0% found this document useful (0 votes)
14 views30 pages

Finite Fields

Uploaded by

pappukapapita
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views30 pages

Finite Fields

Uploaded by

pappukapapita
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 30

NOTES ON FINITE FIELDS

AARON LANDESMAN

C ONTENTS
1. Introduction to finite fields 2
2. Definition and constructions of fields 3
2.1. The definition of a field 3
2.2. Constructing field extensions by adjoining elements 4
3. A quick intro to field theory 7
3.1. Maps of fields 7
3.2. Characteristic of a field 8
3.3. Showing the characteristic of any finite field is a prime 8
4. Characterization of finite fields 10
5. Properties of finite fields 13
5.1. The multiplicative group of a finite field 13
5.2. Frobenius 14
5.3. Containments of finite fields 15
6. Counting Linear Algebraic Data 17
6.1. Matrices over Finite Fields 17
6.2. Counting independent sets 18
6.3. Counting Matrices 19
6.4. Counting subspaces 19
6.5. A Further exercise on Grassmannians 21
7. Counting Polynomials 22
7.1. The average number of roots of a polynomial 22
7.2. Squarefree polynomials 23
Appendix A. Existence of algebraic closures 25
Appendix B. Basics of rings 29
B.1. Quotients 30
References 30

1
2 AARON LANDESMAN

1. I NTRODUCTION TO FINITE FIELDS


In this course, we’ll discuss the theory of finite fields. Along the
way, we’ll learn a bit about field theory more generally. So, the nat-
ural place to start is: what is a field? Many fields appear in nature,
such as the real numbers, the complex numbers the rational num-
bers, and even finite fields! Before giving a formal definition, let’s
see some examples.
Example 1.1. The rational numbers Q = ba : a, b ∈ Z, b 6= 0 are

a field. The key properties are that we can multiply rational num-
bers, add rational numbers (via addition of fractions) and further
that nonzero rational numbers have inverses. That is, ba · ba = 1 when-
ever a 6= 0.
Now, let’s see some examples of finite fields.
Example 1.2. Consider the field F2 , the finite field with two ele-
ments. Call these elements 0, 1. The addition law is given by 0 + a =
a + 0 = a and 1 + 1 = 0. The multiplication law is given by 1 · a = a
and 0 · a = 0. 1 is invertible and its inverse is given by 1 since
1 · 1 = 1. This can succinctly be described by Z/2Z.
Example 1.3. Next, let’s consider the finite field with 3 elements. As
above, we can consider Z/3Z. Elements can be added and multi-
plied by reducing addition and multiplication in Z modulo 3. The
key property to check is that nonzero elements have inverses (mean-
ing that for any nonzero a there is some b with ab = 1). Indeed,
1 · 1 = 1 and 2 · 2 = 1.
Warning 1.4. So far, we have seen that Z/2Z and Z/3Z are fields.
However, Z/4Z is not a field! The way to see this is that there is no
element a ∈ Z/4Z with 2a = 1. Indeed, either 2a = 2 or 2a = 0. So,
Z/nZ is not in general a field.
Question 1.5. Do you think there exists a finite field of order 4? Do
you think there exists a finite field of order 5? Do you think there
exists a finite field of order 6? For which n ∈ Z does there exist a
finite field with n elements?
NOTES ON FINITE FIELDS 3

2. D EFINITION AND CONSTRUCTIONS OF FIELDS


Before understanding finite fields, we first need to understand
what a field is in general. To this end, we first define fields. After
defining fields, if we have one field K, we give a way to construct
many fields from K by adjoining elements.
2.1. The definition of a field. A field is a special type of ring. So, we
first define a ring:
Definition 2.1. A commutative ring with unit is a set R together
with two operations (+, ·) satisfying the following properties:
(1) Associativity: a + (b + c) = ( a + b) + c, a · (b · c) = ( a · b) · c
(2) Commutativity: a + b = b + a, a · b = b · a
(3) Additive identity: there exists 0 ∈ R so that a + 0 = a
(4) Multiplicative identity: there exists 1 6= 0 ∈ R so that 1 · a = a
(5) Additive inverses: For every a ∈ R, there is a additive in-
verse, denoted − a satisfying a + (− a) = 0
(6) Distributivity of multiplication over addition: a · (b + c) =
( a · b) + ( a · c) .
Remark 2.2. Any mention of “ring” in what follows implicitly means
“commutative ring with unit.” There will be no noncommutative
rings or rings without units.
Definition 2.3. A field is a nonzero ring K, such that every nonzero
element has a multiplicative inverse. That is, 1 6= 0, and for each
a ∈ K with a 6= 0, there is some a−1 ∈ K so that a · a−1 = 1.
Definition 2.4. A finite field is simply a field whose underlying set
is finite.
Example 2.5. Given any prime number p, the set Z/pZ forms a field
under addition and multiplication. This field is denoted F p . Nearly
all the axioms are immediate, except possibly for the existence of
multiplicative inverses.
Exercise 2.6. Verify that every nonzero element has a multiplicative
inverse in two ways:
(1) Use the Euclidean algorithm to show that for any a < p there
exists some b with ab ≡ 1 mod p and conclude that b is an
inverse for a. Hint: Use that gcd( a, p) = 1.
(2) Show that a p−1 = 1, so a p−2 is an inverse for a. This is also
known as “Fermat’s Little theorem,” not to be confused with
“Fermat’s Last theorem,” which is much more difficult. Hint:
4 AARON LANDESMAN

Show that the powers of any element form a subgroup of


(Z/pZ)× := Z/pZ − {0} under multiplication. Use La-
grange’s theorem (i.e., the order of a subgroup divides the
order of the ambient group) to deduce that this subgroup gen-
erated by a has order dividing #(Z/pZ)× = p − 1. Conclude
that am = 1 for some m dividing p − 1 and hence a p−1 = 1.
2.2. Constructing field extensions by adjoining elements. We now
explain how to construct extensions of fields by adjoining elements.
Here is a prototypical example:
√ 
Example 2.7. Consider the field Q 2 . How should we interpret
this? The elements of this field are of the form
√  n √ o
Q 2 = a + b 2 : a, b ∈ Q .
Multiplication works by
 √  √  √
a + b 2 c + d 2 = ( ac + 2bd) + ( ad + bc) 2.

Here is another perspective on this field: What is 2? It is simply a
root of the polynomial x2 − 2. Therefore, we could instead consider
the field
Q[ x ] / ( x 2 − 2),
where this means the ring where we adjoin a root of the polynomial
x2 − 2. Concretely, Q[ x ] means polynomials with coefficients in Q,
and the notation Q[ x ]/( x2 − 2) means that in any polynomial f ( x ),
we can replace x2 by 2. So for example, if we had the polynomial x3 +
2x2 + 3 this would be considered equivalent to ( x2 ) · x + 2 · ( x2 ) +
3 = 2x + 4 + 3 = 2x + 7. In this way, we can replace any polynomial
with a polynomial of degree 1 of the form a + bx. Identifying x with
√ √ 
2 gives the isomorphism of this ring with the above field Q 2 .

Exercise 2.8. Describe the elements of the fields K as in Example 2.7


for K one of the following fields
√ 
(1) K = Q 3 ,
(2) K = Q 71/5 ,


(3) K = Q (ζ 3 ), for ζ 3 a primitive cube root of unity.


In each of the above cases, write K = Q[ x ]/ f ( x ) for an appropriate
polynomial f . In each of the above cases, what is the dimension of K
over Q, when K is viewed as a Q vector space?
NOTES ON FINITE FIELDS 5

Definition 2.9. Let K be a field. Define the polynomial ring


( )
n
K [ x ] := ∑ ai x i : ai ∈ K .
i =0

For f ∈ K [ x ], define
K [ x ]/( f ) := K [ x ]/ ∼
where ∼ is the equivalence relation defined by g ∼ h if f | g − h.
Exercise 2.10. Show that K [ x ]/( x ) ' K, where the map is given by
sending a polynomial to its constant coefficient.
Lemma 2.11. Let K be a field and let f ∈ K [ x ] be a monic irreducible
polynomial. Then K [ x ]/( f ) is a field.
Proof. Note that K [ x ]/( f ) is a ring as it inherits multiplication and
addition and all the resulting properties of a ring from K [ x ]. (Check
this!) Therefore, it suffices to check that if f is monic and irreducible,
then every element has an inverse. In other words, given any g ∈
K [ x ]/( f ), we need to show there is some h with gh = 1. We can
consider g ∈ K [ x ] as a polynomial of degree less than f . Since f
is irreducible, and deg g < deg f , it follows that the two polynomi-
als share no common factors. Then, by the Euclidean algorithm for
polynomials (if you have only seen the euclidean algorithm over the
integers, check that the natural analog to the Euclidean algorithm for
the integers works equally well in polynomial rings over arbitrary
fields, where the remainder is then a polynomial of degree less than
the polynomial you are dividing by) we obtain some h, α ∈ K [ x ] with
gh + f α = 1 as elements of K [ x ]. It follows that gh ∼ 1 in K [ x ]/( f )
because gh − 1 = f α in K [ x ]. 
Exercise 2.12. Let K be a field and f ∈ K [ x ] a monic irreducible poly-
nomial. Suppose L = K [ x ]/( f ). Show that dimK L = deg f , where
deg f denotes the degree of the polynomial f and dimK L denotes
the dimension of L as a K vector space.
Example 2.13. Consider the field F2 [ x ]/( x2 + x + 1). We claim this
is a finite field of order 4. Indeed, this holds because the polynomial
x2 + x + 1 is irreducible. To check this, we only need to check it has
no linear factors. It has a linear factor if and only if x2 + x + 1 has a
root in F2 . But, when we evaluate it at 0 we get 1 mod 2 and when
we evaluate it at 1, we get 1 mod 2. So it has no roots, and the claim
follows from Lemma 2.11.
6 AARON LANDESMAN

Exercise 2.14 (Squares in finite fields). (1) For any p > 2, show
p +1
that there are exactly 2 elements x ∈ F p with x = y2 for
some y ∈ F p . We call such x squares.
(2) Conclude that there is some x ∈ F p which is not a square
whenever p > 2. Hint: Show that if x = y2 then we also have
x = (−y)2 and further that there y and −y are the only two
elements of F p squaring to x.
(Challenge) Show that for any a ∈ F p , the equation x2 + y2 = a has a
solution x, y ∈ F p with x and y not both 0.
Example 2.15. Let p > 2 be a prime and let ε ∈ F p be an element
which is not a square (which exists by Exercise 2.14). Then,
F p [ x ]/( x2 − ε)
is a finite field of order p2 . It is order p2 because it is a two dimen-
sional vector space over F p spanned by the basis 1 and x. It is a field
because x2 − ε is irreducible in F p [ x ]. Indeed, to see this, note that
if it were not irreducible, it would factor as a product of two linear
factors, which means it would have a root. But, if it had a root, there
would be some y ∈ F p so that y2 = ε. However, we chose ε not to be
a square, and so no root exists.
NOTES ON FINITE FIELDS 7

3. A QUICK INTRO TO FIELD THEORY


In order to classify finite fields, we’ll need some inputs from field
theory. In particular, we’ll need to understand maps of fields and the
characteristic of a field, which we discuss in this section.
3.1. Maps of fields.
Definition 3.1. Given two fields K and L, a map φ : K → L is a map
of sets sending 1 7→ 1, 0 7→ 0 such that φ( a +K b) = φ( a) + L φ(b) and
φ ( a · K b ) = φ ( a ) · L φ ( b ).
Remark 3.2. Sometimes, a map of fields is referred to as a homomor-
phism or extension. Whenever we have a map of fields, it is required
to be compatible with the addition and multiplication operations, as
defined above. If we do not wish to require such compatibility, we
will call the map “a map of sets”
Remark 3.3. We shall typically drop the subscript +K , ·K on addition
and multiplication when it is clear from context.
Exercise 3.4. Verify from the definition of map that
φ ( a −1 ) = φ ( a ) −1
and
φ(− a) = −φ( a).
We next prove that maps of fields are injective. If you have not
worked much with the notion of injectivity before, you may want to
try the following exercise first.
Exercise 3.5. Show that a map of rings is injective (using the defini-
tion that f : R → S is injective if f ( a) = f (b) implies a = b) if and
only if the only element mapping to 0 is 0. Hint: Consider f ( a − b).
Lemma 3.6. Any map of fields is injective.
Proof. By Exercise 3.5 it suffices to show that any c 6= 0 satisfies
φ(c) 6= 0. Indeed, 1 = φ(1) = φ(cc−1 ) = φ(c)φ(c−1 ), and so
φ(c) 6= 0 because its product with φ(c−1 ) is nonzero. 
Remark 3.7. Because of Lemma 3.6, a map of fields is also typically
called an extension of fields or a field extension.
Remark 3.8. The property that maps of fields are injective is very
special to fields. Indeed, it is not true for groups. For example, the
map Z → {1} is not injective!
8 AARON LANDESMAN

Remark 3.9. Using Lemma 3.6, whenever we have a map of fields


φ : K → L, we can consider L as a vector space over K. The map
K × L → L corresponding to scalar multiplication is given by
K×L→ L
( a, b) 7→ φ( a) · b
3.2. Characteristic of a field.
Definition 3.10. Let K be a field. If there is some n so that
(3.1) n := 1| + 1 +{z· · · + 1}
n
is equal to 0 in K, the the minimal such n is defined to be the char-
acteristic of K, denoted char(K ). If no such n ∈ Z>0 exists, then we
say K has characteristic 0.
Example 3.11. The rational numbers Q has characteristic 0, but the
field F p has characteristic p.
Exercise 3.12 (Important exercise). Let p be a prime number and sup-
pose K is a field of characteristic p. Show that for any x, y ∈ K, we
have
( x + y) p = x p + y p .
Hint: Expand the left hand side using binomial coefficients, and
show that p divides nearly all of the binomial coefficients.
3.3. Showing the characteristic of any finite field is a prime.
Lemma 3.13. The characteristic of any field is either 0 or prime.
Proof. Note that the characteristic cannot be 1 because 1 6= 0 by defi-
nition of a field. So, we have to show that the characteristic is never
composite.
Let n be a composite number with n = f g for f , g > 1 two factors
of n.
Exercise 3.14. Suppose a, b ∈ K with ab = 0. Then show either a = 0
or b = 0.
By the above exercise, if n = f g = 0, then either f = 0 or g = 0.
Say f = 0. But then, we obtain that f < n, and so K does not have
characteristic n. 
Definition 3.15. For K a field, we say a subset K 0 ⊂ K is a subfield
if it is a field and the inclusion K 0 ⊂ K is a map of fields (meaning
1 7→ 1, 0 7→ 0 and the multiplication and addition are compatible).
NOTES ON FINITE FIELDS 9

Exercise 3.16. Verify similarly that any field of characteristic 0 con-


tains Q as a subfield. Hint: Define a map of fields
φ: Q → K
a
7→ ab−1 .
b
Use that b ∈ K is nonzero by the assumption that K has characteristic
0 to show this is well defined.
Lemma 3.17. The characteristic of any finite field is prime (and, in partic-
ular, never 0).
Proof. By Lemma 3.13, we only need to show the characteristic of
a finite field is nonzero. So, it suffices to show every characteristic
0 field is infinite. But, by Exercise 3.16, every characteristic 0 field
contains Q as a subfield, and is therefore infinite. 
Lemma 3.18. Any field K of characteristic p > 0 (for p a prime) contains
F p as a subfield.
Proof. Inside K, consider the subset {0, 1, 2, . . . , p − 1}. These form p
distinct elements because char K = p. By definition, of n = 1| + 1 +{z· · · + 1},
n
the elements 0, 1, . . . , p − 1 satisfy the same addition and multiplica-
tion rules as F p ' (Z/pZ). Therefore, when we restrict the multi-
plication and addition from K to {0, 1, 2, . . . , p − 1}, we realize F p as
this subfield. 
Lemma 3.19. Any finite field K has order pn for p a prime and n ∈ Z.
Proof. By Lemma 3.18, K contains F p as a subfield, so we have an
inclusion F p → K. Observe that this makes K into a vector space
over F p . Because K is finite, it must even be a finite dimensional
vector space over F p . Say it has dimension n. Then we see that as a
vector space, K ' Fnp , so |K | = |F p |n = pn , as desired. 
So, we have established that every finite field has order a prime
power. Our next goal is to show that there is a unique finite field
of order pn . However, for this, we will need to introduce algebraic
closures, which we do now.
10 AARON LANDESMAN

4. C HARACTERIZATION OF FINITE FIELDS


We are nearly ready to show there is a unique finite field of order
pn ,for every prime p and every n ≥ 1.
In order to construct finite fields, we need the existence of an alge-
braic closure of F p , which we call F p . The proof of this is relegated
to various parts of an appendix.
Theorem 4.1. There exists a field F p with the following properties:
(1) Every element x ∈ F p satisfies a monic irreducible polynomial re-
lation of the form ∑in=0 ai xi for ai ∈ F p for n > 0,
(2) any polynomial with coefficients in F p factors as a product of linear
factors
(3) any finite extension of F p is a subfield of F p .
Proof. The first part follows from Theorem A.9. The second part fol-
lows from Lemma A.4. The last part amounts to Exercise A.8. 
Remark 4.2. The basic idea for the construction of F p is to adjoin all
roots of all irreducible polynomials to F p , and show that after doing
so we obtain a field.
First, we need a preparatory definition and lemma.
Definition 4.3. If K is a field and f := ∑in=0 ai xi ∈ K [ x ] is a polyno-
mial, we define the derivative of f , denoted f 0 , to be ∑in=1 iai xi−1 .
Lemma 4.4. Let K be a field and let f ∈ K [ x ] be a polynomial which factors
as a product of linear factors. Then, if gcd( f , f 0 ) = 1, f has no repeated
roots. That is, there is no a ∈ K with ( x − a)2 | f .
Proof. Suppose f has a repated root. Call that root r ∈ K. Then since
( x − r )2 | f , it follows from the product rule that x − r | f 0 . Therefore,
x − r | gcd( f , f 0 ), and so gcd( f , f 0 ) 6= 1, as desired. 
We can now state and prove our main result.
Theorem 4.5. Let p be a prime and n ≥ 1.
(1) There exists a finite field of order pn , notated F pn . Further, F pn is
n
realized as the set of elements of F p satisfying x p = x.
(2) Any two finite fields of order pn are isomorphic.
Proof. First let us show there exists a finite field of order pn . Let F p
denote an algebraic closure of F p . Define
n n
o
F pn : = x ∈ F p : x p = x .
NOTES ON FINITE FIELDS 11

We claim F pn is a field. To check this, the essential points to verify


are that F pn is closed under multiplication, addition, and inversion.
n n
(1) Addition: We need to show that if x p = x, y p = y then
n
( x + y) p = x + y. Indeed, this follows from Exercise 3.12.
n n
(2) Multiplication: We need to show that if x p = x, y p = y then
n
( xy) p = xy. Indeed, this is clear by commutativity of F p .
n
(3) Inversion: Given x 6= 0 with x p = x, we want to show x −1
n
(which exists as an element of F p ) satisfies ( x −1 ) p = x −1 . But
indeed,
n n
( x −1 ) p = ( x p ) −1 = x −1 ,
as desired.
Exercise 4.6. Verify the remaining properties such as distributivity
and commutativity to show that F pn is indeed a field. Hint: You may
be able to inherit many of these properties from F p .
To complete the proof of existence, we have to check that |F pn | = pn .
Indeed, by construction, the elements of F pn are the set of roots to
n
the polynomial f (t) = t p − t. So, it suffices to show this has pn
distinct roots in F p (using that Theorem 4.1 implies F p contains F pn ).
Note that f (t) has at most pn roots because it has degree pn . By
Theorem 4.1, f (t) factors as a product of pn distinct linear factors
over F pn . Further, the roots of f (t) are all distinct by Lemma 4.4
n
because f 0 (t) = pn · t p −1 − 1 = −1.
So, we have shown existence of finite field over order pn . It re-
mains to show uniqueness up to isomorphism.
Let K be some finite field of size pn . We want to construct an iso-
morphism K ' F pn . Because K is finite, it is a subfield of F p by
Theorem 4.1. Choose an extension φ : K → F p . We will show
im φ ⊂ F pn ⊂ F p . This will complete the proof as it will imply
that φ defines a map between two fields of size pn . It is then injective
by Lemma 3.6 and hence it will be surjective because it is an injective
map between two sets of the same finite size.
So, we will now show im φ ⊂ F pn . For this, it suffices to show
n
that any x ∈ K satisfies φ( x ) p = φ( x ). For this, it suffices to show
n
x p = x. This clearly holds for x = 0, so after dividing by x, it
n −1
suffices to show x p = 1. But now, note that K × := K − {0} is a
multiplicative group of size |K × | = |K | − 1 = pn − 1. Lagrange’s
theorem tells us that the order of any element of a group divides the
12 AARON LANDESMAN

order of the group. This implies that x m = 1 for some m | pn − 1,


n
which implies that x p −1 = 1. 
Exercise 4.7 (Tricky, but fun). (1) Let p be a prime number. How
many degree d monic irreducible polynomials are there over
Fp?
(2) Let q be a positive power of a prime number. How many
monic degree d irreducible polynomials are there over Fq ?
NOTES ON FINITE FIELDS 13

5. P ROPERTIES OF FINITE FIELDS


We next develop several interesting properties of finite fields.
5.1. The multiplicative group of a finite field. As we saw near the
end of the proof of Theorem 4.5, because multiplicative inverses ex-
ist, for any field K, the nonzero elements K × form a group under
multiplication. The identity element is 1. It turns out that finite fields
have a particularly nice multiplicative structure.
Proposition 5.1. There is an isomorphism F×
pn ' (Z/ ( p − 1)Z). That
n

is, F×
pn is cyclic.

Proof. To show F× n
pn is cyclic of order p − 1, since we know it has
order pn − 1 as a group, it suffices to show there is some element of
order pn − 1.
Exercise 5.2 (Tricky exercise). Verify using that all finite abelian groups
are products of cyclic groups (the fundamental theorem for finite
abelian groups) that if there is no element of order pn − 1 then there
is some m < pn − 1 with x m = 1 for all x ∈ F× pn . Hint: Show that if
ni
G ' ∏i Z/pi Z, (where the isomorphism holds by the fundamental
theorem of finite abelian groups) has some pi = p j for i 6= j then then
n
every element of G has order strictly less than | G | = ∏i pi i . For this
n nj
it may help to consider the subgroup Z/pi i × Z/p j . Then, show
using the Chinese Remainder theorem that if pi 6= p j for any i 6= j
then G is cyclic.
However, we cannot have x m = 1 for all x ∈ F× n
pn with m < p − 1
because x m − 1 only has m < pn − 1 roots in F p . Hence, there is
some element of F× n
pn of order exactly p − 1, and so it is isomorphic
to Z/( pn − 1)Z× . 
Exercise 5.3. Using Proposition 5.1 we can now prove results about
roots of unity modulo primes.
(1) Let p be an odd prime. Using Proposition 5.1, show that −1
is a square mod p if and only if p ≡ 1 mod 4.
(2) Let p be an odd prime. Show that there is some x 6≡ 1 mod p
so that x3 ≡ 1 mod p if and only if p ≡ 1 mod 3.
(3) Let p be an odd prime. Determine a necessary and sufficient
condition on p modulo n such that there will be n distinct
roots of unity modulo p, i.e., there are n distinct residues
x1 , . . . , xn mod p with xin ≡ 1 mod p.
14 AARON LANDESMAN

(4) Given n and p, determine the number of nth roots of unity


mod p. That is, determine the number of residues x so that
x n ≡ 1 mod p.
Exercise 5.4. Prove Wilson’s theorem: show that ( p − 1)! ≡ −1 mod
p.
5.2. Frobenius. In what follows, we will let q denote a power of p,
say q = pn .
Definition 5.5. The map
Frob p : Fq → Fq
x 7→ x p
is the Frobenius map.
Exercise 5.6. Verify that Frob p is a map Fq → Fq over F p . That is,
show that the natural inclusion i : F p → Fq respects Frob p in the
sense that Frob p ◦i = i. Hint: Show that for any x ∈ F p , x p = x. See
Exercise 2.6(2) for further help.
Our next goal is to show that the automorphisms of Fq over F p
(i.e., maps Fq → Fq as vectors spaces over F p ) are precisely id, Frob p , . . . , Frobnp−1 ,
where q = pn . First, we show that these are all distinct:
Lemma 5.7. Let q = pn . The maps id, Frob p , . . . , Frobnp−1 are distinct as
maps Fq → Fq .
Proof. We wish to show Frobap 6= Frobbp for 0 ≤ a < b ≤ n − 1.
After composing with Frob− a b− a
p , it suffices to show that Frob p is not
the identity for any 0 < b − a ≤ n − 1. That is, we have to show
Frob p , . . . , Frobnp−1 are all distinct from id.
To show that Frobcp is not the identity, we have to show there is
c c
some x ∈ Fq with x p 6= x. However, since x p − x is a polynomial of
degree pc , there are at most pc such elements. Since |Fq | = pn > pc ,
c
there is some element x ∈ Fq with x p 6= x. 
Hence, we have produced that there are at least p distinct auto-
morphisms of Fq given by powers of Frobenius. We next want to
show that these are all the automorphisms of Fq . For this, we will
now give an explicit construction of Fq as a field, by adjoining an
element to F p , via the method in subsection 2.2.
Lemma 5.8. We can express F pn in the form F p [ x ]/( f ) for f ∈ F p [ x ] of
degree n.
NOTES ON FINITE FIELDS 15

Proof. Pick y ∈ F pn to be a generator of F× pn (which is possible by


Proposition 5.1). Note that y generates F pn over F p because all nonzero
elements of F pn are powers of y. Further, y satisfies an irreducible
monic polynomial f over F p , where f expresses a linear dependence
relation between powers of f . We obtain that F p [ x ]/( f ) is a field by
Lemma 2.11. We obtain a map
φ : F p [ x ] / ( f ) → Fq
x 7→ y.

Exercise 5.9. Verify this is a well defined map.


This map is necessarily injective by Lemma 3.6 but it is also sur-
jective because y generates Fq . Therefore it is an isomorphism. It
follows that f must have degree n since F p [ x ] is a dimension deg f
vector space over F p , but it is also isomorphic to F pn , which is a di-
mension n vector space over F p . 

Corollary 5.10. The automorphisms of Fq over F p are precisely id, Frob p , . . . , Frobnp−1 .
Proof. We have seen in Lemma 5.7 that these are all distinct, so it
suffices to show there are at most n automorphisms of Fq over F p .
However, by Lemma 5.8, we have Fq = F p [ x ]/( f ). Note that any
map F p [ x ]/( f ) → F p [ x ]/( f ) must send x to some root of f , and
further the map is determined by where it sends x. Since deg f = n,
there are at most deg f = n roots of f and hence at most n such maps,
as we wanted to show. 

5.3. Containments of finite fields. Let us now determine when F pm ⊂


F pn for m, n > 0. First, we establish this containment when m | n.
Lemma 5.11. If m | n then F pm ⊂ F pn .
Proof. Recall from Theorem 4.5 that F pm was precisely the set of ele-
m 2m m
ments in F p with x p = x. We also have x p = x p = x. If m | n, say
n dm ( d −1) m
m = dn then iterating this d times we obtain x p = x p = xp =
m
· · · = x p = x, so x ∈ F pn . 
In fact, the above case is the only case that n | m, as we will now
see.
Proposition 5.12. For p a prime and n, m > 0, we have F pm ⊂ F pn if and
only if m | n.
16 AARON LANDESMAN

Proof. If m | n, then the inclusion holds by Lemma 5.11. Conversely,


if F pm ⊂ F pn then F pn is a vector space over F pm . Say F pn has dimen-
sion d over F pm . It follows that |F pn | = |F pm |d , so pn = ( pm )d = pmd
and so m | n. 
NOTES ON FINITE FIELDS 17

6. C OUNTING L INEAR A LGEBRAIC D ATA


Today, we’ll aim to get acquainted with how to count objects over
finite fields. In order to guide our study, we’ll begin by trying to
answer the following question:

Question 6.1. How many invertible matrices are there over Fq ?

6.1. Matrices over Finite Fields. As a motivating question, for to-


day, we’ll count the number of invertible matrices over Fq .

Definition 6.2. For L a field, Let GLn ( L) denote the group of invert-
ible n × n matrices over L. In particular, GLn (Fq ) denotes the set of
n × n matrices invertible over Fq .

Let’s start with some examples of cases when n is small.

Example 6.3. How many 1 × 1 matrices are there over Fq . That is,
what is # GL1 (Fq )? Well, a 1 × 1 matrix is the same as an element
of Fq . Since Fq is a field, any nonzero element of Fq is invertible.
Therefore, we are trying to count the number of nonzero elements of
Fq , which is q − 1.

Example 6.4. Let’s next try to count # GL2 (Fq ). Now, we are consid-
ering
 
a b
c d

with a, b, c, d ∈ Fq and ad − bc 6= 0. We can count these directly by


splitting into 2 cases.
(1) If ad = 0, then we need bc 6= 0. In this case, there are 2q − 1
possibilities for ( a, d), depending on whether a = 0, d = 0, or
both equal 0. Then, b and c each have q − 1 possibilities, and
so this case contributes (q − 1)2 (2q − 1).
(2) If ad 6= 0, then there are (q − 1)2 possibilities for ( a, d). In
this case, if bc = 0, as above, there are 2q − 1 possibilities for
(b, c). If bc 6= 0, then we must avoid the case that bc = ad.
Choosing b to be an arbitrary element of Fq × := Fq − {0},
we must avoid the unique nonzero value of c so that bc = ad.
There are therefore q − 2 possibilities for c. Altogether, this
gives (q − 1)2 ((2q − 1) + (q − 1)(q − 2)) possibilities.
18 AARON LANDESMAN

Adding up the above cases yields


GL2 (Fq ) = (q − 1)2 (2q − 1) + (q − 1)2 ((2q − 1) + (q − 1)(q − 2))
= ( q − 1)2 ( q2 + q )
= ( q + 1) q ( q − 1)2 .
Whew, that was a lot of work! Now how are we ever going to
compute the number of 3 × 3 matrices? But given the fairly simple
answer, we might expect there is a nicer solution than dividing into
cases based on which entries are 0.
6.2. Counting independent sets. To approach a nice way to count
matrices, we’ll first have to learn to count the number of subspaces
of a vector space. Before counting subspaces, let’s count the number
of independent sets in a vector space.
Lemma 6.5. For V an n-dimensional vector space over Fq , V has qn ele-
ments.
Proof. Let e1 , . . . , en be a basis for V. Then an arbitrary element of V
can be written uniquely in the form a1 e1 + · · · + an en with ai ∈ Fq .
There are q possibilities for each ai and hence qn total elements. 
Let V be a vector space over L with v1 , . . . , vk ∈ V. We say a tuple
(v1 , . . . , vk ) is independent if the associated set {v1 , . . . , vk } is inde-
pendent. Concretely, this means that if ∑ik=1 ai vi = 0 for some ai ∈ L
then a1 = · · · = ak = 0. Recall that a set is independent if and only if
it can be extended to a basis.
Proposition 6.6. The number of ordered tuples of independent elements
(e1 , . . . , ek ) of size k in Fq n is (qn − 1)(qn − q)(qn − q2 ) · · · (qn − qk−1 ).
Proof. To count the number of independent tuples, we will work
inductively. We will choose each ei one at a time, and count the
number of possible choices for ei assuming we have already picked
e1 , . . . , e i −1 .
To start this procedure, note that for {e1 } to be independent is the
same as e1 6= ~0, so there are #Fq n − 1 = qn − 1 possibilities.
If we have already chosen e1 , in order for e2 to be independent
from e1 , we need to choose it outside of the span of e1 . There are q
vectors in the span of e1 , and hence #Fq n − q = qn − q possibilities
for e2 .
In general, for e1 , . . . , ei−1 an independent set, there will be qi−1
vectors in their span. Therefore, the number of choices for ei such
that e1 , . . . , ei forms an independent set is #Fq n − qi−1 = qn − qi−1 .
NOTES ON FINITE FIELDS 19

Hence, there are qn − qi−1 ways to choose the ith basis vector, and
hence the total number of size k independent tuples is (qn − 1)(qn −
q ) · · · ( q n − q k −1 ). 

6.3. Counting Matrices. We can now count the number of invertible


matrices. In particular, this agrees with our answers for n = 1 and
n = 2 in Example 6.3 and Example 6.4, but the proof is much simpler
than dividing into cases as in Example 6.4.
The key to counting # GLn (Fq ) is the following useful linear alge-
braic fact:
Exercise 6.7. Show that the set of n × n invertible matrices over a
field k is in bijection with the set of ordered bases e1 , . . . , en of kn .
In particular, conclude # GLn (Fq ) is equal to the number of ordered
bases e1 , . . . , en of Fq n .
Corollary 6.8. We have
# GLn (Fq ) = (qn − 1)(qn − q)(qn − q2 ) · · · (qn − qn−1 ).
Proof. By Exercise 6.7, it suffices to count the number of ordered
bases of an n-dimensional vector space over Fq . An ordered basis is
the same as an ordered linearly independent set of size n. By Propo-
sition 6.6 there are # GLn (Fq ) = (qn − 1)(qn − q)(qn − q2 ) · · · (qn −
qn−1 ) such sets. 

6.4. Counting subspaces. We’ve just seen how counting bases re-
lates to counting tuples of independent vectors. Let’s next use this to
count the number of subspaces of a vector space of a given dimen-
sion. We could answer this question fairly quickly, but we prefer to
take our time, introducing a space parameterizing these objects.
Definition 6.9. Let n and k be integers with 0 ≤ k ≤ n. For L a
field, let Grassmannian Gr(k, n)( L) denote the set of k-dimensional
subspaces of Ln . In particular, Gr(k, n)(Fq ) denotes the set of k di-
mensional subspaces of Fq n .
In order to get comfortable with Grassmannians, let’s go over a
few examples.
Example 6.10. If L is any field, n is arbitrary, and k = 0, Gr(0, n)( L) =
{0}, since 0 is the only 0-dimensional subspace. Similarly, if k =
n, we find Gr(n, n)( L) = { Ln } has size 1, since Ln is the only n-
dimensional subspace of Ln .
20 AARON LANDESMAN

Exercise 6.11 (Useful exercise). (1) In the case k = 1 and n = 2,


identify Gr(1, 2)( L) with the set of lines in L2 . Directly count
Gr(1, 2)(Fq ).
(2) More generally, identify Gr(1, n)( L) with the set of lines in Ln ,
and compute a formula for # Gr(1, n)(Fq ). Hint: How many
points are on a line?
Remark 6.12. This Gr(1, n)( L) will play an important role toward
the end of the course, and is also referred to as Pn−1 ( L), or n − 1
dimensional projective space.
Using this idea, let’s count Gr(k, n)(Fq ).
Lemma 6.13. The number of pairs (S, (e1 , . . . , ek )) where S is a k-dimensional
subspace of Fq n and e1 , . . . , ek is an ordered basis for S is (qn − 1)(qn −
q ) · · · ( q n − q k −1 ).
Proof. Observe that S is uniquely determined as the span of e1 , . . . , ek ,
with the only requirement on the ei being that e1 , . . . , ek are inde-
pendent. Hence, the number of such pairs is the same as the num-
ber of k-tuples of independent vectors, which we found in Proposi-
tion 6.6 
Proposition 6.14. We have
(qn − 1)(qn − q) · · · (qn − qk−1 )
Gr(k, n)(Fq ) = .
(qk − 1)(qk − q) · · · (qk − qk−1 )
Proof. To count the number of k-dimensional subspaces, we can count
the number of k-dimensional subspaces together with an ordered ba-
sis, and then divide out by the number of ordered bases of a fixed
subspace. By Lemma 6.13, there are (qn − 1)(qn − q) · · · (qn − qk−1 )
subspaces with an ordered basis and there are (qk − 1)(qk − q) · · · (qk −
qk−1 ) ways to choose an ordered basis of a fixed subspace (we saw
this above when counting the number of invertible matrices).
(qn −1)(qn −q)···(qn −qk−1 )
Therefore, the number of k-dimensional subspaces of Fq n is (qk −1)(qk −q)···(qk −qk−1 )
.

Exercise 6.15. Let n (F ) denote the group of n × n matrices with de-
terminant equal to 1. (If you haven’t seen the notion of determinants
before, look it up!) What is #n (F )?
Exercise 6.16. A subgroup H ⊂ GLn (Fq ) is called Borel if it is con-
jugate the subgroup of all upper triangular matrices. Recall that a
subgroup H ⊂ G is conjugate to H 0 ⊂ G if there is some g ∈ G with
NOTES ON FINITE FIELDS 21

gHg−1 = H 0 . For example, in the case n = 2, any Borel subgroup of


GL2 (Fq ) is conjugate to
  
a b ×
: a, c ∈ Fq , b ∈ Fq .
0 c
(1) How many Borel subgroups of GL2 (Fq ) are there?
(2) How many Borel subgroups of GLn (Fq ) are there? Hint: Use
the orbit stabilizer theorem: When B is the upper triangular
Borel subgroup, for which g ∈ GLn (Fq ) does gBg−1 = B?
(3) Can you find a direct bijection between Borel subgroups over
Fq and elements of Gr(1, 2)(Fq ). Can you describe a param-
eter space for certain types of subspaces of Fq n which is nat-
urally in bijection with the set of Borel subgroups of GLn (Fq )
and explicitly describe this bijection?
6.5. A Further exercise on Grassmannians. If you look at the for-
mula for # Gr(k, n)(Fq ), you may notice # Gr(k, n)(Fq ) = # Gr(n −
k, n)(Fq ). This is not a coincidence, and the reason they are the same
is ultimately due to the fact that dim V ∨ = V, for V ∨ the dual vector
space of V. We now investigate this.
Exercise 6.17. Fix a vector space V of dimension n over a field k.
(1) Show there is a bijection between subspaces W ⊂ V and quo-
tient vector spaces V → Q. Show these are further in bijection
with exact sequences of vector spaces
(6.1) 0 W V Q 0.
Recall that being an exact sequence means V → Q is a surjec-
tion of vector spaces with W the kernel of the map V → Q.
(2) Construct a natural bijection between subspaces W ⊂ V and
quotients V ∨ → Q.
(3) Use the previous two parts to identify k-dimensional sub-
spaces of V with n − k dimensional subspaces of V ∨ .
(4) Conclude from the previous parts that # Gr(k, n)(Fq ) = # Gr(n −
k, n)(Fq ). (Do this using the previous parts, although you can
certainly conclude it immediately from Proposition 6.14.)
22 AARON LANDESMAN

7. C OUNTING P OLYNOMIALS
7.1. The average number of roots of a polynomial. Let’s compute
the average number of roots of a monic degree d polynomial over a
finite field. As a warm up, we’ll need the following:
Let Polyd (Fq ) denote the set of degree d monic polynomials over
Fq .

Exercise 7.1 (Easy exercise). Show # Polyd (Fq ) = qd .


Proposition 7.2. The average number of roots of a degree d monic polyno-
mial over Fq is 1.
Proof. Let
Ψd := {( f , a) : f ∈ Polyd , a ∈ Fq , f ( a) = 0}

The average number of roots of a monic polynomial can be expressed


as the ratio
#Ψd
#{ f : f ∈ Polyd }

We know the denominator has size qd . So it suffices to show the


numerator also has size qd . So, we want to count the size of Ψd . For
this, we use the map
π : Ψ d → Fq
( f , a) 7→ a.

We claim that for every b ∈ Fq , #π −1 (b) = qd−1 . The result will then
follow because
#Ψd = #{( f , a) : f ∈ Polyd , a ∈ Fq , f ( a) = 0} = ∑ #{ f : f ∈ Polyd , f (b) = 0}
b ∈Fq

= ∑ #π −1 (b)
b ∈Fq

= ∑ q d −1
b ∈Fq

= qd ,
as we wanted to show. So, it suffices to prove:

Lemma 7.3. For every b ∈ Fq , #π −1 (b) = qd−1 .


NOTES ON FINITE FIELDS 23

Proof. By division, any monic degree d polynomial f ( x ) which has b


as a root can be written uniquely as ( x − b) g( x ) for a unique degree
d − 1 monic polynomial g( x ). So,
#π −1 (b) = # Polyd−1 (Fq ) = qd−1 .


Exercise 7.4. Another way of stating Proposition 7.2 is that the num-
ber of pairs ( f , α) ∈ Polyd (Fq ) × Fq with x − α | f is qd . Show that
the number of ordered tuples ( f , α1 , . . . , αd ) ∈ Polyd (Fq ) × Fq d with
( x − α1 ) · · · ( x − αd ) | f is also qd . In other words, the number of
degree d polynomials with an ordered choice of all d roots is qd .
Exercise 7.5. Give a common generalization of the preceding exer-
cise and Proposition 7.2 by showing that for any k ≤ d, the number of
tuples ( f , α1 , . . . , αk ) ∈ Polyd (Fq ) × Fq k with ( x − α1 ) · · · ( x − αk ) | p
is qd . Hint: Follow the method of proof of Proposition 7.2.
7.2. Squarefree polynomials.
Proposition 7.6. Let Polysf d (Fq ) denote the set of squarefree polynomials
of degree d over Fq .

d d −1
q − q
 if d > 1
# Poly d (Fq ) = q
sf
if d = 1

1 if d = 0
Proof. By unique factorization, we can write any monic polynomial
h( x ) uniquely as a product h( x ) = f ( x ) g( x )2 for f ( x ), g( x ) monic
polynomials where f ( x ) is squarefree. This gives the recurrence re-
lation
   
Polyd (Fq ) = Polysf d (Fq ) ä Polysf d−2 (Fq ) × Poly1 ä Polysf d−4 (Fq ) × Poly2 (Fq ) ä
 
· · · ä Polysf d mod 2 (Fq ) × Polybd/2c
Now, we take sizes of both sides, and assume, by induction, that
the formula holds for k < d. The base cases d = 0 and d = 1 are
elementary. For the inductive step, with d > 1,
qd = # Polysf d (Fq ) + (qd−2 − qd−3 )q + (qd−4 − qd−5 )q2 + · · · + qdd/2e
= # Polysf d (Fq ) + (qd−1 − qd−2 ) + (qd−2 − qd−3 ) + · · · + (qdd/2e+1 − qdd/2e ) + qdd/2e
= # Polysf d (Fq ) + qd−1
24 AARON LANDESMAN

Therefore, # Polysf d (Fq ) = qd − qd−1 , as desired. 


Exercise 7.7. Using a similar method to that of Proposition 7.6, count
the number of pairs of degree n polynomials ( P, Q) for P( x ) = x n +
an−1 x n−1 + · · · + a0 and Q( x ) = x n + bn−1 x n−1 + · · · + b0 with ai ∈
Fq , bi ∈ Fq so that P and Q have no common irreducible factor. Hint:
Use a method similar to that of Proposition 7.6. As a further hint,
show that every pair of degree n polynomials ( P, Q) can be written
uniquely as ( f · p, f · q) where p and q have no common factors. Now
count the number of such pairs by induction on the degree of f .
Remark 7.8. If you understood Proposition 7.6 and Exercise 7.7 cor-
rectly, you may notice a striking similarity between the two answers.
There is indeed a connection, but the answer lies deep. Loosely
speaking, if you take a polynomial P( x ) with no repeated factors,
you can send it to the pair of polynomials ( P( x ) + P0 ( x ), P( x )). Here
P0 ( x ) denotes the derivative of P( x ). If P( x ) = ∑i ai xi then P0 ( x ) =
∑ i i · a i x i −1 .
Exercise 7.9. Verify that the above indeed defines a map from the
space of polynomials with no repeated factors to the space of pairs
of polynomials with no common factors.
In some sense (which we do not explain) this map explains why
the counts from Proposition 7.6 and Exercise 7.7 are so similar.
NOTES ON FINITE FIELDS 25

A PPENDIX A. E XISTENCE OF ALGEBRAIC CLOSURES


In this section, we now guide the reader through a proof of the
existence of algebraic closures in series of exercises. Before that, we
prove Theorem A.9, Lemma A.4, and Exercise A.8. which together
provide a proof of Theorem 4.1.
Definition A.1. An extension of fields φ : K → L is finite if φ makes L
into a finite dimensional vector space over K. An extension of fields
φ : K → L is algebraic if for every a ∈ L, there is a finite extension
K → L a with L a ⊂ L a subfield containing a.
In order to get the theory of finite fields off the ground, we will
need the existence of an algebraic closure. It is not too difficult to
show this exists, but to jump to the interesting stuff, we will defer it
for later:
Definition A.2. A field K is algebraically closed if any finite field
extension K → L is an isomorphism.
Exercise A.3. Show that the real numbers are not algebraically closed.
Show that the rational numbers are not algebraically closed.
Lemma A.4. Let K be a field. The following are equivalent.
(1) K is algebraically closed.
(2) Every monic irreducible polynomial over K has a root.
(3) Every monic irreducible polynomial over K factors as a product of
linear polyomials.
Proof. For (1) =⇒ (2), we suppose K is algebraically closed and
show every monic irreducible polynomial over K has a root. Let f be
any monic irreducible polynomial over K. Then, K [ x ]/( f ) is a field
extension of K. Because K is algebraically closed, the natural map
K → K [ x ]/( f ) is an isomorphism. Therefore, dimK K [ x ]/( f ) = 1
and so f has degree 1 by Exercise 2.12 (which says deg f = dimK K [ x ]/( f )),
and hence has a root.
Next, if (2) holds, one can prove (3) by induction on the degree of
the polynomial.
Finally, for (3) =⇒ (1), suppose K is not algebraically closed.
We want to show there is some irreducible polynomial over K which
does not factor completely. Let L be a finite extension of K with the
inclusion K → L not an isomorphism. Since K → L is an injection it
is not a surjection, so we may take some y ∈ L \ K. We claim there is
some monic irreducible polynomial f ∈ K [ x ] with f (y) = 0. Indeed,
this is the content of the following exercise.
26 AARON LANDESMAN

Exercise A.5. Let K → L be an algebraic extension. Show that any


element x ∈ L satisfies some monic irreducible polynomial f ( x ) =
x n + k n−1 x n−1 + · · · + k0 , for k i ∈ K. Hint: By definition of an alge-
braic extension, show that the powers of x satisfy some linear depen-
dence relation, and obtain the monic irreducible polynomial from
this relation.
Note that since y ∈
/ K, the polynomial f with f (y) = 0 has degree
more than 1. Since f is irreducible and has degree more than 1, f
does not have a root in K, as we wanted to show. 
Exercise A.6. Show that the complex numbers are algebraically closed
(you may assume that every polynomial over the complex numbers
has a root).
Definition A.7. A field extension K → K is an algebraic closure if
(1) K → K is algebraic and
(2) K is algebraically closed.
Exercise A.8. Let K → L be an algebraic extension and let L denote
an algebraic closure of L. Show that L is also an algebraic closure of
K.
Theorem A.9 (Existence of algebraic closures). Let K be a field.
(1) K has an algebraic closure.
(2) Any two algebraic closures of K are isomorphic as field extensions
0
(meaning that for two algebraic closures K, K , with K as a subfield
0
via the maps φ : K → K, φ0 : K → K , there is an isomorphism
0
f : K → K so that f ◦ φ = φ0 ).
We first prove the existence of an algebraic closure Theorem A.9(1),
and then show it is unique up to (non-unique) isomorphism. The
key to proving the existence of an algebraic closure will be Zorn’s
lemma, which we now recall:
Lemma A.10. Suppose I is a partially ordered set. Suppose any totally
ordered subset I 0 ⊂ I has a maximum element, i.e., there is some i ∈ I with
i ≥ j for all j ∈ I 0 . Then I contains a maximal element, i.e., there is some
i ∈ I so that for any j ∈ I, j 6> i.
Remark A.11. Zorn’s lemma is not a lemma in the conventional sense
because it is equivalent to the axiom of choice. Therefore, we will not
prove it, but rather take it as an axiom.
We next aim to prove existence of algebraic closures. Logically, if
you’d like, you can skip directly to Exercise A.13. However, it may
NOTES ON FINITE FIELDS 27

help your understanding of that exercise if you do the prior exercises


first.
Exercise A.12. We now prove some basic properties about cardinal-
ities of field extensions.
(1) Show that if L is an algebraic extension of a finite field K, then
| L| ≤ |Z|. Here |S| denotes the set-theoretic cardinality of a
set S.
(2) Show that if L is an algebraic extension of an infinite field K,
then | L| = |K |. Hint: Show that K has the same cardinality
as K [ x ] and defined a map of sets L → K [ x ] by sending an
element to its minimal polynomial. Show that there are only
finitely many elements with a given minimal polynomial and
deduce |K | = | L|.
(3) Conclude that for any infinite field K, if T is a set with | T | >
|K | then for any algebraic extension L of K, we have | T | > | L|.
(4) Conclude that for any field K if T is an infinite set with | T | >
|K |, then | T | ≥ | L| for any algebraic extension L of K. (By the
above, the only interesting case is the case that K is finite.)
Exercise A.13 (Difficult exercise). Use Zorn’s lemma to show an al-
gebraic closure of a field K exists as follows: Let T be an infinite set
with | T | > |K | and let S be a set with |S| > | T |.
(1) Consider the partially ordered set
R := {( L, φ) : L is an algebraic extension of K and φ : L ,→ S is a subset }
Check that one can define a partial ordering on R by declaring
( L1 , φ1 ) ≤ ( L2 , φ2 ) if i : L1 → L2 is an algebraic extension, and
φ2 ◦ i = φ1 .
(2) Use Zorn’s lemma, Lemma A.10, to show that R has a maxi-
mal element, call it ( M, φ).
(3) Show that M is algebraically closed by showing that if i :
M → N is any algebraic extension then there is a map ψ :
N → S with ψ ◦ i ( x ) = φ( x ). Hint: Use that | N | ≤ | M | ≤
| T | < S and |S − M| = |S| > | N − M|.
Exercise A.14. Assume K is an infinite field. Using Exercise A.12,
solve a slightly simplified version of Exercise A.13 with the modifi-
cation that S is any set so that |S| > |K | (so that there is no intermedi-
ate set T in the picture). Therefore, the addition of T is only needed
to deal with finite fields. Why did we need to introduce this set T in
Exercise A.13 to deal with the case that K is finite?
28 AARON LANDESMAN

Exercise A.15. Suppose we have an algebraic extension K ⊂ L and


K ⊂ K with K algebraically closed. Show that there is a map of
extensions L → K in the following steps:
(1) Consider the partially ordered set I of pairs ( M, φ) with K ⊂
M ⊂ L and φ : M → K a map of fields. Check that the relation
( M1 , φ1 ) ≤ ( M2 , φ2 )
if M1 ⊂ M2 and φ2 | M1 = φ1 defines a partial ordering on such
pairs ( M, φ).
(2) Show that any totally ordered subset I 0 ⊂ I corresponding to
a collection {( Mi , φi )}i∈ I 0 has a maximum element given by
taking (∪i Mi , ∪i φi ), with ∪i φi interpreted suitably.
(3) Using Zorn’s lemma obtain a maximal element ( M, φ) of I.
(4) Verify that the maximum element ( M, φ) has M = L and con-
clude there is a map L → K Hint: Suppose L 6= M. Then
there is some x ∈ L − M. Show that x satisfies some minimal
polynomial over L. Deduce there is a map M( x ) → K restrict-
ing to the given map φ : M → K, and hence ( M, φ) was not
maximal.
Exercise A.16. Prove Theorem A.9(2) using Exercise A.15 as follows:
(1) Show that for any two algebraic closures K1 , K2 of the same
field K there is an injective map between φ : K1 → K2 .
(2) Show that the injective map φ is an algebraic extension.
(3) Conclude that the map produced K1 → K2 is an isomorphism
from the definition of algebraic closure.
NOTES ON FINITE FIELDS 29

A PPENDIX B. B ASICS OF RINGS


In this appendix, we review some basic definitions relating to rings.
Recall our definition of a (commutative) ring (with unit), Defini-
tion 2.1. We repeat this now for your convenience.
Definition B.1. A commutative ring with unit is a set R together
with two operations (+, ·) satisfying the following properties:
(1) Associativity: a + (b + c) = ( a + b) + c, a · (b · c) = ( a · b) · c
(2) Commutativity: a + b = b + a, a · b = b · a
(3) Additive identity: there exists 0 ∈ R so that a + 0 = a
(4) Multiplicative identity: there exists 1 6= 0 ∈ R so that 1 · a = a
(5) Additive inverses: For every a ∈ R, there is a additive in-
verse, denoted − a satisfying a + (− a) = 0
(6) Distributivity of multiplication over addition: a · (b + c) =
( a · b) + ( a · c) .
For us, all rings will be commutative rings with unit, and so we
will simply refer to them as rings from now on. We now recall some
elementary properties of rings. Many of these follow directly from
the analogous properties for groups.
Exercise B.2. Verify, directly from the definition that every ring has
a unique 0 and 1. Show that for any a ∈ R, a has a unique additive
inverse, and so the name − a is justified.
Definition B.3. A map of rings f : R → S is a map of sets such that
f (1R ) = 1S , f (0R ) = 0S , f ( a + R b) = f ( a) +S f (b) and f ( a · R b) =
f ( a) ·S f (b). where the subscripts denote the identity, multiplication,
and addition in the corresponding ring.
Definition B.4. A ring map f : R → S is injective if f ( a) = f (b) =⇒
a = b. It is surjective if for every s ∈ S there is some r ∈ R with
f (r ) = s. It is bijective (also known as an isomorphism) if it is both
injective and surjective. If f : R → S is bijective, we write R ' S.
Exercise B.5. Show that a ring map f : R → S is injective if and only
if f −1 (0S ) = 0R .
Exercise B.6. Show that a ring map f : R → S is bijective if and only
if there is a ring map f −1 : S → R so that f −1 ◦ f = idR , f ◦ f −1 = idS .
Hint: Show that a map is bijective if and only if there is a unique
element of R mapping to any given element of S. Use this to define
an inverse map.
30 AARON LANDESMAN

B.1. Quotients. The following will not be needed in this course. In-
deed, we will construct particular quotient rings in this course, but
these quotients will all be of the form K [ x ]/( f ) for K a field, a situ-
ation which is much more concrete than the general case developed
below. We encourage you to skip the following, but we include it for
completeness.
Definition B.7. An ideal I of R is a subset I ⊂ R so that
(1) 0R ∈ I.
(2) For any r ∈ I we also have −s ∈ I.
(3) If a, b ∈ I then a + b ∈ I.
(4) If r ∈ R and a ∈ I then a · R b ∈ I.
Definition B.8. Let I ⊂ R be a subring. Construct the quotient R/I
as the set of all elements a ∈ R modulo the equivalence relation a ∼ b
if there is some c ∈ I with a = b + c. The equivalence class of a is
called the coset of a and the coset is notated a + I.
Exercise B.9. Verify that the relation ∼ as defined in Definition B.8 is
indeed an equivalence relation.
Exercise B.10. Show that if I ⊂ R is an ideal then R/I is again a
ring. (Under our definition, this includes verifying that the quotient
is commutative and has a unit).
R EFERENCES

You might also like