
name of the Abel–Ruffini theorem. What Galois theory does provides is a way
to decide whether a given polynomial has a solution in terms of radicals, as well
as a nice way to prove this result.
However, Galois theory is more than equation solving. In fact, the funda-
mental theorem of Galois theory, which is obviously an important theorem in
Galois theory, has completely nothing to do with equation solving. Instead, it is
about group theory.
In modern days, Galois theory is often said to be the study of field extensions.
The idea is that we have a field
K
, and then add more elements to get a field
L
.
When we want to study solutions to polynomial equations, what we add is the
roots of the polynomials. We then study the properties of this field extension,
and in some cases, show that this field extension cannot be obtained by just
adding radicals.
For certain “nice” field extensions
K L
, we can assign to it the Galois group
Gal
(
L/K
). In general, given any group
G
, we can find subgroups of
G
. On the
other hand, given a field extension
K L
, we can try to find some intermediate
field
F
that can be fitted into
K F L
. The key idea of Galois theory is
that these two processes are closely related we can establish a one-to-one
correspondence between the subgroups of G and the intermediate fields F .
Moreover, many properties of (intermediate) field extensions correspond to
analogous ideas in group theory. For example, we have the notion of normal
subgroups, and hence there is an analogous notion of normal extensions. Similarly,
we have soluble extensions (i.e. extensions that can be obtained by adding
radicals), and these correspond to “soluble groups”. In Galois theory, we will
study how group-theoretic notions and field-theoretic notions interact.
Nowadays, Galois theory is an important field in mathematics, and finds its
applications in number theory, algebraic geometry and even cryptography.
1 Solving equations
Galois theory grew of the desire to solve equations. In particular, to solve
polynomial equations. To begin with, we will come up with general solutions to
polynomial equations of up to degree 4. However, this is the best we can do, as
we will later show in the course there is no general solution to polynomial
equations of degree 5 or above.
Before we start, we will define some notations that we will frequently use.
If
R
is a ring, then
R
[
t
] is the polynomial ring over
R
in the variable
t
.
Usually, we take
R
=
Q
and consider polynomials
f
(
t
)
Q
[
t
]. The objective
is then to find roots to the equation
f
(
t
) = 0. Often, we want to restrict our
search domain. For example, we might ask if there is a root in
Q
. We will thus
use Root
f
(X) to denote the set of all roots of f in X.
Linear equations
Suppose that
f
=
t
+
a Q
[
t
] (with
a Q
). This is easy to solve we have
Root
f
(Q) = {−a}.
Quadratic equations
Consider a simple quadratic
f
=
t
2
+ 1
Q
[
t
]. Then
Root
f
(
Q
) =
since the
square of all rationals are positive. However, in the complex plane, we have
Root
f
(C) = {
1,
1}.
In general, let
f
=
t
2
+
at
+
b Q
[
t
]. Then as we all know, the roots are
given by
Root
f
(C) =
(
a ±
a
2
4b
2
)
Cubic equations
Let f = t
3
+ c Q[t]. The roots are then
Root
f
(C) = {
3
c, µ
3
c, µ
2
3
c},
where
µ
=
1+
3
2
is the 3rd root of unity. Note that
µ
is defined by the
equation µ
3
1 = 0, and satisfies µ
2
+ µ + 1 = 0.
In general, let
f
=
t
3
+
at
2
+
bt
+
c Q
[
t
], and let
Root
f
(
C
) =
{α
1
, α
2
, α
3
}
,
not necessarily distinct.
Our objective is to solve
f
= 0. Before doing so, we have to make it explicit
what we mean by “solving” the equation. As in solving the quadratic, we want
to express the roots α
1
, α
2
and α
3
in terms of “radicals” involving a, b and c.
Unlike the quadratic case, there is no straightforward means of coming up
with a general formula. The result we currently have is the result of many
many years of hard work, and the substitutions we make seemingly come out of
nowhere. However, after a lot of magic, we will indeed come up with a general
formula for it.
We first simplify our polynomial by assuming
a
= 0. Given any polynomial
f
=
t
3
+
at
2
+
bt
+
c
, we know
a
is the negative of the sum of the roots. So we
can increase each root by
a
3
so that the coefficient of
t
2
vanishes. So we perform
the change of variables
t 7→ t
a
3
, and get rid of the coefficient of
t
2
. So we can
assume a = 0.
Let µ be as above. Define
β = α
1
+ µα
2
+ µ
2
α
3
γ = α
1
+ µ
2
α
2
+ µα
3
These are the Lagrange resolvers. We obtain
βγ = α
2
1
+ α
2
2
+ α
2
3
+ (µ + µ
2
)(α
1
α
2
+ α
2
α
3
+ α
1
α
3
)
Since µ
2
+ µ + 1 = 0, we have µ
2
+ µ = 1. So we can simplify to obtain
= (α
1
+ α
2
+ α
3
)
2
3(α
1
α
2
+ α
2
α
3
+ α
1
α
3
)
We have α
1
+ α
2
+ α
3
= a = 0, while b = α
1
α
2
+ α
2
α
3
+ α
1
α
3
. So
= 3b
Cubing, we obtain
β
3
γ
3
= 27b
3
.
On the other hand, recalling again that α
1
+ α
2
+ α
3
= 0, we have
β
3
+ γ
3
= (α
1
+ µα
2
+ µ
2
α
3
)
3
+ (α
1
+ µ
2
α
2
+ µα
3
)
3
+ (α
1
+ α
2
+ α
3
)
3
= 3(α
3
1
+ α
3
2
+ α
3
3
) + 18α
1
α
2
α
3
We have
α
1
α
2
α
3
=
c
, and since
α
3
i
+
i
+
c
= 0 for all
i
, summing gives
α
3
1
+ α
3
2
+ α
3
3
+ 3c = 0. So
= 27c
Hence, we obtain
(t β
3
)(t γ
3
) = t
2
+ 27ct 27b
3
.
We already know how to solve this equation using the quadratic formula. We
obtain
{β
3
, γ
3
} =
(
27c ±
p
(27c)
2
+ 4 × 27b
3
2
)
We now have
β
3
and
γ
3
in terms of radicals. So we can find
β
and
γ
in terms of
radicals. Finally, we can solve for α
i
using
0 = α
1
+ α
2
+ α
3
β = α
1
+ µα
2
+ µ
2
α
3
γ = α
1
+ µ
2
α
2
+ µα
3
In particular, we obtain
α
1
=
1
3
(β + γ)
α
2
=
1
3
(µ
2
β + µγ)
α
3
=
1
3
(µβ + µ
2
γ)
So we can solve a cubic in terms of radicals.
This was a lot of magic involved, and indeed this was discovered through a
lot of hard work throughout many many years. This is also not a very helpful
result since we have no idea where these substitutions came from and why they
intuitively work.
Quartic equations
Assume
f
=
t
4
+
at
3
+
bt
2
+
ct
+
d Q
[
t
]. Let
Root
f
(
C
) =
{α
1
, α
2
, α
3
, α
4
}
.
Can we express all these in terms of radicals? Again the answer is yes, but the
procedure is much more complicated.
We can perform a similar change of variable to assume a = 0. So α
1
+ α
2
+
α
3
+ α
4
= 0.
This time, define
β = α
1
+ α
2
γ = α
1
+ α
3
λ = α
1
+ α
4
Doing some calculations, we see that
β
2
= (α
1
+ α
2
)(α
3
+ α
4
)
γ
2
= (α
1
+ α
3
)(α
2
+ α
4
)
λ
2
= (α
1
+ α
4
)(α
2
+ α
3
)
Now consider
g = (t β
2
)(t γ
2
)(t λ
2
)
= t
3
+ 2bt
2
+ (b
2
4d)t c
2
This we know how to solve, and so we are done.
Quintics and above
So far so good. But how about polynomials of higher degrees? In general, let
f Q
[
t
]. Can we write down all the roots of
f
in terms of radicals? We know
that the answer is yes if deg f 4.
Unfortunately, for
deg f
5, the answer is no. Of course, this “no” means
no in general. For example,
f
= (
t
1)(
t
2)
···
(
t
5)
Q
[
t
] has the obvious
roots in terms of radicals.
There isn’t an easy proof of this result. The general idea is to first associate
a field extension
F Q
for our polynomial
f
. This field
F
will be obtained
by adding all roots of
f
. Then we associate a Galois group
G
to this field
extension. We will then prove a theorem that says
f
has a solution in terms
of radicals if and only if the Galois group is “soluble”, where “soluble” has a
specific algebraic definition in group theory we will explore later. Finally, we
find specific polynomials whose Galois group is not soluble.
2 Field extensions
After all that (hopefully) fun introduction and motivation, we will now start
Galois theory in a more abstract way. The modern approach is to describe these
in terms of field extensions.
2.1 Field extensions
Definition (Field extension). A field extension is an inclusion of a field
K L
,
where
K
inherits the algebraic operations from
L
. We also write this as
L/K
.
Alternatively, we can define this by a injective homomorphism
K L
. We say
L is an extension of K, and K is a subfield of L.
Example.
(i) R/Q is a field extension.
(ii) C/Q is a field extension.
(iii) Q(
2) = {a + b
2 : a, b Q} R is a field extension over Q.
Given a field extension
L/K
, we want to quantify how much “bigger”
L
is compared to
K
. For example, to get from
Q
to
R
, we need to add a lot of
elements (since
Q
is countable and
R
is uncountable). On the other hand, to get
from R to C, we just need to add a single element
1.
To do so, we can consider
L
as a vector space over
K
. We know that
L
already comes with an additive abelian group structure, and we can define scalar
multiplication by simply multiplying: if
a K, α L
, then
a · α
is defined as
multiplication in L.
Definition (Degree of field extension). The degree of
L
over
K
is [
L
:
K
] is the
dimension of
L
as a vector space over
K
. The extension is finite if the degree is
finite.
In this course, we are mostly concerned with finite extensions.
Example.
(i)
Consider
C/R
. This is a finite extension with degree [
C
:
R
] = 2 since we
have a basis of {1, i}.
(ii) The extension Q(
2)/Q has degree 2 since we have a basis of {1,
2}.
(iii) The extension R/Q is not finite.
We are going to use the following result a lot:
Theorem (Tower Law). Let F/L/K be field extensions. Then
[F : K] = [F : L][L : K]
Proof.
Assume [
F
:
L
] and [
L
:
K
] are finite. Let
{α
1
, ··· , α
m
}
be a basis for
L
over
K
, and
{β
1
, ··· , β
n
}
be a basis for
F
over
L
. Pick
γ F
. Then we can
write
γ =
X
i
b
i
β
i
, b
i
L.
For each b
i
, we can write as
b
i
=
X
j
a
ij
α
j
, a
ij
K.
So we can write
γ =
X
i
X
j
a
ij
α
j
β
i
=
X
i,j
a
ij
α
j
β
i
.
So
T
=
{α
j
β
i
}
i,j
spans
F
over
K
. To show that this is a basis, we have to show
that they are linearly independent. Consider the case where
γ
= 0. Then we
must have
b
i
= 0 since
{β
i
}
is a basis of
F
over
L
. Hence each
a
ij
= 0 since
{α
j
} is a basis of L over K.
This implies that T is a basis of F over K. So
[F : K] = |T | = nm = [F : L][L : K].
Finally, if [
F
:
L
] =
or [
L
:
K
] =
, then clearly [
F
:
K
] =
as well. So
equality holds as well.
Recall that in IA Numbers and Sets, we defined a real number
x
to be
algebraic if it is a root of some polynomial in integer (or rational) coefficients.
We can do this for general field (extensions) as well.
Definition (Algebraic number). Let
L/K
be a field extension,
α L
. We
define
I
α
= {f K[t] : f (α) = 0} K[t]
This is the set of polynomials for which
α
is a root. It is easy to show that
I
α
is
an ideal, since it is the kernel of the ring homomorphism
K
[
t
]
L
by
g 7→ g
(
α
).
We say
α
is algebraic over
K
if
I
α
= 0. Otherwise,
α
is transcendental over
K.
We say L is algebraic over K if every element of L is algebraic.
Example.
(i)
9
7
is algebraic over
Q
because
f
(
9
7
) = 0, where
f
=
t
9
7. In general,
any number written with radicals is algebraic over Q.
(ii) π is not algebraic over Q.
These are rather simple examples, and the following lemma will provide us a
way of generating much more examples.
Lemma. Let L/K be a finite extension. Then L is algebraic over K.
Proof.
Let
n
= [
L
:
K
], and let
α L
. Then 1
, α, α
2
, ··· , α
n
are linearly
dependent over
K
(since there are
n
+ 1 elements). So there exists some
a
i
K
(not all zero) such that
a
n
α
n
+ a
n1
α
n1
+ ··· + a
1
α + a
0
= 0.
So we have a non-trivial polynomial that vanishes at
α
. So
α
is algebraic over
K.
Since α was arbitrary, L itself is algebraic.
If
L/K
is a field extension and
α L
is algebraic, then by definition, there is
some polynomial
f
such that
f
(
α
) = 0. It is a natural question to ask if there is
a “smallest” polynomial that does this job. Obviously we can find a polynomial
of smallest degree (by the well-ordering principle of the natural numbers), but
we can get something even stronger.
Since
K
is a field,
K
[
t
] is a PID (principal ideal domain). This, by definition,
implies we can find some (monic)
P
α
K
[
t
] such that
I
α
=
P
α
. In other words,
every element of I
α
is just a multiple of P
α
.
Definition (Minimal polynomial). Let
L/K
be a field extension,
α L
. The
minimal polynomial of
α
over
K
is a monic polynomial
P
α
such that
I
α
=
P
α
.
Example.
(i) Consider R/Q, α =
3
2. Then the minimal polynomial is P
α
= t
3
2.
(ii) Consider C/R, α =
3
2. Then the minimal polynomial is P
α
= t
3
2.
It should be intuitively obvious that by virtue of being “minimal”, the
minimal polynomial is irreducible.
Proposition. Let
L/K
be a field extension,
α L
algebraic over
K
, and
P
α
the minimal polynomial. Then P
α
is irreducible in K[t].
Proof.
Assume that
P
α
=
QR
in
K
[
t
]. So 0 =
P
α
(
α
) =
Q
(
α
)
R
(
α
). So
Q
(
α
) = 0
or
R
(
α
) = 0. Say
Q
(
α
) = 0. So
Q I
α
. So
Q
is a multiple of
P
α
. However, we
also know that
P
α
is a multiple of
Q
α
. This is possible only if
R
is a unit in
K[t], i.e. R K. So P
α
is irreducible.
It should also be clear that if
f
is irreducible and
f
(
α
) = 0, then
f
is the
minimal polynomial. Often, it is the irreducibility of P
α
that is important.
Apart from the minimal polynomial, we can also ask for the minimal field
containing α.
Definition (Field generated by
α
). Let
L/K
be a field extension,
α L
. We
define
K
(
α
) to be the smallest subfield of
L
containing
K
and
α
. We call
K
(
α
)
the field generated by α over K.
This definition by itself is rather abstract and not very helpful. Intuitively,
K
(
α
) is what we get when we add
α
to
K
, plus all the extra elements needed to
make
K
(
α
) a field (i.e. closed under addition, multiplication and inverse). We
can express this idea more formally by the following result:
Theorem. Let L/K a field extension, α L algebraic. Then
(i) K
(
α
) is the image of the (ring) homomorphism
φ
:
K
[
t
]
L
defined by
f 7→ f (α).
(ii) [K(α) : K] = deg P
α
, where P
α
is the minimal polynomial of α over K.
Note that the kernel of the homomorphism
φ
is (almost) by definition the
ideal P
α
. So this theorem tells us
K[t]
P
α
=
K(α).
Proof.
(i)
Let
F
be the image of
φ
. The first step is to show that
F
is indeed a field.
Since
F
is the image of a ring homomorphism, we know
F
is a subring of
L. Given β F non-zero, we have to find an inverse.
By definition,
β
=
f
(
α
) for some
f K
[
t
]. The idea is to use B´ezout’s
identity. Since
β
= 0,
f
(
α
)
= 0. So
f ∈ I
α
=
P
α
. So
P
α
f
in
K
[
t
]. Since
P
α
is irreducible,
P
α
and
f
are coprime. Then there exists some
g, h K
[
t
]
such that
fg
+
hP
α
= 1. So
f
(
α
)
g
(
α
) =
f
(
α
)
g
(
α
) +
h
(
α
)
P
α
(
α
) = 1. So
βg(α) = 1. So β has an inverse. So F is a field.
From the definition of
F
, we have
K F
and
α F
, using the constant
polynomials f = c K and the identity f = t.
Now, if
K G L
and
α G
, then
G
contains all the polynomial
expressions of α. Hence F G. So K(α) = F .
(ii)
Let
n
=
deg P
α
. We show that
{
1
, α, α
2
, ··· , α
n1
}
is a basis for
K
(
α
)
over K.
First note that since deg P
α
= n, we can write
α
n
=
n1
X
i=0
a
i
α
i
.
So any other higher powers are also linear combinations of the
α
i
s (by
induction). This means that
K
(
α
) is spanned by 1
, ··· , α
n1
as a
K
vector
space.
It remains to show that
{
1
, ··· , α
n1
}
is linearly independent. Assume
not. Then for some b
i
, we have
n1
X
i=0
b
i
α
i
= 0.
Let
f
=
P
b
i
t
i
. Then
f
(
α
) = 0. So
f I
α
=
P
α
. However,
deg f <
deg P
α
. So we must have
f
= 0. So all
b
i
= 0. So
{
1
, ··· , α
n1
}
is a basis
for K(α) over K. So [K(α) : K] = n.
Corollary. Let
L/K
be a field extension,
α L
. Then
α
is algebraic over
K
if
and only if K(α)/K is a finite extension.
Proof.
If
α
is algebraic, then [
K
(
α
) :
K
] =
deg P
α
<
by above. So the
extension is finite.
If
K K
(
α
) is a finite extension, then by previous lemma, the entire
K
(
α
)
is algebraic over K. So α is algebraic over K.
We can extend this definition to allow more elements in the generating set.
Definition (Field generated by elements). Let
L/K
be a field extension,
α
1
, ··· , α
n
L
. We define
K
(
α
1
, ··· , α
n
) to be the smallest subfield of
L
containing K and α
1
, ··· , α
n
.
We call K(α
1
, ··· , α
n
) the field generated by α
1
, ··· , α
n
over K.
And we can prove some similar results.
Theorem. Suppose that L/K is a field extension.
(i)
If
α
1
, ··· , α
n
L
are algebraic over
K
, then
K
(
α
1
, ··· , α
n
)
/K
is a finite
extension.
(ii)
If we have field extensions
L/F/K
and
F/K
is a finite extension, then
F = K(α
1
, ··· , α
n
) for some α
1
, ··· , α
n
L.
Proof.
(i)
We prove this by induction. Since
α
1
is algebraic over
K
,
K K
(
α
1
) is a
finite extension.
For 1
i < n
,
α
i+1
is algebraic over
K
. So
α
i+1
is also algebraic
over
K
(
α
1
, ··· , α
i
). So
K
(
α
1
, ··· , α
i
)
K
(
α
1
, ··· , α
i
)(
α
i+1
) is a finite
extension. But
K
(
α
1
, ··· , α
i
)(
α
i+1
) =
K
(
α
1
, ··· , α
i+1
). By the tower law,
K K(α
i
, ··· , α
i+1
) is a finite extension.
(ii)
Since
F
is a finite dimensional vector space over
K
, we can take a basis
{α
1
, ··· , α
n
}
of
F
over
K
. Then it should be clear that
F
=
K
(
α
1
, ··· , α
n
).
When studying polynomials, the following result from IB Groups, Rings and
Modules is often helpful:
Proposition (Eisenstein’s criterion). Let
f
=
a
n
t
n
+
···
+
a
1
t
+
a
0
Z
[
t
].
Assume that there is some prime number p such that
(i) p | a
i
for all i < n.
(ii) p a
n
(iii) p
2
a
0
.
Then f is irreducible in Q[t].
Example. Consider the field extensions
Q Q(
2) Q(
2,
3
2) R,
Q Q(
3
2) Q(
2,
3
2) R.
We have [Q(
2) : Q] = 2 since {1,
2} is a basis of Q(
2) over Q.
How about [
Q
(
3
2
) :
Q
]? By the Eisenstein criterion, we know that
t
3
2 is
irreducible in
Q
[
t
]. So the minimal polynomial of
3
2
over
Q
is
t
3
2 which has
degree 3. So [Q(
3
2) : Q] = 3.
These results immediately tells that
3
2 ∈ Q
(
2
). Otherwise, this entails
that Q(
3
2) Q(
2). Then the tower law says that
[Q(
2) : Q] = [Q(
2) : Q(
3
2)][Q(
3
2) : Q].
In particular, plugging the numbers in entails that that 3 is a factor of 2, which
is clearly nonsense. Similarly,
2 ∈ Q(
3
2).
How about the inclusion
Q
(
2
)
Q
(
2,
3
2
)? We now show that the
minimal polynomial P
3
2
of
3
2 over Q(
2) is t
3
2.
Suppose not. Then
t
3
2 is reducible, with the real
P
3
2
as one of its factors.
Let t
3
2 = P
3
2
· R for some non-unit polynomial R.
We know that
P
3
2
does not have degree 3 (or else it would be
t
3
2), and
not degree 1, since a degree 1 polynomial has a root. So it has degree 2. So
R
has degree 1. Then
R
has a root, i.e.
R
(
β
) = 0 for some
β Q
(
2
). So
β
3
2 = 0. Hence [Q(β) : Q] = 3. Again, by the tower law, we have
[Q(
2) : Q] = [Q(
2) : Q(β)][Q(β) : Q].
Again, this is nonsense since it entails that 3 is a factor of 2. So the minimal
polynomial is indeed t
3
2. So [Q(
2,
3
2) : Q] = 6 by the tower law.
Alternatively, we can obtain this result by noting that the tower law on
Q Q
(
2
)
Q
(
2,
3
2
) and
Q Q
(
3
2
)
Q
(
2,
3
2
) entails that 2 and 3 are
both factors of [
Q
(
2,
3
2
) :
Q
]. So it is at least 6. Then since
t
3
2
Q
(
2
)[
t
]
has
3
2 as a root, the degree is at most 6. So it is indeed 6.
2.2 Ruler and compass constructions
Before we develop our theory further, we first look into a rather unexpected
application of field extensions. We are going to look at some classic problems in
geometry, and solve them using what we’ve learnt so far. In particular, we want
to show that certain things cannot be constructed using a compass and a ruler
(as usual, we assume the ruler does not have markings on it).
It is often easy to prove that certain things are constructible just exhibit
an explicit construction of it. However, it is much more difficult to show that
things are not constructible. Two classical examples are
(i)
Doubling the cube: Given a cube, can we construct the side of another
cube whose volume is double the volume of the original cube?
(ii)
Trisecting an angle: Given an angle, can we divide the angle into three
equal angles?
The idea here is to associate with each possible construction a field extension,
and then prove certain results about how these field extensions should behave.
We then show that if we could, say, double the cube, then this construction
would inevitable break some of the properties it should have.
Firstly, we want to formulate our problem in a more convenient way. In
particular, we will view the plane as
R
2
, and describe lines and circles by
equations. We also want to describe “compass and ruler” constructions in a
more solid way.
Definition (Constructible points). Let
S R
2
be a set of (usually finite) points
in the plane.
A “ruler” allows us to do the following: if
P, Q S
, then we can draw the
line passing through P and Q.
A “compass” allows us to do the following: if
P, Q, Q
S
, then we can draw
the circle with center at P and radius of length QQ
.
Any point
R R
2
is 1-step constructible from
S
if
R
belongs to the in-
tersection of two distinct lines or circles constructed from
S
using rulers and
compasses.
A point
R R
2
is constructible from
S
if there is some
R
1
, ··· , R
n
=
R R
2
such that R
i+1
is 1-step constructible from S {R
1
, ··· , R
i
} for each i.
Example. Let
S
=
{
(0
,
0)
,
(1
,
0)
}
. What can we construct? It should be easy
to see that (
n,
0) for all
n Z
are all constructible from
S
. In fact, we can show
that all points of the form (m, n) Z are constructible from S.
(0, 0) (1, 0)
Definition (Field of S). Let S R
2
be finite. Define the field of S by
Q(S) = Q({coordinates of points in S}) R,
where we put in the
x
coordinate and
y
coordinate separately into the generating
set.
For example, if S = {(
2,
3)}, then Q(S) = Q(
2,
3).
The key theorem we will use to prove our results is
Theorem. Let S R
2
be finite. Then
(i) If R is 1-step constructible from S, then [Q(S {R}) : Q(S)] = 1 or 2.
(ii)
If
T R
2
is finite,
S T
, and the points in
T
are constructible from
S
,
Then [Q(S T ) : Q(S)] = 2
k
for some k (where k can be 0).
Proof.
By assumption, there are distinct lines or circles
C, C
constructed from
S
using ruler and compass, such that
R C C
. By elementary geometry,
C
and C
can be given by the equations
C : a(x
2
+ y
2
) + bx + cy + d = 0,
C
: a
(x
2
+ y
2
) + b
x + c
y + d
= 0.
where
a, b, c, d, a
, b
, c
, d
Q
(
S
). In particular, if we have a line, then we can
take a = 0.
Let
R
= (
r
1
, r
2
). If
a
=
a
= 0 (i.e.
C
and
C
are lines), then solving the two
linear equations gives r
1
, r
2
Q(S). So [Q(S {R}) : Q(S)] = 1.
So we can now assume wlog that a = 0. We let
p = a
b ab
, q = a
c ac
, = a
d ad
,
which are the coefficients when we perform
a
×C a ×C
. Then by assumption,
p
= 0 or
q
= 0. Otherwise,
c
and
c
would be the same curve. wlog
p
= 0. Then
since (r
1
, r
2
) satisfy both equations of C and C
, they satisfy
px + qy + = 0.
In other words, pr
1
+ qr
2
+ = 0. This tells us that
r
1
=
qr
2
+
p
. ()
If we put
r
1
, r
2
into the equations of
C
and
C
and use (
), we get an equation
of the form
αr
2
2
+ βr
2
+ γ = 0,
where
α, β, γ Q
(
S
). So we can find
r
2
(and hence
r
1
using linear relations)
using only a single radical of degree 2. So
[Q(S {R}) : Q(S)] = [Q(S)(r
2
) : Q(S)] = 1 or 2,
since the minimal polynomial of r
2
over Q(S) has degree 1 or 2.
Then (ii) follows directly from induction, using the tower law.
Corollary. It is impossible to “double the cube”.
Proof.
Consider the cube with unit side length, i.e. we are given the set
S
=
{
(0
,
0)
,
(1
,
0)
}
. Then doubling the cube would correspond to constructing a side
of length
such that
3
= 2, i.e.
=
3
2
. Thus we need to construct a point
R = (
3
2, 0) from S.
If we can indeed construct this R, then we need
[Q(S {R}) : Q(S)] = 2
k
for some k. But we know that Q(S) = Q and Q(S {R}) = Q(
3
2), and that
[Q(
3
2) : Q] = 3.
This is a contradiction since 3 is not a power of 2.
2.3 K-homomorphisms and the Galois Group
Usually in mathematics, we not only want to study objects, but maps between
objects. Suppose we have two field extensions
K L
and
K L
. What should
a map between these two objects look like? Obviously, we would like this map
to be a field homomorphisms between
L
and
L
. Moreover, since this is a map
between the two field extensions, and not just the fields themselves, we would
like this map to preserve things in
K
, and is just a map between the “extended
parts” of L and L
.
Definition (
K
-homomorphism). Let
L/K
and
L
/K
be field extensions. A
K
-homomorphism
φ
:
L L
is a ring homomorphism such that
φ|
K
=
id
, i.e. it
fixes everything in
K
. We write
Hom
K
(
L, L
) for the set of all
K
-homomorphisms
L L
.
A
K
-isomorphism is a
K
-homomorphism which is an isomorphism of rings.
A
K
-automorphism is a
K
-isomorphism
L L
. We write
Aut
K
(
L
) for the set
of all K-automorphism L L.
There are a couple of things to take note of
(i) Given any φ Hom
K
(L, L
), we know that
(a)
Since
φ|
K
=
id
, we know that
ker φ
=
L
. Since we know that
ker φ
is
an ideal, and a field only has two ideals, we must have
ker φ
= 0. So
φ
is injective. It is, in fact, true that any homomorphisms of fields is
injective.
(b) φ
gives an isomorphism
L φ
(
L
). So
φ
(
L
) is a field and we get the
field extensions K φ(L) L
.
(ii)
If [
L
:
K
] = [
L
:
K
]
<
, then any homomorphism in
Hom
K
(
L, L
) is in
fact an isomorphism. So
{K-homomorphisms : L L
} = {K-isomorphisms : L L
},
This is since any
K
-homomorphism
φ
:
L L
is an injection. So
[
L
:
K
] = [
φ
(
L
) :
K
]. Hence we know that [
L
:
K
] = [
φ
(
L
) :
K
]. But we
know that
φ
(
L
) is a subfield of
L
. This is possible only if
L
=
φ
(
L
). So
φ is a surjection, and hence an isomorphism.
In particular, Aut
K
(L) = Hom
K
(L, L).
Example. We want to determine Aut
R
(C). If we pick any ψ Aut
R
(C), then
(ψ(
1))
2
+ 1 = ψ(
1
2
+ 1) = ψ(0) = 0.
So under any automorphism
ψ
, the image of
1
is a root of
t
2
+ 1. Therefore
ψ
(
1
) =
1
or
1
. In the first case,
ψ
is the identity. In the second
case, the automorphism is
φ
:
a
+
b
1 7→ a b
1
, i.e. the complex conjugate.
So Aut
R
(C) = {id, φ}.
Similarly, we can show that
Aut
Q
(
Q
(
2
)) =
{id, φ}
, where
φ
swaps
2
with
2.
Example. Let
µ
3
= 1 but
µ
= 1 (i.e.
µ
is a third root of unity). We want to
determine A = Hom
Q
(Q(
3
2), C).
First define φ, ψ by
φ(
3
2) =
3
2µ
ψ(
3
2) =
3
2µ
2
,
We have φ, ψ A. Are there more?
Let λ A. Then we must have
(λ(
3
2))
3
2 = 0.
So
λ
(
3
2
) is a root of
t
3
2. So it is either
3
2,
3
2µ
or
3
2µ
2
. So
λ
is either
id
,
φ or ψ. So A = {id, φ, ψ}.
Note that in general, if
α
is algebraic over
Q
, then
Q
(
α
)
=
Q
[
t
]
/P
α
. Hence
to specify a
Q
-homomorphism from
Q
(
α
), it suffices to specify the image of
t
, or
just the image of α.
We will later see that the number of automorphisms
|Aut
K
(
L
)
|
is bounded
above by the degree of the extension [
L
:
K
]. However, we need not always have
[
L
:
K
] many automorphisms. When we do have enough automorphisms, we call
it a Galois extension.
Definition (Galois extension). Let
L/K
be a finite field extension. This is a
Galois extension if |Aut
K
(L)| = [L : K].
Definition (Galois group). The Galois group of a Galois extension
L/K
is
defined as
Gal
(
L/K
) =
Aut
K
(
L
). The group operation is defined by function
composition. It is easy to see that this is indeed a group.
Example. The extension
Q
(
7
)
/Q
is Galois. The degree [
Q
(
7
) :
Q
] = 2, and
the automorphism group is
Aut
Q
(
Q
(
7
)) =
{id, φ}
, where
φ
swaps
7
with
7.
Example. The extension
Q
(
3
2
)
/Q
is not Galois. The degree is [
Q
(
3
2
) :
Q
] = 3,
but the automorphism group is Aut
Q
(Q(
3
2)) = {id}.
To show that there is no other automorphism, note that the automorphism
group can be viewed as a subset of
Hom
Q
(
Q
(
3
2
)
, C
). We have just seen that
Hom
Q
(
Q
(
3
2
)
, C
) has three elements, but only the identity maps
Q
(
3
2
) to itself,
while the others map
3
2 to
3
2µ
i
∈ Q(
3
2). So this is the only automorphism.
The way we should think about this is that there is something missing in
Q
(
3
2
), namely
µ
. Without the
µ
, we cannot get the other automorphisms we
need. In fact, in the next example, we will show that Q Q(
3
2, µ) is Galois.
Example.
Q
(
3
2, µ
)
/Q
is a Galois extension. Firstly, we know that [
Q
(
3
2, µ
) :
Q
(
3
2
)] = 2 because
µ
3
1 = 0 implies
µ
2
+
µ
+ 1 = 0. So the minimal
polynomial has degree 2. This also means that
µ ∈ Q
(
3
2
). We also know that
[Q(
3
2) : Q] = 3. So we have
[Q(
3
2, µ) : Q] = 6
by the Tower law.
Now denote
α
=
3
2
,
β
=
3
2µ
and
γ
=
3
2µ
2
. Then
Q
(
3
2, µ
) =
Q
(
α, β, γ
).
Now let
φ Aut
Q
(
Q
(
3
2, µ
)), then
φ
(
α
),
φ
(
β
) and
φ
(
γ
) are roots of
t
3
2. These
roots are exactly α, β, γ. So
{φ(α), φ(β), φ(γ)} = {α, β, γ}.
Hence
φ
is completely determined by a permutation of the roots of
t
3
2. So
Aut
Q
(
3
2, µ)
=
S
3
and |Aut
Q
(
3
2, µ)| = 6.
Most of the time, we will only be interested in Galois extensions. The main
reason is that Galois extensions satisfy the fundamental theorem of Galois theory,
which roughly says: if
L/K
is a finite Galois extension, then there is a one-to-one
correspondence of the set of subgroups
H Gal
(
L/K
) and the intermediate
fields
K F L
. In particular, the normal subgroups corresponds to the
“normal extensions”, which is something we will define later.
However, just as we have seen, it is not straightforward to check if an extension
is Galois, even in specific cases like the examples above. Fortunately, by the
time we reach the proper statement of the fundamental theorem, we would have
developed enough machinery to decide easily whether certain extensions are
Galois.
2.4 Splitting fields
As mentioned in the introduction, one major motivation for Galois theory is to
study the roots of polynomials. So far, we have just been talking about field
extensions. The idea here is given a field
K
and a polynomial
f K
[
t
], we
would like to study the field extension obtained by adding all roots of
f
. This is
known as the splitting field of f (over K).
Notation. Let
L/K
be a field extension,
f K
[
t
]. We write
Root
f
(
L
) for the
roots of f in L.
First, we establish a correspondence between the roots of a polynomial and
K-homomorphisms.
Lemma. Let
L/K
be a field extension,
f K
[
t
] irreducible,
deg f >
0. Then
there is a 1-to-1 correspondence
Root
f
(L) Hom
K
(K[t]/f, L).
Proof.
Since
f
is irreducible,
f
is a maximal ideal. So
K
[
t
]
/f
is a field. Also,
there is a natural inclusion
K K
[
t
]
/f
. So it makes sense to talk about
Hom
K
(K[t]/f, L).
To any
β Root
f
(
L
), we assign
φ
:
K
[
t
]
/f L
where we map
¯
t 7→ β
(
¯
t
is
the equivalence class of
t
). This is well defined since if
¯
t
=
¯g
, then
g
=
t
+
hf
for
some h K[t]. So φg) = φ(t + hf) = β + h(β)f(β) = β.
Conversely, given any
K
-homomorphism
φ
:
K
[
t
]
/f L
, we assign
β
=
φ(
¯
t). This is a root since f(β) = f (φ(
¯
t)) = φ(f(
¯
t)) = φ(0) = 0.
This assignments are inverses to each other. So we get a one-to-one corre-
spondence.
Recall that if
K F
is a field extension, then for any
α F
with minimal
polynomial
P
α
, we have
K
[
t
]
/P
α
=
K
(
α
). Since an irreducible
f
is the minimal
polynomial of its roots, we can view the above lemma as telling us something
about Hom
K
(K(α), L).
Corollary. Let
L/K
be a field extension,
f K
[
t
] irreducible,
deg f >
0. Then
|Hom
K
(K[t]/f, L)| deg f.
In particular, if E = K[t]/f , then
|Aut
K
(E)| = |Root
f
(E)| deg f = [E : K].
So E/K is a Galois extension iff |Root
f
(E)| = deg f.
Proof. This follows directly from the following three facts:
|Root
f
(L)| deg f
Aut
K
(E) = Hom
K
(E, E)
deg f = [K(α) : K] = [E : K].
Definition (Splitting field). Let
L/K
be a field extensions,
f K
[
t
]. We say
f
splits over L if we can factor f as
f = a(t α
1
) ···(t α
n
)
for some
a K
and
α
j
L
. Alternatively, this says that
L
contains all roots of
f.
We say
L
is a splitting field of
f
if
L
=
K
(
α
1
, ··· , α
n
). This is the smallest
field where f has all its roots.
Example.
C is the splitting field of t
2
+ 1 R[t].
Q
(
3
2, µ
) is a splitting field of
t
3
2
Q
[
t
], where
µ
is a third root of
unity.
By the fundamental theorem of algebra, for any
K C
and
f K
[
t
],
there is a splitting field L C of f.
Note that the degree of the splitting field need not be (bounded by) the
degree of the polynomial. In the second example, we have [
Q
(
3
2, µ
) :
Q
] = 6,
but t
3
2 only has degree 3.
More generally, we can show that every polynomial has a splitting field, and
this is unique up to isomorphism. This is important, since we would like to talk
about the splitting field of a polynomial all the time.
Theorem. Let K be a field, f K[t]. Then
(i) There is a splitting field of f.
(ii) The splitting field is unique (up to K-isomorphism).
Proof.
(i)
If
deg f
= 0, then
K
is a splitting field of
f
. Otherwise, we add the roots
of f one by one.
Pick
g | f
in
K
[
t
], where
g
is irreducible and
deg g >
0. We have the field
extension
K K
[
t
]
/g
. Let
α
1
=
¯
t
. Then
g
(
α
1
) = 0 which implies that
f
(
α
1
) = 0. Hence we can write
f
= (
t α
1
)
h
in
K
(
α
1
)[
t
]. Note that
deg h < deg f
. So we can repeat the process on
h
iteratively to get a field
extensions
K K
(
α
1
, ··· , α
n
). This
K
(
α
1
, ··· , α
n
) is a splitting field of
f.
(ii)
Assume
L
and
L
are both splitting fields of
f
over
K
. We want to find a
K-isomorphism from L to L
.
Pick largest
F, F
such that
K F L
and
K F
L
are field
extensions and there is a
K
-isomorphism from
ψ
:
F F
. By “largest”,
we mean we want to maximize [F : K].
We want to show that we must have
F
=
L
. Then we are done because
this means that F
is a splitting field, and hence F
= L
.
So suppose
F
=
L
. We will try to produce a larger
˜
F
with
K
-isomorphism
˜
F
˜
F
L
.
Since
F
=
L
, we know that there is some
α Root
f
(
L
) such that
α ∈ F
.
Then there is some irreducible
g K
[
t
] with
deg g >
0 such that
g
(
α
) = 0
and g | f . Say f = gh.
Now we know there is an isomorphism
F
[
t
]
/g F
(
α
) by
¯
t 7→ α
. The
isomorphism
ψ
:
F F
extends to a isomorphism
µ
:
F
[
t
]
F
[
t
].
Then since the coefficients of
f
are in
K
, we have
f
=
µ
(
f
) =
µ
(
g
)
µ
(
h
).
So
µ
(
g
)
| f
in
F
[
t
]. Since
g
is irreducible in
F
[
t
],
µ
(
g
) is irreducible in
F
[
t
]. So there is some
α
Root
µ(g)
(
L
)
Root
f
(
L
) and isomorphism
F
[t]/µ(g) F
(α
).
Now
µ
induces a
K
-isomorphism
F
[
t
]
/g F
[
t
]
/µ
(
g
)
, which in turn
induces a
K
-isomorphism
F
(
α
)
F
(
α
). This contradicts the maximality
of F . So we must have had F = L.
Note that the splitting is unique just up to isomorphism. We could be
quotienting by different polynomials and still get the same splitting field.
Example.
Q
(
7
) is a splitting field of
t
2
7
Q
[
t
]. At the same time,
Q
(
7
)
is also a splitting field of t
2
+ 3t +
1
2
Q[t].
2.5 Algebraic closures
The splitting field gives us the field with the root of one particular polynomial.
We could be greedy and ask for the roots for all polynomials, and get the
algebraic closure. The algebraic closure will not be of much use in this course,
but is a nice thing to know about. The major theorems would be the existence
and uniqueness of algebraic closures.
Definition (Algebraically closed field). A field
L
is algebraically closed if for all
f L[t], we have
f = a(t α
1
)(t α
2
) ···(t α
n
)
for some a, α
i
L. In other words, L contains all roots of its polynomials.
Let L/K be a field extension. We say L is an algebraic closure of K if
L is algebraic over K
L is algebraically closed.
Example.
L
is an algebraically closed field iff (
L E
is a finite extension
implies E = L).
This is since if
L E
is finite, then
E
is algebraic over
L
, and hence must
be L.
Example.
C
is algebraically closed by the fundamental theorem of algebra, and
is the algebraic closure of R (but not Q).
Before we prove our next theorem, we need the following technical lemma:
Lemma. If
R
is a commutative ring, then it has a maximal ideal. In particular,
if I is an ideal of R, then there is a maximal ideal that contains I.
Proof. Let
P = {I : I is an ideal of R, I = R}.
If
I
1
I
2
···
is any chain of
I
i
P
, then
I
=
S
I
i
P
. By Zorn’s lemma,
there is a maximal element of
P
(containing
I
). So
R
has at least one maximal
ideal (containing I).
Theorem (Existence of algebraic closure). Any field
K
has an algebraic closure.
Proof. Let
A = {λ = (f, j) : f K[t] irreducible monic, 1 j deg f}.
We can think of
j
as labelling which root of
f
we want. For each
λ A
, we
assign a variable t
λ
. We take
R = K[t
λ
: λ A]
to be the polynomial ring over
K
with variables
t
λ
. This
R
contains all the
“roots” of the polynomials in
K
. However, we’ve got a bit too much. For example,
(if
K
=
Q
), in
R
,
3
and
3
+ 1 would be put down as separate, unrelated
variables. So we want to quotient this R by something.
For every monic and irreducible f K[t], we define
˜
f = f
deg f
Y
j=1
(t t
(f,j)
) R[t].
If we want the
t
(f,j)
to be roots of
f
, then
˜
f
should vanish for all
f
. Denote the
coefficient of t
in
˜
f by b
(f,ℓ)
. Then we want b
(f,ℓ)
= 0 for all f, .
To do so, let
I R
be the ideal generated by all such coefficients. We now
want to quotient R by I. We first have to check that I = R.
Suppose not. So there are
b
(f
1
,ℓ
1
)
, ··· , b
(f
r
,ℓ
r
)
with
g
1
, ··· , g
r
R
such that
g
1
b
(f
1
,ℓ
1
)
+ ··· + g
r
b
(f
r
,ℓ
r
)
= 1. ()
We will attempt to reach a contradiction by constructing a homomorphism
φ
that sends each b
(f
i
,ℓ
i
)
to 0.
Let E be a splitting field of f
1
f
2
···f
r
. So in E[t], for each i, we can write
f
i
=
deg f
i
Y
j=1
(t α
i,j
).
Then we define a homomorphism φ : R E by
(
φ(t
(f
i
,j)
) = α
i,j
φ(t
λ
) = 0 otherwise
This induces a homomorphism
˜
φ : R[t] E[t].
Now apply
˜
φ(
˜
f
i
) =
˜
φ(f
i
)
deg f
i
Y
j=1
˜
φ(t t
(f
i
,j)
)
= f
i
deg f
i
Y
j=1
(t α
i,j
)
= 0
So φ(b
(f
i
,ℓ
i
)
) = 0 as b
(f
i
,ℓ
i
)
is a coefficient of f
i
.
Now we apply φ to () to obtain
φ(g
1
b
(f
1
,ℓ
1
)
+ ··· + g
r
b
(f
r
,ℓ
r
)
) = φ(1).
But this is a contradiction since the left had side is 0 while the right is 1. Hence
we must have I = R.
We would like to quotient by
I
, but we have to be a bit more careful, since
the quotient need not be a field. Instead, pick a maximal ideal
M
containing
I
, and consider
L
=
R/M
. Then
L
is a field. Moreover, since we couldn’t
have quotiented out anything in
K
(any ideal containing anything in
K
would
automatically contain all of
R
), this is a field extension
L/K
. We want to show
that L is an algebraic closure.
Now we show that
L
is algebraic over
K
. This should all work out smoothly,
since that’s how we constructed
L
. First we pick
α L
. Since
L
=
R/M
and
R
is generated by the terms t
λ
, there is some (f
1
, j
1
), ··· , (f
r
, j
r
) such that
α K(
¯
t
(f
i
,j
i
)
, ··· ,
¯
t
(f
r
,j
r
)
).
So
α
is algebraic over
K
if each
¯
t
(f
i
,j
i
)
is algebraic over
K
. To show this, note
that
˜
f
i
= 0, since we’ve quotiented out each of its coefficients. So by definition,
0 = f
i
(t)
deg f
i
Y
j=1
(t
¯
t
(f
i
,j)
).
So f
i
(
¯
t
(f
i
,j
i
)
) = 0. So done.
Finally, we have to show that
L
is algebraically closed. Suppose
L E
is a
finite (and hence algebraic) extension. We want to show that L = E.
Consider arbitrary
β E
. Then
β
is algebraic over
L
, say a root of
f L
[
t
]. Since every coefficient of
f
can be found in some finite extension
K
(
¯
t
(f
i
,j
i
)
, ··· ,
¯
t
(f
r
,j
r
)
), there is a finite extension
F
of
K
that contains all coeffi-
cients of
f
. Since
F
(
β
) is a finite extension of
F
, we know
F
(
β
) is a finite and
hence algebraic extension of K. In particular, β is algebraic in K.
Let
P
β
be the minimal polynomial of
β
over
K
. Since all polynomials in
K
split over
L
by construction (
f
(
t
) =
Q
(
t
¯
t
(f,j)
)), its roots must in
L
. In
particular, β L. So L = E.
Theorem (Uniqueness of algebraic closure). Any field
K
has a unique algebraic
closure up to K-isomorphism.
This is the same proof as the proof that the splitting field is unique given
two algebraic closures, we take the largest subfield of the algebraic closures that
biject with each other. However, since there could be infinitely many subfields,
we have to apply Zorn’s lemma to obtain the maximal such subfield.
Proof. (sketch) Suppose L, L
are both algebraic closures of K. Let
H = {(F, ψ) : K F L, ψ Hom
K
(F, L
)}.
We define a partial order on
H
by (
F
1
, ψ
1
)
(
F
2
, ψ
2
) if
F
1
F
2
and
ψ
1
=
ψ
2
|
F
1
.
We have to show that chains have upper bounds. Given a chain
{
(
F
α
, ψ
α
)
}
,
we define
F =
[
F
α
, ψ(x) = ψ
α
(x) for x F
α
.
Then (
F, ψ
)
H
. Then applying Zorn’s lemma, there is a maximal element of
H, say (F, ψ).
Finally, we have to prove that
F
=
L
, and that
ψ
(
L
) =
L
. Suppose
F
=
L
.
Then we attempt to produce a larger
˜
F
and a
K
-isomorphism
˜
F
˜
F
L
.
Since
F
=
L
, there is some
α L \ F
. Since
L
is an algebraic extension of
K
,
there is some irreducible g K[t] such that deg g > 0 and g(α) = 0.
Now there is an isomorphism
F
[
t
]
/g F
(
α
) defined by
¯
t 7→ α
. The
isomorphism
ψ
:
F F
then extends to an isomorphism
µ
:
F
[
t
]
F
[
t
]
and thus to
F
[
t
]
/g F
[
t
]
/µ
(
g
)
. Then if
α
is a root of
µ
(
g
), then we have
F
[
t
]
//µ
(
g
)
=
F
(
α
). So this gives an isomorphism
F
(
α
)
F
(
α
). This
contradicts the maximality of φ.
By doing the argument the other way round, we must have
ψ
(
L
) =
L
. So
done.
2.6 Separable extensions
Here we will define what it means for an extension to be separable. This is
done via defining separable polynomials, and then an extension is separable if
all minimal polynomials are separable.
At first, the definition of separability might seem absurd surely every
polynomial should be separable. Indeed, polynomials that are not separable
tend to be weird, and our theories often break without separability. Hence it is
important to figure out when polynomials are separable, and when they are not.
Fortunately, we will end up with a result that tells us exactly when a polynomial
is not separable, and this is just a very small, specific class. In particular, in
fields of characteristic zero, all polynomials are separable.
Definition (Separable polynomial). Let
K
be a field,
f K
[
t
] non-zero, and
L
a splitting field of
f
. For an irreducible
f
, we say it is separable if
f
has no
repeated roots, i.e.
|Root
f
(
L
)
|
=
deg f
. For a general polynomial
f
, we say it is
separable if all its irreducible factors in K[t] are separable.
It should be obvious from definition that if
P
is separable and
Q | P
, then
Q
is also separable.
Note that some people instead define a separable polynomial to be one with
no repeated roots, so (
x
2)
2
over
Q
would not be separable under this definition.
Example. Any linear polynomial t a (with a K) is separable.
This is, however, not a very interesting example. To get to more interesting
examples, we need even more preparation.
Definition (Formal derivative). Let
K
be a field,
f K
[
t
]. (Formal) differenti-
ation the K-linear map K[t] K[t] defined by t
n
7→ nt
n1
.
The image of a polynomial f is the derivative of f , written f
.
This is similar to how we differentiate real or complex polynomials (in case
that isn’t obvious).
The following lemma summarizes the properties of the derivative we need.
Lemma. Let K be a field, f, g K[t]. Then
(i) (f + g)
= f
+ g
, (f g)
= fg
+ f
g.
(ii)
Assume
f
= 0 and
L
is a splitting field of
f
. Then
f
has a repeated root in
L
if and only if
f
and
f
have a common (non-constant) irreducible factor
in K[t] (if and only if f and f
have a common root in L).
This will allow us to show when irreducible polynomials are separable.
Proof.
(i) (f + g)
= f
+ g
is true by linearity.
To show that (
fg
)
=
fg
+
f
g
, we use linearity to reduce to the case
where
f
=
t
n
, g
=
t
m
. Then both sides are (
n
+
m
)
t
n+m1
. So this holds.
(ii)
First assume that
f
has a repeated root. So let
f
= (
t α
)
2
h L
[
t
] where
α L
. Then
f
= 2(
t α
)
h
+ (
t α
)
2
h
= (
t α
)(2
h
+ (
t α
)
h
). So
f
(
α
) =
f
(
α
) = 0. So
f
and
f
have common roots. However, we want a
common irreducible factor in
K
[
t
], not
L
[
t
]. So we let
P
α
be the minimal
polynomial of α over K. Then P
α
| f and P
α
| f
. So done.
Conversely, suppose
e
is a common irreducible factor of
f
and
f
in
K
[
t
],
with deg e > 0. Pick α Root
e
(L). Then α Root
f
(L) Root
f
(L).
Since α is a root of f, we can write f = (t α)q L[t] for some q. Then
f
= (t α)q
+ q.
Since (t α) | f
, we must have (t α) | q. So (t α)
2
| f.
Recall that the characteristic of a field
char K
is the minimum
p
such that
p ·
1
K
= 0. If no such
p
exists, we say
char K
= 0. For example,
Q
has
characteristic 0 while Z
p
has characteristic p.
Corollary. Let K be a field, f K[t] non-zero irreducible. Then
(i) If char K = 0, then f is separable.
(ii)
If
char K
=
p >
0, then
f
is not separable iff
deg f >
0 and
f K
[
t
p
]. For
example, t
2p
+ 3t
p
+ 1 is not separable.
Proof.
By definition, for irreducible
f
,
f
is not separable iff
f
has a repeated
root. So by our previous lemma,
f
is not separable if and only if
f
and
f
have a common irreducible factor of positive degree in
K
[
t
]. However, since
f
is
irreducible, its only factors are 1 and itself. So this can happen if and only if
f
= 0.
To make it more explicit, we can write
f = a
n
t
n
+ ··· + a
1
t + a
0
.
Then we can write
f
= na
n
t
n1
+ ··· + a
1
.
Now f
= 0 if and only if all coefficients ia
i
= 0 for all i.
(i)
Suppose
char K
= 0, then if
deg f
= 0, then
f
is trivially separable. If
deg f >
0, then
f
is not separable iff
f
= 0 iff
ia
i
= 0 for all
i
iff
a
i
= 0
for
i
1. But we cannot have a polynomial of positive degree with all its
coefficients zero (apart from the constant term). So f must be separable.
(ii) If deg f = 0, then f is trivially separable. So assume deg f > 0.
Then
f
is not separable
f
= 0
ia
i
= 0 for
i
0
a
i
= 0 for all
i 1 not multiples of p f K[t
p
].
Using this, it should be easy to find lots of examples of separable polynomials.
Definition (Separable elements and extensions). Let
K L
be an algebraic
field extension. We say
α L
is separable over
K
if
P
α
is separable, where
P
α
is the minimal polynomial of α over K.
We say
L
is separable over
K
(or
K L
is separable) if all
α L
are
separable.
Example.
The extensions
Q Q
(
2
) and
R C
are separable because
char Q
=
char R = 0. So we can apply our previous corollary.
Let
L
=
F
p
(
s
) be the field of rational functions in
s
over
F
p
(which is the
fraction field of
F
p
[
s
]), and
K
=
F
p
(
s
p
). We have
K L
, and
L
=
K
(
s
).
Since
s
p
K
,
s
is a root of
t
p
s
p
K
[
t
]. So
s
is algebraic over
K
and
hence
L
is algebraic over
K
. In fact
P
s
=
t
p
s
p
is the minimal polynomial
of s over K.
Now
t
p
s
p
= (
ts
)
p
since the field has characteristic
p
. So
Root
t
p
s
p
(
L
) =
{s}. So P
s
is not separable.
As mentioned in the beginning, separable extensions are nice, or at least
non-weird. One particular nice result about separable extensions is that all finite
separable extensions are simple, i.e. if
K L
is finite separable, then
L
=
K
(
α
)
for some
α L
. This is what we will be working towards for the remaining of
the section.
Example. Consider
Q Q
(
2,
3
). This is a separable finite extension. So
we should be able to generate
Q
(
2,
3
) by just one element, not just two. In
fact, we can use α =
2 +
3, since we have
α
3
= 11
2 + 9
3 = 2
2 + 9α.
So since α
3
Q(α), we know that
2 Q(α). So we also have
3 Q(α).
In general, it is not easy to find an
α
that works, but we our later result will
show that such an α exists.
Before that, we will prove some results about the K-homomorphisms.
Lemma. Let
L/F/K
be finite extensions, and
E/K
be a field extension. Then
for all α L, we have
|Hom
K
(F (α), E)| [F (α) : F ]|Hom
K
(F, E)|.
Note that if
P
α
is the minimal polynomial of
α
over
F
, then [
F
(
α
) :
F
] =
deg P
α
. So we can interpret this intuitively as follows: for each
ψ Hom
K
(
F, E
),
we can obtain a
K
-homomorphism in
Hom
K
(
F
(
α
)
, E
) by sending things in
F
according to
ψ
, and then send
α
to any root of
P
α
. Then there are at
most [
F
(
α
) :
F
]
K
-homomorphisms generated this way. Moreover, each
K
-
homomorphism in
Hom
K
(
F
(
α
)
, E
) can be created this way. So we get this
result.
Proof.
We show that for each
ψ Hom
K
(
F, E
), there are at most [
F
(
α
) :
F
]
K
-isomorphisms in
Hom
K
(
F
(
α
)
, E
) that restrict to
ψ
in
F
. Since each
K
-
isomorphism in
Hom
K
(
F
(
α
)
, E
) has to restrict to something, it follows that
there are at most [
F
(
α
) :
F
]
|Hom
K
(
F, E
)
| K
-homomorphisms from
F
(
α
) to
E
.
Now let
P
α
be the minimal polynomial for
α
in
F
, and let
ψ Hom
K
(
F, E
).
To extend ψ to a morphism F (α) E, we need to decide where to send α. So
there should be some sort of correspondence
Root
P
α
(E) {φ Hom
K
(F (α), E) : φ|
F
= ψ}.
Except that the previous sentence makes no sense, since
P
α
F
[
t
] but we are
not told that F is a subfield of E. So we use our ψ to “move” our things to E.
We let
M
=
ψ
(
F
)
E
, and
q M
[
t
] be the image of
P
α
under the
homomorphism
F
[
t
]
M
[
t
] induced by
ψ
. As we have previously shown, there
is a one-to-one correspondence
Root
q
(E) Hom
M
(M[t]/q, E).
What we really want to show is the correspondence between
Root
q
(
E
) and the
K
-homomorphisms
F
[
t
]
/P
α
E
that restrict to
ψ
on
F
. Let’s ignore the
quotient for the moment and think: what does it mean for
φ Hom
K
(
F
[
t
]
, E
) to
restrict to
ψ
on
F
? We know that any
φ Hom
L
(
F
[
t
]
, E
) is uniquely determined
by the values it takes on
F
and
t
. Hence if
φ|
F
=
ψ
, then our
φ
must send
F
to
ψ
(
F
) =
M
, and can send
t
to anything in
E
. This corresponds exactly to
the
M
-homomorphisms
M
[
t
]
E
that does nothing to
M
and sends
t
to that
“anything” in E.
The situation does not change when we put back the quotient. Changing
from
M
[
t
]
E
to
M
[
t
]
/q E
just requires that the image of
t
must be
a root of
q
. On the other hand, using
F
[
t
]
/P
α
instead of
F
[
t
] requires that
φ
(
P
α
(
t
)) = 0. But we know that
φ
(
P
α
) =
ψ
(
P
α
) =
q
. So this just requires
q(t) = 0 as well. So we get the one-to-one correspondence
Hom
M
(M[t]/q, E) {φ Hom
K
(F [t]/P
α
, E) : φ|
F
= ψ}.
Since F [t]/P
α
= F (α), there is a one-to-one correspondence
Root
q
(E) {φ Hom
K
(F (α), E) : φ|
F
= ψ}.
So done.
Theorem. Let L/K and E/K be field extensions. Then
(i) |Hom
K
(L, E)| [L : K]. In particular, |Aut
K
(L)| [L : K].
(ii) If equality holds in (i), then for any intermediate field K F L:
(a) We also have |Hom
K
(F, E)| = [F : K].
(b) The map Hom
K
(L, E) Hom
K
(F, E) by restriction is surjective.
Proof.
(i) We have previously shown we can find a sequence of field extensions
K = F
0
F
1
··· F
n
= L
such that for each
i
, there is some
α
i
such that
F
i
=
F
i1
(
α
i
). Then by
our previous lemma, we have
|Hom
K
(L, E)| [F
n
: F
n1
]|Hom
K
(F
n1
, E)|
[F
n
: F
n1
][F
n1
: F
n2
]|Hom
K
(F
n2
, E)|
.
.
.
[F
n
: F
n1
][F
n1
: F
n2
] ···[F
1
: F
0
]|Hom
K
(F
0
, E)|
= [F
n
: F
0
]
= [L : K]
(ii) (a)
If equality holds in (i), then every inequality in the proof above has
to an equality. Instead of directly decomposing
K L
as a chain
above, we can first decompose
K F
, then
F L
, then join them
together. Then we can assume that F = F
i
for some i. Then we get
|Hom
K
(L, E)| = [L : F ]|Hom
K
(F, E)| = [L : K].
Then the tower law says
|Hom
K
(F, E)| = [F : K].
(b)
By the proof of the lemma, for each
ψ Hom
K
(
F, E
), we know that
{φ : Hom
K
(L, E) : φ|
F
= ψ} [L : F ]. ()
As we know that
|Hom
K
(F, E)| = [F : K], |Hom
K
(L, E)| = [L : K]
we must have had equality in (
), or else we won’t have enough
elements. So in particular
{φ
:
Hom
K
(
L, E
) :
φ|
F
=
ψ}
1. So the
map is surjective.
With this result, we can prove prove the following result characterizing
separable extensions.
Theorem. Let
L/K
be a finite field extension. Then the following are equivalent:
(i) There is some extension E of K such that |Hom
K
(L, E)| = [L : K].
(ii) L/K is separable.
(iii) L
=
K
(
α
1
, ··· , α
n
) such that
P
α
i
, the minimal polynomial of
α
i
over
K
,
is separable for all i.
(iv) L
=
K
(
α
1
, ··· , α
n
) such that
R
α
i
, the minimal polynomial of
α
i
over
K(α
1
, ··· , α
i1
) is separable for all i.
Proof.
(i)
(ii): For all
α L
, if
P
α
is the minimal polynomial of
α
over
K
,
then since K(α) is a subfield of L, by our previous theorem, we have
|Hom
K
(K(α), E)| = [K(α) : K].
We also know that
|Root
P
α
(
E
)
|
=
|Hom
K
(
K
(
α
)
, E
)
|
, and that [
K
(
α
) :
K
] =
deg P
α
. So we know that
P
α
has no repeated roots in any splitting
field. So P
α
is a separable. So L/K is a separable extension.
(ii) (iii): Obvious from definition
(iii)
(iv): Since
R
α
i
is a minimal polynomial in
K
(
α
1
, ··· , α
i1
), we
know that R
α
i
| P
α
i
. So R
α
i
is separable as P
α
i
is separable.
(iv)
(i): Let
E
be the splitting field of
P
α
1
, ··· , P
α
n
. We do induction
on
n
to show that this satisfies the properties we want. If
n
= 1, then
L = K(α
1
). Then we have
|Hom
K
(L, E)| = |Root
P
α
i
(E)| = deg P
α
1
= [K(α
1
) : K] = [L : K].
We now induct on
n
. So we can assume that (iv)
(i) holds for smaller
number of generators. For convenience, we write
K
i
=
K
(
α
1
, ··· , α
i
).
Then we have
|Hom
K
(K
n1
, E)| = [K
n1
: K].
We also know that
|Hom
K
(K
n
, E)| [K
n
: K
n1
]|Hom
K
(K
n1
, E)|.
What we actually want is equality. We now re-do (parts of) the proof of
this result, and see that separability guarantees that equality holds. If
we pick
ψ Hom
K
(
K
n1
, E
), then there is a one-to-one correspondence
between
{φ Hom
K
(
K
n
, E
) :
φ|
K
n1
=
ψ}
and
Root
q
(
E
), where
q M
[
t
]
is defined as the image of
R
α
n
under
K
n1
[
t
]
M
[
t
], and
M
is the image
of ψ.
Since
P
α
n
K
[
t
] and
R
α
n
| P
α
n
, then
q | P
α
n
. So
q
splits over
E
. By
separability assumption , we get that
|Root
q
(E)| = deg q = deg R
α
n
= [K
n
: K
n1
].
Hence we know that
|Hom
K
(L, E)| = [K
n
: K
n1
]|Hom
K
(K
n1
, E)|
= [K
n
: K
n1
][K
n1
: K]
= [K
n
: K].
So done.
Before we finally get to the primitive element theorem, we prove the following
lemma. This will enable us to prove the trivial case of the primitive element
theorem, and will also be very useful later on.
Lemma. Let
L
be a field,
L
=
L \ {
0
}
be the multiplicative group of
L
. If
G
is a finite subgroup of L
, then G is cyclic.
Proof.
Since
L
is abelian,
G
is also abelian. Then by the structure theorem on
finite abelian groups,
G
=
Z
n
1
× ··· ×
Z
n
r
,
for some
n
i
N
. Let
m
be the least common multiple of
n
1
, ··· , n
r
, and let
f = t
m
1.
If α G, then α
m
= 1. So f(α) = 0 for all α G. Therefore
|G| = n
1
···n
r
|Root
f
(L)| deg f = m.
Since
m
is the least common multiple of
n
1
, ··· , n
r
, we must have
m
=
n
1
···n
r
and thus (
n
i
, n
j
) = 1 for all
i
=
j
. Then by the Chinese remainder theorem, we
have
G
=
Z
n
1
× ··· ×
Z
n
r
=
Z
n
1
···n
r
.
So G is cyclic.
We now come to the main theorem of the lecture:
Theorem (Primitive element theorem). Assume
L/K
is a finite and separable
extension. Then L/K is simple, i.e. there is some α L such that L = K(α).
Proof.
At some point in our proof, we will require that
L
is infinite. So we
first do the finite case first. If
K
is finite, then
L
is also finite, which in turns
implies
L
is finite too. So by the lemma,
L
is a cyclic group (since it is a finite
subgroup of itself). So there is some
α L
such that every element in
L
is a
power of α. So L = K(α).
So focus on the case where
K
is infinite. Also, assume
K
=
L
. Then since
L/K
is a finite extension, there is some intermediate field
K F L
such that
L
=
F
(
β
) for some
β
. Now
L/K
is separable. So
F/K
is also separable, and
[
F
:
K
]
<
[
L
:
K
]. Then by induction on degree of extension, we can assume
F/K
is simple. In other words, there is some
λ F
such that
F
=
K
(
λ
). Now
L = K(λ, β). In the rest of the proof, we will try to replace the two generators
λ, β with just a single generator.
Unsurprisingly, the generator of
L
will be chosen to be a linear combination
of β and λ. We set
α = β +
for some
a K
to be chosen later. We will show that
K
(
α
) =
L
. Actually,
almost any choice of
a
will do, but at the end of the proof, we will see which
ones are the bad ones.
Let
P
β
and
P
λ
be the minimal polynomial of
β
and
λ
over
K
respectively.
Consider the polynomial f = P
β
(α at) K(α)[t]. Then we have
f(λ) = P
β
(α ) = P
β
(β) = 0.
On the other hand, P
λ
(λ) = 0. So λ is a common root of P
λ
and f .
We now want to pick an
a
such that
λ
is the only common root of
f
and
P
λ
(in
E
). If so, then the gcd of
f
and
P
α
in
K
(
α
) must only have
λ
as a root.
But since
P
λ
is separable, it has no double roots. So the gcd must be
t λ
. In
particular, we must have
λ K
(
α
). Since
α
=
β
+
, it follows that
β K
(
α
)
as well, and so K(α) = L.
Thus, it remains to choose an
a
such that there are no other common roots.
We work in a splitting field of P
β
P
λ
, and write
P
β
= (t β
1
) ···(t β
m
)
P
λ
= (t λ
1
) ···(t λ
n
).
We wlog β
1
= β and λ
1
= λ.
Now suppose θ is a common root of f and P
λ
. Then
(
f(θ) = 0
P
λ
(θ) = 0
(
P
β
(α ) = 0
P
λ
(θ) = 0
(
α = β
i
θ = λ
j
for some i, j. Then we know that
α = β
i
+
j
.
However, by definition, we also know that
α = β +
Now we see how we need to choose
a
. We need to choose
a
such that the elements
β + = β
i
+
j
for all i, j. But if they were equal, then we have
a =
λ λ
j
β
i
β
,
and there are only finitely many elements of this form. So we just have to pick
an a not in this list.
Corollary. Any finite extension
L/K
of field of characteristic 0 is simple, i.e.
L = K(α) for some α L.
Proof.
This follows from the fact that all extensions of fields of characteristic
zero are separable.
We have previously seen that
Q
(
2,
3
)
/Q
is a simple extension, but that
is of course true from this theorem. A more interesting example would be one in
which this fails. We will need a field with non-zero characteristic.
Example. Let
L
=
F
p
(
s, u
), the fraction field of
F
p
[
s, u
]. Let
K
=
F
p
(
s
p
, u
p
).
We have L/K. We want to show this is not simple.
If
α L
, then
α
p
K
. So
α
is a root of
t
p
α
p
K
[
t
]. Thus the minimal
polynomial
P
α
has degree at most
p
. So [
K
(
α
) :
K
] =
deg P
α
p
. On the other
hand, we have [
L
:
K
] =
p
2
, since
{s
i
u
j
: 0
i, j < p}
is a basis. So for any
α
,
we have
K
(
α
)
=
L
. So
L/K
is not a simple extension. This then implies
L/K
is not separable.
At this point, one might suspect that all fields with positive characteristic
are not separable. This is not true by considering a rather silly example.
Example. Consider
K
=
F
2
and
L
=
F
2
[
s
]
/s
2
+
s
+ 1
. We can check manually
that
s
2
+
s
+ 1 has no roots and hence irreducible. So
L
is a field. So
L/F
2
is a
finite extension. Note that L only has 4 elements.
Now if
α L \ F
2
, and
P
α
is the minimal polynomial of
α
over
F
2
, then
P
α
| t
2
+ t + 1. So P
α
is separable as a polynomial. So L/F
2
is separable.
In fact, we have
Proposition. Let
L/K
be an extension of finite fields. Then the extension is
separable.
Proof.
Let the characteristic of the fields be
p
. Suppose the extension were not
separable. Then there is some non-separable element
α L
. Then its minimal
polynomial must be of the form P
α
=
P
a
i
t
pi
.
Now note that the map
K K
given by
x 7→ x
p
is injective, hence surjective.
So we can write a
i
= b
p
i
for all i. Then we have
P
α
=
X
a
i
t
pi
=
X
b
i
t
i
p
,
and so P
α
is not irreducible, which is a contradiction.
2.7 Normal extensions
We are almost there. We will now move on to study normal extensions. Normal
extensions are very closely related to Galois extensions. In fact, we will show
that if an extension is normal and separable, then it is Galois. The advantage
of introducing the idea of normality is that normality is a much more concrete
definition to work with. It is much easier to check if an extension is normal than
to check if
|Aut
K
(
L
)
|
= [
K
:
L
]. In particular, we will shortly prove that the
splitting field of any polynomial is normal.
This is an important result, since we are going to use the splitting field to
study the roots of a polynomial, and since we mostly care about polynomials
over
Q
, this means all these splitting fields are automatically Galois extensions
of Q.
It is not immediately obvious why these extensions are called “normal” (just
like most other names in Galois theory). We will later see that normal extensions
are extensions that correspond to normal subgroups, in some precise sense given
by the fundamental theorem of Galois theory.
Definition (Normal extension). Let
K L
be an algebraic extension. We say
L/K
is normal if for all
α L
, the minimal polynomial of
α
over
K
splits over
L.
In other words, given any minimal polynomial, L should have all its roots.
Example. The extension
Q
(
3
2
)
/Q
is not normal since the minimal polynomial
t
3
2 does not split over Q(
3
2).
In some sense, extensions that are not “normal” are missing something. This
is somewhat similar to how Galois extensions work. Before we go deeper into
this, we need a lemma.
Lemma. Let
L/F/K
be finite extensions, and
¯
K
is the algebraic closure of
K
.
Then any ψ Hom
K
(F,
¯
K) extends to some φ Hom
K
(L,
¯
K).
Proof.
Let
ψ Hom
K
(
F,
¯
K
). If
F
=
L
, then the statement is trivial. So assume
L = F .
Pick
α L \ F
. Let
q
α
F
[
t
] be the minimal polynomial of
α
over
F
.
Consider
ψ
(
q
α
)
¯
K
[
t
]. Let
β
be any root of
q
α
, which exists since
¯
K
is
algebraically closed. Then as before, we can extend
ψ
to
F
(
α
) by sending
α
to
β. More explicitly, we send
N
X
i=0
a
i
α
i
7→
X
ψ(a
i
)β
i
,
which is well-defined since any polynomial relation satisfied by
α
in
F
is also
satisfied by β.
Repeat this process finitely many times to get some element in
Hom
K
(
L,
¯
K
).
We will use this lemma to characterize normal extensions.
Theorem. Let
L/K
be a finite extension. Then
L/K
is a normal extension if
and only if L is the splitting field of some f K[t].
Proof.
Suppose
L/K
is normal. Since
L
is finite, let
L
=
K
(
α
1
, ··· , α
n
) for some
α
i
L
. Let
P
α
i
be the minimal polynomial of
α
i
over
K
. Take
f
=
P
α
1
···P
α
n
.
Since
L/K
is normal, each
P
α
i
splits over
L
. So
f
splits over
L
, and
L
is a
splitting field of f.
For the other direction, suppose that
L
is the splitting field of some
f K
[
t
].
First we wlog assume
L
¯
K
. This is possible since the natural injection
K
¯
K
extends to some
φ
:
L
¯
K
by our previous lemma, and we can replace
L
with
φ(L).
Now suppose
β L
, and let
P
β
be its minimal polynomial. Let
β
be another
root. We want to show it lives in L.
Now consider
K
(
β
). By the proof of the lemma, we can produce an embedding
ι
:
K
(
β
)
¯
K
that sends
β
to
β
. By the lemma again, this extends to an
embedding of
L
into
¯
K
. But any such embedding must send a root of
f
to a
root of
f
. So it must send
L
to
L
. In particular,
ι
(
β
) =
β
L
. So
P
β
splits
over L.
This allows us to identify normal extensions easily. The following theorem
then allows us to identify Galois extensions using this convenient tool.
Theorem. Let L/K be a finite extension. Then the following are equivalent:
(i) L/K is a Galois extension.
(ii) L/K is separable and normal.
(iii) L
=
K
(
α
1
, ··· , α
n
) and
P
α
i
, the minimal polynomial of
α
i
over
K
, is
separable and splits over L for all i.
Proof.
(i)
(ii): Suppose
L/K
is a Galois extension. Then by definition, this
means
|Hom
K
(L, L)| = |Aut
K
(L)| = [L : K].
To show that
L/K
is separable, recall that we proved that an extension is
separable if and only if there is some
E
such that
|Hom
K
(
L, E
)
|
= [
L
:
K
].
In this case, just pick
E
=
L
. Then we know that the extension is separable.
To check normality, let
α L
, and let
P
α
be its minimal polynomial over
K. We know that
|Root
P
α
(L)| = |Hom
K
(K[t]/P
α
, L)| = |Hom
K
(K(α), L)|.
But since
|Hom
K
(
L, L
)
|
= [
L
:
K
] and
K
(
α
) is a subfield of
L
, this implies
|Hom
K
(K(α), L)| = [K(α) : K] = deg P
α
.
Hence we know that
|Root
P
α
(L)| = deg P
α
.
So P
α
splits over L.
(ii)
(iii): Just pick
α
1
, ··· , α
n
such that
L
=
K
(
α
1
, ··· , α
n
). Then these
polynomials are separable since the extension is separable, and they split
since
L/K
is normal. In fact, by the primitive element theorem, we can
pick these such that n = 1.
(iii)
(i): Since
L
=
K
(
α
1
, ··· , α
n
) and the minimal polynomials
P
α
i
over
K
are separable, by a previous theorem, there are some extension
E
of K such that
|Hom
K
(L, E)| = [L : K].
To simplify notation, we first replace
L
with its image inside
E
under some
K
-homomorphism
L E
, which exists since
|Hom
K
(
L, E
)
|
= [
L
:
K
]
>
0.
So we can assume L E.
We now claim that the inclusion
Hom
K
(L, L) Hom
K
(L, E)
is a surjection, hence a bijection. Indeed, if
φ
:
L E
, then
φ
takes
α
i
to
φ
(
α
i
), which is a root of
P
α
i
. Since
P
α
i
splits over
L
, we know
φ
(
α
i
)
L
for all i. Since L is generated by these α
i
, it follows that φ(L) L.
Thus, we have
[L : K] = |Hom
K
(L, E)| = |Hom
K
(L, L)|,
and the extension is Galois.
From this, it follows that if
L/K
is Galois, and we have an intermediate field
K F L, then L/F is also Galois.
Corollary. Let
K
be a field and
f K
[
t
] be a separable polynomial. Then the
splitting field of f is Galois.
This is one of the most crucial examples.
2.8 The fundamental theorem of Galois theory
Finally, we can get to the fundamental theorem of Galois theory. Roughly, given
a Galois extension
K L
, the fundamental theorem tell us there is a one-to-one
correspondence between intermediate field extensions
K F L
and subgroups
of the automorphism group Gal(L/K).
Given an intermediate field
F
, we can obtain a subgroup of
Gal
(
L/K
) by
looking at the automorphisms that fix
F
. To go the other way round, given a
subgroup
H Gal
(
L/K
), we can obtain a corresponding field by looking at the
field of elements that are fixed by everything in
H
. This is known as the fixed
field, and can in general be defined even for non-Galois extensions.
Definition (Fixed field). Let
L/K
be a field extension,
H Aut
K
(
L
) a
subgroup. We define the fixed field of H as
L
H
= {α L : φ(α) = α for all φ H}.
It is easy to see that L
H
is an intermediate field K L
H
L.
Before we get to the fundamental theorem, we first prove Artin’s lemma.
This in fact proves part of the results in the fundamental theorem, but is also
useful on its own right.
Lemma (Artin’s lemma). Let
L/K
be a field extension and
H Aut
K
(
L
) a
finite subgroup. Then L/L
H
is a Galois extension with Aut
L
H
(L) = H.
Note that we are not assuming that L/K is Galois, or even finite!
Proof. Pick any α L. We set
{α
1
, ··· , α
n
} = {φ(α) : φ H},
where
α
i
are distinct. Here we are allowing for the possibility that
φ
(
α
) =
ψ
(
α
)
for some distinct φ, ψ H.
By definition, we clearly have n < |H|. Let
f =
n
Y
1
(t α
i
) L[t].
We know that any
φ H
gives an homomorphism
L
[
t
]
L
[
t
], and any such
map fixes
f
because
φ
just permutes the
α
i
. Thus, the coefficients of
f
are in
L
H
, and thus f L
H
[t].
Since
id H
, we know that
f
(
α
) = 0. So
α
is algebraic over
L
H
. Moreover,
if q
α
is the minimal polynomial of α over L
H
, then q
α
| f in L
H
[t]. Hence
[L
H
(α) : L
H
] = deg q
α
deg f |H|.
Further, we know that
f
has distinct roots. So
q
α
is separable, and so
α
is
separable. So it follows that L/L
H
is a separable extension.
We next show that
L/L
H
is simple. This doesn’t immediately follow from
the primitive element theorem, because we don’t know it is a finite extension
yet, but we can still apply the theorem cleverly.
Pick
α L
such that [
L
H
(
α
) :
L
H
] is maximal. This is possible since
[L
H
(α) : L
H
] is bounded by |H|. The claim is that L = L
H
(α).
We pick an arbitrary
β L
, and will show that this is in
L
H
(
α
). By the
above arguments,
L
H
L
H
(
α, β
) is a finite separable extension. So by the
primitive element theorem, there is some
λ L
such that
L
H
(
α, β
) =
L
H
(
λ
).
Note that we must have
[L
H
(λ) : L
H
] [L
H
(α) : L
H
].
By maximality of [
L
H
(
α
) :
L
H
], we must have equality. So
L
H
(
λ
) =
L
H
(
α
). So
β L
H
(α). So L = L
H
(α).
Finally, we show it is a Galois extension. Let L = L
H
(α). Then
[L : L
H
] = [L
H
(α) : L
H
] |H| |Aut
L
H
(L)|
Recall that we have previously shown that for any extension
L/L
H
, we have
|Aut
L
H
(L)| [L : L
H
]. Hence we must have equality above. So
[L : L
H
] = |Aut
L
H
(L)|.
So the extension is Galois. Also, since we know that
H Aut
L
H
(
L
), we must
have H = Aut
L
H
(L).
Theorem. Let
L/K
be a finite field extension. Then
L/K
is Galois if and only
if L
H
= K, where H = Aut
K
(L).
Proof.
(
) Suppose
L/K
is a Galois extension. We want to show
L
H
=
K
.
Using Artin’s lemma (and the definition of H), we have
[L : K] = |Aut
K
(L)| = |H| = |Aut
L
H
(L)| = [L : L
H
]
So [L : K] = [L : L
H
]. So we must have L
H
= K.
() By the lemma, K = L
H
L is Galois.
This is an important theorem. Given a Galois extension
L/K
, this gives us
a very useful test of when elements of
α L
are in fact in
K
. We will use this a
lot.
Finally, we get to the fundamental theorem.
Theorem (Fundamental theorem of Galois theory). Assume
L/K
is a (finite)
Galois extension. Then
(i) There is a one-to-one correspondence
H Aut
K
(L) intermediate fields K F L.
This is given by the maps
H 7→ L
H
and
F 7→ Aut
F
(
L
) respectively.
Moreover, |Aut
K
(L) : H| = [L
H
: K].
(ii) H Aut
K
(
L
) is normal (as a subgroup) if and only if
L
H
/K
is a normal
extension if and only if L
H
/K is a Galois extension.
(iii)
If
H Aut
K
(
L
), then the map
Aut
K
(
L
)
Aut
K
(
L
H
) by the restriction
map is well-defined and surjective with kernel isomorphic to H, i.e.
Aut
K
(L)
H
= Aut
K
(L
H
).
Proof. Note that since L/K is a Galois extension, we know
|Aut
K
(L)| = |Hom
K
(L, L)| = [L : K],
By a previous theorem, for any intermediate field
K F L
, we know
|Hom
K
(
F, L
)
|
= [
F
:
K
] and the restriction map
Hom
K
(
L, L
)
Hom
K
(
F, L
)
is surjective.
(i)
The maps are already well-defined, so we just have to show that the maps
are inverses to each other. By Artin’s lemma, we know that
H
=
Aut
L
H
(
L
),
and since
L/F
is a Galois extension, the previous theorem tells that
L
Aut
F
(L)
=
F
. So they are indeed inverses. The formula relating the index
and the degree follows from Artin’s lemma.
(ii)
Note that for every
φ Aut
K
(
L
), we have that
L
ϕHϕ
1
=
φL
H
, since
α L
ϕHϕ
1
iff
φ
(
ψ
(
φ
1
(
α
))) =
α
for all
ψ H
iff
ψ
(
φ
1
(
α
)) =
φ
1
(
α
)
for all ψ H iff α φL
H
. Hence H is a normal subgroup if and only if
φ(L
H
) = L
H
for all φ Aut
K
(L). ()
Assume (
). We want to first show that
Hom
K
(
L
H
, L
H
) =
Hom
K
(
L
H
, L
).
Let
ψ Hom
K
(
L
H
, L
). Then by the surjectivity of the restriction map
Hom
K
(
L, L
)
Hom
K
(
L
H
, L
),
ψ
must be the restriction of some
˜
ψ
Hom
K
(
L, L
). So
˜
ψ
fixes
L
H
by (
). So
ψ
sends
L
H
to
L
H
. So
ψ
Hom
K
(L
H
, L
H
). So we have
|Aut
K
(L
H
)| = |Hom
K
(L
H
, L
H
)| = |Hom
K
(L
H
, L)| = [L
H
: K].
So L
H
/K is Galois, and hence normal.
Now suppose
L
H
/K
is a normal extension. We want to show this implies
(
). Pick any
α L
H
and
φ Aut
K
(
L
). Let
P
α
be the minimal polynomial
of
α
over
K
. So
φ
(
α
) is a root of
P
α
(since
φ
fixes
P
α
K
, and hence
maps roots to roots). Since
L
H
/K
is normal,
P
α
splits over
L
H
. This
implies that φ(α) L
H
. So φ(L
H
) = L
H
.
Hence,
H
is a normal subgroup if and only if
φ
(
L
H
) =
L
H
if and only if
L
H
/K is a Galois extension.
(iii)
Suppose
H
is normal. We know that
Aut
K
(
L
) =
Hom
K
(
L, L
) restricts
to
Hom
K
(
L
H
, L
) surjectively. To show that we in fact have restriction
to
Aut
K
(
L
H
), by the proof above, we know that
φ
(
L
H
) =
L
H
for all
φ Aut
K
(
L
H
). So this does restrict to an automorphism of
L
H
. In other
words, the map
Aut
K
(
L
)
Aut
K
(
L
H
) is well-defined. It is easy to see
this is a group homomorphism.
Finally, we have to calculate the kernel of this homomorphism. Let
E
be the kernel. Then by definition,
E H
. So it suffices to show that
|E|
=
|H|
. By surjectivity of the map and the first isomorphism theorem
of groups, we have
|Aut
K
(L)|
|E|
= |Aut
K
(L
H
)| = [L
H
: K] =
[L : K]
[L : L
H
]
=
|Aut
K
(L)|
|H|
,
noting that
L
H
/K
and
L/K
are both Galois extensions, and
|H|
= [
L
H
:
K] by Artin’s lemma. So |E| = |H|. So we must have E = H.
Example. Let
p
be an odd prime, and
ζ
p
be a primitive
p
th root of unity. Given
a (square-free) integer
n
, when is
n
in
Q
(
ζ
p
)? We know that
n Q
(
ζ
p
) if and
only if
Q
(
n
)
Q
(
ζ
p
). Moreover, [
Q
(
n
) :
Q
] = 2, i.e.
Q
(
n
) is a quadratic
extension.
We will later show that
Gal
(
Q
(
ζ
p
)
/Q
)
=
(
Z/pZ
)
=
C
p1
. Then by the
fundamental theorem of Galois theory, quadratic extensions contained in
Q
(
ζ
p
)
correspond to index 2-subgroups of
Gal
(
Q
(
ζ
p
)
/Q
). By general group theory,
there is exactly one such subgroup. So there is exactly one square-free
n
such
that
Q
(
n
)
Q
(
ζ
p
) (since all quadratic extensions are of the form
Q
(
n
)),
given by the fixed field of the index 2 subgroup of (Z/pZ)
.
Now we shall try to find some square root lying in
Q
(
ζ
p
). We will not fully
justify the derivation, since we can just square the resulting number to see that
it is correct. We know the general element of Q(ζ
p
) looks like
p1
X
k=0
c
k
ζ
k
p
.
We know
Gal
(
Q
(
ζ
p
)
/Q
)
=
(
Z/pZ
)
acts by sending
ζ
p
7→ ζ
n
p
for each
n
(
Z/pZ
)
,
and the index 2 subgroup consists of the quadratic residues. Thus, if an element
is fixed under the action of the quadratic residues, the quadratic residue powers
all have the same coefficient, and similarly for the non-residue powers.
If we wanted this to be a square root, then the action of the remaining
elements of
Gal
(
Q
(
ζ
p
)
/Q
) should negate this object. Since these elements swap
the residues and the non-residues, we would want to have something like
c
k
= 1 if
k
is a quadratic residue, and
1 if it is a non-residue, which is just the Legendre
symbol! So we are led to try to square
τ =
p1
X
k=1
k
p
ζ
k
p
.
It is an exercise in the Number Theory example sheet to show that the square
of this is in fact
τ
2
=
1
p
p.
So we have
p Q(ζ
p
) if p 1 (mod 4), and
p Q(ζ
p
) if p 3 (mod 4).
2.9 Finite fields
We’ll have a slight digression and look at finite fields. We adopt the notation
where
p
is always a prime number, and
Z
p
=
Z/p
. It turns out finite fields are
rather simple, as described in the lemma below:
Lemma. Let K be a finite field with q = |K| element. Then
(i) q = p
d
for some d N, where p = char K > 0.
(ii)
Let
f
=
t
q
t
. Then
f
(
α
) = 0 for all
α K
. Moreover,
K
is the splitting
field of f over F
p
.
This means that a finite field is completely determined by the number of
elements.
Proof.
(i)
Consider the set
{m·
1
K
}
mZ
, where 1
K
is the unit in
K
and
m·
represents
repeated addition. We can identify this with
F
p
. So we have the extension
F
p
K. Let d = [K : F
p
]. Then q = |K| = p
d
.
(ii)
Note that
K
=
K \ {
0
}
is a finite multiplicative group with order
q
1.
Then by Lagrange’s theorem, α
q1
= 1 for all α K
. So α
q
α = 0 for
all α = 0. The α = 0 case is trivial.
Now every element in
K
is a root of
f
. So we need to check that all roots
of
f
are in
K
. Note that the derivative
f
=
qt
q1
1 =
1 (since
q
is a
power of the characteristic). So
f
(
α
) =
1
= 0 for all
α K
. So
f
and
f
have no common roots. So
f
has no repeated roots. So
K
contains
q
distinct roots of f. So K is a splitting field.
Lemma. Let q = p
d
, q
= p
d
, where d, d
N. Then
(i)
There is a finite field
K
with exactly
q
elements, which is unique up to
isomorphism. We write this as F
q
.
(ii) We can embed F
q
F
q
iff d | d
.
Proof.
(i)
Let
f
=
t
q
t
, and let
K
be a splitting field of
f
over
F
p
. Let
L
=
Root
f
(
K
). The objective is to show that
L
=
K
. Then we will have
|K|
=
|L|
=
|Root
f
(
K
)
|
=
deg f
=
q
, because the proof of the previous
lemma shows that f has no repeated roots.
To show that
L
=
K
, by definition, we have
L K
. So we need to show
every element in
K
is in
L
. We do so by showing that
L
itself is a field.
Then since
L
contains all the roots of
f
and is a subfield of the splitting
field K, we must have K = L.
It is straightforward to show that L is a field: if α, β L, then
(α + β)
q
= α
q
+ β
q
= α + β.
So α + β L. Similarly, we have
(αβ)
q
= α
q
β
q
= αβ.
So αβ L. Also, we have
(α
1
)
q
= (α
q
)
1
= α
1
.
So α
1
L. So L is in fact a field.
Since any field of size
q
is a splitting field of
f
, and splitting fields are
unique to isomorphism, we know that K is unique.
(ii)
Suppose
F
q
F
q
. Then let
n
= [
F
q
:
F
q
]. So
q
=
q
n
. So
d
=
nd
. So
d | d
.
On the other hand, suppose
d | d
. Let
d
=
dn
. We let
f
=
t
q
t
. Then
for any α F
q
, we have
f(α) = α
q
α = α
q
n
α = (···((α
q
)
q
)
q
···)
q
α = α α = 0.
Since
F
q
is the splitting field of
f
, all roots of
f
are in
F
q
. So we know
that F
q
F
q
.
Note that if
¯
F
p
is the algebraic closure of
F
p
, then
F
q
¯
F
p
for every
q
=
p
d
.
We then have
[
kN
F
p
k
=
¯
F
p
,
because any α
¯
F
p
is algebraic over F
p
, and so belongs to some F
q
.
Definition. Consider the extension
F
q
n
/F
q
, where
q
is a power of
p
. The
Frobenius Fr
q
: F
q
n
F
q
n
is defined by α 7→ α
q
.
This is a homomorphism precisely because the field is of characteristic zero.
In fact, Fr
q
Aut
F
q
(F
q
n
), since α
q
= α for all α F
q
.
The following two theorems tells us why we care about the Frobenius.
Theorem. Consider
F
q
n
/F
q
. Then
Fr
q
is an element of order
n
as an element
of Aut
F
q
(F
q
n
).
Proof.
For all
α F
q
n
, we have
Fr
n
q
(
α
) =
α
q
n
=
α
. So the order of
Fr
q
divides
n.
If m | n, then the set
{α F
q
n
: Fr
m
q
(α) = α} = {α F
q
n
: α
q
m
= α} = F
q
m
.
So if m is the order of Fr
q
, then F
q
m
= F
q
n
. So m = n.
Theorem. The extension
F
q
n
/F
q
is Galois with Galois group
Gal
(
F
q
n
/F
q
) =
Aut
F
q
(F
q
n
)
=
Z/nZ, generated by Fr
q
.
Proof.
The multiplicative group
F
q
n
=
F
q
n
\ {
0
}
is finite. We have previously
seen that multiplicative groups of finite fields are cyclic. So let
α
be a generator
of this group. Then
F
q
n
=
F
q
(
α
). Let
P
α
be the minimal polynomial of
α
over
F
q
. Then since Aut
F
q
(F
q
n
) has an element of order n, we get
n |Aut
F
q
(F
q
n
)| = |Hom
F
q
(F
q
(α), F
q
n
)|.
Since F
q
(α) is generated by one element, we know
|Hom
F
q
(F
q
(α), F
q
n
)| = |Root
P
α
(F
q
n
)|
So we have
n |Root
P
α
(F
q
n
)| deg P
α
= [F
q
n
: F
q
] = n.
So we know that
|Aut
F
q
(F
q
n
)| = [F
q
n
: F
q
] = n.
So F
q
n
/F
q
is a Galois extension.
Since
|Aut
F
q
(
F
q
n
)
|
, it has to be generated by
Fr
q
, since this has order
n
. In
particular, this group is cyclic.
We see that finite fields are rather nice there is exactly one field of order
p
d
for each
d
and prime
p
, and these are all of the finite fields. All extensions
are Galois and the Galois group is a simple cyclic group.
Example. Consider F
4
/F
2
. We can write
F
2
= {0, 1} F
4
= {0, 1, α, α
2
},
where α is a generator of F
4
. Define φ Aut
F
2
(F
4
) by φ(α) = α
2
. Then
Aut
F
2
(F
4
) = {id, φ}
since it has order 2.
Note that we can also define the Frobenius
Fr
p
:
¯
F
p
¯
F
p
, where
α 7→ α
p
.
Then
F
p
d
is the elements of
¯
F
p
fixed by
Fr
d
p
. So we can recover this subfield by
just looking at the Frobenius.
3 Solutions to polynomial equations
We have now proved the fundamental theorem of Galois theory, and this gives a
one-to-one correspondence between (intermediate) field extensions and subgroups
of the Galois group. That is our first goal achieved. Our next big goal is to use
this Galois correspondence to show that, in general, polynomials of degree 5 or
more cannot be solved by radicals.
First of all, we want to make this notion of “solving by radicals” precise.
We all know what this means if we are working over
Q
, but we need to be very
precise when working with arbitrary fields.
For example, we know that the polynomial
f
=
t
3
5
Q
[
t
] can be “solved
by radicals”. In this case, we have
Root
f
(C) = {
3
5, µ
3
5, µ
2
3
5},
where
µ
3
= 1
, µ
= 1. In general fields, we want to properly define the analogues
of µ and
3
5.
These will correspond to two different concepts. The first is cyclotomic
extensions, where the extension adds the analogues of
µ
, and the second is
Kummer extensions, where we add things like
3
5.
Then, we would say a polynomial is soluble by radicals if the splitting field
of the polynomial can be obtained by repeatedly taking cyclotomic and Kummer
extensions.
3.1 Cyclotomic extensions
Definition (Cyclotomic extension). For a field
K
, we define the
n
th cyclotomic
extension to be the splitting field of t
n
1.
Note that if
K
is a field and
L
is the
n
th cyclotomic extension, then
Root
t
n
1
(
L
) is a subgroup of multiplicative group
L
=
L \ {
0
}
. Since this is a
finite subgroup of L
, it is a cyclic group.
Moreover, if
char K
= 0 or 0
< char K n
, then (
t
n
1)
=
nt
n1
and this
has no common roots with
t
n
1. So
t
n
1 has no repeated roots. In other
words, t
n
1 has n distinct roots. So as a group,
Root
t
n
1
(L)
=
Z/nZ.
In particular, this group has at least one element µ of order n.
Definition (Primitive root of unity). The
n
th primitive root of unity is an
element of order n in Root
t
n
1
(L).
These elements correspond to the elements of the multiplicative group of
units in Z/nZ, written (Z/nZ)
×
.
The next theorem tells us some interesting information about these roots
and some related polynomials.
Theorem. For each
d N
, there exists a
d
th cyclotomic monic polynomial
φ
d
Z[t] satisfying:
(i) For each n N, we have
t
n
1 =
Y
d|n
φ
d
.
(ii) Assume char K = 0 or 0 < char K n. Then
Root
ϕ
n
(L) = {nth primitive roots of unity}.
Note that here we have an abuse of notation, since
φ
n
is a polynomial in
Z
[
t
], not
K
[
t
], but we can just use the canonical map
Z
[
t
]
K
[
t
] mapping
1 to 1 and t to t.
Proof.
We do induction on
n
to construct
φ
n
. When
n
= 1, let
φ
1
=
t
1. Then
(i) and (ii) hold in this case, trivially.
Assume now that (i) and (ii) hold for smaller values of n. Let
f =
Y
d|n,d<n
φ
d
.
By induction,
f Z
[
t
]. Moreover, if
d | n
and
d < n
, then
φ
d
|
(
t
n
1) because
(
t
d
1)
|
(
t
n
1). We would like to say that
f
also divides
t
n
1. However, we
have to be careful, since to make this conclusion, we need to show that
φ
d
and
φ
d
have no common roots for distinct d, d
| n (and d, , d
< n).
Indeed, by induction, φ
d
and φ
d
have no common roots because
Root
ϕ
d
(L) = {dth primitive roots of unity},
Root
ϕ
d
(L) = {d
th primitive roots of unity},
and these two sets are disjoint (or else the roots would not be primitive).
Therefore
φ
d
and
φ
d
have no common irreducible factors. Hence
f | t
n
1. So
we can write
t
n
1 = fφ
n
,
where
φ
n
Q
[
t
]. Since
f
is monic,
φ
n
has integer coefficients. So indeed
φ
n
Z[t]. So the first part is proven.
To prove the second part, note that by induction,
Root
f
(L) = {non-primitive nth roots of unit},
since all nth roots of unity are dth primitive roots of unity for some smaller d.
Since
fφ
n
=
t
n
1,
φ
n
contains the remaining, primitive
n
th roots of unit.
Since
t
n
1 has no repeated roots, we know that
φ
n
does not contain any extra
roots. So
Root
ϕ
n
(L) = {nth primitive roots of unity}.
These
φ
n
are what we use to “build up” the polynomials
t
n
1. These
will later serve as a technical tool to characterize the Galois group of the
n
th
cyclotomic extension of Q.
Before we an reach that, we first take a tiny step, and prove something that
works for arbitrary fields first.
Theorem. Let
K
be a field with
char K
= 0 or 0
< char K n
. Let
L
be the
n
th cyclotomic extension of
K
. Then
L/K
is a Galois extension, and there is an
injective homomorphism θ : Gal(L/K) (Z/nZ)
×
.
In addition, every irreducible factor of φ
n
(in K[t]) has degree [L : K].
The important thing about our theorem is the homomorphism
θ : Gal(L/K) (Z/nZ)
×
.
In general, we don’t necessarily know much about
Gal
(
L/K
), but the group
(
Z/nZ
)
×
is well-understood. In particular, we now know that
Gal
(
L/K
) is
abelian.
Proof. Let µ be an nth primitive root of unity. Then
Root
t
n
1
(L) = {1, µ, µ
2
, ··· , µ
n1
}
is a cyclic group of order
n
generated by
µ
. We first construct the homomorphism
θ
:
Aut
K
(
L
)
(
Z/nZ
)
×
as follows: for each
φ Aut
K
(
L
),
φ
is completely
determined by the value of
φ
(
µ
) since
L
=
K
(
µ
). Since
φ
is an automorphism, it
must take an
n
th primitive root of unity to another
n
th primitive root of unity.
So
φ
(
µ
) =
µ
i
for some
i
such that (
i, n
) = 1. Now let
θ
(
φ
) =
¯
i
(
Z/nZ
)
×
. Note
that this is well-defined since if µ
i
= µ
j
, then i j has to be a multiple of n.
Now it is easy to see that if
φ, ψ Aut
K
(
L
) are given by
φ
(
µ
) =
µ
i
, and
ψ
(
µ
) =
µ
j
, then
φ ψ
(
µ
) =
φ
(
µ
j
) =
µ
ij
. So
θ
(
φψ
) =
¯
ij
=
θ
(
φ
)
θ
(
ψ
). So
θ
is a
group homomorphism.
Now we check that
θ
is injective. If
θ
(
φ
) =
¯
1
(note that (
Z/nZ
)
×
is a
multiplicative group with unit 1), then φ(µ) = µ. So φ = id.
Now we show that
L/K
is Galois. Recall that
L
=
K
(
µ
), and let
P
µ
be
a minimal polynomial of
µ
over
K
. Since
µ
is a root of
t
n
1, we know that
P
µ
| t
n
1. Since
t
n
1 has no repeated roots,
P
µ
has no repeated roots. So
P
µ
is separable. Moreover,
P
µ
splits over
L
as
t
n
1 splits over
L
. So the extension
is separable and normal, and hence Galois.
Applying the previous theorem, each irreducible factor
g
of
φ
n
is a minimal
polynomial of some nth primitive root of unity, say λ. Then L = K(λ). So
deg g = deg P
λ
= [K(λ) : K] = [L : K].
Example. We can calculate the following in Q[t].
(i) φ
1
= t 1.
(ii) φ
2
= t + 1 since t
2
1 = φ
1
φ
2
.
(iii) φ
3
= t
2
+ t + 1.
(iv) φ
4
= t
2
+ 1.
These are rather expected. Now take
K
=
F
2
. Then 1 =
1. So we might be
able to further decompose these polynomials. For example,
t
+ 1 =
t
1 in
F
2
.
So we have
φ
4
= t
2
+ 1 = t
2
1 = φ
1
φ
2
.
So in
F
2
,
φ
4
is not irreducible. Similarly, if we have too much time, we can show
that
φ
15
= (t
4
+ t + 1)(t
4
+ t
3
+ 1).
So
φ
15
is not irreducible. However, they are irreducible over the rationals, as we
will soon see.
So far, we know
Gal
(
L/K
) is an abelian group, isomorphic to a subgroup of
(
Z/nZ
)
×
. However, we are greedy and we want to know more. The following
lemma tells us when this θ is an isomorphism.
Lemma. Under the notation and assumptions of the previous theorem,
φ
n
is
irreducible in K[t] if and only if θ is an isomorphism.
Proof.
(
) Suppose
φ
n
is irreducible. Recall that
Root
ϕ
n
(
L
) is exactly the
n
th
primitive roots of unity. So if
µ
is an
n
th primitive root of unity, then
P
µ
, the
minimal polynomial of
µ
over
K
is
φ
n
. In particular, if
λ
is also an
n
th primitive
root of unity, then
P
µ
=
P
λ
. This implies that there is some
φ
λ
Aut
K
(
L
) such
that φ
λ
(µ) = λ.
Now if
¯
i
(
Z/nZ
)
×
, then taking
λ
=
µ
i
, this shows that we have
φ
λ
Aut
K
(L) such that θ(φ
λ
) =
¯
i. So θ is surjective, and hence an isomorphism.
(
) Suppose that
θ
is an isomorphism. We will reverse the above argument
and show that all roots have the same minimal polynomial. Let
µ
be a
n
th
primitive root of unity, and pick
¯
i
(
Z/nZ
)
×
, and let
λ
=
µ
i
. Since
θ
is an
isomorphism, there is some
φ
λ
Aut
K
(
L
) such that
θ
(
φ
λ
) =
¯
i
, i.e.
φ
λ
(
µ
) =
µ
i
= λ. Then we must have P
µ
= P
λ
.
Since every
n
th primitive root of unity is of the form
µ
i
(with (
i, n
) = 1), this
implies that all
n
th primitive roots have the same minimal polynomial. Since
the roots of
φ
n
are all the
n
th primitive roots of unity, its irreducible factors are
exactly the minimal polynomials of the primitive roots. Moreover, φ
n
does not
have repeated roots. So φ
n
= P
µ
. In particular, φ
n
is irreducible.
We want to apply this lemma to the case of rational numbers. We want to
show that
θ
is an isomorphism. So we have to show that
φ
n
is irreducible in
Q[t].
Theorem. φ
n
is irreducible in Q[t]. In particular, it is also irreducible in Z[t].
Proof.
As before, this can be achieved by showing that all
n
th primitive roots
have the same minimal polynomial. Moreover, let
µ
be our favorite
n
th primitive
root. Then all other primitive roots
λ
are of the form
λ
=
µ
i
, where (
i, n
) = 1. By
the fundamental theorem of arithmetic, we can write
i
as a product
i
=
q
1
···q
n
.
Hence it suffices to show that for all primes
q n
, we have
P
µ
=
P
µ
q
. Noting
that µ
q
is also an nth primitive root, this gives
P
µ
= P
µ
q
1
= P
(µ
q
1
)
q
2
= P
µ
q
1
q
2
= ··· = P
µ
q
1
···q
r
= P
µ
i
.
So we now let
µ
be an
n
th primitive root,
P
µ
be its minimal polynomial. Since
µ is a root of φ
n
, we can write P
µ
| φ
n
inside Q[t]. So we can write
φ
n
= P
µ
R,
Since
φ
n
and
P
µ
are monic,
R
is also monic. By Gauss’ lemma, we must have
P
µ
, R Z[t].
Note that showing
P
µ
=
P
µ
q
is the same as showing
µ
q
is a root of
P
µ
, since
deg P
µ
=
deg P
µ
q
. So suppose it’s not. Since
µ
q
is an
n
th primitive root of unity,
it is a root of
φ
n
. So
µ
q
must be a root of
R
. Now let
S
=
R
(
t
q
). Then
µ
is a
root of S, and so P
µ
| S.
We now reduce mod
q
. For any polynomial
f Z
[
t
], we write the result of
reducing the coefficients mod
q
as
¯
f
. Then we have
¯
S
=
R(t
q
)
=
R(t)
q
. Since
¯
P
µ
divides
¯
S
(by Gauss’ lemma), we know
¯
P
µ
and
R(t)
have common roots. But
¯
φ
n
=
¯
P
µ
¯
R
, and so this implies
¯
φ
n
has repeated roots. This is impossible since
¯
φ
n
divides
t
n
1, and since
q n
, we know the derivative of
t
n
1 does not
vanish at the roots. So we are done.
Corollary. Let
K
=
Q
and
L
be the
n
th cyclotomic extension of
Q
. Then the
injection θ : Gal(L/Q) (Z/nZ)
×
is an isomorphism.
Example. Let
p
be a prime number, and
q
=
p
d
,
d N
. Consider
F
q
, a field
with
q
elements, and let
L
be the
n
th cyclotomic extension of
F
q
(where
p n
).
Then we have a homomorphism θ : Gal(L/F
q
) (Z/nZ)
×
.
We have previously shown that
Gal
(
L/F
q
) must be a cyclic group. So if
(
Z/nZ
)
×
is non-cyclic, then
θ
is not an isomorphism, and
φ
n
is not irreducible
in F
q
[t].
For example, take p = q = 7 and n = 8. Then
(Z/8Z)
×
= {
¯
1,
¯
3,
¯
5,
¯
7}
is not cyclic, because manual checking shows that there is no element of order 4.
Hence
θ
:
Gal
(
L/F
7
)
(
Z/
8
Z
)
×
is not an isomorphism, and
φ
8
is not irreducible
in F
7
[t].
3.2 Kummer extensions
We shall now consider a more general case, and study the splitting field of
t
n
λ K
[
t
]. As we have previously seen, we will need to make use of the
n
th
primitive roots of unity.
The definition of a Kummer extension will involve a bit more that it being
the splitting field of
t
n
λ
. So before we reach the definition, we first studying
some properties of an arbitrary splitting field of
t
n
λ
, and use this to motivate
the definition of a Kummer extension.
Definition (Cyclic extension). We say a Galois extension
L/K
is cyclic is
Gal(L/K) is a cyclic group.
Theorem. Let
K
be a field,
λ K
non-zero,
n N
,
char K
= 0 or 0
< char K
n. Let L be the splitting field of t
n
λ. Then
(i) L contains an nth primitive root of unity, say µ.
(ii) L/K
(
µ
) is a cyclic (and in particular Galois) extension with degree [
L
:
K(µ)] | n.
(iii) [L : K(µ)] = n if and only if t
n
λ is irreducible in K(µ)[t].
Proof.
(i)
Under our assumptions,
t
n
λ
and (
t
n
λ
)
=
nt
n1
have no common
roots in L. So t
n
λ has distinct roots in L, say α
1
, ··· , α
n
L.
It then follows by direct computation that
α
1
α
1
1
, α
2
α
1
1
, ··· , α
n
α
1
1
are
distinct roots of unity, i.e. roots of
t
n
1. Then one of these, say
µ
must
be an nth primitive root of unity.
(ii)
We know
L/K
(
µ
) is a Galois extension because it is the splitting field of
the separable polynomial t
n
λ.
To understand the Galois group, we need to know how this field exactly
looks like. We let
α
be any root of
t
n
λ
. Then the set of all roots can be
written as
{α, µα, µ
2
α, ··· , µ
n1
α}
Then
L = K(α
1
, ··· , α
n
) = K(µ, α) = K(µ)(α).
Thus, any element of
Gal
(
L/K
(
µ
)) is uniquely determined by what it sends
α
to, and any homomorphism must send
α
to one of the other roots of
t
n
λ, namely µ
i
α for some i.
Define a homomorphism
σ
:
Gal
(
L/K
(
µ
))
Z/nZ
that sends
φ
to the
corresponding i (as an element of Z/nZ, so that it is well-defined).
It is easy to see that σ is an injective group homomorphism. So we know
Gal
(
L/K
(
µ
)) is isomorphic to a subgroup of
Z/nZ
. Since the subgroup of
any cyclic group is cyclic, we know that
Gal
(
L/K
(
µ
)) is cyclic, and its size
is a factor of
n
by Lagrange’s theorem. Since
|Gal
(
L/K
(
µ
))
|
= [
L
:
K
(
µ
)]
by definition of a Galois extension, it follows that [L : K(µ)] divides n.
(iii)
We know that [
L
:
K
(
µ
)] = [
K
(
µ, α
) :
K
(
µ
)] =
deg q
α
. So [
L
:
K
(
µ
)] =
n
if and only if
deg q
α
=
n
. Since
q
α
is a factor of
t
n
λ
,
deg q
α
=
n
if and
only if
q
α
=
t
n
λ
. This is true if and only if
t
n
λ
is irreducible
K
(
µ
)[
t
].
So done.
Example. Consider
t
4
+ 2
Q
[
t
]. Let
µ
=
1
, which is a 4th primitive root
of unity. Now
t
4
+ 2 = (t α)(t + α)(t µα)(t + µα),
where
α
=
4
2
is one of the roots of
t
4
+ 2. Then we have the field extension
Q Q(µ) Q(µ, α), where Q(µ, α) is a splitting field of t
4
+ 2.
Since
2 ∈ Q
(
µ
), we know that
t
4
+ 2 is irreducible in
Q
(
µ
)[
t
] by looking at
the factorization above. So by our theorem,
Q
(
µ
)
Q
(
µ, α
) is a cyclic extension
of degree exactly 4.
Definition (Kummer extension). Let
K
be a field,
λ K
non-zero,
n N
,
char K
= 0 or 0
< char K n
. Suppose
K
contains an
n
th primitive root of
unity, and
L
is a splitting field of
t
n
λ
. If
deg
[
L
:
K
] =
n
, we say
L/K
is a
Kummer extension.
Note that we used to have extensions
K K
(
µ
)
L
. But if
K
already
contains a primitive root of unity, then
K
=
K
(
µ
). So we are left with the cyclic
extension K L.
To following technical lemma will be useful:
Lemma. Assume
L/K
is a field extension. Then
Hom
K
(
L, L
) is linearly in-
dependent. More concretely, let
λ
1
, ··· , λ
n
L
and
φ
1
, ··· , φ
n
Hom
K
(
L, L
)
distinct. Suppose for all α L, we have
λ
1
φ
1
(α) + ··· + λ
n
φ
n
(α) = 0.
Then λ
i
= 0 for all i.
Proof. We perform induction on n.
Suppose we have some λ
i
L and φ
i
Hom
K
(L, L) such that
λ
1
φ
1
(α) + ··· + λ
n
φ
n
(α) = 0.
The
n
= 1 case is trivial, since
λ
1
φ
1
= 0 implies
λ
1
= 0 (the zero homomorphism
does not fix K).
Otherwise, since the homomorphisms are distinct, pick
β L
such that
φ
1
(β) = φ
n
(β). Then we know that
λ
1
φ
1
(αβ) + ··· + λ
n
φ
n
(αβ) = 0
for all α L. Since φ
i
are homomorphisms, we can write this as
λ
1
φ
1
(α)φ
1
(β) + ··· + λ
n
φ
n
(α)φ
n
(β) = 0.
On the other hand, by just multiplying the original equation by φ
n
(β), we get
λ
1
φ
1
(α)φ
n
(β) + ··· + λ
n
φ
n
(α)φ
n
(β) = 0.
Subtracting the equations gives
λ
1
φ
1
(α)(φ
1
(β) φ
n
(β)) + ··· + λ
n1
φ
n1
(α)(φ
n1
(β) φ
n
(β)) = 0
for all
α L
. By induction,
λ
i
(
φ
i
(
β
)
φ
n
(
β
)) = 0 for all 1
i n
1. In
particular, since φ
1
(β) φ
n
(β) = 0, we have λ
1
= 0. Then we are left with
λ
2
φ
2
(α) + ··· + λ
n
φ
n
(α) = 0.
Then by induction again, we know that all coefficients are zero.
Theorem. Let
K
be a field,
n N
,
char K
= 0 or 0
< char K n
. Suppose
K
contains an
n
th primitive root of unity, and
L/K
is a cyclic extension of degree
[L : K] = n. Then L/K is a Kummer extension.
This is a rather useful result. If we look at the splitting field of a polynomial
t
n
λ
, even if the ground field includes the right roots of unity, a priori, this
doesn’t have to be a Kummer extension if it doesn’t have degree
n
. But we
previously showed that the extension must be cyclic. And so this theorem shows
that it is still a Kummer extension of some sort.
This is perhaps not too surprising. For example, if, say,
n
= 4 and
λ
is
secretly a square, then the splitting field of
t
4
λ
is just the splitting field of
t
2
λ.
Proof.
Our objective here is to find a clever
λ K
such that
L
is the splitting
field of t
n
λ. To do so, we will have to hunt for a root β of t
n
λ in L.
Pick
φ
a generator of
Gal
(
L/K
). We know that if
β
were a root of
t
n
λ
,
then
φ
(
β
) =
µ
1
β
for some primitive
n
th root of unity
µ
. Thus, we want to find
an element that satisfies such a property.
By the previous lemma, we can find some α L such that
β = α + µφ(α) + µ
2
φ
2
(α) + ··· + µ
n1
φ
n1
(α) = 0.
Then, noting that
φ
n
is the identity and
φ
fixes
µ K
, we see that
β
trivially
satisfies
φ(β) = φ(α) + µφ
2
α + ··· + µ
n1
φ
n
(α) = µ
1
β,
In particular, we know that φ(β) K(β).
Now pick
λ
=
β
n
. Then
φ
(
β
n
) =
µ
n
β
n
=
β
n
. So
φ
fixes
β
n
. Since
φ
generates Gal(L/K), we know all automorphisms of L/K fixes β
n
. So β
n
K.
Now the roots of
t
n
λ
are
β, µβ, ··· , µ
n1
β
. Since these are all in
β
, we
know K(β) is the splitting field of t
n
λ.
Finally, to show that
K
(
β
) =
L
, we observe that
id, φ|
K(β)
, . . . ,
φ
n
|
K(β)
are
distinct elements of
Aut
K
(
K
(
β
)) since they do different things to
β
. Recall our
previous theorem that
[K(β) : K] |Aut
K
(K(β))|.
So we know that n = [L : K] = [K(β) : K]. So L = K(β). So done.
Example. Consider
t
3
2
Q
[
t
], and
µ
a third primitive root of unity. Then
we have the extension
Q Q
(
µ
)
Q
(
µ,
3
2
). Then
Q Q
(
µ
) is a cyclotomic
extension of degree 2, and
Q
(
µ
)
Q
(
µ,
3
2
) is a Kummer extension of degree 3.
3.3 Radical extensions
We are going to put these together and look at radical extensions, which allows
us to characterize what it means to “solve a polynomial with radicals”.
Definition (Radical extension). A field extension
L/K
is radical if there is
some further extension E/L and with a sequence
K = E
0
E
1
··· E
r
= E,
such that each
E
i
E
i+1
is a cyclotomic or Kummer extension, i.e.
E
i+1
is a
splitting field of t
n
λ
i+1
over E
i
for some λ
i+1
E
i
.
Informally, we say
E
i+1
is obtained by adding the roots
n
p
λ
i+1
to
E
i
.
Hence we interpret a radical extension as an extension that only adds radicals.
Definition (Solubility by radicals). Let
K
be a field, and
f K
[
t
].
f
. We say
f is soluble by radicals if the splitting field of f is a radical extension of K.
This means that f can be solved by radicals of the form
n
λ
i
.
Let’s go back to our first lecture and describe what we’ve done in the language
we’ve developed in the course.
Example. As we have shown in lecture 1, any polynomial
f Q
[
t
] of degree at
most 4 can be solved by radicals.
For example, assume
deg f
= 3. So
f
=
t
3
+
at
2
+
bt
+
c
. Let
L
be the
splitting field of
f
. Recall we reduced the problem of “solving”
f
to the case
a
= 0 by the substitution
x 7→ x
a
3
. Then we found our
β, γ C
such that
each root
α
i
can be written as a linear combination of
β
and
γ
(and
µ
), i.e.
L Q(β, γ, µ).
Then we showed that
{β
3
, γ
3
} =
(
27c ±
p
(27c)
2
+ 4 × 27b
3
2
)
.
We now let
λ =
p
(27c)
2
+ 4 × 27b
3
.
Then we have the extensions
Q Q(λ) Q(λ, µ) Q(λ, µ, β),
and also
Q L Q(λ, µ, β).
Note that the first extension
Q Q
(
λ
) is a Kummer extension since it is a
splitting field of
t
2
λ
2
. Then
Q
(
λ
)
Q
(
λ, µ
) is the third cyclotomic extension.
Finally,
Q
(
λ, µ
)
Q
(
λ, µβ
) is a Kummer extension, the splitting field of
t
3
β
3
.
So Q L is a radical extension.
Let’s go back to the definition of a radical extension. We said
L/K
is radical
if there is a further extension
E/L
that satisfies certain nice properties. It would
be great if
E/K
is actually a Galois extensions. To show this, we first need a
technical lemma.
Lemma. Let
L/K
be a Galois extension,
char K
= 0,
γ L
and
F
the splitting
field of
t
n
γ
over
L
. Then there exists a further extension
E/F
such that
E/L
is radical and E/K is Galois.
Here we have the inclusions
K L F E,
where
K, L
and
F
are given and
E
is what we need to find. The idea of the proof
is that we just add in the “missing roots” to obtain
E
so that
E/K
is Galois,
and doing so only requires performing cyclotomic and Kummer extensions.
Proof.
Since we know that
L/K
is Galois, we would rather work in
K
than in
L
.
However, our
γ
is in
L
, not
K
. Hence we will employ a trick we’ve used before,
where we introduce a new polynomial
f
, and show that its coefficients are fixed
by
Gal
(
L/K
), and hence in
K
. Then we can look at the splitting field of
f
or
its close relatives.
Let
f =
Y
ϕGal(L/K)
(t
n
φ(γ)).
Each
φ Gal
(
L/K
) induces a homomorphism
L
[
t
]
L
[
t
]. Since each
φ Gal
(
L/K
) just rotates the roots of
f
around, we know that this induced
homomorphism fixes
f
. Since all automorphisms in
Gal
(
L/K
) fix the coefficients
of f , the coefficients must all be in K. So f K[t].
Now since
L/K
is Galois, we know that
L/K
is normal. So
L
is the splitting
field of some
g K
[
t
]. Let
E
be the splitting field of
fg
over
K
. Then
K E
is normal. Since the characteristic is zero, this is automatically separable. So
the extension K E is Galois.
We have to show that
L E
is a radical extension. We pick our fields as
follows:
E
0
= L
E
1
= splitting field of t
n
1 over E
0
E
2
= splitting field of t
n
γ over E
1
E
3
= splitting field of t
n
φ
1
(γ) over E
2
. . .
E
r
= E,
where we enumerate Gal(L/K) as {id, φ
1
, φ
2
, ···}.
We then have the sequence of extensions
L = E
0
E
1
E
2
··· E
r
Here
E
0
E
1
is a cyclotomic extension, and
E
1
E
2
,
E
2
E
3
etc. are
Kummer extensions since they contain enough roots of unity and are cyclic. By
construction, F E
2
. So F E.
Theorem. Suppose
L/K
is a radical extension and
char K
= 0. Then there is
an extension E/L such that E/K is Galois and there is a sequence
K = E
0
E
1
··· E,
where E
i
E
i+1
is cyclotomic or Kummer.
Proof. Note that this is equivalent to proving the following statement: Let
K = L
0
L
1
···L
s
be a sequence of cyclotomic or Kummer extensions. Then there exists an
extension
L
s
E
such that
K E
is Galois and can be written as a sequence
of cyclotomic or Kummer extensions.
We perform induction on s. The s = 0 case is trivial.
If
s >
0, then by induction, there is an extension
M/L
s1
such that
M/K
is
Galois and is a sequence of cyclotomic and Kummer extensions. Now
L
s
is a
splitting field of
t
n
γ
over
L
s1
for some
γ L
s1
. Let
F
be the splitting field
of
t
n
γ
over
M
. Then by the lemma and its proof, there exists an extension
E/M
that is a sequence of cyclotomic or Kummer extensions, and
E/K
is Galois.
K
L
s1
L
s
= L
s1
(
n
γ)
M
F = M(
n
γ)
E
However, we already know that
M/K
is a sequence of cyclotomic and Kummer
extensions. So
E/K
is a sequence of cyclotomic and Kummer extension. So
done.
3.4 Solubility of groups, extensions and polynomials
Let
f K
[
t
]. We defined the notion of solubility of
f
in terms of radical
extensions. However, can we decide whether
f
is soluble or not without resorting
to the definition? In particular, is it possible to decide whether its soluble by
just looking at
Gal
(
L/K
), where
L
is the splitting field of
f
over
K
? It would
be great if we could do so, since groups are easier to understand than fields.
The answer is yes. It turns out the solubility of
f
corresponds to the solubility
of
Gal
(
L/K
). Of course, we will have to first define what it means for a group
to be soluble. After that, we will find examples of polynomials
f
of degree at
least 5 such that
Gal
(
L/K
) is not soluble. In other words, there are polynomials
that cannot be solved by radicals.
Definition (Soluble group). A finite group
G
is soluble if there exists a sequence
of subgroups
G
r
= {1} ··· G
1
G
0
= G,
where G
i+1
is normal in G
i
and G
i
/G
i+1
is cyclic.
Example. Any finite abelian group is solvable by the structure theorem of finite
abelian groups:
G
=
Z
n
1
× ··· ×
Z
n
r
.
Example. Let
S
n
be the symmetric group of permutations of
n
letters. We
know that G
3
is soluble since
{1} A
3
S
3
,
where S
3
/A
3
=
Z/2 and A
3
/{0}
=
Z/3.
S
4
is also soluble by
{1} {e, (1 2)(3 4)} {e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} A
4
S
4
.
We can show that the quotients are
Z/
2
,
Z/
3
,
Z/
2
and
Z/
2
respectively.
How about
S
n
for higher
n
? It turns out they are no longer soluble for
n
5.
To prove this, we first need a lemma.
Lemma. Let G be a finite group. Then
(i) If G is soluble, then any subgroup of G is soluble.
(ii)
If
A G
is a normal subgroup, then
G
is soluble if and only if
A
and
G/A
are both soluble.
Proof.
(i) If G is soluble, then by definition, there is a sequence
G
r
= {1} ··· G
1
G
0
= G,
such that G
i+1
is normal in G
i
and G
i
/G
i+1
is cyclic.
Let
H
i
=
H G
i
. Note that
H
i+1
is just the kernel of the obvious homo-
morphism
H
i
G
i
/G
i+1
. So
H
i+1
H
i
. Also, by the first isomorphism
theorem, this gives an injective homomorphism
H
i
/H
i+1
G
i
/G
i+1
. So
H
i
/H
i+1
is cyclic. So H is soluble.
(ii)
(
) By (i), we know that
A
is solvable. To show the quotient is soluble,
by assumption, we have the sequence
G
r
= {1} ··· G
1
G
0
= G,
such that
G
i+1
is normal in
G
i
and
G
i
/G
i+1
is cyclic. We construct the
sequence for the quotient in the obvious way. We want to define
E
i
as the
quotient
G
i
/A
, but since
A
is not necessarily a subgroup of
E
, we instead
define
E
i
to be the image of quotient map
G
i
G/A
. Then we have a
sequence
E
r
= {1} ··· E
0
= G/A.
The quotient map induces a surjective homomorphism
G
i
/G
i+1
E
i
/E
i+1
, showing that E
i
/E
i+1
are cyclic.
() From the assumptions, we get the sequences
A
m
= {1} ··· A
0
= A
F
n
= A ··· F
0
= G
where each quotient is cyclic. So we get a sequence
A
m
= {1} A
1
··· A
0
= F
n
F
n1
··· F
0
= G,
and each quotient is cyclic. So done.
Example. We want to show that
S
n
is not soluble if
n
5. It is a well-known
fact that
A
n
is a simple non-abelian group, i.e. it has no non-trivial subgroup.
So A
n
is not soluble. So S
n
is not soluble.
The key observation in Galois theory is that solubility of polynomials is
related to solubility of the Galois group.
Definition (Soluble extension). A finite field extension
L/K
is soluble if there
is some extension L E such that K E is Galois and Gal(E/K) is soluble.
Note that this definition is rather like the definition of a radical extension,
since we do not require the extension itself to be “nice”, but just for there to be
a further extension that is nice. In fact, we will soon see they are the same.
Lemma. Let
L/K
be a Galois extension. Then
L/K
is soluble if and only if
Gal(L/K) is soluble.
This means that the whole purpose of extending to
E
is just to make it a
Galois group, and it isn’t used to introduce extra solubility.
Proof. () is clear from definition.
(
) By definition, there is some
E L
such that
E/K
is Galois and
Gal
(
E/K
) is soluble. By the fundamental theorem of Galois theory,
Gal
(
L/K
) is
a quotient of
Gal
(
E/K
). So by our previous lemma,
Gal
(
L/K
) is also soluble.
We now come to the main result of the lecture:
Theorem. Let
K
be a field with
char K
= 0, and
L/K
is a radical extension.
Then L/K is a soluble extension.
Proof.
We have already shown that if we have a radical extension
L/K
, then
there is a finite extension
K E
such that
K E
is a Galois extension, and
there is a sequence of cyclotomic or Kummer extensions
E
0
= K E
1
··· E
r
= E.
Let
G
i
=
Gal
(
E/E
i
). By the fundamental theorem of Galois theory, inclusion of
subfields induces an inclusion of subgroups
G
0
= Gal(E/K) G
1
··· G
r
= {1}.
In fact,
G
i
G
i+1
because
E
i
E
i+1
are Galois (since cyclotomic and Kummer
extensions are). So in fact we have
G
0
= Gal(E/K) G
1
··· G
r
= {1}.
Finally, note that by the fundamental theorem of Galois theory,
G
i
/G
i+1
= Gal(E
i+1
/E
i
).
We also know that the Galois groups of cyclotomic and Kummer extensions are
abelian. Since abelian groups are soluble, our previous lemma implies that
L/K
is soluble.
In fact, we will later show that the converse is also true. So an extension is
soluble if and only if it is radical.
Corollary. Let
K
be a field with
char K
= 0, and
f K
[
t
]. If
f
can be solved
by radicals, then
Gal
(
L/K
) is soluble, where
L
is the splitting field of
f
over
K
.
Again, we will later show that the converse is also true. However, to construct
polynomials that cannot be solved by radicals, this suffices. In fact, this corollary
is all we really need.
Proof.
We have seen that
L/K
is a Galois extension. By assumption,
L/K
is
thus a radical extension. By the theorem,
L/K
is also a soluble extension. So
Gal(L/K) is soluble.
To find an
f Q
[
t
] which cannot be solved by radicals, it suffices to find an
f
such that the Galois group is not soluble. We don’t know many non-soluble
groups so far. So in fact, we will find an f such that Gal(L/Q) = S
5
.
To do so, we want to relate Galois groups to permutation groups.
Lemma. Let
K
be a field,
f K
[
t
] of degree
n
with no repeated roots. Let
L
be the splitting field of
f
over
K
. Then
L/K
is Galois and there exist an
injective group homomorphism
Gal(L/K) S
n
.
Proof.
Let
Root
f
(
L
) =
{α
1
, ··· , α
n
}
. Let
P
α
i
be the minimal polynomial of
α
i
over
K
. Then
P
α
i
| f
implies that
P
α
i
is separable and splits over
L
. So
L/K
is
Galois.
Now each
φ Gal
(
L/K
) permutes the
α
i
, which gives a map
Gal
(
L/K
)
S
n
.
It is easy to show this is an injective group homomorphism.
Note that there is not a unique or naturally-defined injective group homo-
morphism to
S
n
. This homomorphism, obviously, depends on how we decide to
number our roots.
Example. Let
f
= (
t
2
2)(
t
2
3)
Q
[
t
]. Let
L
be the splitting field of
f
over
Q. Then the roots are
Root
f
(L) = {
2,
2,
3,
3}.
We label these roots as
α
1
, α
2
, α
3
, α
4
in order. Now note that
L
=
Q
(
2,
3
),
and thus [
L
:
Q
] = 4. Hence
|Gal
(
L/Q
)
|
= 4 as well. We can let
Gal
(
L/Q
) =
{id, ϕ, ψ, λ}, where
id(
2) =
2 id(
3) =
3
ϕ(
2) =
2 ϕ(
3) =
3
ψ(
2) =
2 ψ(
3) =
3
λ(
2) =
2 λ(
3) =
3
Then the image of Gal(L/Q) S
4
is given by
{e, (1 2), (3 4), (1 2)(3 4)}.
What we really want to know is if there are polynomials in which this map is
in fact an isomorphism, i.e. the Galois group is the symmetric group. If so, then
we can use this to produce a polynomial that is not soluble by polynomials.
To find this, we first note a group-theoretic fact.
Lemma. Let p be a prime, and σ S
p
have order p. Then σ is a p-cycle.
Proof. By IA Groups, we can decompose σ into a product of disjoint cycles:
σ = σ
1
···σ
r
.
Let σ
i
have order m
i
> 1. Again by IA Groups, we know that
p = order of σ = lcm(m
1
, ··· , m
r
).
Since
p
is a prime number, we know that
p
=
m
i
for all
i
. Hence we must have
r
= 1, since the cycles are disjoint and there are only
p
elements. So
σ
=
σ
1
.
Hence σ is indeed an p cycle.
We will use these to find an example where the Galois group is the symmetric
group. The conditions for this to happen are slightly awkward, but the necessity
of these will become apparent in the proof.
Theorem. Let
f Q
[
t
] be irreducible and
deg f
=
p
prime. Let
L C
be the
splitting field of f over Q. Let
Root
f
(L) = {α
1
, α
2
, ··· , α
p2
, α
p1
, α
p
}.
Suppose that
α
1
, α
2
, ··· , α
p2
are all real numbers, but
α
p1
and
α
p
are not.
In particular,
α
p1
=
¯α
p
. Then the homomorphism
β
:
Gal
(
L/Q
)
S
n
is an
isomorphism.
Proof.
From IA groups, we know that the cycles (1 2
··· p
) and (
p
1
p
)
generate the whole of
S
n
. So we show that these two are both in the image of
β
.
As
f
is irreducible, we know that
f
=
P
α
1
, the minimal polynomial of
α
1
over Q. Then
p = deg P
α
i
= [Q(α
1
) : Q].
By the tower law, this divides [
L
:
Q
], which is equal to
|Gal
(
L/Q
)
|
since the
extension is Galois. Since
p
divides the order of
Gal
(
L/Q
), by Cauchy’s theorem
of groups, there must be an element of
Gal
(
L/Q
) that is of order
p
. This maps
to an element σ im β of order exactly p. So σ is a p-cycle.
On the other hand, the isomorphism
C C
given by
z 7→ ¯z
restricted to
L
gives an automorphism in
Gal
(
L/Q
). This simply permutes
α
p1
and
α
p
, since
it fixes the real numbers and
α
p1
and
α
p
must be complex conjugate pairs. So
τ = (p 1 p) im β.
Now for every 1
i < p
, we know that
σ
i
again has order
p
, and hence
is a
p
-cycle. So if we change the labels of the roots
α
1
, ··· , α
p
and replace
σ
with
σ
i
, and then waffle something about combinatorics, we can assume
σ = (1 2 ··· p 1 p). So done.
Example. Let t
5
4t + 2 Q[t]. Let L be the splitting field of f over Q.
First note that
deg f
= 5 is a prime. Also, by Eisenstein’s criterion,
f
is
irreducible. By finding the local maximum and minimum points, we find that
f
has exactly three real roots. So by the theorem, there is an isomorphism
Gal(L/Q) S
5
. In other words, Gal(L/Q)
=
S
5
.
We know S
5
is not a soluble group. So f cannot be solved by radicals.
After spending 19 lectures, we have found an example of a polynomial that
cannot be solved by radicals. Finally.
Note that there are, of course, many examples of
f Q
[
t
] irreducible of
degree at least 5 that can be solved by radicals, such as f = t
5
2.
3.5 Insolubility of general equations of degree 5 or more
We now want to do something more interesting. Instead of looking at a particular
example, we want to say there is no general formula for solving polynomial
equations of degree 5 or above. First we want to define certain helpful notions.
Definition (Field of symmetric rational functions). Let
K
be a field,
L
=
K
(
x
1
, ··· , x
n
), the field of rational functions over
K
. Then there is an injective
homomorphism S
n
Aut
K
(L) given by permutations of x
i
.
We define the field of symmetric rational functions
F
=
L
S
n
to be the fixed
field of S
n
.
There are a few important symmetric rational functions that we care about
more.
Definition (Elementary symmetric polynomials). The elementary symmetric
polynomials are e
1
, e
2
, ··· , e
n
defined by
e
i
=
X
1l
1
<l
2
<···<l
i
n
x
1
···x
i
.
It is easy to see that
e
1
= x
1
+ x
2
+ ··· + x
n
e
2
= x
1
x
2
+ x
1
x
3
+ ··· + x
n1
x
n
e
n
= x
1
···x
n
.
Obviously, e
1
, ··· , e
n
F .
Theorem (Symmetric rational function theorem). Let
K
be a field,
L
=
K
(
x
1
, ··· , x
n
). Let
F
the field fixed by the automorphisms that permute the
x
i
.
Then
(i) L is the splitting field of
f = t
n
e
1
t
n1
+ ··· + (1)
n
e
n
over F .
(ii) F = L
S
n
L is a Galois group with Gal(L/F ) isomorphic to S
n
.
(iii) F = K(e
1
, ··· , e
n
).
Proof.
(i) In L[t], we have
f = (t x
1
) ···(t x
n
).
So L is the splitting field of f over F .
(ii) By Artin’s lemma, L/K is Galois and Gal(L/F )
=
S
n
.
(iii)
Let
E
=
K
(
e
1
, ··· , e
n
). Clearly,
E F
. Now
E L
is a Galois extension,
since L is the splitting field of f over E and f has no repeated roots.
By the fundamental theorem of Galois theory, since we have the Galois ex-
tensions
E F L
, we have
Gal
(
L/F
)
Gal
(
L/E
). So
S
n
Gal
(
L/E
).
However, we also know that
Gal
(
L/E
) is a subgroup of
S
n
, we must have
Gal(L/E) = Gal(L/F ) = S
n
. So we must have E = F .
Definition (General polynomial). Let
K
be a field,
u
1
, ··· , u
n
variables. The
general polynomial over K of degree n is
f = t
n
+ u
1
t
n1
+ ··· + u
n
.
Technically, this is a polynomial in the polynomial ring
K
(
u
1
, ··· , u
n
)[
t
]. How-
ever, we say this is the general polynomial over
K
be cause we tend to think of
these u
i
as representing actual elements of K.
We say the general polynomial over
K
of degree
n
can be solved by radicals
if f can be solved by radicals over K(u
1
, ··· , u
n
).
Example. The general polynomial of degree 2 over Q is
t
2
+ u
1
t + u
2
.
This can be solved by radicals because its roots are
u
1
±
p
u
2
1
4u
2
2
.
Theorem. Let
K
be a field with
char K
= 0. Then the general polynomial
polynomial over K of degree n cannot be solved by radicals if n 5.
Proof. Let
f = t
n
+ u
1
t
n1
+ ··· + u
n
.
be our general polynomial of degree
n
5. Let
N
be a splitting field of
f
over
K(u
1
, ··· , u
n
). Let
Root
f
(N) = {α
1
, ··· , α
n
}.
We know the roots are distinct because
f
is irreducible and the field has charac-
teristic 0. So we can write
f = (t α
1
) ···(t α
n
) N[t].
We can expand this to get
u
1
= (α
1
+ ··· + α
n
)
u
2
= α
1
α
2
+ α
1
α
3
+ ··· + α
n1
α
n
.
.
.
u
i
= (1)
i
(ith elementary symmetric polynomial in α
1
, ··· , α
n
).
Now let
x
1
, ··· , x
n
be new variables, and
e
i
the
i
th elementary symmetric
polynomial in
x
1
, ··· , x
n
. Let
L
=
K
(
x
1
, ··· , x
n
), and
F
=
K
(
e
1
, ··· , e
n
). We
know that F L is a Galois extension with Galois group isomorphic to S
n
.
We define a ring homomorphism
θ : K[u
1
, ··· , u
n
] K[e
1
, ··· , e
n
] K[x
1
, ··· , x
n
]
u
i
7→ (1)
i
e
i
.
This is our equations of u
i
in terms α
i
, but with x
i
instead of α
i
.
We want to show that
θ
is an isomorphism. Note that since the homomorphism
just renames
u
i
into
e
i
, the fact that
θ
is an isomorphism means there are no
“hidden relations” between the
e
i
. It is clear that
θ
is a surjection. So it suffices
to show θ is injective. Suppose θ(h) = 0. Then
h(e
1
, ··· , (1)
n
e
n
) = 0.
Since the x
i
are just arbitrary variables, we now replace x
i
with α
i
. So we get
h(e
1
(α
1
, ··· , α
n
), ··· , (1)
n
(e
n
(α
1
, ··· , α
n
))) = 0.
Using our expressions for u
i
in terms of e
i
, we have
h(u
1
, ··· , u
n
) = 0,
But
h
(
u
1
, ··· , u
n
) is just
h
itself. So
h
= 0. Hence
θ
is injective. So it is an
isomorphism. This in turns gives an isomorphism between
K(u
1
, ··· , u
n
) K(e
1
, ··· , e
n
) = F.
We can extend this to their polynomial rings to get isomorphisms between
K(u
1
, ··· , u
n
)[t] F [t].
In particular, this map sends our original f to
f 7→ t
n
e
1
t
n1
+ ··· + (1)
n
e
n
= g.
Thus, we get an isomorphism between the splitting field of
f
over
K
(
u
1
, ··· , u
n
)
and the splitting field g over F .
The splitting field of
f
over
K
(
u
1
, ··· , u
n
) is just
N
by definition. From the
symmetric rational function theorem, we know that the splitting field of
g
over
F is just L, and So N
=
L. So we have an isomorphism
Gal(N/K(u
1
, ··· , u
n
)) Gal(L/F )
=
S
n
.
Since S
n
is not soluble, f is not soluble.
This is our second main goal of the course, to prove that general polynomials
of degree 5 or more are not soluble by radicals.
Recall that we proved that all radical extensions are soluble. We now prove
the converse.
Theorem. Let
K
be a field with
char K
= 0. If
L/K
is a soluble extension,
then it is a radical extension.
Proof.
Let
L E
be such that
K E
is Galois and
Gal
(
E/K
) is soluble. We
can replace
L
with
E
, and assume that in fact
L/K
is a soluble Galois extension.
So there is a sequence of groups
{0} = G
r
··· G
1
G
0
= Gal(L/K)
such that G
i
/G
i+1
is cyclic.
By the fundamental theorem of Galois theory, we get a sequence of field
extension given by L
i
= L
G
i
:
K = L
0
··· L
r
= L.
Moreover, we know that
L
i
L
i+1
is a Galois extension with Galois group
Gal(L
i+1
/L
i
)
=
G
i
/G
i+1
. So Gal(L
i+1
/L
i
) is cyclic.
Let
n
= [
L
:
K
]. Recall that we proved a previous theorem that if
Gal
(
L
i+1
/L
i
) is cyclic, and
L
i
contains a primitive
k
th root of unity (with
k
= [
L
i+1
:
L
i
]), then
L
i
L
i+1
is a Kummer extension. However, we do not
know of
L
i
contains the right root of unity. Hence, the trick here is to add an
nth primitive root of unity to each field in the sequence.
Let
µ
an
n
th primitive root of unity. Then if we add the
n
th primitive root
to each item of the sequence, we have
L
0
(µ) ··· L
i
(µ) L
i+1
(µ) ··· L
r
(µ)
K = L
0
··· L
i
L
i+1
··· L
r
= L
We know that
L
0
L
0
(
µ
) is a cyclotomic extension by definition. We will now
show that
L
i
(
µ
)
L
i+1
(
µ
) is a Kummer extension for all
i
. Then
L/K
is radical
since L L
r
(µ).
Before we do anything, we have to show
L
i
(
µ
)
L
i+1
(
µ
) is a Galois extension.
To show this, it suffices to show L
i
L
i+1
(µ) is a Galois extension.
Since
L
i
L
i+1
is Galois,
L
i
L
i+1
is normal. So
L
i+1
is the splitting of
some
h
over
L
i
. Then
L
i+1
(
µ
) is just the splitting field of (
t
n
1)
h
. So
L
i
L
i+1
(
µ
) is normal. Also,
L
i
L
i+1
(
µ
) is separable since
char K
=
char L
i
= 0.
Hence L
i
L
i+1
(µ) is Galois, which implies that L
i
(µ) L
i+1
(µ) is Galois.
We define a homomorphism of groups
Gal(L
i+1
(µ)/L
i
(µ)) Gal(L
i+1
/L
i
)
by restriction. This is well-defined because L
i+1
is the splitting field of some h
over
L
i
, and hence any automorphism of
L
i+1
(
µ
) must send roots of
h
to roots
of h, i.e. L
i+1
to L
i+1
.
Moreover, we can see that this homomorphism is injective. If
φ 7→ φ|
L
i+1
=
id
,
then it fixes everything in
L
i+1
. Also, since it is in
Gal
(
L
i+1
(
µ
)
/L
i
(
µ
)), it fixes
L
i
(µ). In particular, it fixes µ. So φ must fix the whole of L
i+1
(µ). So φ = id.
By injectivity, we know that
Gal
(
L
i+1
(
µ
)
/L
i
(
µ
)) is isomorphic to a subgroup
of
Gal
(
L
i+1
/L
i
). Hence it is cyclic. By our previous theorem, it follows that
L
i
(µ) L
i+1
(µ) is a Kummer extension. So L/K is radical.
Corollary. Let
K
be a field with
char K
= 0 and
h K
[
t
]. Let
L
be the
splitting of
h
over
K
. Then
h
can be solved by radicals if and only if
Gal
(
L/K
)
is soluble.
Proof. () Proved before.
(
) Since
L/K
is a Galois extension,
L/K
is a soluble extension. So it is a
radial extension. So h can be solved by radicals.
Corollary. Let
K
be a field with
char K
= 0. Let
f K
[
t
] have
deg f
4.
Then f can be solved by radicals.
Proof. Exercise.
Note that in the case where
K
=
Q
, we have proven this already by given
explicit solutions in terms of radicals in the first lecture.
4 Computational techniques
In the last three lectures, we will look at some techniques that allow us to
actually compute the Galois group of polynomials (i.e. Galois groups of their
splitting fields).
4.1 Reduction mod p
The goal of this chapter is to see what happens when we reduce a polynomial
f Z[t] to the corresponding polynomial
¯
f F
p
[t].
More precisely, suppose we have a polynomial
f Z
[
t
], and
E
is its splitting
field over
Q
. We then reduce
f
to
¯
f F
p
[
t
] by reducing the coefficients mod
p
,
and let
¯
E be the splitting field of
¯
f over F
p
.
The ultimate goal is to show that under mild assumptions, there is an
injection
Gal(E/F
p
) Gal(E/Q).
To do this, we will go through a lot of algebraic fluff to obtain an alternative
characterization of the Galois group, and obtain the result as an easy corollary.
This section will be notationally heavy. First, in the background, we have a
polynomial
f
of degree
n
(whose field we shall specify later). Then we will have
three distinct set of variables, namely (
x
1
, ··· , x
n
), (
u
1
, ··· , u
n
), plus a
t
. They
will play different roles.
The
x
i
will be placeholders. After establishing our definitions, we will then
map each x
i
to α
i
, a root of our f.
The u
i
will stay as “general coefficients” all the time.
t
will be the actual variable we think our polynomial is in, i.e. all polyno-
mials will be variables in
t
, and
u
i
and
x
i
will form part of the coefficients.
To begin with, let
L = Q(x
1
, ··· , x
n
)
F = Q(e
1
, ··· , e
n
).
where
x
i
are variables and
e
i
are the symmetric polynomials in the
x
1
, ··· , x
n
.
We have seen that Gal(L/F )
=
S
n
.
Now let
B = Z[x
1
, ··· , x
n
]
A = Z[e
1
, ··· , e
n
].
It is an exercise on example sheet 4 to show that
B F = A. ()
We will for now take this for granted.
We now add it new variables
u
1
, ··· , u
n
, t
. We previously mentioned that
S
n
can act on, say
L
[
u
1
, ··· , u
n
, t
] by permuting the variables. Here there are two
ways in which this can happen a permutation can either permute the
x
i
, or
permute the u
i
. We will have to keep this in mind.
Now for each σ S
n
, we define the linear polynomial
R
σ
= t x
σ(1)
u
1
··· x
σ(n)
u
n
.
For example, we have
R
(1)
= t x
1
u
1
··· x
n
u
n
.
As mentioned, an element
ρ S
n
can act on
R
ρ
in two ways: it either sends
R
σ
7→ R
ρσ
or R
σ
7→ R
σρ
1
.
It should be clear that the first action permutes the
x
i
. What the second
action does is permute the
u
i
. To see this, we can consider a simple case where
n = 2. Then the action ρ acting on R
(1)
sends
t x
1
u
1
x
2
u
2
7→ t x
ρ
1
(1)
u
1
x
ρ
2
(2)
u
2
= t x
1
u
ρ(1)
x
2
u
ρ(2)
.
Finally, we define the following big scary polynomial:
R =
Y
σS
n
R
σ
B[u
1
, ··· , u
n
, t].
We see that this is fixed by any permutation in
σ S
n
under both actions.
Considering the first action and using (), we see that in fact
R A[u
1
, ··· , u
n
, t].
This is since if we view
R
as a polynomial over
B
in the variables
u
1
, ··· , u
n
, t
,
then its coefficients will be invariant under permuting the
x
i
. So the coefficients
must be a function of the e
i
, i.e. lie in A.
With these definitions in place, we can focus on a concrete polynomial.
Let K be a field, and let
f = t
n
+ a
1
t
n1
+ ··· + a
n
K[t]
have no repeated roots. We let E be the splitting field of f over K. Write
Root
f
(E) = {α
1
, ··· , α
n
}.
Note that this is the setting we had at the beginning of the chapter, but with an
arbitrary field K instead of Q and F
p
.
We define a ring homomorphism
θ
:
B E
by
x
i
7→ α
i
. This extends to a
ring homomorphism
θ : B[u
1
, ··· , u
n
, t] E[u
1
, ··· , u
n
, t].
Note that the ring homomorphism
θ
send
e
i
7→
(
1)
i
a
i
. So in particular, if
we restrict the homomorphism
θ
to
A
, then the image is restricted to the field
generated by
a
i
. But we already have
a
i
K
. So
θ
(
A
) =
K
. In particular, since
R A[u
1
, ··· , u
n
, t], we have
θ(R) K[u
1
, ··· , u
n
, t].
Now let
P
be an irreducible factor of
θ
(
R
) in
K
[
u
1
, ··· , u
n
, t
]. We want to say
each such irreducible polynomial is related to the Galois group
G
=
Gal
(
E/K
).
Since
f
has no repeated roots, we can consider
G
as a subgroup of
S
n
, where
the elements of
G
are just the permutations of the roots
α
i
. We will then show
that each irreducible polynomial corresponds to a coset of G.
Recall that at the beginning, we said
S
n
can act on our polynomial rings
by permuting the
x
i
or
u
i
. However, once we have mapped the
x
i
to the
α
i
and focus on a specific field,
S
n
as a whole can no longer act on the
α
i
, since
there might be non-trivial relations between the
α
i
. Instead, only the subgroup
G S
n
can act on α
i
. On the other hand, S
n
can still act on u
i
.
Recall that
R
is defined as a product of linear factors
R
σ
’s. So we can find a
subset Λ S
n
such that
P =
Y
σΛ
R
σ
.
We will later see this Λ is just a coset of the Galois group G.
Pick σ Λ. Then by definition of P ,
R
σ
| P
in
E
[
u
1
, ··· , u
n
, t
]. Now if
ρ G
, then we can let
ρ
act on both sides by
permuting the
x
i
(i.e. the
α
i
). This does not change
P
because
P
has coefficients
in K and the action of G has to fix K. Hence we have
R
ρσ
| P.
More generally, if we let
H =
Y
ρG
R
ρσ
E[u
1
, ··· , u
n
, t],
then
H | P
by the irreducibility of P .
Since
H
is also invariant under the action of
G
, we know
H K
[
u
1
, ··· , u
n
, t
].
By the irreducibility of P , we know H = P. Hence, we know
Λ = Gσ.
We have thus proved that the irreducible factors of
θ
(
R
) in
K
[
u
1
, ··· , u
n
, t
] are
in one-to-one correspondence with the cosets of
G
in
S
n
. In particular, if
P
corresponds to G itself, then
P =
Y
τG
R
τ
.
In general, if
P
corresponds to a coset
, we can let
λ S
n
act on
P
by
permuting the u
i
’s. Then this sends
P =
Y
ρG
R
ρσ
7→ Q =
Y
ρG
R
ρσλ
1
.
So this corresponds to the coset
λ
1
. In particular,
P
=
Q
if and only if
=
λ
1
. So we can use this to figure out what permutations preserve an
irreducible factor. In particular, taking σ = (1), we have
Theorem.
G = {λ S
n
: λ preserves the irreducible factor corresponding to G}. ()
This is the key result of this chapter, and we will apply this as follows:
Theorem. Let
f Z
[
t
] be monic with no repeated roots. Let
E
be the splitting
field of
f
over
Q
, and take
¯
f F
p
[
t
] be the obvious polynomial obtained by
reducing the coefficients of
f
mod
p
. We also assume this has no repeated roots,
and let
¯
E be the splitting field of
¯
f.
Then there is an injective homomorphism
¯
G = Gal(
¯
E/F
p
) G = Gal(E/Q).
Moreover, if
¯
f
factors as a product of irreducibles of length
n
1
, n
2
, ··· , n
r
, then
Gal(f) contains an element of cycle type (n
1
, ··· , n
r
).
Proof. We apply the previous theorem twice. First, we take K = Q. Then
θ(R) Z[u
1
, ··· , u
n
, t].
Let
P
be the irreducible factor of
θ
(
R
) corresponding to the Galois group
G
.
Applying Gauss’ lemma, we know P has integer coefficients.
Applying the theorem again, taking
K
=
F
p
. Denote the ring homomorphism
as
¯
θ
. Then
¯
θ
(
R
)
F
p
[
u
1
, ··· , u
n
, t
]. Now let
Q
be the irreducible factor
¯
θ
(
R
)
corresponding to
¯
G.
Now note that
θ
(
R
(1)
)
| P
and
¯
θ
(
R
(1)
)
| Q
, since the identity is in
G
and
¯
G
.
Also, note that
¯
θ
(
R
) =
θ(R)
, where the bar again denotes reduction mod
p
. So
Q |
¯
P .
Considering the second action of
S
n
(i.e. permuting the
u
i
), we can show
¯
G G, using the characterization (). Details are left as an exercise.
This is incredibly useful for computing Galois groups, as it allows us to
explicitly write down some cycles in Gal(E, Q).
4.2 Trace, norm and discriminant
We are going to change direction a bit and look at traces and norms. These will
help us understand the field better, and perhaps prove some useful facts from
it. They will also lead to the notion of the discriminant, which is again another
tool that can be used to compute Galois groups, amongst many other things.
Definition (Trace). Let
K
be a field. If
A
= [
a
ij
] is an
n × n
matrix over
K
,
we define the trace of A to be
tr(A) =
n
X
i=1
a
ii
,
i.e. we take the sum of the diagonal terms.
It is a well-known fact that if B is an invertible n × n matrix, then
tr(B
1
AB) = tr(A).
Hence given a finite-dimensional vector space
V
over
K
and
σ
:
V V
a
K-linear map, then we can define the trace for the linear map as well.
Definition (Trace of linear map). Let
V
be a finite-dimensional vector space
over K, and σ : V V a K-linear map. Then we can define
tr(σ) = tr(any matrix representing σ).
Definition (Trace of element). Let
K L
be a finite field extension, and
α L
. Consider the
K
-linear map
σ
:
L L
given by multiplication with
α
, i.e.
β 7→ αβ. Then we define the trace of α to be
tr
L/K
(α) = tr(σ).
Similarly, we can consider the determinant, and obtain the norm.
Definition (Norm of element). We define the norm of α to be
N
L/K
(α) = det(σ),
where σ is, again, the multiplication-by-α map.
This construction gives us two functions
tr
L/K
, N
L/K
:
L K
. It is easy to
see from definition that tr
L/K
is additive while N
L/K
is multiplicative.
Example. Let
L/K
be a finite field extension, and
x K
. Then the matrix of
x is represented by xI, where I is the identity matrix. So
N
L/K
(x) = x
[L:K]
, tr
L/K
(x) = [L : K]x.
Example. Let
K
=
Q
,
L
=
Q
(
i
). Consider an element
a
+
bi Q
(
i
), and pick
the basis {1, i} for Q(i). Then the matrix of a + bi is
a b
b a
.
So we find that tr
L/K
(a + bi) = 2a and N(a + bi) = a
2
+ b
2
= |a + bi|
2
.
In general, if
K
=
Q
and
L
=
Q
(
d
) where
d >
0 is square-free, then
N
(
a
+
b
d
) =
a
2
+
b
2
d
=
|a
+
b
d|
2
. However, for other fields, the norm is
not at all related to the absolute value.
In general, computing norms and traces with the definition directly is not
fun. It turns out we can easily find the trace and norm of
α
from the minimal
polynomial of
α
, just like how we can find usual traces and determinants from
the characteristic polynomial.
To do so, we first prove the transitivity of trace and norm.
Lemma. Let L/F/K be finite field extensions. Then
tr
L/K
= tr
F/K
tr
L/F
, N
L/K
= N
F/K
N
L/F
.
To prove this directly is not difficult, but involves some confusing notation.
Purely for the sake of notational convenience, we shall prove the following more
general fact:
Lemma. Let
F/K
be a field extension, and
V
an
F
-vector space. Let
T
:
V V
be an F -linear map. Then it is in particular a K-linear map. Then
det
K
T = N
F/K
(det
F
T ), tr
K
T = tr
F/K
(tr
F
T ).
Taking
V
to be
L
and
T
to be multiplication by
α F
clearly gives the
original intended result.
Proof.
For
α F
, we will write
m
α
:
F F
for multiplication by
α
map viewed
as a K-linear map.
By IB Groups, Rings and Modules, there exists a basis
{e
i
}
such that
T
is
in rational canonical form, i.e. such that
T
is block diagonal with each diagonal
looking like
0 0 ··· 0 a
0
1 0 ··· 0 a
1
0 1 ··· 0 a
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· 1 a
r1
.
Since the norm is multiplicative and trace is additive, and
det
A 0
0 B
= det A det B, tr
A 0
0 B
= tr A + tr B,
we may wlog T is represented by a single block as above.
From the rational canonical form, we can read off
det
F
T = (1)
r1
a
0
, tr
F
T = a
r1
.
We now pick a basis
{f
j
}
of
F
over
K
, and then
{e
i
f
j
}
is a basis for
V
over
K. Then in this basis, the matrix of T over K is given by
0 0 ··· 0 m
a
0
1 0 ··· 0 m
a
1
0 1 ··· 0 m
a
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· 1 m
a
r1
.
It is clear that this has trace
tr
K
(m
a
r1
) = tr
F/K
(a
r1
) = tr
F/K
(tr
F
T ).
Moreover, writing n = [L : K], we have
det
K
0 0 ··· 0 m
a
0
1 0 ··· 0 m
a
1
0 1 ··· 0 m
a
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· 1 m
a
r1
= (1)
n(r1)
det
K
m
a
0
0 0 ··· 0
m
a
1
1 0 ··· 0
m
a
2
0 1 ··· 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
m
a
r1
0 0 ··· 1
= (1)
n(r1)
det
K
(m
a
0
)
= det
K
((1)
r1
m
a
0
)
= N
F/K
((1)
r1
a
0
)
= N
F/K
(det
F
T ).
So the result follows.
As a corollary, we have the following very powerful tool for computing norms
and traces.
Corollary. Let
L/K
be a finite field extension, and
α L
. Let
r
= [
L
:
K
(
α
)]
and let P
α
be the minimal polynomial of α over K, say
P
α
= t
n
+ a
n1
t
n1
+ ··· + a
0
.
with a
i
K. Then
tr
L/K
(α) = ra
n1
and
N
L/K
(α) = (1)
nr
a
r
0
.
Note how this resembles the relation between the characteristic polynomial
and trace/determinants in linear algebra.
Proof.
We first consider the case
r
= 1. Write
m
α
for the matrix representing
multiplication by
α
. Then
P
α
is the minimal polynomial of
m
α
. But since
deg P
α
=
n
=
dim
K
K
(
α
), it follows that this is also the characteristic polynomial.
So the result follows.
Now if
r
= 1, we can consider the tower of extensions
L/K
(
α
)
/K
. Then we
have
N
L/K
(α) = N
K(α)/K
(N
L/K(α)
(α)) = N
K(α)/K
(α
r
)
= (N
K(α)/K
(α))
r
= (1)
nr
a
r
0
.
The computation for trace is similar.
It is also instructive to prove this directly. In the case
r
= 1, we can pick the
basis {1, α, α
2
, ··· , α
n1
} of L over K. Then the multiplication map sends
1 7→ α
α 7→ α
2
.
.
.
α
n1
7→ α
n
= a
n1
α
n1
··· a
0
So the matrix is just
A =
0 0 ··· a
0
1 0 ··· a
1
0 1 ··· a
2
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· a
n1
The characteristic polynomial of this matrix is
det(tI A) = det
t 0 ··· a
0
1 t ··· a
1
0 1 ··· a
2
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· t + a
n1
By adding t
i
multiples of the ith row to the first row for each i, this gives
det(tI A) = det
0 0 ··· P
α
1 t ··· a
1
0 1 ··· a
2
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· t + a
n1
= P
α
.
Then we notice that for
r
= 1, in an appropriate choice of basis, the matrix
looks like
C =
A 0 ··· 0
0 A ··· 0
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· A
.
Theorem. Let
L/K
be a finite but not separable extension. Then
tr
L/K
(
α
) = 0
for all α L.
Proof.
Pick
β L
such that
P
β
, the minimal polynomial of
β
over
K
, is not
separable. Then by the previous characterization of separable polynomials, we
know p = char K > 0 with P
β
= q(t
p
) for some q K[t].
Now consider
K K(β
p
) K(β) L.
To show
tr
L/K
= 0, by the previous proposition, it suffices to show
tr
K(β)/K(β
p
)
=
0.
Note that the minimal polynomial of
β
p
over
K
is
q
because
q
(
β
p
) = 0 and
q
is irreducible. Then [
K
(
β
) :
K
] =
deg P
β
=
p deg q
and
deg
[
K
(
β
p
) :
K
] =
deg q
.
So [K(β) : K(β
p
)] = p.
Now
{
1
, β, β
2
, ··· , β
p1
}
is a basis of
K
(
β
) over
K
(
β
p
). Let
R
β
i
be the
minimal polynomial of β
i
over K(β
p
). Then
R
β
i
=
(
t 1 i = 0
t
p
β
ir
i = 0
,
We get the second case using the fact that
p
is a prime number, and hence
K
(
β
p
)(
β
i
) =
K
(
β
) if 1
i < p
. So [
K
(
β
p
)(
β
i
) :
K
(
β
p
)] =
p
and hence the
minimal polynomial has degree p. Hence tr
K(β)/K(β
p
)
(β
i
) = 0 for all i.
Thus, tr
K(β)/K(β
p
)
= 0. Hence
tr
L/K
= tr
K(β
p
)/K
tr
K(β)/K(β
p
)
tr
L/K(β)
= 0.
Note that if L/K is a finite extension, and char K = 0, then
tr
L/K
(1) = [L : K] = 0.
So
tr
L/K
= 0. It is in fact true that all separable extensions have
tr
L/K
= 0, not
only when the field has characteristic 0.
Example. We want to show
3
3 ∈ Q
(
3
2
). Suppose not. Then we have
L
=
Q
(
3
3
) =
Q
(
3
2
), since both extensions of
Q
have degree 3. Then there
exists some a, b, c Q such that
3
3 = a + b
3
2 + c
3
2
2
.
We now compute the traces over Q. The minimal polynomials over Q are
P
3
3
= t
3
3, P
3
2
= t
3
2, P
3
4
= t
3
4.
So we have
tr
L/Q
(
3
3) = a tr
L/Q
(1) + b tr
L/Q
(
3
2) + c tr
L/Q
(
3
4).
Since the minimal polynomials above do not have coefficients in
t
2
, the traces of
the cube roots are zero. So we need a = 0. Then we are left with
3
3 = b
3
2 + c
3
4.
We apply the same trick again. We multiply by
3
2 to obtain
3
6 = b
3
4 + 2c.
We note that the minimal polynomial of
3
6 is t
3
6. Taking the trace gives
tr
L/Q
(
3
6) = b tr
L/Q
(
3
4) + 6c.
Again, the traces are zero. So c = 0. So we have
3
3 = b
3
2.
In other words,
b
3
=
3
2
,
which is clearly nonsense. This is a contradiction. So
3
3 ∈ Q(
3
2).
We can obtain another formula for the trace and norm as follows:
Theorem. Let
L/K
be a finite separable extension. Pick a further extension
E/L such that E/K is normal and
|Hom
K
(L, E)| = [L : K].
Write Hom
K
(L, E) = {ϕ
1
, ··· , ϕ
n
}. Then
tr
L/K
(α) =
n
X
i=1
ϕ
i
(α), N
L/K
(α) =
n
Y
i=1
ϕ
i
(α)
for all α L.
Proof.
Let
α L
. Let
P
α
be the minimal polynomial of
α
over
K
. Then there
is a one-to-one correspondence between
Hom
K
(K(α), E) Root
P
α
(E) = {α
1
, ··· , α
d
}.
wlog we let α = α
1
.
Also, since
|Hom
K
(L, E)| = [L : K],
we get
|Hom
K
(K(α), E)| = [K(α) : K] = deg P
α
.
Moreover, the restriction map
Hom
K
(
L, E
)
Hom
K
(
K
(
α
)
, E
) (defined by
ϕ 7→ ϕ|
K(α)
) is surjective and sends exactly [
K
(
α
) :
K
] elements to any particular
element in Hom
K
(K(α), E).
Therefore
X
ϕ
i
(α) = [L : K(α)]
X
ψHom
K
(K(α),E)
ψ(α) = [L : K(α)]
d
X
i=1
α
i
.
Moreover, we can read the sum of roots of a polynomial is the (negative of the)
coefficient of t
d1
, where
P
α
= t
d
+ a
d1
t
d1
+ ··· + a
0
.
So
X
ϕ
i
(α) = [L : K(α)](a
d1
) = tr
L/K
(α).
Similarly, we have
Y
ϕ
i
(α) =
Y
ψHom
K
(K(α),E)
ψ(α)
[L:K(α)]
=
d
Y
i=1
α
i
!
[L:K(α)]
= ((1)
d
a
0
)
[L:K(α)]
= N
L/K
(α).
Corollary. Let
L/K
be a finite separable extension. Then there is some
α L
such that tr
L/K
(α) = 0.
Proof. Using the notation of the previous theorem, we have
tr
L/K
(α) =
X
ϕ
i
(α).
Similar to a previous lemma, we can show that
ϕ
1
, ··· , ϕ
n
are “linearly indepen-
dent” over
E
, and hence
P
ϕ
i
cannot be identically zero. Hence there is some
α
such that
tr
L/K
(α) =
X
ϕ
i
(α) = 0.
Example. Let
K
=
F
q
L
=
F
q
n
, with
q
is a power of some prime number
p
.
By a previous theorem on finite fields, we know L/K is Galois and
Gal(L/K) =
Z
nZ
and is generated by the Frobenius ϕ = Fr
q
.
To apply the theorem, we had to pick an
E
such that
E/K
is normal and
Hom
K
(
L, E
) = [
L
:
K
]. However, since
L/K
is Galois, we can simply pick
E = L.
Then we know
tr
L/K
(α) =
X
ψGal(L/K)
ψ(α)
=
n1
X
i=0
ϕ
i
(α)
= α + α
q
+ α
q
2
+ ··· + α
q
n1
.
Similarly, the norm is
N
L/K
(α) =
n1
Y
i=0
ϕ
i
(α) = α · α
q
· ··· · α
q
n1
.
Recall that when solving quadratic equations
f
=
t
2
+
bt
+
c
, we defined the
discriminant as
b
2
4
c
. This discriminant then determined the types of roots of
f
. In general, we can define the discriminant of a polynomial of any degree, in a
scary way.
Definition (Discriminant). Let
K
be a field and
f K
[
t
],
L
the splitting field
of f over K. So we have
f = a(t α
1
) ···(t α
n
)
for some a, α
1
, ··· , α
n
L. We define
f
=
Y
i<j
(α
i
α
j
), D
f
=
2
f
= (1)
n(n1)/2
Y
i=j
(α
i
α
j
).
We call D
f
the discriminant of f .
Clearly, D
f
= 0 if and only if f has no repeated roots.
Theorem. Let
K
be a field and
f K
[
t
],
L
is the splitting field of
f
over
K
.
Suppose D
f
= 0 and char K = 2. Then
(i) D
f
K.
(ii)
Let
G
=
Gal
(
L/K
), and
θ
:
G S
n
be the embedding given by the
permutation of the roots. Then
im θ A
n
if and only if
f
K
(if and
only if D
f
is a square in K).
Proof.
(i) It is clear that D
f
is fixed by Gal(L/K) since it only permutes the roots.
(ii)
Consider a permutation
σ S
n
of the form
σ
= (
m
), and let it act on
the roots. Then we claim that
σ(∆
f
) =
f
. ()
So in general, odd elements in
S
n
negate
f
while even elements fix it.
Thus,
f
K
iff
f
is fixed by
Gal
(
L/K
) iff every element of
Gal
(
L/K
)
is even.
To prove (
), we have to painstakingly check all terms in the product. We
wlog
< m
. If
k < , m
. Then this swaps (
α
k
α
) with
α
k
α
m
), which
has no effect. The
k > m
case is similar. If
< k < m
, then this sends
(
α
α
k
)
7→
(
α
m
α
k
) and (
α
k
α
m
)
7→
(
α
α
m
). This introduces two
negative signs, which has no net effect. Finally, this sends (
α
k
α
m
) to
its negation, and so introduces a negative sign.
We will later use this result to compute certain Galois groups. Before that,
we see how this discriminant is related to the norm.
Theorem. Let
K
be a field, and
f K
[
t
] be an
n
-degree monic irreducible
polynomial with no repeated roots. Let
L
be the splitting field of
f
over
K
, and
let α Root
F
(L). Then
D
f
= (1)
n(n1)/2
N
K(α)/K
(f
(α)).
Proof.
Let
Hom
K
(
K
(
α
)
, L
) =
{ϕ
1
, ··· , ϕ
n
}
. Recall these are in one-to-one
correspondence with Root
f
(L) = {α
1
, ··· , α
n
}. Then we can compute
Y
i=j
(α
i
α
j
) =
Y
i
Y
j=i
(α
i
α
j
).
Note that since f is just monic, we have
f = (t α
1
) ···(t α
n
).
Computing the derivative directly, we find
Y
j=i
(α
i
α
j
) = f
(α
i
).
So we have
Y
i=j
(α
i
α
j
) =
Y
i
f
(α
i
).
Now since the ϕ
i
just maps α to α
i
, we have
Y
i=j
(α
i
α
j
) =
Y
i
ϕ
i
(f
(α)) = N
K(α)/K
(f
(α)).
Finally, multiplying the factor of (1)
n(n1)/2
gives the desired result.
Example. Let
K
be a field with
char K
= 2
,
3. Let
f K
[
t
] have degree 3, say
f = t
3
+ bt + c
where we have gotten rid of the
t
2
term as in the first lecture. We further assume
f is irreducible with no repeated roots, and let L be the splitting field of f.
We want to compute the discriminant of this polynomial. Let
α Root
f
(
L
).
Then
β = f
(α) = 3α
2
+ b.
Then we can see
β = 2b
3c
α
.
Alternatively, we have
α =
3c
β + 2b
. ()
Putting (
) into
α
3
+
+
c
= 0, we find the minimal polynomial of
β
has
constant term 4b
3
27c
2
. This then gives us the norm, and we get
D
f
= N
K(α)/K
(β) = 4b
3
27c
2
.
This is the discriminant of a cubic.
We can take a specific example, where
f = t
3
31t + 62.
Then
f
is irreducible over
Q
. We can compute
D
f
, and find that it is a square.
So the previous theorem says the image of the Galois group
Gal
(
L/K
) is a
subgroup of
A
3
. However, we also know
Gal
(
L/K
) has three elements since
deg f = 3. So we know Gal(L/K)
=
A
3
.