Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Finite fields} | |
In this short chapter, we classify all fields with finitely many elements | |
and compute the Galois groups. | |
Nothing in here is very hard, and so most of the proofs are just sketches; | |
if you like, you should check the details yourself. | |
The whole point of this chapter is to prove: | |
\begin{itemize} | |
\ii A finite field $F$ must have order $p^n$, with $p$ prime and $n$ an integer. | |
\ii In this case, $F$ has characteristic $p$. | |
\ii All such fields are isomorphic, | |
so it's customary to use the notation $\FF_{p^n}$ | |
for ``the'' finite field of order $p^n$ if we only care up to isomorphism. | |
\ii The extension $F/\FF_p$ is Galois, and $\Gal(F/\FF_p)$ is a cyclic group of order $n$. | |
The generator is the automorphism \[ \sigma : F \to F \quad\text{by}\quad x \mapsto x^p. \] | |
\end{itemize} | |
If you're in a hurry you can just remember these results and skip to the next chapter. | |
\section{Example of a finite field} | |
Before diving in, we give some examples. | |
Recall that the \emph{characteristic} of a field $F$ | |
is the smallest positive integer $p$ such that | |
\[ \underbrace{1_F + \dots + 1_F}_{\text{$p$ times}} = 0 \] | |
or $0$ if no such integer $p$ exists. | |
\begin{example}[Base field] | |
Let $\FF_p$ denote the field of integers modulo $p$. | |
This is a field with $p$ elements, with characteristic $p$. | |
\end{example} | |
\begin{example}[The finite field of nine elements] | |
Let | |
\[ F \cong \FF_3[X]/(X^2+1) \cong \ZZ[i] / (3). \] | |
We can think of its elements as \[ \left\{ a + bi \mid 0 \le a,b \le 2 \right\}. \] | |
Since $(3)$ is prime in $\ZZ[i]$, the ring of integers of $\QQ(i)$, | |
we see $F$ is a field with $3^2 = 9$ elements inside it. | |
Note that, although this field has $9$ elements, every element $x$ has the property that | |
\[ 3x = \underbrace{x + \dots + x}_{\text{$3$ times}} = 0. \] | |
In particular, $F$ has characteristic $3$. | |
\end{example} | |
\section{Finite fields have prime power order} | |
\begin{lemma} | |
If the characteristic of a field $F$ isn't zero, | |
it must be a prime number. | |
\end{lemma} | |
\begin{proof} | |
Assume not, so $n = ab$ for $a,b < n$. | |
Then let | |
\[ A = \underbrace{1_F + \dots + 1_F}_{\text{$a$ times}} \neq 0 \] | |
and | |
\[ B = \underbrace{1_F + \dots + 1_F}_{\text{$b$ times}} \neq 0. \] | |
Then $AB = 0$, contradicting the fact that $F$ is a field. | |
\end{proof} | |
We like fields of characteristic zero, but unfortunately for finite fields | |
we are doomed to have nonzero characteristic. | |
\begin{lemma} | |
[Finite fields have prime power orders] | |
Let $F$ be a finite field. | |
Then | |
\begin{enumerate}[(a)] | |
\ii Its characteristic is nonzero, and hence some prime $p$. | |
\ii The field $F$ is a finite extension of $\FF_p$, | |
and in particular it is an $\FF_p$-vector space. | |
\ii We have $\left\lvert F \right\rvert = p^n$ for some prime $p$, integer $n$. | |
\end{enumerate} | |
\end{lemma} | |
\begin{proof} | |
Very briefly, since this is easy: | |
\begin{enumerate}[(a)] | |
\ii Apply Lagrange's theorem (or pigeonhole principle!) | |
to $(F, +)$ to get the characteristic isn't zero. | |
\ii The additive subgroup of $(F,+)$ generated | |
by $1_F$ is an isomorphic copy of $\FF_p$. | |
\ii Since it's a field extension, | |
$F$ is a finite-dimensional vector space over $\FF_p$, | |
with some basis $e_1, \dots, e_n$. | |
It follows that there are $p^n$ elements of $F$. \qedhere | |
\end{enumerate} | |
\end{proof} | |
\begin{remark} | |
An amusing alternate proof of (c) by contradiction: | |
if a prime $q \neq p$ divides $\left\lvert F \right\rvert$, then | |
by Cauchy's theorem (\Cref{thm:cauchy_group}) on $(F, +)$ | |
there's a (nonzero) element $x$ of order $q$. | |
Evidently \[ x \cdot ( \underbrace{1_F + \dots + 1_F}_{\text{$q$ times}} ) = 0 \] | |
then, but $x \neq 0$, and hence the characteristic of $F$ also divides $q$, | |
which is impossible. | |
\end{remark} | |
An important point in the above proof is that | |
\begin{lemma}[Finite fields are field extensions of $\FF_p$] | |
If $\left\lvert F \right\rvert = p^n$ is a finite field, | |
then there is an isomorphic copy of $\FF_p$ sitting inside $F$. | |
Thus $F$ is a field extension of $\FF_p$. | |
\end{lemma} | |
We want to refer a lot to this copy of $\FF_p$, so in what follows: | |
\begin{abuse} | |
Every integer $n$ can be identified as an element of $F$, namely | |
\[ n \defeq \underbrace{1_F + \dots + 1_F}_{\text{$n$ times}}. \] | |
Note that (as expected) this depends only on $n \pmod p$. | |
\end{abuse} | |
This notation makes it easier to think about statements like the following. | |
\begin{theorem} | |
[Freshman's dream] | |
For any $a,b \in F$ we have | |
\[ (a+b)^p = a^p + b^p. \] | |
\end{theorem} | |
\begin{proof} | |
Use the Binomial theorem, and the fact that $\binom pi$ is divisible by $p$ for $0 < i < p$. | |
\end{proof} | |
\begin{exercise} | |
Convince yourself that this proof works. | |
\end{exercise} | |
\section{All finite fields are isomorphic} | |
We next proceed to prove ``Fermat's little theorem'': | |
\begin{theorem} | |
[Fermat's little theorem in finite fields] | |
Let $F$ be a finite field of order $p^n$. | |
Then every element $x \in F$ satisfies | |
\[ x^{p^n} - x = 0. \] | |
\end{theorem} | |
\begin{proof} | |
If $x = 0$ it's true; otherwise, use Lagrange's theorem | |
on the abelian group $(F, \times)$ to get $x^{p^n-1} = 1_F$. | |
\end{proof} | |
We can now prove the following result, | |
which is the ``main surprise'' about finite fields: | |
that there is a unique one up to isomorphism for each size. | |
\begin{theorem}[Complete classification of finite fields] | |
A field $F$ is a finite field with $p^n$ elements if and only if | |
it is a splitting field of $x^{p^n}-x$ over $\FF_p$. | |
\end{theorem} | |
\begin{proof} | |
By ``Fermat's little theorem'', all the elements of $F$ satisfy this polynomial. | |
So we just have to show that the roots of this polynomial are distinct | |
(i.e.\ that it is separable). | |
To do this, we use the derivative trick again: the derivative of this polynomial is | |
\[ p^n \cdot x^{p^n-1} - 1 = -1 \] | |
which has no roots at all, so the polynomial cannot have any double roots. \qedhere | |
\end{proof} | |
\begin{definition} | |
For this reason, it's customary to denote \emph{the} | |
field with $p^n$ elements by $\FF_{p^n}$. | |
\end{definition} | |
Note that the polynomial $x^{p^n}-x \pmod p$ is far from irreducible, but | |
the computation above shows that it's separable. | |
\begin{example}[The finite field of order nine again] | |
The polynomial $x^9-x$ is separable modulo $3$ and has factorization | |
\[ x(x+1)(x+2)(x^2+1)(x^2+x+2)(x^2+2x+2) \pmod 3. \] | |
So if $F$ has order $9$, then we intuitively expect it to be the field | |
generated by adjoining all the roots: $0$, $1$, $2$, as well as | |
$\pm i$, $1 \pm i$, $2 \pm i$. | |
Indeed, that's the example we had at the beginning of this chapter. | |
(Here $i$ denotes \emph{an} element of $\FF_9$ satisfying $i^2=-1$. | |
The notation is deliberately similar to the usual imaginary unit.) | |
\end{example} | |
\section{The Galois theory of finite fields} | |
Retain the notation $\FF_{p^n}$ now (instead of $F$ like before). | |
By the above theorem, it's the splitting field of a separable polynomial, | |
hence we know that $\FF_{p^n} /\FF_p$ is a Galois extension. | |
We would like to find the Galois group. | |
In fact, we are very lucky: it is cyclic. | |
First, we exhibit one such element $\sigma_p \in \Gal(\FF_{p^n} /\FF_p)$: | |
\begin{theorem}[The $p$th power automorphism] | |
The map $\sigma_p : \FF_{p^n} \to \FF_{p^n}$ defined by | |
\[ \sigma_p(x) = x^p \] | |
is an automorphism, and moreover fixes $\FF_p$. | |
\end{theorem} | |
\begin{proof} | |
It's a homomorphism since it fixes $1$, | |
respects multiplication, | |
and respects addition. | |
\begin{ques} | |
Why does it respect addition? | |
\end{ques} | |
Next, we claim that it is injective. To see this, note that | |
\[ x^p = y^p | |
\iff x^p - y^p = 0 | |
\iff (x-y)^p = 0 | |
\iff x=y. | |
\] | |
Here we have again used the Freshman's Dream. | |
Since $\FF_{p^n}$ is finite, this injective map is automatically bijective. | |
The fact that it fixes $\FF_p$ is Fermat's little theorem. | |
\end{proof} | |
Now we're done: | |
\begin{theorem} | |
[Galois group of the extension $\FF_{p^n}/\FF_p$] | |
We have $\Gal(\FF_{p^n}/\FF_p) \cong \Zc n$ with generator $\sigma_p$. | |
\end{theorem} | |
\begin{proof} | |
Since $[\FF_{p^n}:\FF_p] = n$, the Galois group $G$ has order $n$. | |
So we just need to show $\sigma_p \in G$ has order $n$. | |
Note that $\sigma_p$ applied $k$ times gives $x \mapsto x^{p^k}$. | |
Hence, $\sigma_p$ applied $n$ times is the identity, | |
as all elements of $\FF_{p^n}$ satisfy $x^{p^n}=x$. | |
But if $k < n$, then $\sigma_p$ applied $k$ times | |
cannot be the identity or $x^{p^k}-x$ would have too many roots. | |
\end{proof} | |
We can see an example of this again with the finite field of order $9$. | |
\begin{example} | |
[Galois group of finite field of order $9$] | |
Let $\FF_9$ be the finite field of order $9$, | |
and represent it concretely by $\FF_9 = \ZZ[i]/(3)$. | |
Let $\sigma_3 : \FF_9 \to \FF_9$ be $x \mapsto x^3$. | |
We can witness the fate of all nine elements: | |
\begin{center} | |
\begin{tikzcd} | |
0 & 1 & 2 | |
& i \ar[d, leftrightarrow, "\sigma"] | |
& 1+i \ar[d, leftrightarrow, "\sigma"] | |
& 2+i \ar[d, leftrightarrow, "\sigma"] \\ | |
&&& -i & 1-i & 2-i | |
\end{tikzcd} | |
\end{center} | |
(As claimed, $0$, $1$, $2$ are the fixed points, | |
so I haven't drawn arrows for them.) | |
As predicted, the Galois group has order two: | |
\[ \Gal(\FF_9/\FF_3) = \left\{ \id, \sigma_3 \right\} \cong \Zc 2. \] | |
\end{example} | |
This concludes the proof of all results stated at the beginning of this chapter. | |
\section\problemhead | |
\begin{dproblem}[HMMT 2017] | |
\gim | |
What is the period of the Fibonacci sequence modulo $127$? | |
\begin{hint} | |
The Fibonacci sequence is given by | |
$F_n = \frac{\alpha^n - \beta^n}{\alpha - \beta}$ | |
where $\alpha = \frac{1 + \sqrt5}{2}$ and $\beta = \frac{1 - \sqrt5}{2}$ | |
are the two roots of $P(X) \overset{\text{def}}{=} X^2-X-1$. | |
Show the polynomial $P(X)$ is irreducible modulo $127$; | |
then work in the splitting field of $P$, namely $\FF_{p^2}$. | |
Show that $\FF_p=-1$, $\FF_{p+1}=0$, $\FF_{2p+1}=1$, $\FF_{2p+2}=0$. | |
(Look at the action of $\Gal(\FF_{p^2}/\FF_p)$ | |
on the roots of $P$.) | |
\end{hint} | |
\begin{sol} | |
Recall that the Fibonacci sequence is given by | |
\[ F_n = \frac{\alpha^n - \beta^n}{\alpha - \beta} \] | |
where $\alpha = \frac{1 + \sqrt5}{2}$ and $\beta = \frac{1 - \sqrt5}{2}$ | |
are the two roots of $P(X) \defeq X^2-X-1$. | |
Let $p = 127$ and work modulo $p$. | |
As \[ \left( \frac5p \right) = \left( \frac p5 \right) = \left( \frac 25 | |
\right) = -1 \] | |
we see $5$ is not a quadratic residue mod $127$. | |
Thus the polynomial $P(X)$, | |
viewed as a polynomial in $\FF_p[X]$, is irreducible | |
(intuitively, $\alpha$ and $\beta$ are not elements of $\FF_p$). | |
Accordingly we will work in the finite field $\FF_{p^2}$, | |
which is the $\FF_p$-splitting field of $P(X)$. | |
In other words we interpret $\alpha$ and $\beta$ as elements | |
of $\FF_{p^2}$ which do not lie in $\FF_p$. | |
Let $\sigma \colon \FF_{p^2} \to \FF_{p^2}$ by $t \mapsto t^p$ | |
be the nontrivial element of | |
$\opname{Gal}\left( \FF_{p^2} / \FF_p \right)$; | |
in other words, $\sigma$ is the non-identity automorphism of $\FF_{p^2}$. | |
Since the fixed points of $\sigma$ are the elements of $\FF_p$, | |
this means $\sigma$ does not fix either root of $P$; thus we must have | |
\begin{align*} | |
\alpha^p = \sigma(\alpha) &= \beta \\ | |
\beta^p = \sigma(\beta) &= \alpha. | |
\end{align*} | |
Now, compute | |
\begin{align*} | |
F_{p} &= \frac{\alpha^{p} - \beta^{p}}{\alpha-\beta} = | |
\frac{\beta-\alpha}{\alpha-\beta} = -1. \\ | |
F_{p+1} &= \frac{\alpha^{p+1} - \beta^{p+1}}{\alpha-\beta} = | |
\frac{\alpha\beta-\beta\alpha}{\alpha-\beta} = 0. \\ | |
F_{2p+1} &= \frac{\alpha^{2p+1} - \beta^{2p+1}}{\alpha-\beta} | |
= \frac{\beta^2\alpha-\alpha^2\beta}{\alpha-\beta} | |
= - \alpha \beta = 1. \\ | |
F_{2p+2} &= \frac{\alpha^{2p+2} - \beta^{2p+2}}{\alpha-\beta} | |
= \frac{\beta^2\alpha^2-\alpha^2\beta^2}{\alpha-\beta} = 0. | |
\end{align*} | |
Consequently, the period must divide $2p+2$ but not $p+1$. | |
We now use for the first time the exact numerical value $p=127$ | |
to see the period divides $2p+2 = 256 = 2^8$, | |
but not $p+1 = 128 = 2^7$. | |
(Previously we only used the fact that $(5/p)=-1$.) | |
Thus the period must be exactly $256$. | |
\end{sol} | |
\end{dproblem} | |