We begin this chapter with some basic facts about vector spaces. These will be familiar (at least in the case of real vector spaces) to those readers who have studied linear algebra. We then focus our attention on the particular case of a field extension. A number of properties of field extensions are discussed.
Let F
be a field and a nonconstant polynomial. We
demonstrate how to create a field extension in which f(x)
splits into a product of polynomials of degree 1. This leads to a
classification of all finite fields.
12.1 Vector Spaces
We begin with the definition of a vector space. In most linear
algebra courses, vector spaces are defined over or, occasionally,
. But we can do the same thing over
any field.
If F
is a field and V is a set, then
a scalar multiplication
on V is a function from
to V. If
,
, then we write av for the image of (a, v) under such a function.
Definition 12.1.
- 1.
V is an abelian group under addition;
- 2.
for all
and all
;
- 3.
for all
and all
;
- 4.
for all
and all
;
- 5.
for all
and all
; and
- 6.
for all
.
Of course, condition (2) is redundant, given the definition of a scalar multiplication, but we include it, because it must be checked.
Certainly the most familiar vector space
over is
. We can generalize this.
Example 12.1.
Let F be a field. For any positive integer
n, let . Then
is a vector over F with the usual addition operation and
scalar multiplication
, for any
.
Example 12.2.
Let F be any field. Then F[x]
is a vector space over F with
the usual polynomial addition and , for any
.
Example 12.3.
Let m and n be any positive integers, and let
V be the set of matrices with entries in a field
F. Then V is a vector space over F using matrix addition and scalar
multiplication.
The least exciting example of a vector space is the following.
Example 12.4.
Let F be any field and V the trivial additive group,
. Then V is a vector space using the only
available addition and scalar multiplication options,
and
, for all
.
The most important example for our purposes is the following.
Definition 12.2.
If F and K are fields, with F a subfield of K, then we say that K is an extension field of F.
Example 12.5.
Any extension field K of F is a vector space over F, using the addition operation in
K and multiplication in
K as the scalar multiplication.
(All the properties are immediate, except that for all
. To be sure of that, we must know
that the identity of F is the
identity of K. But this follows
from Theorem 8.12.) For example,
and
are vector spaces over
.
Let us mention a few basic properties of vector spaces.
Theorem 12.1.
- 1.
for all
;
- 2.
for all
; and
- 3.
for all
.
Proof.
(1) Note that . Adding
to both sides, we see that
.
(2) We have . Adding
to both sides, we obtain the desired
conclusion.
(3) Observe that . Thus,
.
We do have to be a bit careful about
which 0 we are using. For example, when we write in the theorem above, the first 0 is
in F and the second is in
V.
Definition 12.3.
Let V be a vector space over a field F. Then a subset W of V is said to be a subspace of V if it is a vector space over F using the same addition and scalar multiplication.
Example 12.6.
If F is a subfield of K, and K is a subfield of L, then L is a vector space over F having K and F as subspaces.
Example 12.7.
Regarding as a vector space over
, we note that
is a subspace.
There is a simple test for a subspace.
Theorem 12.2.
- 1.
;
- 2.
for all
(closure under addition); and
- 3.
for all
and
(closure under scalar multiplication).
Proof.
If W
is a subspace then, in particular, it is an additive subgroup, so
(1) and (2) hold. Part (3) is one of the conditions for a vector
space. Conversely, suppose that (1), (2) and (3) hold. Noting that
(3) tells us that , for all
, we see from Theorem 3.10 that W is an additive subgroup of V. We are given closure under scalar
multiplication. The remaining vector space properties hold in
V, and therefore in any subset
of V.
Note that in the preceding theorem,
condition (1) could be replaced with the condition that W is not the empty set, for if
, then
, and therefore
.
Example 12.8.









Exercises
12.1.
Let F be a field and n a positive integer. If V is the set of all polynomials of degree n in F[x], together with the zero polynomial, is V a subspace of F[x]?
12.2.
Let F be a field and n a positive integer. If V is the set of all polynomials of degree at most n in F[x], together with the zero polynomial, show that V is a subspace of F[x].
12.3.
Let V be a vector space having subspaces
U and W. Show that is a subspace of V. Extend this to the intersection of an
arbitrary collection of subspaces.
12.4.
Let V be a vector space having subspaces
U and W. Show that (regarding U and W as additive subgroups of V) is a subspace of V.
12.5.
Let V and W be vector spaces over a field
F. A function is said to be a linear transformation if
and
for all
,
. If U is a subspace of V, show that
is a subspace of W.
12.6.
Let F, V,
W and be as in the preceding exercise. Show
that the kernel of
(regarding
as a homomorphism of additive groups)
is a subspace of V. Further
show that
is one-to-one if and only if the
kernel is
.
12.7.
Let and
. If
, is W a subspace of V?
12.8.
Let F be a field with vector spaces
V and W. Let be the direct product of the additive
groups V and W. Define a scalar multiplication on
U via
for all
,
and
. Is U a vector space over F?
12.9.
Let F be a field of characteristic 3 and
V a vector space over
F. Show that for all
.
12.10.
Suppose that V is a vector space over an infinite field F. Show that V is not the union of a finite number of proper subspaces.
12.2 Basis and Dimension
In order to define a basis for a vector space, we must first discuss linear combinations of vectors.
Definition 12.4.
Let V be a vector
space over a field F. If
, then a vector
is said to be a linear combination of the
if
, for some
.
Example 12.9.
Let and
. If
and
, then
is a linear combination of
and
, since
.
Definition 12.5.
Let F be a field and V a vector space
over F. Let . We say that the
are linearly dependent if there exist
, not all zero, such that
. Otherwise, the
are linearly independent.
Example 12.10.
Let and
. The vectors (2, 1, 3),
(1, 3, 0) and (2, 1, 4) are linearly dependent,
since
. On the other hand,
(1, 0, 4), (3, 2, 1) and (2, 0, 2)
are linearly independent. Indeed, if
, then looking at the middle entry, we
see immediately that
. Then
. This yields
, and hence
and, finally,
.
Here is a handy test for linear dependence.
Theorem 12.3.


- 1.
; or
- 2.
there exists an
such that
is a linear combination of
.
Proof.
Suppose that the are linearly dependent. Choose
, not all zero, such that
. Let m be the largest positive integer such
that
. Then
. If
, then
, with
. Thus,
, giving case (1). If
, then
, and so
is a linear combination of
, which proves case (2).








Linear independence is most useful when combined with another property.
Definition 12.6.
Let V be a vector
space over a field F, and let
. Then we say that the
span V if every
is a linear combination of the
.
Example 12.11.
Regarding as a vector space over
, we note that 1 and i span
, as
.
Example 12.12.
Let and
. Then the vectors
(1, 0, 0), (0, 1, 0) and (0, 0, 1)
span V, since
.
The following lemma describes a very nice relationship between linear independence and spanning.
Lemma 12.1.
Let V be a vector space over a field
F. Suppose that span V. If
, and
, then the
are linearly dependent.
Proof.















Now consider . It is a linear combination of
. Let us say that
, with
. If
for all
, then
is a linear combination of
, proving that the
are linearly dependent. Thus, we may
assume that there exists an
with
. Without loss of generality, say
. But then
is a linear combination of
. And just as before, we now deduce
that
span V.
Repeat this argument. We will conclude
either that the are linearly dependent or,
eventually, that
span V. But then
is a linear combination of
. By Theorem 12.3, the
are linearly dependent.
What we really need is a basis for a vector space.
Definition 12.7.
Let V be a vector
space over a field F. We say
that form a basis for V if they are linearly independent and
span V.
Example 12.13.
Regarding as a vector space over
, we can see that 1 and i form a basis for
.
Example 12.14.


Example 12.15.
Let F be any field and . Then V has no finite basis. Indeed, if
, then any linear combination of these
vectors must have degree no larger than the maximum of the degrees
of the
. On the other hand, for any positive
integer n, let W be the set of all polynomials having
degree at most n (including the
zero polynomial). By Exercise 12.2, W is
a subspace of V, and the
polynomials
form a basis.
Theorem 12.4.
Let V be a vector space over a field
F. If form a basis for V, then every element of V can be written uniquely in the form
, with
.
Proof.





Bases are not unique. For instance,
(1, 0) and (0, 1) form a basis for over
, but so do (1, 3) and
(5, 2). However, any two bases for a vector space must have
the same number of vectors.
Theorem 12.5.
Let V be a vector space over a field
F. If and
are bases for V, then
.
Proof.
Suppose the theorem is false. Without
loss of generality, say . Then
span V. Since
, Lemma 12.1 tells us that
are linearly dependent. We have a
contradiction.
Definition 12.8.
Let V be a vector space over a field
F. If is a basis for V, then we say
that V has dimension k, and write
(or
, if the field is unclear from the context). We also stipulate that
. In either
of these cases, V is
finite-dimensional. If
V has no finite basis, then
V is infinite-dimensional.
Example 12.16.
For any field F and positive
integer n, . See Example 12.14.
Example 12.17.
The dimension of over
is 2. See Example 12.13.
Example 12.18.
If F
is any field, then F[x]
is infinite-dimensional. The vector space consisting of the
polynomials of degree at most n
over F, including the zero
polynomial, has dimension . See Example 12.15.
In a finite-dimensional space, we can discard vectors from a spanning set to obtain a basis, or add vectors to a linearly independent set to obtain a basis.
Theorem 12.6.


- 1.
if
span V, then some subset of
is a basis for V; and
- 2.
if
are linearly independent, and
, then there exist
such that
form a basis for V.
Proof.


































Example 12.19.
Let and
. The vectors
and (1, 2, 0) are easily
seen to be linearly independent. Furthermore, (2, 5, 8)
is not a linearly combination of these two vectors. Thus, since
, we see that the vectors
form a basis for V.
Exercises
12.11.


- 1.
(1, 3, 5), (2, 1, 4), (7, 11, 23)
- 2.
(1, 3, 4), (2, 2, 1), (3, 6, 3)
12.12.


- 1.
- 2.
12.13.


- 1.
(1, 0, 2), (2, 5, 3), (3, 5, 5)
- 2.
(1, 0, 2), (2, 3, 5), (0, 0, 4)
12.14.
Do the following matrices span
(as a vector space over
), namely
?
12.15.
Let . Find the dimension of V as a vector space over
, and as a vector space over
.
12.16.
Let and
. Find the dimension of V over F.
12.17.
Let F be a field and V a finite-dimensional vector space. If
W is a subspace of V, show that , with equality if and only if
. (Do not assume, to begin with, that
W is finite-dimensional.)
12.18.
Suppose that a vector space V with dimension n has subspaces U and W with dimensions m and k, respectively. If , show that
.
12.19.
Let F, V,
W and be as in Exercise 12.5. Suppose that
are linearly independent and
is one-to-one. Show that
are linearly independent.
12.20.
Let F be a field and V a finite-dimensional vector space over
F. Say . Show that there exists a bijective
linear transformation (see Exercise 12.5 for the definition)
.
12.3 Field Extensions
Let us now focus on our main vector space of interest: the field extension.
Definition 12.9.
Let K be a field
extension of F. Then the
degree of the extension
is the dimension of K over
F. We write . The extension is finite if
and, in particular, quadratic if
.
Example 12.20.
As we observed in Example 12.17, is a quadratic extension of
.
Example 12.21.








![$$\mathbb Q[x]$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq318.png)

![$$u(x),v(x)\in \mathbb Q[x]$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq319.png)



In fact, . To see this, we observe that
is a basis for K over
. Clearly, these numbers span
K. If they are linearly
dependent, then there are rational numbers a, b, c, not all zero, such that
is a root of
. But again,
, and we write
, for some
. Evaluating at
, we obtain
, giving a contradiction and
establishing that we have a basis.
We are, in fact, engaging in a small abuse of notation here. If K is an extension field of F then, of course, F is also an additive subgroup of K. We could also use the notation [K : F] to mean the index of F in K as additive subgroup. This is not the same as the degree of the extension! For the remainder of the book, when we write [K : F], we will mean the degree of the extension.
In the particular case of a finite
field, we can illustrate the difference. By Lagrange’s theorem, the
index of the additive groups would be . However, the degree is calculated as
follows.
Theorem 12.7.
Let K be a field extension of F, such that K is a finite field. Then .
Proof.
First, we note that K must be finite-dimensional over
F. Indeed, the elements of
K must span K, and by Theorem 12.6, we can obtain a
finite basis. Let , and suppose that
is a basis for K over F. By Theorem 12.4, the elements of
K are uniquely of the form
, with
. As there are |F| choices for each
, the total number of elements of
K is
. Taking the base |F| logarithm, we obtain our
result.
Degrees of extensions behave in a nice way.
Theorem 12.8.
Let K be a finite extension of F and L a finite extension of K. Then .
Proof.
Let be a basis for K over F, and let
be a basis for L over K. We claim that
is a basis for L over F. This will complete the proof.
















Example 12.22.
Let . By Example 8.30, K is an extension field of
. Clearly, 1 and
span K over
. If they were linearly dependent,
then
would lie in
, which is not the case. Thus,
. Let
. We claim that L is a subfield of
and, hence, an extension of
K. All of the subfield
properties are easy to check except perhaps the existence of
inverses. Let
. Then
. Suppose that this is 0. Then
. If
, then so is c, giving us a contradiction. Otherwise,
. Thus, we can write
, with
. Then
. If
, then
, which is not true. If
, then
. But then
has a rational root which, by Theorem
11.5, is not the case. Thus,
, and
, giving us a contradiction.
Therefore,
. Now, 1 and
span L over K. If they were linearly dependent, then
we would have
which, as we have just seen, is not
the case. Therefore,
. By the theorem above,
.
One particular type of extension is especially important.
Definition 12.10.
Let K be a field extension of F. If , then we write F(a)
for the intersection of all subfields of
K containing F and a. We say that K is a simple extension of F if
for some
.
By Exercise 8.33, the intersection of some set of fields is a field. Thus, F(a) is always a field. Indeed, it is the smallest subfield of K containing F and a.
Example 12.23.
By Example 8.30, is a subfield of
. Thus, since any field including
and
would surely contain this field, it
is
.
Example 12.24.
In a similar manner, we note that
would have to contain
and
. Example 12.21 shows us that
.
Let us concentrate on simple extensions. In fact, we need to break them down into two types, depending upon one specific property of the element a.
Definition 12.11.
Let K be a field
extension of F and . We say that a is algebraic over F if there exists a nonzero polynomial
such that
. Otherwise, a is transcendental over F.
Example 12.25.
The number is algebraic over
, since it is a root of
.
Example 12.26.
The number is algebraic over
, since it is a root of
.
Finding examples of real numbers that
are transcendental over is a bit tricky. As it happens, the
constants e and
are both transcendental. (This is a
difficult result. For a proof, see the advanced monograph of Baker
[1].) Of course, the underlying field is
important! If we let
, then
is algebraic over F, as
is a root of
.
We are primarily interested in algebraic elements. However, we can mention one important fact about transcendental elements. If F is a field, then F[x] is an integral domain, and so we can consider its field of fractions. Denote this field of fractions by F(x).
Theorem 12.9.
Let K be an extension field of F, and let be transcendental over F. Then F(a)
is isomorphic to F(x). In particular, F(a)
is of infinite degree over F.
Proof.
Define via
. By Lemma 11.1,
is a homomorphism. If
, then
. Since a is transcendental, f(x)
is the zero polynomial. Thus,
is one-to-one, and F[x]
is isomorphic to
. Also,
for all
; thus,
is a subring of F(a).
By Theorem 9.15, there is a subfield L of F(a)
such that L is isomorphic to
F(x) and contains
. Clearly
for all
and
; thus,
contains both F and a. But F(a)
is the smallest subfield of K
containing both F and
a; thus,
.
Suppose that . Then according to Lemma 12.1, the elements
must be linearly dependent over
F. But then there exist
, not all zero, such that a is a root of
. That is, a is algebraic, giving us a
contradiction.
Now suppose that a is algebraic over F. We know that it satisfies a nonzero polynomial in F[x]. But one particular such polynomial is key.
Definition 12.12.
Let K be an extension field of F and let be algebraic over F. Then the minimal polynomial of a over F is the monic irreducible polynomial
such that
.
Example 12.27.
The minimal
polynomial of over
is
. Indeed,
is a root, and the polynomial is irreducible by Example 11.10.
Example 12.28.
The minimal polynomial of over
is
. As we noted in Example 12.26,
is a root. Suppose it were reducible
over
. The Rational Roots Theorem shows us
that it has no roots in
. Thus, it would have to factor as a
product of two polynomials of degree 2. By Theorem 11.4, these polynomials may be
assumed to be in
. Looking at the coefficients, we see
immediately that (up to multiplying both factors by
) the only possibilities are
and
, for some
. But then
or
. Neither of these has a solution in
.
We were a bit bold in our definition of the minimal polynomial. Indeed, we assumed that such a polynomial exists, and that there is only one. Fortunately, our presumptuousness was justified; in fact, we can say more.
Theorem 12.10.

- 1.
the minimal polynomial m(x) of a over F exists, and is the unique monic polynomial of smallest degree in F[x] of which a is a root; and
- 2.
if
, then
if and only if m(x)|f(x).
Proof.
Let . We claim that I is an ideal of F[x].
Surely
. If
, then
, and hence
. Also, if
, then
, and hence
, proving the claim.
We know that F[x]
is a Euclidean domain and hence, by Theorem 10.8, a PID. Thus, let . Since a is algebraic, m(x)
is not the zero polynomial. As
if
, we may as well assume that
m(x) is monic. Now,
if and only if m(x)|f(x),
as required by (2). As such,
, unless
. If
, then f(x)
is simply m(x) multiplied by an element of F. If f(x)
is also monic, then
. Thus, m(x)
satisfies condition (1) as well.
We must still establish that m(x)
is actually the minimal polynomial of a over F. To demonstrate this, we must show that
m(x) is irreducible. But if , with
, then
. Thus,
or
. Without loss of generality, say
. Then m(x)|f(x).
But also f(x)|m(x).
It now follows that
, and hence g(x)
is a constant polynomial. Thus, m(x)
is irreducible, and hence a minimal polynomial for a. If g(x)
is another minimal polynomial, then
, and hence m(x)|g(x).
But g(x) is irreducible, and therefore
for some
. As m(x)
and g(x) are both monic,
, and the proof is
complete.
We can use the minimal polynomial to describe the simple extension.
Theorem 12.11.


- 1.
;
- 2.
is a basis for F(a) over F; and
- 3.
F(a) is isomorphic to F[x] / (m(x)).
Proof.
Of course, (1) follows immediately from
(2), so let us prove (2). Suppose that are linearly dependent. Then there
exist
, not all zero, such that
. That is, a is a root of
. But this polynomial has degree
smaller than that of m(x),
contradicting Theorem 12.10. Thus,
are linearly independent.
We claim that they span F(a).
We know that F(a) is the smallest field containing
F and a (and, hence, all of the ). Therefore, it is sufficient to show
that
is a field. Clearly, it contains 1
and is closed under subtraction. To show that it is closed under
multiplication, it is enough to show that
for all positive integers i. Our proof is by strong induction upon
i. If
, there is nothing to do, so let
and suppose it is true for smaller
exponents. Writing
, we have
. But this is a linear combination of
terms of the form
, with
. Thus, by our inductive hypothesis,
. Finally, we must check that every
nonzero element of K has an
inverse in K. But a nonzero
element of K has the form
f(a), for some
, with
. Now, (f(x), m(x))|m(x).
As m(x) is irreducible, (f(x), m(x))
is either 1 or an associate of m(x).
However,
. Thus,
. By Theorem 10.6, there exist
such that
. Since
, we have
. Furthermore, as we noted above,
, so f(a)
has an inverse in K. Therefore,
K is a field, and (2) is
proved.
(3) Define via
. By Lemma 11.1,
is a homomorphism. In view of (2), it
is onto. The kernel is the set of all polynomials in F[x]
of which a is a root. By
Theorem 12.10, this is (m(x)). Apply the First Isomorphism Theorem.
Example 12.29.
As is the minimal polynomial of
i over
, we see that
is isomorphic to
.
Example 12.30.




![$$\mathbb Q[x]/(x^4-10x^2+1).$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_Equ15.png)









Our last theorem has an interesting immediate consequence.
Corollary 12.1.
Let K be an extension field of F. If , and a and b have the same minimal polynomial over
F, then F(a)
is isomorphic to F(b).
Proof.
If m(x)
is the minimal polynomial, then by Theorem 12.11, both F(a)
and F(b) are isomorphic to F[x] / (m(x)).
Example 12.31.
Let be a primitive cube root of unity in
. (That is,
but
.) Then
and
are both roots of
. As we have observed,
is irreducible over
, so it is the minimal polynomial of
both
and
. Thus,
is isomorphic to
. These fields are clearly distinct,
as
is a subfield of
, but
.
Exercises
12.21.
Find the minimal polynomial of
over
.
12.22.
Find the minimal polynomial of
over
.
12.23.
Let K be a finite extension field of F. Show that every element of K is algebraic over F.
12.24.
Let K be an extension field of F and L an extension field of K. If is algebraic over F, show that
.
12.25.
Suppose that we have subfields
of K with
. Show that
is a field.
12.26.
For each positive integer n, let . If
, show that K is an infinite field extension of
, but every element of K is algebraic over
.
12.27.
Let K be a field extension of F. Show that for every ,
. Also, give an explicit example
illustrating that we do not have
in general.
12.28.
Let K be a field extension of F and . Show that a is algebraic over F if and only if
is algebraic over F.
12.29.
Let K be an extension field of . If
is algebraic over
, show that
.
12.30.
Let K be a finite extension of F. If R is a subring of K containing F, show that R is a field.
12.4 Splitting Fields
Let us now take a slightly different
perspective from the preceding section. Given a field F, instead of looking at elements of
extension fields and finding their minimal polynomials, let us
instead take a nonconstant polynomial and see if we can find a field
containing F and a root of
f(x). For instance, suppose that we only
knew about the rational numbers, and we wanted to construct a field
having a root of
.
Definition 12.13.
Let F be a field and let be a nonconstant polynomial. If
K is an extension field of
F, then we say that f(x)
splits over K if there exist
such that
. In particular, K is a splitting field for f(x)
if f(x) splits over K, and if L is any subfield of K with
, then f(x)
does not split over L.
To put this another way, if K is an extension field of F and , write
for the intersection of all subfields of K containing F and all of the
. If
splits over K, and
are the roots of f(x)
in K, then K is a splitting field of f(x)
if and only if
.
But how to construct such a field? The following observation is helpful.
Lemma 12.2.
Every nonzero prime ideal in a PID is maximal.
Proof.
Let I be a nonzero prime ideal in a PID
R. Then , for some
. By Lemma 10.2, a is prime. In particular, by Theorem
10.10, a is irreducible. Since I is prime,
. Suppose that J is an ideal of R with
. Let
. Then
, so b|a.
As a is irreducible, b is a unit or an associate of a. In the former case,
. In the latter, a|b,
and hence
. Either way, we have a
contradiction.
The next lemma is the key to our construction.
Lemma 12.3.
Let F be a field and f(x)
an irreducible polynomial in F[x].
Let . Then K is a field containing (an isomorphic
copy of) F and a root
a of f(x).
In fact,
.
Proof.
We know that F[x]
is a Euclidean domain and hence, by Theorem 10.8, a PID. By Theorem 10.11, f(x)
is prime. Thus, by Lemma 10.2, (f(x))
is a prime ideal. The preceding lemma tells us that (f(x))
is maximal. By Theorem 9.20, K is indeed a field. Define via
. It is immediate that
is a homomorphism. If
, then
, which means that f(x)|b.
As b is a constant,
, and hence
is one-to-one. Thus, K contains an isomorphic copy of
F, namely
. Finally, let us show that K contains a root of f(x).
But this root is
. Indeed,
, as required. Clearly, F(a)
would have to contain
for all
. Thus,
.
Let us combine the preceding lemma with
Theorem 12.11. We see that if is irreducible of degree n, then the field K has, as a basis over F, the terms
, with
. This allows us for the first time to
create finite fields other than
, where p is a prime.
Example 12.32.

![$$f(x)=x^3+3x^2+x+2\in \mathbb Z_5[x]$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq621.png)






We can now construct splitting fields.
Theorem 12.12.
Let F be a field and a nonconstant polynomial. Then there
is a splitting field of f(x)
over F.
Proof.
First, let us prove the existence of a
field extension in which f(x)
splits. We proceed by induction on . If
, then F will suffice. Assume that
and the
case holds. We know that F[x]
is a UFD. Thus, write
, where the
are irreducible in F[x].
By Lemma 12.3, there is an extension field K of F in which
has a root, a. Then by Theorem 11.2,
, for some
. Thus, in K[x],
we have
. Now,
has degree
. Thus, by our inductive hypothesis,
it splits in some extension field L of K. Hence, L is an extension field of F, and f(x)
splits over L.
Let us write , with
. Then
is a splitting field for f(x)
over F.
But we can go one step further. We want
to show that splitting fields are unique up to isomorphism. (The
proof is a bit technical, but the result will pay dividends when we
classify the finite fields.) To this end, we need to sharpen
Corollary 12.1 a bit. If is a ring homomorphism, and
, then we write
.
Lemma 12.4.
Let be an isomorphism of fields. Let
be an irreducible polynomial. Suppose
that a is a root of f(x)
in some extension field of F
and b is a root of
in some extension field of K. Then there exists an isomorphism
such that
for all
and
.
Proof.
Define via
. By Lemma 11.1,
is a homomorphism. By Theorem
12.10,
. (We assumed that f(x)
was monic in that theorem, but that is immaterial here.) In view of
Theorem 12.11,
is onto. Thus, the proof of the First
Isomorphism Theorem shows us that the map
given by
is an isomorphism. We also note that
if
, then
and
. In precisely the same manner, the
map
given by
is an isomorphism,
for all
and
.

![$$K[x]/(\alpha (f(x)))$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq670.png)
![$$K[x]/(\alpha (f(x)))$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq671.png)
![$$\sigma :F[x]/(f(x))\rightarrow K[x]/(\alpha (f(x)))$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq672.png)










This allows us to prove the uniqueness of splitting fields.
Theorem 12.13.
Let be a field isomorphism, and let
be a nonconstant polynomial. If
L is a splitting field of
f(x) over F, and M is a splitting field of
over K, then there is an isomorphism
such that
and
agree on F.
Proof.
We proceed by induction on . If
, then we can only have
and
. Thus, letting
will suffice. Assume that the result
is true for polynomials of degree
. As f(x)
is a product of irreducibles in F[x],
let us say that
, where g(x)
is irreducible and
. Let a be a root of g(x)
in L and b a root of
in M. By the preceding lemma, there is an
isomorphism
such that
agrees with
on F and
. We have
, for some
, by Theorem 11.2. Also,
in K(b)[x]. Now, L is a splitting field for u(x)
over F(a) and M is a splitting field for
over K(b).
Since
, our inductive hypothesis completes
the proof.
Corollary 12.2.
Let F be a field and a nonconstant polynomial. Then any
two splitting fields of f(x)
over F are isomorphic.
Proof.
In the preceding theorem, let
be the identity automorphism.
Exercises
12.31.
Construct an extension field F of having order
. In particular, if
, what do all of the elements of
F look like? To which of these
elements is
equal?
12.32.
Construct an extension field F of having order 81. In particular, if
, what do all of the elements of
F look like? To which of these
elements is
equal?
12.33.
Show that is a splitting field of
over
, where
,
, but
.
12.34.
Let F be a field and a nonconstant polynomial. If
K is a splitting field of
f(x) over F and L is any extension field of F, suppose that
is a homomorphism satisfying
for all
. If
is a root of f(x),
show that
is also a root of f(x).
12.35.
Find every automorphism of .
12.36.
Construct a splitting field for
over
. Show that it has degree 3 over
.
12.37.
Let F be any field and a nonconstant polynomial. If we let
, show that f(x)
and g(x) have the same splitting fields over
F.
12.38.
Let F be a field and a polynomial with
. Show that f(x)
has a splitting field K over
F with
.
12.5 Applications to Finite Fields
Let us see what we can deduce about
finite fields. If F is a finite
field, we know that its prime subfield must be isomorphic to
, for some prime p. By Theorem 12.7, F must have order
, for some positive integer n. We will construct a field of order
and show that, up to isomorphism,
there is only one such field.
The following concept looks suspiciously like calculus, but is not.
Definition 12.14.
Let F be a field and . Then the formal derivative of f(x)
is
.
Note that this has nothing whatsoever to do with limits, as limits do not necessarily make sense in an arbitrary field. The formula happens to agree with the one used for the derivative of real polynomials. We will also not be disturbed by the fact that the following lemma extends the similarity to calculus.
Lemma 12.5.
![$$f(x),g(x)\in F[x]$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq741.png)

- 1.
;
- 2.
; and
- 3.
.
Proof.
The first two parts follow immediately from the definition. The third is left
as Exercise 12.40.
Definition 12.15.
Let F be a field, and
. We say that a is a multiple root of f(x)
if
.
Example 12.33.
In , 2 is a multiple root of
, since the polynomial factors as
.
Theorem 12.14.
Let F be a field, and let
be a root of f(x).
Then a is a multiple root of
f(x) if and only if
.
Proof.
Suppose that a is a multiple root of f(x),
say , with
. Then by Lemma 12.5,
. Thus,
. Conversely, suppose that
. By Theorem 11.2,
, for some
. Thus,
. As
, we have
. By Theorem 11.2,
, and hence
.
Corollary 12.3.
Let F be a field and let be irreducible. Let K be a splitting field of f(x)
over F. If f(x)
has a multiple root in K, then
is the zero polynomial.
Proof.
Let a be the multiple root. Then (multiplying
f(x) by a suitable element of F to make it monic), we see that
f(x) is the minimal polynomial of
a over F. By Theorem 12.10, . But if
, then
, which is impossible. Therefore,
is the zero
polynomial.
Definition 12.16.
A field F is said to be perfect if no irreducible has multiple roots in any
splitting field of f(x) over F.
We digress from our discussion of finite fields to mention the following.
Theorem 12.15.
Every field of characteristic zero is perfect.
Proof.
If , with
and
, then
has leading coefficient
. Thus,
is not the zero polynomial. Apply
Corollary 12.3.
Actually, finite fields are perfect too! Let us see why.
Lemma 12.6.
Let F be a finite field of characteristic
p. Then the function
given by
is an automorphism.
Proof.
Since F is commutative, , for all
. By Theorem 8.14,
. If
, then since F is a field,
. Thus,
is one-to-one. Since F is finite,
is onto as
well.
Theorem 12.16.
Every finite field is perfect.
Proof.
![$$f(x)\in F[x]$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq794.png)














What would an imperfect field look like? Clearly, it would have to be an infinite field of prime characteristic. Exercise 12.44 shows how to construct an imperfect field.
Back to the finite fields!
Lemma 12.7.
Let F be a field of prime characteristic
p and let n be a positive integer. If , then K is a subfield of F.
Proof.
See Exercise 8.40.
Theorem 12.17.
Let p be a prime and n a positive integer. Then a field
F has order if and only if it is a splitting
field of
over the prime subfield, (an
isomorphic copy of)
.
Proof.
Let F have order . Then U(F)
has order
. Thus, if
, then
, and hence
. Clearly,
as well. Thus, every element of
F is a root of
. By Corollary 11.3,
can only have
roots. Thus,
splits over F, and surely it cannot split over any
smaller field, as all of the roots must be present. Therefore,
F is a splitting field of
over
.
Conversely, let F be a splitting field of over
. By Lemma 12.7, the roots of
form a subfield K of F. Since
splits over K, we must have
. Furthermore, the formal derivative
of
is
, which has no roots. Therefore, by
Theorem 12.14,
has no multiple roots. In particular,
, as
required.
Theorem 12.18.
If k is a positive integer, then there is a
field of order k if and only if
for some prime p and positive
integer n. All fields of order
are isomorphic.
Proof.
By Theorem 12.7, a finite field
must have order . Theorem 12.12 tells us that
has a splitting field over
. By Theorem 12.17, this splitting
field has order
. But Theorem 12.17 also says that
every field of order
is such a splitting field. By
Corollary 12.2, these splitting fields are isomorphic.
The unique (up to isomorphism)
field of order is called the Galois field of order
.
We can also determine the subfields of a finite field. In order to do so, we will need the following theorem, which is of interest on its own.
Theorem 12.19.
Let F be a field. Then any finite subgroup G of U(F) is cyclic.
Proof.
Since G is a finite abelian group, Theorem
5.3 tells us that it is a direct
product of cyclic groups. If all of these cyclic groups have
relatively prime orders, then by Theorem 5.4, G is cyclic, and we are done. Otherwise,
we may assume that G has a
subgroup , and there exists a prime p dividing the orders of a and b. By Cauchy’s theorem,
and
each contain an element of order
p. Thus, G has a subgroup isomorphic to
. But every element of
has order 1 or p. That is, we have at least
roots for the polynomial
, which has degree p, giving us a contradiction and
completing the proof.
Theorem 12.20.
Let F be a field of order , for some prime p and positive integer n. Then every subfield of F has order
, for some positive divisor m of n. Furthermore, for each positive divisor
m of n, F
has exactly one subfield of order
, namely
.
Proof.

![$$n=[F:\mathbb Z_p]=[F:K][K:\mathbb Z_p].$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_Equ22.png)
![$$[K:\mathbb Z_p]=m$$](/epubstore/L/G-T-Lee/Abstract-Algebra/OEBPS/images/429924_1_En_12_Chapter/429924_1_En_12_Chapter_TeX_IEq856.png)













To prove the uniqueness of this
subfield, suppose that L is
another subfield of F with
elements. Then U(K)
and U(L) are both subgroups of order
in U(F).
However, Corollary 3.3 tells us that U(F)
has only one such subgroup. Therefore,
. As the unit group of a field
consists of everything except 0, we have
, as
required.
Exercises
12.39.
Find the smallest field containing exactly 3 proper subfields.
12.40.
Let F be a field and . Show that
.
12.41.
Let be an irreducible polynomial of
degree 3. If K is a splitting
field of f(x) over
, show that
or
.
12.42.
Let K be a field of order for some prime p and positive integer n, having subfields F and L of orders
and
, respectively. Find the order of
.
12.43.
Let F be a field and an irreducible polynomial having a
multiple root in some extension field of F. Show that char
for some prime p,
for some
, and that at least one of the
is transcendental over the prime
subfield of F.
12.44.
Let be a polynomial ring over
and
its field of fractions. Show that the
polynomial
is irreducible over F, but that it has a multiple root in
some extension field of F. In
particular, conclude that F is
not a perfect field.
12.45.
Theorem 12.19 tells us that the
unit group of a finite field is cyclic. If char , show that the unit group of an
infinite field is not cyclic.
12.46.
Suppose char . Let us prove that the preceding
exercise still holds. Suppose, to the contrary, that U(F)
is cyclic. Let
.
- 1.
Show that
.
- 2.
If a is algebraic over
, show that F is finite, and we are done.
- 3.
If a is transcendental over
, show that there exists an integer n such that
, and obtain a contradiction.
12.47.
Suppose we wrote as a product of irreducibles over
. Show that each of these irreducible
polynomials has degree 1 or 3. (Please do not actually write the
polynomials!)
12.48.
Show that for every prime p and positive integer n, there exists an irreducible polynomial
of degree n in .