Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Some applications} | |
With all this setup, we now take the time to develop some nice | |
results which are of independent interest. | |
\section{Frobenius divisibility} | |
\begin{theorem} | |
[Frobenius divisibility] | |
Let $V$ be a complex irrep of a finite group $G$. | |
Then $\dim V$ divides $|G|$. | |
\end{theorem} | |
The proof of this will require algebraic integers | |
(developed in the algebraic number theory chapter). | |
Recall that an \emph{algebraic integer} is a complex number | |
which is the root of a monic polynomial with integer coefficients, | |
and that these algebraic integers form a ring $\ol\ZZ$ | |
under addition and multiplication, and that $\ol\ZZ \cap \QQ = \ZZ$. | |
First, we prove: | |
\newcommand{\tempuuujbhtkx}{$\ZZ[G]$} | |
\begin{lemma}[Elements of \tempuuujbhtkx\ are integral] | |
\label{lem:group_ring_integral} | |
Let $\alpha \in \ZZ[G]$. | |
Then there exists a monic polynomial $P$ with integer coefficients | |
such that $P(\alpha) = 0$. | |
\end{lemma} | |
\begin{proof} | |
Let $A_k$ be the $\ZZ$-span of $1, \alpha^1, \dots, \alpha^k$. | |
Since $\ZZ[G]$ is Noetherian, | |
the inclusions $A_0 \subseteq A_1 \subseteq A_2 \subseteq \dots$ | |
cannot all be strict, hence $A_k = A_{k+1}$ for some $k$, | |
which means $\alpha^{k+1}$ can be expressed in terms of | |
lower powers of $\alpha$. | |
\end{proof} | |
\begin{proof} | |
[Proof of Frobenius divisibility] | |
Let $C_1$, \dots, $C_m$ denote the conjugacy classes of $G$. | |
Then consider the rational number \[ \frac{|G|}{\dim V}; \] | |
we will show it is an algebraic integer, which will prove the theorem. | |
Observe that we can rewrite it as | |
\[ | |
\frac{|G|}{\dim V} | |
= \frac{|G| \left< \chi_V, \chi_V \right>}{\dim V} | |
= \sum_{g \in G} \frac{\chi_V(g) \ol{\chi_V(g)}}{\dim V}. | |
\] | |
We split the sum over conjugacy classes, so | |
\[ | |
\frac{|G|}{\dim V} | |
= | |
\sum_{i=1}^m \ol{\chi_V(C_i)} \cdot \frac{|C_i| \chi_V(C_i)}{\dim V}. | |
\] | |
We claim that for every $i$, | |
\[ \frac{|C_i| \chi_V(C_i)}{\dim V} | |
= \frac{1}{\dim V} \Tr T_i \] | |
is an algebraic integer, | |
where \[ T_i \defeq \rho\left(\sum_{h \in C_i} h\right). \] | |
To see this, note that $T_i$ commutes with elements of $G$, | |
and hence is an intertwining operator $T_i : V \to V$. | |
Thus by Schur's lemma, $T_i = \lambda_i \cdot \id_V$ | |
and $\Tr T = \lambda_i \dim V$. | |
By \Cref{lem:group_ring_integral}, $\lambda_i \in \ol\ZZ$, as desired. | |
Now we are done, since $\ol{\chi_V(C_i)} \in \ol\ZZ$ too | |
(it is the sum of conjugates of roots of unity), | |
so $\frac{|G|}{\dim V}$ is the sum of products of algebraic integers, | |
hence itself an algebraic integer. | |
\end{proof} | |
\section{Burnside's theorem} | |
We now prove a group-theoretic result. | |
This is the famous poster child for representation theory | |
(in the same way that RSA is the poster child of number theory) | |
because the result is purely group theoretic. | |
Recall that a group is \vocab{simple} if it has no normal subgroups. | |
In fact, we will prove: | |
\begin{theorem}[Burnside] | |
Let $G$ be a nonabelian group of order $p^a q^b$ (where $p,q$ are distinct primes and $a,b \ge 0$). | |
Then $G$ is not simple. | |
\end{theorem} | |
In what follows $p$ and $q$ will always denote prime numbers. | |
\begin{lemma}[On $\gcd(|C|, \dim V) = 1$] | |
\label{lem:burnside_ant_lemma} | |
Let $V = (V, \rho)$ be an complex irrep of $G$. | |
Assume $C$ is a conjugacy class of $G$ with $\gcd(|C|, \dim V) = 1$. | |
Then for any $g \in C$, either | |
\begin{itemize} | |
\ii $\rho(g)$ is multiplication by a scalar, or | |
\ii $\chi_V(g) = \Tr \rho(g) = 0$. | |
\end{itemize} | |
\end{lemma} | |
\begin{proof} | |
If $\eps_i$ are the $n$ eigenvalues of $\rho(g)$ (which are roots of unity), | |
then from the proof of Frobenius divisibility we know | |
$\frac{|C|}{n} \chi_V(g) \in \ol\ZZ$, | |
thus from $\gcd(|C|, n) = 1$ we get | |
\[ \frac1n \chi_V(g) = \frac1n(\eps_1 + \dots + \eps_n) \in \ol\ZZ. \] | |
So this follows readily from a fact from algebraic number theory, | |
namely \Cref{prob:rep_lemma}: | |
either $\eps_1 = \dots = \eps_n$ (first case) or | |
$\eps_1 + \dots + \eps_n = 0$ (second case). | |
\end{proof} | |
\begin{lemma} | |
[Simple groups don't have prime power conjugacy classes] | |
Let $G$ be a finite simple group. | |
Then $G$ cannot have a conjugacy class of order $p^k$ (where $k > 0$). | |
\end{lemma} | |
\begin{proof} | |
By contradiction. | |
Assume $C$ is such a conjugacy class, and fix any $g \in C$. | |
By the second orthogonality formula (\Cref{prob:second_orthog}) | |
applied $g$ and $1_G$ (which are not conjugate since $g \neq 1_G$) we have | |
\[ \sum_{i=1}^r \dim V_i \chi_{V_i}(g) = 0 \] | |
where $V_i$ are as usual all irreps of $G$. | |
\begin{exercise} | |
Show that there exists a nontrivial irrep $V$ | |
such that $p \nmid \dim V$ and $\chi_V(g) \neq 0$. | |
(Proceed by contradiction to show that $-\frac1p \in \ol\ZZ$ if not.) | |
\end{exercise} | |
Let $V = (V, \rho)$ be the irrep mentioned. | |
By the previous lemma, we now know that $\rho(g)$ acts as a scalar in $V$. | |
Now consider the subgroup | |
\[ H = \left< ab\inv \mid a,b \in C \right> \subseteq G. \] | |
We claim this is a nontrivial normal subgroup of $G$. | |
It is easy to check $H$ is normal, | |
and since $|C| > 1$ we have that $H$ is nontrivial. | |
As represented by $V$ each element of $H$ acts trivially in $G$, | |
so since $V$ is nontrivial and irreducible, $H \neq G$. | |
This contradicts the assumption that $G$ was simple. | |
\end{proof} | |
With this lemma, Burnside's theorem follows by partitioning | |
the $|G|$ elements of our group into conjugacy classes. | |
Assume for contradiction $G$ is simple. | |
Each conjugacy class must have order either $1$ (of which there are $|Z(G)|$ by \Cref{prob:class_eq}) | |
or divisible by $pq$ (by the previous lemma), but on the other hand the sum equals $|G| = p^aq^b$. | |
Consequently, we must have $|Z(G)| > 1$. | |
But $G$ is not abelian, hence $Z(G) \neq G$, | |
thus the center $Z(G)$ is a nontrivial normal subgroup, | |
contradicting the assumption that $G$ was simple. | |
\section{Frobenius determinant} | |
We finish with the following result, | |
the problem that started the branch of representation theory. | |
Given a finite group $G$, | |
we create $n$ variables $\{x_g\}_{g \in G}$, | |
and an $n \times n$ matrix $M_G$ whose $(g,h)$th entry is $x_{gh}$. | |
\begin{example}[Frobenius determinants] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii If $G = \Zc 2 = \left< T \mid T^2 = 1\right>$ | |
then the matrix would be \[ M_G = | |
\begin{bmatrix} x_{\id} & x_T \\ x_T & x_{\id} \end{bmatrix}. \] | |
Then $\det M_G = (x_{\id}-x_T)(x_{\id}+x_T)$. | |
\ii If $G = S_3$, a long computation gives | |
the irreducible factorization of $\det M_G$ is | |
\[ | |
\left( \sum_{\sigma \in S_3} x_{\sigma} \right) | |
\left( \sum_{\sigma \in S_3} \sign(\sigma)x_\sigma \right) | |
\Big( F\left(x_\id, x_{(123)}, x_{(321)}\right) | |
- F\left(x_{(12)}, x_{(23)}, x_{(31)}\right) \Big)^2 \] | |
where $F(a,b,c) = a^2+b^2+c^2-ab-bc-ca$; | |
the latter factor is irreducible. | |
\end{enumerate} | |
\end{example} | |
\begin{theorem} | |
[Frobenius determinant] | |
The polynomial $\det M_G$ (in $|G|$ variables) factors | |
into a product of irreducible polynomials such that | |
\begin{enumerate}[(i)] | |
\ii The number of polynomials equals the number | |
of conjugacy classes of $G$, and | |
\ii The multiplicity of each polynomial | |
equals its degree. | |
\end{enumerate} | |
\end{theorem} | |
You may already be able to guess how the ``sum of squares'' result | |
is related! (Indeed, look at $\deg\det M_G$.) | |
Legend has it that Dedekind observed this behavior first in 1896. | |
He didn't know how to prove it in general, | |
so he sent it in a letter to Frobenius, | |
who created representation theory to solve the problem. | |
With all the tools we've built, it is now fairly straightforward | |
to prove the result. | |
\begin{proof} | |
Let $V = (V, \rho) = \Reg(\CC[G])$ and let $V_1$, \dots, $V_r$ | |
be the irreps of $G$. | |
Let's consider the map $T \colon \CC[G] \to \CC[G]$ | |
which has matrix $M_G$ in the usual basis of $\CC[G]$, namely | |
\[ T : T(\{x_g\}_{g \in G}) = \sum_{g \in G} x_g \rho(g) \in \Mat(V). \] | |
Thus we want to examine $\det T$. | |
But we know that $V = \bigoplus_{i=1}^r V_i^{\oplus \dim V_i}$ | |
as before, and so breaking down $T$ over its subspaces we know | |
\[ | |
\det T | |
= \prod_{i=1}^r \left( \det (T \restrict{V_i}) \right)^{\dim V_i}. | |
\] | |
So we only have to show two things: | |
the polynomials $\det T_{V_i}$ are irreducible, | |
and they are pairwise different for different $i$. | |
Let $V_i = (V_i, \rho)$, and pick $k = \dim V_i$. | |
\begin{itemize} | |
\ii \emph{Irreducible}: | |
By the density theorem, for any $M \in \Mat(V_i)$ there exists | |
a \emph{particular} choice of complex numbers $x_g \in G$ such that | |
\[ | |
M = \sum_{g \in G} x_g | |
\cdot \rho_i(g) | |
= (T \restrict{V_i})(\{x_g\}). | |
\] | |
View $\rho_i(g)$ as a $k \times k$ matrix with complex coefficients. | |
Thus the ``generic'' $(T \restrict{V_i})(\{x_g\})$, viewed as a matrix with | |
polynomial entries, must have linearly independent entries | |
(or there would be some matrix in $\Mat(V_i)$ that we can't achieve). | |
Then, the assertion follows (by a linear variable change) | |
from the simple fact that the polynomial | |
$\det (y_{ij})_{1 \le i, j \le m}$ in $m^2$ variables | |
is always irreducible. | |
\ii \emph{Pairwise distinct}: | |
We show that from $\det T|_{V_i}(\{x_g\})$ we can read | |
off the character $\chi_{V_i}$, which proves the claim. | |
In fact | |
\begin{exercise} | |
Pick \emph{any} basis for $V_i$. | |
If $\dim V_i = k$, and $1_G \neq g \in G$, then | |
\[ | |
\chi_{V_i} (g) | |
\text{ is the coefficient of } x_g x_{1_G}^{k-1}. | |
\] | |
\end{exercise} | |
Thus, we are done. \qedhere | |
\end{itemize} | |
\end{proof} | |