% vim: tw=50 % 28/01/2023 12AM \newpage \section{$p$-groups and $p$-subgroups} \begin{definition*} Let $p$ be a prime. A finite group $G$ is a $p$-group if $|G| = p^n$, $n \ge 1$. \end{definition*} \begin{theorem} If $G$ is a $p$-group, then $Z(G) \neq 1$. \end{theorem} \begin{proof} For $g \in G$, we have $|\ccl_G(g)||C_G(g)| = |G| = p^n$, so each conjugacy class has size a power of $p$. Since $G$ is a union of conjugacy classes: \[ |G| = \#(\text{conjugacy classes of size $1$}) \pmod{p} \] Note that \begin{align*} g \in Z(G) &\iff gxg^{-1} = x ~\forall x \in G \\ &\iff x^{-1}gx = g ~\forall x \in G \\ &\iff \ccl_G(g) = \{g\} \end{align*} So $|Z(G)| = \#(\text{conjugacy classes of size 1})$. So $0 \equiv |Z(G)| \pmod{p}$. We know $|Z(G)| \ge 1$ since $e \in Z(G)$, so therefore $|Z(G)| \ge p > 1$. \end{proof} \begin{corollary} The only simple $p$-group is $C_p$. \end{corollary} \begin{proof} Let $G$ be a simple $p$-group. Since $Z(G) \normalsub G$ we have $Z(G) = 1$ or $G$. But by the previous theorem, $Z(G) \neq 1$, so $Z(G) = G$, so $G$ is abelian. Conclude by Lemma 1.3. \end{proof} \begin{flashcard}[p-group-has-subgroups-of-every-order] \begin{corollary*} Let $G$ be a $p$-group of order $p^n$. Then $G$ has \cloze{a subgroup of order $p^n$ for all $0 \le r \le n$. } \end{corollary*} \begin{proof} \cloze{ By Lemma 1.4, $G$ has a composition series \[ 1 = G_0 \normalsub G_1 \normalsub \cdots \normalsub G_{m - 1} \normalsub G_m = G ,\] with each $G_i / G_{i - 1}$ being simple, and also since $G$ is a $p$-group, $G_i / G_{i - 1}$ is a $p$-group, so $G_i / G_{i = 1} \cong C_p$ by Corollary 4.2. \myskip Thus $|G_i| = p^i$ for $0 \le i \le m$ and $m = n$. } \end{proof} \end{flashcard} \begin{lemma} For $G$ a group, if $G / Z(G)$ is cyclic, then $G$ is abelian (and so $G / Z(G)$ is trivial). \end{lemma} \begin{proof} Let $gZ(G)$ be a generator for $G / Z(G)$. Then each coset is of the form $g^r Z(G)$ for some $r \in \ZZ$. Thus $G = \{g^r z \colon r \in \ZZ, z \in G(Z)\}$. Then \begin{align*} (g^{r_1} z_1) \cdot (g^{r_2} z_2) &= g^{r_1 + r_2} z_1 z_2 \\ &= g^{r_1 + r_2} z_2 z_1 \\ &= (g^{r_2} z_2) \cdot (g^{r_1} z_1) \end{align*} So $G$ is abelian. \end{proof} \begin{corollary} If $|G| = p^2$, then $G$ is abelian. \end{corollary} \begin{proof} We consider the 3 possible cases for $|Z(G)|$ ($|Z(G)| ~\big|~ p^2$ by Lagrange's theorem) \begin{itemize} \item If $|Z(G)| = 1$, then this contradicts Theorem 4.1. \item If $|Z(G)| = p$, then $|G / Z(G)| = p$. Apply Lemma 4.1, contradiction. \item $|Z(G)| = p^2$, then $Z(G) = G$ so $G$ is abelian. \qedhere \end{itemize} \end{proof} \myskip See example sheet for case $|G| = p^3$. \subsection{Sylow Theorems} \begin{flashcard}[sylow-theorems] \begin{theorem*}[Sylow] Let $G$ be a finite group of order $p^a m$ where $p$ is a prime with $p \nmid m$. Then \begin{enumerate}[(i)] \item \cloze{The set $\Syl_p(G) = \{P \le G \colon |P| = p^a\}$ of Sylow $p$-subgroups is non-empty.} \item \cloze{All elements of $\Syl_p(G)$ are conjugate.} \item \cloze{$n_p \defeq |\Syl_p(G)|$ satisfies $n_p \equiv 1 \pmod{p}$ and $n_p ~\big|~ |G|$ (and hence $n_p \mid m$).} \end{enumerate} \end{theorem*} \end{flashcard} \begin{corollary} If $n_p = 1$, then the unique Sylow $p$-subgroup is normal. \end{corollary} \begin{proof} Let $g \in G$ and $P \in \Syl_p(G)$. Then $gPg^{-1} \in \Syl_p(G)$ and so $gPg^{-1} = P$. Thus $p \normalsub G$. \end{proof} \begin{example*} Let $|G| = 1000 = 2^3 \times 5^3$. Then $n_5 \equiv 1 \pmod{5}$ and $n_5 \mid 8$, so $n_5 = 1$. Thus the unique Sylow $5$-subgroup is normal, and hence $G$ is not simple. \end{example*} \begin{example*} $|G| = 132 = 2^3 \times 3 \times 11$. $n_{11} \equiv 1 \pmod{11}$ and $n_{11} \mid 12$, so $n_{11} = 1$ or $n_{11} = 12$. Suppose $G$ is simple. Then $n_{11} \neq 1$ (otherwise the Sylow 11 subgroup is normal) and hence $n_{11} = 12$. Now $n_3 \equiv 1 \pmod{3}$ and $n_3 \mid 44$. So $n_3 = 4, 22$ ($n_3 \neq 1$ if $G$ is simple). \myskip Suppose $n_3 = 4$. Then letting $G$ act on $\Syl_3(G)$ by conjugation gives a group homomorphism $\phi \colon G \to S_4$. Since $G$ is simple, we must have $\Ker(\phi) =1$ or $\Ker(\phi) = G$. But $\Ker(\phi) = G$ contradicts Sylow (ii). So $\Ker(\phi) = G$, so $G \injto S_4$. But this is not possible since $|G| > |S_4|$. \myskip Thus $n_3 = 22$ and $n_{11} = 12$. So $G$ has $22 \times (3 - 1) = 44$ elements of order 3 and $12 \times (11 - 1) = 120$ elements of order 11. But $44 + 120 > 132 = |G|$. \myskip Hence there does not exist a simple group of order $132$. \end{example*} \subsubsection*{Proof of Sylow Theorems} Let $|G| = p^a m$, $p$ prime, $p \nmid m$. \begin{enumerate}[(i)] \item \begin{flashcard}[proof-of-sylow-i] \prompt{Proof of Sylow(i)? \\} \cloze{ Let $\Omega$ be the set of all subsets of $G$ of size $p^a$. \[ |\Omega| = {p^a m \choose p^a} = \frac{p^a m}{p^a} \cdot \frac{p^a m - 1}{p^a - 1} \cdots \frac{p^a m - p^a + 1}{1} \] For $0 \le k < p^a$, the numbers $p^a m - k$ and $p^a - k$ are divisible by the same power of $p$. Therefore $|\Omega|$ is coprime to $p$ (\dag). \myskip Let $G$ act on $\Omega$ by left multiplication, i.e. for $g \in G$ and $X \in \Omega$ \[ g * X = \{gx \colon x \in X\} \in \Omega \] For any $X \in \Omega$ we have $|G_X| |\orb_G(X)| = |G| = p^a m$. By (\dag) there exists $X$ such that $|\orb_G(X)|$ is coprime to $p$. Thus $p^a ~\big|~ |G_X|$ (1). On the other hand, if $g \in G$ and $x \in X$, then $g \in (gx^{-1}) * X$ and hence \[ G = \bigcup_{g \in G} g * X = \bigcup_{Y \in \orb_G(X)} Y \] \[ \implies |G| \le |\orb_G(X)| |X| \] \[ \implies |G_X| = \frac{|G|}{|\orb_G(X)|} \le |X| = p^a \tag{2} \] (1) and (2) implies \[ |G_X| = p^a \] i.e. $G_X \in \Syl_p(G)$. } \end{flashcard} \end{enumerate}