Group Theory
This page contains my notes for MTH 611M, an introductory graduate class in abstract algebra, which I took in 2023. The notes roughly follow the outline of Hungerford’s Algebra. A complete list of references can be found at the end of the page.
Contents
Some preliminaries
Set theory
A set is a collection of mathematical objects called its elements. We write $a \in X$ if $a$ is an element of the set $X$ and $a \notin X$ otherwise. A set is completely determined by its elements (axiom of extensionality), that is, if $X$ and $Y$ are sets then $$\forall a\ (a \in X \iff a \in Y) \implies X = Y.$$ The converse of this axiom is also true and defines the notion of equality.
We say that a set $B$ is a subset of another set $A$ and write $B \subseteq A$ if and only if $$a \in B \implies a \in A.$$ Set equality can thus be restated as $$A = B \iff A \subseteq B \text{ and } B \subseteq A.$$ This fact is a useful tool in proving set equalities.
The union of two sets $A$ and $B$ is the set $A \cup B = \{x : x \in A \text{ or } x \in B\}$, while the intersection of $A$ and $B$ is the set $A \cap B = \{x : x \in A \text{ and } x \in B\}$. We can extend this definition for an arbitrary collection $\mathfrak{S}$ of sets as follows: $$x \in \bigcup_{S \in \mathfrak{S}} S \iff \exists S \in \mathfrak{S}\ (x \in S)$$ and $$x \in \bigcap_{S \in \mathfrak{S}} S \iff \forall S \in \mathfrak{S}\ (x \in S).$$ Finally, the set difference of $A$ and $B$ is the set $A \setminus B = \{x : x \in A \text{ and } x \notin B\}$.
An ordered pair $(a, b)$ is a set $\{\{a\}, \{a, b\}\}$. We can extend this definition to ordered $n$-tuples $(a_1, a_2, \ldots, a_n)$, which we define to be the set $\{\{a_1\}, \{a_1, a_2\}, \ldots, \{a_1, a_2, \ldots, a_n\}\}$. We say that two ordered $n$-tuples $(a_1, a_2, \ldots, a_n)$ and $(b_1, b_2, \ldots, b_n)$ are equal if and only if $a_i = b_i$ for all $1 \leq i \leq n$.
The Cartesian product of two sets $A$ and $B$ is the set of all ordered pairs $(a, b)$ such that $a \in A$ and $b \in B$. We denote this set by $A \times B$.
A relation on a set $X$ is a subset $R$ of $X \times X$. If $(a, b) \in R$, we say that $a$ is related to $b$ by $R$ and write $R(a, b)$ or $aRb$. If $R$ is a relation on $X$, then we say that $R$ is reflexive if $aRa$ for all $a \in X$. We say that $R$ is symmetric if $aRb$ implies $bRa$ for all $a, b \in X$. We say that $R$ is transitive if $aRb$ and $bRc$ implies $aRc$ for all $a, b, c \in X$. If $R$ is reflexive, symmetric, and transitive, then we say that $R$ is an equivalence relation and we write $a \sim b$ instead of $aRb$. Given a set $X$ and an equivalence relation $\sim$ on $X$, the equivalence class of an element $a$ of $X$ (written $[x]$) is the set $$[a] = \{x \in X : a \sim x\}.$$
Examples. Consider the set $\mathbb{R}$ of real numbers and the equality relation ($=$). Observe that if $x$ is a real number then $x$ is equal to itself, i.e., $x = x$. Moreover, $x = y$ implies $y = x$, and vice versa, for any other real number real number $y$. Finally, if $z$ is a third real number, then $x = y$ and $y = z$ implies $x = z$. Thus, equality is an equivalence relation on $\mathbb{R}$.
On the other hand, consider the relation $<$ on $\mathbb{R}$. By the trichotomy law in $\mathbb{R}$, $x < y$ and $y < x$ cannot both be true, and thus $<$ is not symmetric and hence not an equivalence relation.
Let $\mathbb{Z}$ be the set of integers; then for all integers $a$, $b$, and $n$, with $n > 1$, we define the equivalence relation $\sim$ as $a \sim b \iff n \mid (a - b).$ (The reader can verify that this is indeed an equivalence relation.) We call this relation congruence modulo $n$ and write $a \equiv b \pmod{n}$ whenever $a \sim b$.
Mappings
A mapping (or a function) $f$ from a set $A$ to a set $B$ written $f : A \to B$ is a rule that assigns to each element $a \in A$ a unique element $b \in B$. We would often write $f(a)$ instead of $b$ and say that $f$ maps $a$ to $b$ (or in symbols $a \mapsto b$ or $a \mapsto f(a)$). The set $A$ is called the domain of $f$ and the set $B$ is called the codomain of $f$. The image of $f$ is the set $$f(A) = \{f(a) : a \in A\}.$$ If $X \subseteq A$, then the image of $X$ under $f$ is the set $$f(X) = \{f(x) : x \in X\}.$$ If $Y \subseteq B$, then the inverse image of $Y$ under $f$ is the set $$f^{-1}(Y) = \{x \in A : f(x) \in Y\}.$$ The restriction of $f$ to $X$ is the mapping $f|_X : X \to B$ defined by $f|_X(x) = f(x)$ for all $x \in X$.
A mapping $f : A \to B$ is injective if $f(a) = f(b)$ implies $a = b$ for all $a, b \in A$. A mapping $f : A \to B$ is surjective if $f(A) = B$. A mapping $f : A \to B$ is bijective if it is both injective and surjective. If $f : A \to B$ is bijective, then there exists a mapping $g : B \to A$ such that $g \circ f = \operatorname{id}_A$ and $f \circ g = \operatorname{id}_B$, where $\operatorname{id}_A$ and $\operatorname{id}_B$ are the identity mappings on $A$ and $B$, respectively. The mapping $g$ is called the inverse of $f$ and is denoted by $f^{-1}$.
1. The definition of a group
Definition 1.1.
A binary operation on a set $X$ is a map $X \times X \to X$. Let $\varphi$ be a binary operation on a set $G$ and let $a, b, c$ be elements of $G$. We say that $G$ is a semigroup under $\varphi$ if $\varphi$ is associative, that is, $$\varphi(a, \varphi(b, c)) = \varphi(\varphi(a, b), c)$$
We say that $G$ is a monoid under $\varphi$ if $G$ is a semigroup under $\varphi$ and there exists an element $e \in G$ (called the identity element) such that $$\varphi(e, a) = \varphi(a, e) = a$$ for all $a \in G$.
Finally, we say that $G$ is a group under $\varphi$ if $G$ is a monoid under $\varphi$ and for every $a \in G$ there exists an element $a^{-1} \in G$ (called the inverse of $a$) such that $$\varphi(a, a^{-1}) = \varphi(a^{-1}, a) = e.$$
Observe that commutativity is not required for a group. That is, we do not require that $\varphi(a, b) = \varphi(b, a)$ for all $a, b \in G$. If $\varphi$ does, however, satisfy this property, then we say that the underlying set $G$ is an abelian group under $\varphi$.
For the sake of brevity we will often be writing $\varphi(a, b)$ as $a \cdot b$, or, more succinctly, as $ab$. (In the case of abelian groups we shall instead write $a+b$, although this decision is only a matter of convention.) This shorthand also extends to calling the binary operation ‘multiplication’ (or ‘addition’ in the case of abelian groups) without implying that the operation is commutative or is the usual multiplication (or addition) of numbers. We will also be saying that ‘$G$ is a group’ instead of ‘$G$ is a group under $\varphi$’, unless ambiguity forces us to be more precise.
Examples. The development of group theory has been motivated in large part by the study of symmetry. Consider a square centered at the origin.
The set $\mathbb{Z}$ of integers is a group under the usual addition of integers. The identity element is $0$ and the inverse of an integer $n$ is $-n$. The set $\mathbb{Q}$ of rational numbers is a group under the usual addition of rational numbers, but fails to be a group under the usual multiplication of rational numbers since $0$ has no inverse.
Linear groups. Consider the set of $n \times n$ invertible matrices with entries from a field $\mathbb{K}$. Then matrix multiplication is a binary operation on this set since the product of two $n \times n$ matrices is again an $n \times n$ matrix. The identity matrix $I_n$ is the identity element of this group, and the inverse of a matrix $A$ is the matrix $A^{-1}$ such that $AA^{-1} = A^{-1}A = I_n$. (Note that we require the matrices to be invertible.) We call this group the general linear group of degree $n$ over $\mathbb{K}$, $\operatorname{GL}_n(\mathbb{K})$. If we consider only the set of $n \times n$ matrices with entries from $\mathbb{K}$ with determinant $1$, then we obtain the special linear group of degree $n$ over $\mathbb{K}$, $\operatorname{SL}_n(\mathbb{K})$.
Klein four-group. Consider the matrices $$ a = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix},\ b = \begin{pmatrix} -1 & 0 \\ 0 & 1 \end{pmatrix},\ c = \begin{pmatrix} -1 & 0 \\ 0 & -1 \end{pmatrix}.$$ When taken together with the $2 \times 2$ identity matrix $e$, these matrices form a group under matrix multiplication. This group is called the Klein four-group and is denoted by $V_4$ (after the German Viergruppe). Every element of $V_4$ is its own inverse, and multiplying any two non-identity elements of $V_4$ yields the third element. The group operation is commutative and thus the group is abelian.
Quaternion group.
Theorem 1.2
Let $G$ be a group, and $a, b, c \in G$. Then:
- $aa = a \implies a = e$
- Left-cancellation: $ab = ac \implies b = c$
- Right-cancellation: $ba = ca \implies b = c$
Definition 1.3
The order of a group $G$ (written $|G|$) is the cardinality of its underlying set, i.e., the number of elements of the group. If $|G| \lt \infty$, then we say that $G$ is a finite group; otherwise, we say that $G$ is an infinite group.
We can describe the structure of a finite group by arranging all possible products in the group in a grid, with the factors of each product on either end of the grid’s rows and columns. This grid is called the Cayley table after British mathematician Arthur Cayley, who introduced it in his paper On the Theory of Groups as depending on the Symbolic Equation $\theta^n = 1$ (1854). Cayley tables are particularly useful for non-abelian groups, where the product $ab$ is not necessarily equal to $ba$. Some examples of Cayley tables are shown below
- Cayley table for $V_4$.
$\times$ $e$ $a$ $b$ $c$ $e$ $e$ $a$ $b$ $c$ $a$ $a$ $e$ $c$ $b$ $b$ $b$ $c$ $e$ $a$ $c$ $c$ $b$ $a$ $e$ -
Cayley table for $Q_8$.
$\times$ $e$ $a$ $a^2$ $a^3$ $b$ $ab$ $a^2b$ $a^3b$ $e$ $e$ $a$ $a^2$ $a^3$ $b$ $ab$ $a^2b$ $a^3b$ $a$ $a$ $a^2$ $a^3$ $e$ $ab$ $a^2b$ $a^3b$ $b$ $a^2$ $a^2$ $a^3$ $e$ $a$ $a^2b$ $a^3b$ $b$ $ab$ $a^3$ $a^3$ $e$ $a$ $a^2$ $a^3b$ $b$ $ab$ $a^2b$ $b$ $b$ $a^3b$ $a^2b$ $ab$ $a^2$ $a$ $e$ $a^3$ $ab$ $ab$ $b$ $a^3b$ $a^2b$ $a^3$ $a^2$ $a$ $e$ $a^2b$ $a^2b$ $ab$ $b$ $a^3b$ $e$ $a^3$ $a^2$ $a$ $a^3b$ $a^3b$ $a^2b$ $ab$ $b$ $a$ $e$ $a^3$ $a^2$
2. Homomorphisms and subgroups
In general, our interest in groups is usually not on the groups themselves, but rather on the maps between them. We will be interested in maps that preserve the group structure, which we call homomorphisms.
Definition 2.1. Let $G$ and $H$ be groups under $\varphi$ and $\psi$ respectively. A map $\Phi: G \to H$ is called a homomorphism if $$\Phi(\varphi(a, b)) = \psi(\Phi(a), \Phi(b))$$ for all $a, b \in G$.
A more ambiguous, though admittedly less cumbersome, way of saying this is as follows: $\Phi$ is a homomorphism if $$\Phi(ab) = \Phi(a)\Phi(b).$$ In this alternative notation, we do not distinguish between the multiplication in $G$ and the multiplication in $H$, and instead rely on context to determine which group operation we are referring to.
A (group) homomorphism that is injective as a map between sets is called a monomorphism. A homomorphism that is surjective as a map between sets is called an epimorphism. A homomorphism that is both a monomorphism and an epimorphism is called an isomorphism. If there exists an isomorphism between $G$ and $H$, then we say that $G$ and $H$ are isomorphic and write $G \cong H$.
A homomorphism from a group $G$ to itself is called an endomorphism. An endomorphism that is also an isomorphism is called an automorphism.
Theorem 2.2.
Let $G$ and $H$ be groups with identities $e$ and $e'$ respectively, and let $a$ be an element of $G$. Let $\Phi: G \to H$ be a homomorphism. Then:
- $\Phi(e) = e'$
- $\Phi(a^{-1}) = \Phi(a)^{-1}$
That is, a homomorphism preserves identities and inverses.
Proof. (1) $\Phi(e) = \Phi(ee) = \Phi(e)\Phi(e)$, and so by Theorem 1.2, $\Phi(e) = e'$.
Let $a \in G$. Then $\Phi(a)\Phi(a^{-1}) = \Phi(aa^{-1}) = \Phi(e) = e'$, so that $\Phi(a^{-1}) = \Phi(a)^{-1}$.
Definition 2.3
Let $\Phi: G \to H$ be a homomorphism. The kernel of $\Phi$ is the set $$\operatorname{Ker} \Phi = \{a \in G : \Phi(a) = e\}.$$ where $e$ is the identity element of $H$. If $A \subseteq G$, then the image of $A$ under $\Phi$ is the set $\Phi(A) = \{\Phi(a) : a \in A\}.$. The set $\Phi(G)$ is called the image of $\Phi$, and is denoted by $\operatorname{Im} \Phi$.
Theorem 2.4
A homomorphism $\Phi: G \to H$ is a monomorphism if and only if $\operatorname{Ker} \Phi = \{e\}$.
Proof. Let $\Phi$ be a monomorphism. Then $\operatorname{Ker} \Phi = \{a \in G : \Phi(a) = e\} = \{e\}$, since $\Phi$ is injective. Conversely, suppose that $\operatorname{Ker} \Phi = \{e\}$. Let $a, b \in G$ such that $\Phi(a) = \Phi(b)$. Then $\Phi(a)\Phi(b)^{-1} = \Phi(a)\Phi(b^{-1}) = \Phi(ab^{-1})$. Thus $ab^{-1} \in \operatorname{Ker} \Phi$, so that $ab^{-1} = e$. Hence $a = b$, and therefore $\Phi$ is a monomorphism.
Theorem 2.5
Let $\Phi: G \to H$ be a homomorphism. Then $\Phi$ is an isomorphism if and only if there exists a homomorphism $\Phi^{-1}: H \to G$ such that $\Phi^{-1} \circ \Phi = e$ and $\Phi \circ \Phi^{-1} = e'$, where $e$ and $e'$ are the identity elements of $G$ and $H$ respectively.
Definition 2.6
A subset $H$ of a group $G$ is called a subgroup of $G$ if $H$ is itself a group under the operation of $G$. If $H$ is a subgroup of $G$, then we write $H \leq G$.
3. Cyclic groups
Theorem 3.1
Every subgroup of the additive group $\mathbb{Z}$ of integers is a cyclic group. In particular, if $H \leq \mathbb{Z}$, then $H = \langle 0 \rangle$ or $H = \langle m \rangle$ where $m$ is the smallest positive integer in $H$. If $H \neq \langle 0 \rangle$, then $H$ is infinite.
Proof. Either $H = \langle 0 \rangle \leq \mathbb{Z}$ or $H$ contains a least positive integer $m$ (by the well-ordering principle). We claim that $H = \langle m \rangle$ and prove it by showing that $H \subseteq \langle m \rangle$ and $\langle m \rangle \subseteq H$.
Since $H$ is a subgroup and is thus closed under the binary operation of $\mathbb{Z}$, it follows that $$\langle m \rangle = \{km : k \in \mathbb{Z}\} \subseteq H$$ (here the notation is additive). Now let $h \in H$. By the division algorithm, there exist unique integers $q$ and $r$ such that $h = qm + r$ and $0 \leq r \lt m$. Since $H$ is a subgroup, $h - qm = r \in H$ (because $h, qm \in H$). But $m$ is the least positive integer in $H$, so $0 \leq r \lt m$ implies $r = 0$. Thus $h = qm \in \langle m \rangle$, and so $H \subseteq \langle m \rangle$.
If $H \neq \langle 0 \rangle$, then for integers $k_1, k_2$ such that $k_1 \neq k_2$, we have $k_1m \neq k_2m$. Thus $\langle m \rangle$ is infinite.
Theorem 3.2
Every cyclic group of infinite order is isomorphic to the additive group $\mathbb{Z}$ of integers. Every cyclic group of finite order $n$ is isomorphic to the additive group $\mathbb{Z}_n$ of integers modulo $n$.
Proof. Let $G$ be a cyclic group generated by $a$. If $G$ is infinite, then the map $\Phi: \mathbb{Z} \to G$ defined by $\Phi(k) = a^k$ is a homomorphism since for all integers $\kappa, \lambda$, we have $$\Phi(\kappa + \lambda) = a^{\kappa + \lambda} = a^{\kappa}a^{\lambda} = \Phi(\kappa)\Phi(\lambda).$$ Since $G$ is cyclic, every element of $G$ is a power of $a$ and therefore $\Phi$ is surjective. By Theorem 2.4, $\Phi$ is a monomorphism if and only if $\operatorname{Ker} \Phi = \{0\}$
Now consider the case where $G$ is finite of order $n$. For $r, s \in \mathbb{Z}$, we have $a^r = a^s$ if and only if $a^{r-s} = e$, i.e., if and only if $r - s \in \operatorname{Ker} \Phi$.
Theorem 3.3 Order of elements in a cyclic group
Let $G$ be a cyclic group and $a$ an element of $G$. If $G$ has infinite order then
- $a^k = e$ if and only if $k = 0$.
- $a^k = a^m$ if and only if $k = m$.
If $G$ has finite order $n$, then
- $n$ is the least positive integer such that $a^n = e$.
- $a^k = e$ if and only if $n$ divides $k$.
- $a^k = a^m$ if and only if $k \equiv m \pmod{n}$.
- for each $k$ such that $k \mid n$, then $|a^k| = \frac{n}{k}$.
4. Cosets and Lagrange's theorem
We introduce the notion of cosets, both as the congruence classes modulo a subgroup and geometrically as the translation of a subgroup by an element of the group. Through this we will be able to prove Lagrange's theorem, a fundamental result in group theory, which will help us describe the structure of finite groups.
Definition 4.1
Let $G$ be a group and let $H$ be a subgroup of $G$. For any $a, b \in G$, we say that $a$ is right (resp. left) congruent to $b$ modulo $H$ if $ab^{-1} \in H$ (resp. $a^{-1}b \in H$). We denote this by $a \equiv_R b \pmod{H}$ (resp. $a \equiv_L b \pmod{H}$). We say that $a$ is congruent to $b$ modulo $H$ if it is both right and left congruent to $b$ modulo $H$, and denote this by dropping the subscript, i.e., $a \equiv b \pmod{H}$.
The next result is essential in linking the algebraic and geometric interpretations of cosets.
Theorem 4.2
Let $G$ be a group and let $H \leq G$.
- Right (resp. left) congruence modulo $H$ is an equivalence relation on $G$.
- The equivalence class of $a \in G$ under right (resp. left) congruence modulo $H$ is the set $Ha = \{ha : h \in H\}$ (resp. $aH = \{ah : h \in H\}$).
Proof.
(1) Let $a, b, c \in G$ and let $\sim$ be right congruence modulo $H$ for brevity. Then $aa^{-1} = e \in H$ because $H$ is a subgroup and thus contains the identity, and so $a \sim a$. Since $a \sim b$ if and only if $ab^{-1}$, we have $(ab^{-1})^{-1} = ba^{-1} \in H$ because $H$ is closed under inverses, so that $b \sim a$. Finally, if $a \sim b$ and $b \sim c$, then $ab^{-1}, bc^{-1} \in H$, and so $ac^{-1} = (ab^{-1})(bc^{-1}) \in H$, and thus $a \sim c$. Therefore $\sim$ is an equivalence relation. A similar argument shows that left congruence modulo $H$ is also an equivalence relation.
(2) The equivalence class of $a$ under right congruence modulo $H$ is the set $$[a]_{\equiv_R} = \{b \in G : b \equiv_R a\} = \{b \in G : ba^{-1} \in H\}.$$ Suppose $\xi \in [a]_{\equiv_R}$. Then $\xi a^{-1} \in H$, and so $\xi = (\xi a^{-1})a \in Ha$. Conversely, if $\xi \in Ha$, then $\xi = ha$ for some $h \in H$, and so $\xi a^{-1} = h \in H$. Thus $[a]_{\equiv_R} = Ha$. A similar argument shows that $[a]_{\equiv_L}$ is the set $aH$.
The equivalence classes of $G$ under right (resp. left) congruence modulo $H$ are called the right (resp. left) cosets of $H$ in $G$. We call the element $a \in G$ the coset representative of $Ha$ (and $aH$).
Example. Consider the group $S_3$ and the subgroup $H = \{e, (1\ 2)\}$. Then the right and left cosets of $H$ in $S_3$ are:
$a$ | $Ha$ | $aH$ |
---|---|---|
$e$ | $\{e, (1\ 2)\}$ | $\{e, (1\ 2)\}$ |
$(1\ 2)$ | $\{(1\ 2), e\}$ | $\{(1\ 2), e\}$ |
$(1\ 3)$ | $\{(1\ 3), (1\ 3\ 2)\}$ | $\{(1\ 3), (1\ 2\ 3)\}$ |
$(2\ 3)$ | $\{(2\ 3), (1\ 2\ 3)\}$ | $\{(2\ 3), (1\ 3\ 2)\}$ |
$(1\ 2\ 3)$ | $\{(1\ 2\ 3), (2\ 3)\}$ | $\{(1\ 2\ 3), (1\ 3)\}$ |
$(1\ 3\ 2)$ | $\{(1\ 3\ 2), (1\ 3)\}$ | $\{(1\ 3\ 2), (2\ 3)\}$ |
Observe that the left and right cosets of $H$ are not necessarily equal.
5. Normal subgroups and quotient groups
Definition 5.1 Normal subgroups
A subgroup $N$ of a group $G$ is called a normal subgroup of $G$ if it is invariant under conjugation, i.e., for all $a \in G$ and for all $n \in N$, $ana^{-1} \in N$. If $N$ is normal in $G$, then we write $N \triangleleft G$.
The next result provides alternative characterizations of normal subgroups.
Theorem 5.2 Equivalent conditions for normality
Let $N$ be a subgroup of a group $G$. Then the following statements are equivalent:
- $N$ is normal in $G$.
- The image of $N$ under conjugation is contained in $N$, i.e., for all $a \in G$, $aNa^{-1} \subseteq N$.
- The image of $N$ under conjugation is equal to $N$, i.e., for all $a \in G$, $aNa^{-1} = N$.
- For all $a \in G$, the left and right cosets $aN$ and $Na$ are equal.
- The left and right cosets of $N$ in $G$ are coincide.
Examples. The trivial subgroups $\{e\}$ and $G$ of any group $G$ are normal subgroups of $G$. Every subgroup of an abelian group is trivially normal.
If $\varphi : G \to H$ is a homomorphism, then $\operatorname{Ker} \varphi$ is a normal subgroup of $G$. Indeed if $a \in G$ and $n \in \operatorname{Ker} \varphi$, then $$ \varphi(ana^{-1}) = \varphi(a)\varphi(n)\varphi(a^{-1}) = \varphi(a)\varphi(a^{-1}) = \varphi(aa^{-1}) = \varphi(e) = e', $$ and so $ana^{-1} \in \operatorname{Ker} \varphi$.
Theorem 5.3 Some properties of normal subgroups
Let $H$ and $N$ be subgroups of a group $G$, with $N \triangleleft G$. Then:
- $H \cap N \triangleleft H$.
- $HN = NH = H \vee N$.
- $N \triangleleft H \vee N$.
- if $H$ is also normal in $G$ and $H \cap N = \{e\}$, then $hn = nh$ for all $h \in H$ and $n \in N$.
Proof.
(1) If $n \in N \cap H$ and $a \in H$ then $ana^{-1} \in N$ because $N$ is normal. Moreover, because $H$ is a subgroup, $ana^{-1} \in H$. Thus $ana^{-1} \in N \cap H$, and therefore $N \cap H \triangleleft H$.
(2) Recall that $N \vee H := \langle N \cup H \rangle$.
(4) Since $N \triangleleft G$, we have $hnh^{-1} \in N$ and since $H \triangleleft G$, $nh^{-1}n^{-1} \in H$. Thus $hnh^{-1}n^{-1} \in H \cap N = \{e\}$, and so $hn = nh$.
Theorem 5.4 Quotient groups
Let $N$ be a normal subgroup of a group $G$. Then the set of distinct (left) cosets of $N$ in $G$ is a group under the operation defined by $$(aN)(bN) = (ab)N,$$ for all $a, b \in G$. This group is called the quotient group (or factor group) of $G$ by $N$, and is denoted by $G/N$ (read "$G$ mod $N$").
Proof.
First we show that the operation (which we shall call coset multiplication) is well-defined. Let $a, b, c, d \in G$ such that $aN = cN$ and $bN = dN$. Then since cosets are simply the equivalence classes of elements of $G$ under congruence modulo $N$, we have $a \equiv c \pmod{N}$ and $b \equiv d \pmod{N}$. Thus $c^{-1}a \in H$ and $d^{-1}b \in H$. To show that the operation is well-defined, we must show that $(ab)N = (cd)N$ and thus $$ (cd)^{-1}(ab) \in N. $$ We have $$ (cd)^{-1}(ab) = d^{-1}c^{-1}(ab) = (d^{-1}c^{-1}ad)(d^{-1}b). $$ Because $c^{-1}a \in N$ by assumption, and because $N$ is normal, $d^{-1}c^{-1}ad \in N$. Similarly, $d^{-1}b \in N$. Because $N$ is a subgroup, it is closed under the binary operation of $G$, and so $(cd)^{-1}(ab) \in N$. Thus coset multiplication is well-defined.
We now compare the operation against the group axioms. Given $a, b, c \in G$, we have $(aNbN)cN = aN(bcN) = a(bc)N$, and because the group operation is associative we have $a(bc)N = (ab)cN = (abN)cN = (aNbN)cN$ so that coset multiplication is associative. The (left) coset $eN = N$ is the identity, and for each $a \in G$ (and thus $aN \in G / N$), the coset whose representative is the inverse is $a^{-1}$ of $a$ is the inverse of $aN$. Thus $G / N$ is a group.
The arguments above work with right cosets as well. Without loss of generality, we will use left cosets when working with quotient groups.
Theorem 5.5 Kernel of a homomorphism and the canonical projection
If $\varphi: G \to H$ is a group homomorphism, then $\operatorname{Ker} \varphi$ is a normal subgroup of $G$. Conversely, if $N$ is a normal subgroup of $G$, then the map $\pi: G \to G / N$ defined by $\pi(a) = aN$ is an epimorphism with kernel $N$.
Proof.
Suppose $x \in \operatorname{Ker} \varphi$ and $a \in G$. Then $$ \varphi(axa^{-1}) = \varphi(a)\varphi(x)\varphi(a^{-1}) = \varphi(a)\varphi(a^{-1}) = \varphi(aa^{-1}) = \varphi(e) = e', $$ where $e$ and $e'$ are the identities of $G$ and $H$ respectively. Thus $axa^{-1} \in \operatorname{Ker} \varphi$, and so $\operatorname{Ker} \varphi \triangleleft G$.
The map $\pi$ is surjective by definition of $G / N$; and since for all $a, b \in G$, $$ \pi(ab) = (ab)N = (aN)(bN) = \pi(a)\pi(b), $$ $\pi$ is an epimorphism. We have further that $$ \operatorname{Ker} \pi = \{a \in G : \pi(a) = N\} = \{a \in G : aN = N\} = \{a \in G : a \in N\} = N. $$
We call $\pi$ the canonical projection of $G$ onto $G / N$.
Theorem 5.6 First isomorphism theorem
Let $\varphi: G \to H$ be a group homomorphism. If $N$ is a normal subgroup of $G$ contained in $\operatorname{Ker} \varphi$, then $\varphi$ induces an isomorphism $\psi : G / N \to H$ such that $\psi \circ \pi = \varphi$, where $\pi$ is the canonical projection of $G$ onto $G / N$. Moreover, $\psi$ is an isomorphism if and only if $\varphi$ is an epimorphism and $\operatorname{Ker} \varphi = N$.
In particular, $G / \operatorname{Ker} \varphi \cong \operatorname{Im} \varphi$.
Proof.
Let $b$ be an element of some coset $aN$ of $G / N$. Then $b = an$ for some $n \in N$. Thus $$ \varphi(b) = \varphi(an) = \varphi(a)\varphi(n) = \varphi(a)e' = \varphi(a), $$ and thus $\varphi$ is constant on each coset of $G / N$. Thus if we define $\psi: G / N \to \operatorname{Im} \varphi$ by $\psi(aN) = \varphi(a)$, then $\psi$ is well-defined. Moreover, $\psi$ is a homomorphism because $$ \psi(aN)\psi(bN) = \varphi(a)\varphi(b) = \varphi(ab) = \psi(abN). $$
Moreover, we have $$ \operatorname{Im} \psi = \{\psi(aN) : a \in G\} = \{\varphi(a) : a \in G\} = \operatorname{Im} \varphi. $$ Thus $\psi$ is surjective if and only if $\varphi$ is surjective. Finally, $\psi$ is injective if and only if $$ \begin{align} \operatorname{Ker} \psi &= \{aN \in G / N : \psi(aN) = e'\} = \{aN \in G / N : \varphi(a) = e'\}\\ &= \{aN \in G / N : a \in \operatorname{Ker} \varphi\} = \operatorname{Ker} \varphi / N. \end{align} $$ is trivial, which is true if and only if $\operatorname{Ker} \varphi = N$. Therefore $\psi$ is an isomorphism if and only if $\varphi$ is an epimorphism and $\operatorname{Ker} \varphi = N$.
Since $\psi$ is uniquely determined by $\varphi$, it follows that it is unique. The last statement follows immediately from the proof.
Theorem 5.7 Second isomorphism theorem
Let $G$ be a group and let $H$ and $N$ be subgroups of $G$ such that $N \triangleleft G$. Then $HN / N \cong H / (H \cap N)$.
6. Symmetric, alternating, and dihedral groups
Definition 6.1
A permutation of a set $\Omega$ is a bijection from $\Omega$ to $\Omega$. The set of all permutations of $\Omega$ is denoted by $S_{\Omega}$; if $\Omega = \{1, 2, \ldots, n\}$, then we write $S_n$ instead of $S_{\Omega}$. The group $S_n$ is called the symmetric group of $n$ letters.
Let $\{\omega_1, \ldots, \omega_r\}$ (with $r \leq n$) be a subset of $\Omega = \{1, 2, \ldots, n\}$. Then we we will write $(\omega_1\ \omega_2\ \ldots\ \omega_r)$ for the permutation $\sigma \in S_n$ such that $\sigma(\omega_i) = \omega_{i+1}$ for $i = 1, \ldots, r-1$ and $\sigma(\omega_r) = \omega_1$, and $\sigma(i) = i$ for all $i \in \Omega \setminus \{\omega_1, \ldots, \omega_r\}$. We call $(\omega_1\ \omega_2\ \ldots\ \omega_r)$ a cycle of length $r$. A $2$-cycle is called a transposition.
Example. Consider the permutation $$ \sigma = \begin{pmatrix} 1 & 2 & 3 & 4 & 5 \\ 3 & 5 & 4 & 2 & 1 \end{pmatrix} $$ which is equivalent to the sequence $$ 1 \mapsto 3 \mapsto 4 \mapsto 2 \mapsto 5 \mapsto 1. $$ Thus in cycle notation, $\sigma = (1\ 3\ 4\ 2\ 5)$.
Definition 6.2
The permutations $\sigma_1, \ldots, \sigma_k \in S_n$ are said to be disjoint if they have no common elements other than the identity. That is, for each $1 \leq i \leq k$ and for each $\xi \in \{1, \ldots, n\}$, $\sigma_i(\xi) = \xi$ implies $\sigma_j(\xi) = \xi$ for all $j \neq i$.
The identity permutation $\iota$ is the permutation such that $\iota(\xi) = \xi$ for all $\xi \in \Omega$. The product $\sigma\tau$ of two permutations $\sigma, \tau \in S_n$ is the composition $\sigma \circ \tau$. Observe that $\sigma\tau$ is also a permutation of $\Omega$ since the composition of two bijections is a bijection. The product of any permutation with the identity permutation is the permutation itself. Moreover because each $\sigma \in S_n$ is a bijection, there exists a unique permutation $\sigma^{-1}$ such that $\sigma\sigma^{-1} = \sigma^{-1}\sigma = \iota$. We call $\sigma^{-1}$ the inverse of $\sigma$ and is the cycle $$ (\omega_r\ \omega_{r-1}\ \ldots\ \omega_1) = (\omega_1\ \omega_r\ \omega_{r-1}\ \ldots\ \omega_2). $$ Therefore we see that $S_n$ is a group under the product (i.e., composition) of permutations.
Observe that $\sigma\tau = \tau\sigma$ is not necessarily true.
7. Group actions
Definition 7.1 Group actions
Given a set $X$ and a group $G$, a left group action of $G$ on $X$ is a map $\Phi: G \times X \to X$ such that for all $g, h \in G$ and for all $x \in X$,
- $\Phi(e, x) = x$
- $\Phi(g, \Phi(h, x)) = \Phi(gh, x)$.
For the sake of brevity, we will write $\Phi(g, x)$ as $gx$. We call $X$ a $G$-set under the action $\Phi$.
Examples. The trivial group action by $G$ on $X$ maps each $x \in X$ to itself. We check this against the axioms in our definition. For all $g \in G$ and $x \in X$, we have $ex = x$ and $g(ex) = gx = (ge)x = g(x)$.
Theorem 7.2 Orbit and stabilizer
Let $G$ be a group acting on a set $X$.
- The relation on $X$ defined by $$ x \sim y \iff gx = y \text{ for some } g \in G $$ is an equivalence relation on $X$. The equivalence class of $x \in X$ is called the orbit of $x$ under the action of $G$ on $X$ and is written $\mathcal{O}_x$.
- For each $x \in X$, the set $$ G_x = \{g \in G : gx = x\} $$ is a subgroup of $G$ called the stabilizer (or isotropy group) of $x$ under the action of $G$ on $X$.
Proof.
(1) Let $x \in X$. Then $ex = x$, so that $x \sim x$. If $x \sim y$, then there exists $g \in G$ such that $gx = y$. But $g^{-1}y = g^{-1}(gx) = (g^{-1}g)x = ex = x$, and so $y \sim x$. Finally, if $x \sim y$ and $y \sim z$, then there exist $g, h \in G$ such that $gx = y$ and $hy = z$. Thus $h(gx) = (hg)x = z$, and so $x \sim z$. Therefore $\sim$ is an equivalence relation.
(2) The set $G_x$ is nonempty because $ex = x$ and $e \in G_x$. Let $g, h \in G_x$. Then $gx = x$ and $hx = x$, and so $g(hx) = (gh)x = x$ so that $gh \in G_x$. Finally, if $g \in G_x$, then $gx = x$, and so $g^{-1}x = g^{-1}(gx) = (g^{-1}g)x = ex = x$, and thus $g^{-1} \in G_x$. Therefore $G_x$ is a subgroup of $G$.
Example. If a group $G$ acts on itself by conjugation, then the orbit of $x \in G$ is called the conjugacy class of $x$. If a subgroup $H \leq G$ acts on $G$ by conjugation, then its stabilizer $$ H_x = \{h \in H : hxh^{-1} = x\} = \{h \in H : hx = xh\} $$ is called the centralizer of $x$ in $H$. If $H = G$, then the centralizer of $x$ in $G$ is simply called the centralizer of $x$. (Observed that $C_G(x)$ is exactly the center $Z(G)$ of $G$.)
Let $\mathfrak{G}$ be the set of all subgroups of $G$. Then if $H \leq G $ acts on $\mathfrak{G}$ by conjugation, the stabilizer of some subgroup $K \in \mathfrak{G}$ is the set $$ G_K = \{g \in G : gKg^{-1} = K\} $$ of all elements of $G$ leave $K$ invariant under conjugation. This is called the normalizer of $K$ in $G$, and is denoted by $N_G(K)$.
These subgroups are related by the following chain of inclusions: $$ Z(G) \subseteq C_G(H) \subseteq N_G(H) \subseteq G, $$ where $H \leq G$.
Theorem 7.3 Orbit-stabilizer theorem
Let $G$ be a group acting on a set $X$. Then for all $x \in X$, $$ |\mathcal{O}_x| = [G : G_x] = \frac{|G|}{|G_x|}. $$
Proof.
We want to show that $[G : G_x]$ has the same cardinality as $|\mathcal{O}_x|$, i.e., there is a bijection between $[G : G_x]$ and $\mathcal{O}_x$. Since $G_x$ is a subgroup, we have for any $g, h \in G$ $$ gx = hx \iff g^{-1}hx = x \iff g^{-1}h \in G_x \iff g \equiv h \pmod{G_x} \iff gG_x = hG_x. $$ Thus the map defined by $\varphi: gG_x \mapsto gx$ is a well-defined bijection between $[G : G_x]$ and $\mathcal{O}_x$.
Theorem 7.4 Conjugacy classes and the class equation
Let $G$ be a finite group acting on itself and $H$ a subgroup of $G$. Then:
- The number of elements in the conjugacy class of $x \in G$ is $[G : C_G(x)]$.
- If $\mathcal{O}(x_1), \ldots, \mathcal{O}(x_k)$ are the distinct conjugacy classes of $G$, then $$ |G| = \sum_{i=1}^k |\mathcal{O}(x_i)| = \sum_{i=1}^k [G : C_G(x_i)]. $$
- The number of subgroups of $G$ conjugate to $H$ is $[G : N_G(H)]$.
Theorem 7.5 Group actions as permutations
Let $G$ be a group acting on a set $X$. Then the action of $G$ on $X$ induces a homomorphism $G \to \mathfrak{S}_X$, where $\mathfrak{S}_X$ is the set of all permutations of $X$.
Proof.
Fix $a \in G$ and let $\tau_a : X \to X$ be a map defined by $a \mapsto ax$. Then $\tau_a$ is surjective because for all $x \in X$, $a^{-1}x \in X$ and $\tau_a(a^{-1}x) = a(a^{-1}x) = x$. Moreover, $\tau_a$ is injective because if $\tau_a(x) = \tau_a(y)$, then $ax = ay$ and so $x = y$. Therefore $\tau_a$ is a permutation. Finally, for any $a, b \in G$ and $x \in X$, we have $$ \tau_{ab}(x) = (ab)x = a(bx) = \tau_a(bx) = \tau_a(\tau_b(x)) = (\tau_a \tau_b)(x), $$ and so $\tau_{ab} = \tau_a \tau_b$ and thus $\tau_a: G \to \mathfrak{S}_X$ is a homomorphism.
Theorem 7.6 Cayley’s theorem
Every group $G$ can be embedded in $\mathfrak{S}_G$. In particular, $G$ is isomorphic to a subgroup of $\mathfrak{S}_G$.
Proof.
Let $G$ act on itself by left translation. Then by Theorem 7.5, there exists a homomorphism $\tau: G \to \mathfrak{S}_G$ such that $\tau_g \in \mathfrak{S}_G$. We claim that $\tau$ is injective. Suppose $\tau_g = \tau_h$ for some $g, h \in G$. Then for all $x \in G$, $\tau_g(x) = gx = hx = \tau_h(x)$, and so $g = h$. Thus $\tau$ is injective, and so $G$ is isomorphic to the image of $\tau$ which is a subgroup of $\mathfrak{S}_G$.
8. Sylow theorems
9. Abelian groups*
We first introduce the notion of a free group, and use it to describe groups in terms of generators and relations.
A word on a set $X$ is a sequence of the form $$ a_1, a_2, \ldots, $$ where each $a_i \in X \cup X^{-1} \cup \{1\}$ (here the set $X^{-1}$ is a set disjoint from $X$ such that $|X| = |X^{-1}|$ and whose elements we denote by the bijection $x \mapsto x^{-1}$ and $\{1\}$ is a singleton disjoint from both $X$ and $X^{-1}$, i.e., the choice of $1$ though suggestive is arbitrary) and such that for some $n \in \mathbb{N}$, $a_i = 1$ for all $i > n$. We call $n$ the length of the word and the elements $a_1, a_2, \ldots$ letters. The sequence $$ 1, 1, \ldots $$ is called the empty word and is denoted by $1$.
We say that a word is reduced provided that
- $a_i = x$ implies $a_{i+1} \neq x^{-1}$ for all $i \in \mathbb{N}$ (i.e., no letter is followed by its inverse)
- $a_k = 1$ implies $a_i = 1$ for all $i > k$
Every reduced word is thus of the form $$ a_1^{\lambda_1} a_2^{\lambda_2} \cdots a_n^{\lambda_n}, 1, 1, \ldots, $$ where $a_i \in X$ and $\lambda_i = \pm 1$ for all $i = 1, \ldots, n$. For the sake of brevity we shall hereafter denote this word as $$ a_1^{\lambda_1} a_2^{\lambda_2} \cdots a_n^{\lambda_n}. $$ We denote the set of all reduced words on $X$ by $F_X$.
Theorem 9.1 Set of reduced words is a group
Let $X$ be a set. Then $F_X$ is a group under the operation defined as follows.
Let $w = a_1^{\lambda_1} a_2^{\lambda_2} \cdots a_n^{\lambda_n}$ and $w' = b_1^{\mu_1} b_2^{\mu_2} \cdots b_m^{\mu_m}$ be reduced words on $X$. Then we define the product $ww'$ as the concatenation of $w$ and $w'$ followed by reduction (i.e., cancellation of adjacent letters of the form $xx^{-1}$ or $x^{-1}x$). That is, $$ ww' = a_1^{\lambda_1} a_2^{\lambda_2} \cdots a_n^{\lambda_n} b_1^{\mu_1} b_2^{\mu_2} \cdots b_m^{\mu_m}. $$ Moreover, $F_X = \langle X \rangle$.
We call $F_X$ the free group on $X$.
Proof.
By our definition of the product in $F_X$, it follows that $ww'$ is also a reduced word. The empty word $1$ is the identity because $1w = w1 = w$ for all $w \in F_X$. Finally, for any $w = a_1^{\lambda_1} a_2^{\lambda_2} \cdots a_n^{\lambda_n}$, we have $$ w^{-1} = a_n^{-\lambda_n} \cdots a_2^{-\lambda_2} a_1^{-\lambda_1}, $$ and so $w^{-1}w = ww^{-1} = 1$. Thus $F_X$ is a group.
It remains to show that the product is associative.
To prove the second statement, recall that $\langle X \rangle$ consists of all finite products of the form $$ x_1^{\n_1} x_2^{\n_2} \cdots x_k^{\n_k}, $$ where $x_i \in X$ and $\n_i = \in \mathbb{Z}$ and thus every element of $\langle X \rangle$ is a reduced word. Moreover, every reduced word is a finite product of elements of $X$ and their inverse. Therefore $F_X = \langle X \rangle$.