3AL Notes
3AL Notes
MAM3000W
Morgan Vandeyar
1 Group Theory 2
1.1 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Cosets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Lagrange’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.5 Normal subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6 Factor groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3 p-Groups 34
3.1 Group actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 Conjugation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3 Cauchy’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 p-Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
1
Group Theory 1
Mon 12 Feb Let’s start the course by reviewing some basic group theory. Hopefully this will all be
very familiar to you from 2IA. The content of this chapter is just for review purposes,
you won’t be expected to memorise any of the definitions, results or proofs. Everything
in this chapter will form part of a toolbox for proving more complex results in the rest of
this course. So it is very important you’re comfortable with all of these ideas! 😊
1.1 Groups
Definition 1.1. A group is a pair (G, ·) consisting of a set G and a binary operation ·
on G satisfying the following properties:
Before we cover basic examples of groups, let’s discuss some notation, basic properties
and related concepts.
We call G the underlying set of the group (G, ·). Often, we will just use G to denote
the group, when it is clear what the binary operation · is from context.
2
1.1. Groups
Often times, we will write groups multiplicatively. This means we use the shorthand
ab for a · b. Moreover, we denote the identity element by 1G or 1 and the inverse of an
element a by a−1 . For a positive integer n, we write
a0 = 1,
| · a ·{z· · · · a},
an = a
n times
−n −1
a =a
| · a−1{z· · · · · a−1} .
n times
Remark 1.2. Let G be a group and a, b ∈ G. Then the following properties hold:
Finally, before we get to examples of groups, here are some more algebra terms that
will allow us discuss groups more robustly.
3
1.1. Groups
Example 1.4.
1. The sets Z, Q, R and C (the integers, rational, real and complex numbers, respec-
tively) form abelian groups under addition. In each group, the identity element is
0 and the inverse of an element a is −a. Moreover, Z ≤ Q ≤ R ≤ C.
Wed 14 Feb 4. For each positive integer n, the integers modulo n, denoted Zn = {[0], [1], . . . , [n−1]},
forms a group under addition (where [m] is the equivalence class of m under modulo
n). Namely, [a]+[b] = [a+b]. The identity is [0] and the inverse of [a] is −[a] = [−a].
Such groups are both abelian and cyclic.
7. For a positive integer n, the symmetric group of degree n is the set of permutations
of n objects equipped with the operation of composition. The group is denoted by
Sn and it has order n!. In the case where n = 3, we use the notation
S3 = {ε, σ, σ 2 , τ, τ σ, τ σ 2 },
4
1.1. Groups
where σ 3 = τ 2 = ε and στ σ = τ . ♦
In the previous example, we mentioned that the special linear group is a subgroup of
the general linear group. In order to prove such a statement, we can use the following:
2. If a, b ∈ H, then ab ∈ H,
3. If a ∈ H, then a−1 ∈ H.
1. 1 ∈ H,
In the next example, we use the Subgroup Test to show that SLn (R) ≤ GLn (R).
Example 1.6. First, note that the identity element of GLn (R) is the n × n identity
matrix, In . We know det(In ) = 1, thus In ∈ SLn (R). Now let A, B ∈ SLn (R). We need
to show that AB −1 ∈ SLn (R). That is, we need to show det(AB −1 ) = 1. From properties
of the determinant, we have
So, AB −1 ∈ SLn (R) and we can now conclude that SLn (R) is a subgroup of GLn (R). ♦
Now we see a new example of a group which we will be using in this course.
Example 1.7. Consider a regular n-gon (n ≥ 3). The symmetries of a regular n-gon are
a collection of rotations and reflections that preserve the shape of our n-gon. For example,
consider a square (n = 4). There are 4 rotations and 4 reflections we can do that keep
the shape of the square the same:
5
1.1. Groups
H V D D0
We can rotate the square (around the centre) by 0 degrees, 90 degrees, 180 degrees or 270
degrees. We can also flip the square along the horizontal axis, vertical axis and the two
diagonals. These rotations and reflections obviously keep the square in the same place
but permute the vertices.
We can also compose these symmetries (i.e. rotations and reflections). For example,
we could rotate by 90 degrees then flip along the horizontal. The set of symmetries under
the operation of composition forms a group called the dihedral group, Dn . A dihedral
group always consists of n rotational symmetries and n reflectional symmetries. Thus
|Dn | = 2n.
We can express the dihedral group Dn in a more abstract way which allow us to do
computations more easily. Namely,
where r and s satisfy |r| = n, |s| = 2 and rsr = s. We call r and s the generators of Dn .
In the case of the square, let r represent the rotation by 90 degrees (R90 in the
diagram) and s represent the horizontal reflection (H). Any other reflection or rotation
can be expressed as a composition of r and s. For example, rotating by 180 degrees (R180 )
is r2 . On the other hand, reflection about the off diagonal (D′ ) is the same as a rotation
of 90 degrees (R90 ) followed by a horizontal reflection (H). That is, the reflection D′ can
be represented by sr. ♦
6
1.2. Homomorphisms
1.2 Homomorphisms
Homomorphisms are a very important concept in algebra. These are maps (functions) that
preserve group structure. Homomorphisms give us a way to find interesting connections
between seemingly unrelated groups.
There are also several important subsets that are related to homomorphisms, as out-
lined in the following definition:
Thu 15 Feb Definition 1.10. Let φ : G → G′ be a homomorphism. The kernel of φ is the following
subset of G:
Example 1.11. Consider the groups Z and Z3 (under addition and addition modulo
3, respectively). Let φ : Z → Z3 be the map defined by φ(n) = [n] for all n ∈ Z,
where [n] represents the equivalence class of n under modulo 3. We claim that this is a
homomorphism. To prove this, we take two elements n, m ∈ Z and show φ(n + m) =
φ(n) + φ(m). Using the definition of our map φ, we have φ(n) + φ(m) = [n] + [m]. From
properties of modular arithmetic, we have [n] + [m] = [n + m]. Thus,
7
1.2. Homomorphisms
Note: In the above example, n+m is addition of integers in the group Z and φ(n)+φ(m)
is addition of equivalence classes modulo 3 in Z3 .
Question: Is φ a monomorphism?
2. φ(an ) = φ(a)n for all a ∈ G and n ∈ Z (for the case where n = −1, homomorphisms
preserve inverses),
4. ker φ ≤ G and im φ ≤ G′ ,
Example 1.14.
8
1.3. Cosets
3. K4 is isomorphic to Z2 ×Z2 (the group whose underlying set is the Cartesian product
of Z2 with itself and who operation is component-wise addition modulo 2),
1.3 Cosets
Definition 1.15. Let H be a subgroup of a group G and a ∈ G. The left coset of H
containing a is the set aH = {ah : h ∈ H}. We call a the representative of aH and
use |aH| to denote the number of elements in the set aH. We similarly define the right
coset of H containing a by Ha = {ha : h ∈ H}. #
Note: Cosets are not necessarily subgroups of G. They are merely subsets. However, H
is always a subgroup. For additive groups, we write left cosets as a + H = {a + h : h ∈ H}.
We write right cosets in a similar manner.
Remark 1.16 (Coset properties). Let G be a group with subgroup H and elements
a, b. Then,
1. H = 1H = H1,
2. |aH| = |H|,
3. a ∈ aH,
4. aH = H if and only if a ∈ H,
9
1.3. Cosets
These properties also hold for right cosets (though the final part of property 5 becomes
ab−1 ∈ H). ✠
Coset property 6 suggests that it the left and right cosets of H containing a are not
necessarily equal, that is, aH 6= Ha. The following example shows us that it is indeed
the case.
Mon 19 Feb Example 1.18. Consider S3 , the group symmetric group of order 3, and its subgroup
H = {ε, τ }. We have
σH = {σ, στ } = {σ, τ σ 2 }
Hσ = {σ, τ σ}.
These two sets are not equal, so we have σH 6= Hσ. Now let’s consider the subgroup
K = {ε, σ, σ 2 }. The left and right cosets of K with representative τ are
τ K = {τ, τ σ, τ σ 2 }
Kτ = {τ, στ, σ 2 τ } = {τ, τ σ 2 , τ σ}.
So, in this case, τ K = Kτ . Coset property 6 says that this can happen only when
K = τ Kτ −1 . Let’s confirm this:
τ Kτ −1 = τ Kτ = {τ 2 , τ στ, τ σ 2 τ } = {ε, σ 2 , σ} = K. ♦
Question: Why are H and K subgroups of S3 ? Why do the following identities hold:
στ = τ σ 2 , τ σ = σ 2 τ , τ στ = σ 2 and τ σ 2 τ = σ?
Exercise 1.19. Compare the left and right cosets of K with representative τ σ. Are the
left and right cosets equal? Do the same for τ σ 2 .
Example 1.20. Consider the group Z under addition. For each n ∈ Z, we have the
subgroup hni generated by n. Namely, hni = nZ = {. . . , −2n, −n, 0, n, 2n, . . .}. Let’s
10
1.3. Cosets
look at the subgroup H = 3Z. Since Z is an additive group, the cosets of 3Z are written
additively, that is, m + 3Z for m ∈ Z. The left cosets of H are
0 + 3Z = {. . . , −6, −3, 0, 3, 6, . . .} = 3 + 3Z = . . . ,
1 + 3Z = {. . . , −5, −2, 1, 4, 7, . . .} = 4 + 3Z = . . . ,
2 + 3Z = {. . . , −4, −1, 2, 5, 8, . . .} = 5 + 3Z = . . . .
Thus the left cosets of 3Z are exactly the equivalence classes [0], [1] and [2] in Z3 . Since
Z is abelian, coset property 7 tells us that the left and right cosets of 3Z are equal, that
is 0 + 3Z = 3Z + 0 = [0], etc.
In fact, the same is true for any nZ where n is some integer. Thus, the left and right
cosets of nZ are precisely the equivalence classes [0], [1], . . . , [n − 1] in Zn . ♦
In the previous example, we saw how the cosets of a subgroup H with different rep-
resentatives can actually be the same sets. For example, 0 + 3Z = 3 + 3Z. This means
the number of cosets of H does not necessarily equal the number of elements in G. In
particular, we found that H = 3Z had 3 left cosets even though Z has infinitely many
elements.
Note: The number of left cosets of H always equals the number of right cosets of H.
This is true regardless of whether aH = Ha for all a ∈ G. Note that it is possible that
the number cosets (i.e. index) is infinite.
Thus, following Example 1.20, we can write |Z : 3Z| = 3. Furthermore, for any n ∈ Z,
we have |Z : nZ| = n.
Example 1.22. Consider the dihedral group D4 and its subgroup K = Z(D4 ) = {1, r2 }.
We can find the index of K in D4 by calculating all the left cosets then counting them.
The left cosets of K are
K = {1, r2 }
rK = {r, r3 } = r3 K
sK = {s, sr2 } = sr2 K
srK = {sr, sr3 } = sr3 K.
Thus, |D4 : K| = 4. ♦
11
1.4. Lagrange’s Theorem
Exercise 1.23. Compute the right cosets of K in D4 . How do they compare to the left
cosets?
aH G
a
bH
ah b
H d
1 h
cH dH
c
Note: While the index of a subgroup is defined for infinite groups G, the quotient |G|/|H|
does not make sense when G is infinite. Though notice that the theorem requires G to
be finite, so we run into no problems here.
Lagrange’s Theorem is quite helpful since if we know the order of a subgroup, we have
an idea what the order of the overall group could be. But what would the converse of this
theorem look like? Given the order of G, is there necessarily a subgroup for every divisor
of that order? For example, does a group of order 12 have to have a subgroup of order 6?
12
1.5. Normal subgroups
In general, no. The most simple example is the alternating group A4 which has order 12
but no subgroup of order 6.
One of the goals for this course is to find a partial converse of Lagrange’s Theorem.
Basically, we are going to look for some conditions we can add to the theorem in order to
make its converse true.
For now, let’s look at some important corollaries of Lagrange’s Theorem.
Corollary 1.26. If G is a finite group, the order of each element in G divides the order
of G.
In a way, the above corollary is even more powerful than Lagrange’s Theorem since
now we only need to know the order of a single element to get an idea of the order of G.
The next corollary shows us how knowing the order of a group can tell us about its
structure.
Example 1.29.
2. Following Example 1.18 and the subsequent exercise, the subgroup K = {ε, σ, σ 2 }
of S3 is normal. On the other hand, H = {ε, τ } is not normal in S3 .
13
1.5. Normal subgroups
For the above examples, we explicitly calculated all of the left and right cosets of a
given subgroup. However, the following theorem gives us a way to simplify the process.
Theorem 1.30 (Normality Test). Let H be a subgroup of a group G. Then the fol-
lowing are equivalent:
(i) H ◁ G,
This theorem tells us that in order to prove a subgroup H is normal, all we need to
do is show that one of the equivalent conditions is true.
Mon 22 Feb Example 1.31. Let R = Q, R or C. We know that SLn (R) is a subgroup of GLn (R). We
claim that SLn (R) is in fact normal in GLn (R). To prove this, we will use the Normality
Test. We will show that for all A ∈ GLn (R), we have A SLn (R)A−1 ⊆ SLn (R). To show
this subset inclusion, let B ∈ SLn (R). We will show ABA−1 ∈ SLn (R). The condition for
this to hold is that det(ABA−1 ) = 1. By properties of the determinant, we have
Thus ABA−1 ∈ SLn (R) for all A ∈ GLn (R), B ∈ SLn (R), and so SLn (R) ◁ GLn (R). ♦
2. If H ⊆ Z(G), then H ◁ G,
3. If H ◁ G and H ⊆ K, then H ◁ K. ✠
The following theorem also gives us a way to prove a subgroup is normal (in the very
special case where the subgroup has index 2).
14
1.6. Factor groups
Example 1.36. Consider the dihedral group D4 and the subgroups H = hsi = {1, s} and
K = hs, r2 i = {1, s, r2 , sr2 }. Clearly H ⊆ K. By Lagrange’s Theorem, |K : H| = 4/2 = 2,
so H ◁ K by Theorem 1.34. Moreover, |D4 : K| = 8/4 = 2, so K ◁ D4 . However, H is not
normal in D4 (consider rsr−1 = sr2 ∈
/ hsi). ♦
This example shows us that we need to be careful when considering nested subgroups
and normality. It’s very important that when we talk about a ‘normal subgroup’, we are
clear which group it is normal in. We’ll be returning to this idea in later chapters!
for a, b ∈ G, the set G/H is a group called the factor group (or quotient group) of G
by H. The identity of G/H is H and the inverse of gH ∈ G/H is g −1 H. #
Question: Does it matter whether we define the quotient group using left cosets instead
of right cosets?
Note: Factor groups are only defined for normal subgroups. So, whenever we see a group
of the form G/H, we immediately know that H must be normal in G.
Remark 1.38 (Factor group properties). Let G be a group with H ◁ G. Then the
following properties hold:
15
1.6. Factor groups
|G|
|G/H| = |G : H| = ,
|H|
Remark 1.40. For each normal subgroup H of G, there is a canonical (or natural)
projection map π between G and the factor group of H. It is defined by
π : G → G/H
g 7→ gH.
Example 1.43. Recall from Example 1.20 that, for each n ∈ Z+ , the cosets of nZ in
Z correspond to the equivalence classes [0], [1], . . . , [n − 1] from Zn . We also saw that
nZ ◁ Z. Thus, we can consider the factor group Z/nZ. In fact, through the isomorphism
φ : a + nZ 7→ [a], we obtain Z/nZ ∼
= Zn . ♦
Recall that the group Zn was obtained by taking the integers modulo n. Factor groups
generalise the “modulo” concept to arbitrary groups.
16
1.6. Factor groups
Finally, we see a theorem that shows us how the structure of a factor group can tell
us about the structure of the original group. It might seem a bit strange that we would
know more about the factor group than the original group, but we will see this in action
later in the course!
Hint: Let the generator of G/Z(G) be denoted by gH (where H = Z(G)). Show that
every element of G can be written as g n z for n ∈ Z and z ∈ Z(G).
Michael Kinyon
@ProfKinyon
I
Ah, group theory, where we can write H/H = {H} and have it
make perfect sense
7:18 pm · 02 Feb 22
retweet 4 🗩9 🧡 117
After going through this chapter, is it clear to you why H/H = {H}? 😊
17
1.6. Factor groups
physicist
@FinitePhysicist
I
6:20 pm · 06 Feb 22
retweet 45 🗩7 🧡 364
Évariste Galois was a French mathematician who lived in the 19th century. He made
many contributions to algebra and helped develop the concept of normal groups. He
notably died at age 20 in a duel. If you continue with algebra in honours, you’ll learn
more about his work!
18
The Isomorphism Theorems 2
Wed 28 Feb The Isomorphism Theorems, also called Noether’s Theorems after mathematician Emmy
Noether who discovered them, are very important and useful tools in algebra. They show
us how homomorphic groups relate to each other and the connection between groups and
their normal subgroups. We will use these theorems a lot during the course and they will
help us build a partial converse to Lagrange’s Theorem.
G/ ker φ ∼
= im φ.
Proof. We start by showing ker φ ◁ G. Take k ∈ ker φ, that is, φ(k) = 1G′ . Then, for all
g ∈ G,
Thus, gkg −1 ∈ ker φ for all g ∈ G, k ∈ ker φ. That is, g(ker φ)g −1 ⊆ ker φ for all g ∈ G.
By the Normality Test, this means that ker φ ◁ G.
19
2.1. The First Isomorphism Theorem
To check that Φ is well-defined, we need to make sure that it gives the same output
regardless of the coset representative. Suppose a ker φ = b ker φ. By the coset properties,
ab−1 ∈ ker φ. That is,
Multiplying both sides by φ(b) on the right, we obtain φ(b) = φ(a). Thus Φ(a ker φ) =
Φ(b ker φ) and so Φ is well-defined.
The mapping preserves operations since, for a, b ∈ G, we have
Φ((a ker φ)(b ker φ)) = Φ((ab) ker φ) = φ(ab) = φ(a)φ(b) = Φ(a ker φ)Φ(b ker φ).
Moreover, Φ is a surjection since for every g ′ ∈ im φ there is some g ∈ G such that φ(g) =
g ′ (by definition of the image) and so Φ(g ker φ) = g ′ . Now to prove injectivity, we take
a ker φ, b ker φ ∈ G/ ker φ and prove that Φ(a ker φ) = Φ(b ker φ) implies a ker φ = b ker φ.
Note that
The First Isomorphism Theorem tells us that each homomorphism from a group G
leads to a factor group with the same structure as the image of the homomorphism. For
example, if we have a group G missing a desirable property but we are able to construct
a homomorphism to a group with that desirable property, we can “give” our group that
20
2.1. The First Isomorphism Theorem
desirable property by taking the quotient of that group with respect to the kernel. Another
use is that it allows us to look at factor groups without delving into cosets.
The following corollary provides a special case of the First Isomorphism Theorem
where φ is not only a homomorphism but an epimorphism (that is, a surjective homomor-
phism).
G/ ker φ ∼
= G′ .
Note: Often this corollary is referred to as the First Isomorphism Theorem as well.
Question: What if the homomorphism in the First Isomorphism Theorem is both sur-
jective and injective? What does the theorem tell us then?
G G0
1
ker ϕ ϕ(1)
x
x ker ϕ ϕ(x)
a
ϕ(a) = ϕ(b)
a ker ϕ b
..
.
21
2.1. The First Isomorphism Theorem
which map to the same element of G. From this, we get a correspondence between the
cosets of a normal subgroup of G and the elements of G′ (or, in the case of φ not being
an epimorphism, the elements of im φ = φ(G)).
Thu 29 Feb Example 2.4. Let G be a group. Take x ∈ G, then the map σx : G → G defined
by σx (g) = xgx−1 is called an inner automorphism of G. The group of all inner
automorphisms of G under composition is denoted by Inn(G) = {σx : x ∈ G}.
Let’s construct a homomorphism between G and Inn(G). The most natural one would
be the map that takes an element x to its corresponding inner automorphism, σx , that is,
φ : x 7→ σx . However, we need to check that it is a homomorphism. That is, φ(x)φ(y) =
φ(xy) for all x, y ∈ G. More specifically, σx ◦ σy = σxy . We need to show the equality of
these two maps σx ◦ σy and σxy . To do so, we show that σx (σy (g)) = σxy (g) for all g ∈ G.
Notice that
22
2.2. The Second Isomorphism Theorem
The above exercise shows how using normal subgroups (that is, groups whose left and
right cosets are equal) can lead to us better understanding the group itself.
HK = {hk : h ∈ H, k ∈ K}. #
Note that the product HK is a set, not necessarily a subgroup itself. For example,
consider G = S3 with H = hτ i and K = hτ σ 2 i. Then
HK = {ε, τ, τ σ 2 , σ 2 }.
We see that |HK| = 4 and |S3 | = 6. Since 4 does not divide 6, it is impossible for HK
to be a subgroup of S3 , via Lagrange’s Theorem. What a nice application of Lagrange’s
Theorem!
Luckily, the following lemma tells us the conditions for HK to be a subgroup of G.
Note that the inverse of b indeed exists since b is in the group G. Let k3 = k1 k2−1 ∈ K and
h3 = h−1
2 ∈ H. Then we have ab
−1
= h1 k3 h3 . Since KH ⊆ HK, we can write k3 h3 = h4 k4
and so ab−1 = h1 h4 k4 ∈ HK. Since HK is closed under products and inverses, it must
be a subgroup.
23
2.2. The Second Isomorphism Theorem
(i) H ∩ K ◁ H,
(ii) HK is a subgroup of G,
(iii) K ◁ HK.
H/(H ∩ K) ∼
= (HK)/K.
Proof.
(ii) Using Lemma 2.8, it suffices to show that KH ⊆ HK. Take kh ∈ KH. Observe
that kh = (hh−1 )kh = h(h−1 kh). Note that since K is normal in G, we have that
gKg −1 ⊆ K for all g ∈ G, thus gkg −1 = k ′ for some k ′ ∈ K. Therefore, we can write
h−1 kh = k ′ ∈ K and so kh = hk ′ ∈ HK. Thus KH ⊆ KH and HK is a subgroup of G.
24
2.2. The Second Isomorphism Theorem
using the factor group operation. Thus, φ preserves the operation and so φ is a homo-
morphism. Notice that φ is surjective since for all cosets (hk)K in HK/K we can write
(hk)K = (hK)(kK) = (hK)(K) = hK = φ(h).
Now we look at the kernel of φ. The identity element in the factor group HK/K is
the coset K. So ker φ = {h ∈ H : hK = K}. By the properties of cosets, hK = K
is equivalent to saying h ∈ K. Thus, ker φ = H ∩ K. Thus, by the First Isomorphism
Theorem, H/(H ∩ K) ∼ = (HK)/K.
HK
K H
H ∩K
{1}
The Second Isomorphism Theorem says that each circled pair forms a factor group iso-
morphic to the other circled pair’s factor group.
Example 2.10. A group G is called metabelian if there exists an abelian normal sub-
group K such that the factor group G/K is abelian. All abelian groups are metabelian.
An example of a non-abelian metabelian group is S3 .
We can use the Second Isomorphism Theorem to prove that every subgroup of a
metabelian group is metabelian. Let H be a subgroup of G and K be the abelian normal
subgroup such that G/K is abelian. Applying the Second Isomorphism Theorem, we
obtain
H/(H ∩ K) ∼
= HK/K.
Note that the factor group HK/K is a subgroup of G/K since HK is a subgroup of G. By
assumption, G/K is abelian, thus the subgroup HK/K is also abelian. Thus, H/(H ∩ K)
25
2.3. The Third Isomorphism Theorem
Question: To prove that all abelian groups are metabelian what can we take the normal
subgroup K to be? Why is S3 metabelian?
(G/H)/(K/H) ∼
= G/K.
Proof. We will use the First Isomorphism Theorem to show this isomorphism of groups.
Let us construct an epimorphism φ : G/H → G/K with ker φ = K/H. A good place to
start is by looking at the form of the elements on either side of our mapping. The elements
of G/H have the form gH and the elements of G/K have the form gK. This suggests we
should try the map φ : gH 7→ gK. Note that φ is well-defined since if gH = g ′ H then
g −1 g ′ ∈ H ⊆ K and so φ(gH) = gK = g ′ K = φ(g ′ H). Taking gH, g ′ H ∈ G/H, we have
Moreover, for all gK ∈ G/K, we have φ(gH) = gK. Thus φ is a well-defined epimorphism
and so we can apply the First Isomorphism Theorem. But let’s first check the kernel.
Remember that the identity of a factor group G/K is K itself. Thus, gH ∈ ker φ if and
only if φ(gH) = gK = K. That is, g ∈ K. If g ∈ K then gH belongs to the factor group
K/H. Thus, ker φ = K/H. Finally, applying the First Isomorphism Theorem, we obtain
(G/H)/(K/H) ∼= G/K as desired.
Note that it follows from the fact that ker φ = K/H that K/H ◁ G/H.
Note: (G/H)/(K/H) is a group of cosets of K/H with representatives from G/H, i.e.
representatives are cosets of H in G. That is,
26
2.4. The Correspondence Theorem
Thankfully, applying the First Isomorphism Theorem gave us a way to avoid dealing with
this jumble of cosets.
A nice thing about the Third Isomorphism Theorem is it goes along with our intuition
of fraction cancellation (note the “cancellation” of the Hs on the left-hand side).
(Z/6Z)/(2Z/6Z) ∼
= Z/2Z.
Thus 2Z/6Z is a group consisting of 3 elements. Recall from Example 1.43 that Z/6Z ∼
= Z6
under the isomorphism φ : a + 6Z 7→ [a]. Under this same isomorphism, 2Z/6Z =
{6Z, 2 + 6Z, 4 + 6Z} is isomorphic to the cyclic subgroup h[2]i = {[0], [2], [4]} of Z6 . Then,
Z6 /h[2]i ∼
= Z2 . ♦
Note: It is true that 2Z/6Z ∼= Z3 since it is a cyclic group containing 3 elements. How-
ever, we cannot write Z6 /Z3 ∼
= Z2 . The reason for this is Z3 is not a subgroup of Z6 and
we cannot factor a group by something that is not a normal subgroup. The only reason
we can go from (Z/6Z)/(2Z/6Z) to Z6 /h[2]i is that Z/6Z ∼
= Z6 and 2Z/6Z ∼
= h[2]i under
the same isomorphism, meaning that the subgroup inclusion remains.
27
2.4. The Correspondence Theorem
Proof.
(i) Under the natural projection π : G → G/K, the image of H is H/K. Since π is a
homomorphism, H/K must be a subgroup of G/K.
(ii) Suppose H ⊆ H1 . Then π(H) ⊆ π(H1 ), that is, H/K ⊆ H1 /K. Now suppose
that H/K ⊆ H1 /K. Take h ∈ H, then hK ∈ H/K ⊆ H1 /K, so hK = h1 K for some
h1 ∈ H1 . Since K ⊆ H1 , we have h ∈ h1 K ⊆ h1 H1 = H1 . Thus H ⊆ H1 .
Mon 11 Mar Theorem 2.14 (Correspondence Theorem). Let G be a group and K ◁ G. Define a
map
(ii) Θ is a bijection.
(vi) Θ preserves products: if H1 H2 is a subgroup of G then (H1 /K)(H2 /K) = (H1 H2 )/K.
Proof.
Notice that for all k ∈ K we have k ∈ H since the identity kK = K of G/K must be
contained in the subgroup H, and then by the definition of H, k ∈ H. Therefore, K ⊆ H.
It follows that K ◁ H since H ⊆ G and K ◁ G.
28
2.4. The Correspondence Theorem
(ii) By (i), Θ is surjective, since for all subgroups H of G/K, we can define a subgroup
H of G with K ⊆ H such that Θ(H) = H/K = H. By Lemma 2.13, Θ is injective. If it
is not clear why, take Θ(H) = Θ(H1 ) and plug H and H1 into (ii) of the lemma, then by
symmetry we have injectivity. Thus Θ is a bijection.
h1 Hh−1
1 ⊆ H and by the Normality Test, H ◁ H1 .
Wed 13 Mar Example 2.15. Let G = hai be a cyclic group of order 12. Consider the subgroups
K1 = ha6 i and K2 = ha4 i.
The Fundamental Theorem of Finite Cyclic Groups says that the subgroups of a cyclic
group of order n are in one-to-one correspondence with the divisors of n. Moreover, each
subgroup is the cyclic group of order d, for some unique divisor of n. Thus the subgroups
of G are the cyclic subgroups generated by elements of order 1, 2, 3, 4, 6 or 12. That is,
ha12 i = {1}, ha6 i, ha4 i, ha2 i and hai = G, respectively. From this, we can construct the
subgroup lattice diagram for G. Note that had i ⊆ hak i if and only if k | d.
29
2.4. The Correspondence Theorem
G = hai
ha2 i ha3 i
K1 = ha4 i K2 = ha6 i
{1}
G/K1
ha2 i/K1
K1 /K1 = {K1 }
Similarly, we can inspect the subgroup structure of G/K2 . This means restriction our
attention to the upper square in the subgroup lattice diagram. By taking the quotient of
each subgroup by K2 , we obtain the subgroup lattice diagram for G/K2 :
G/K2
K2 /K2 = {K2 }
♦
30
2.4. The Correspondence Theorem
Example 2.16. Recall that K = Z(D4 ) = {1, r2 } is a normal subgroup of the dihedral
group D4 . The subgroups of D4 /K are D4 /K, {K} and
D4 /K
H1 H2 H3
{K}
H1 = {g ∈ D4 : gK ∈ H1 } = {1, r, r2 , r3 },
H2 = {g ∈ D4 : gK ∈ H2 } = {1, r2 , s, sr 2 },
H3 = {g ∈ D4 : gK ∈ H3 } = {1, r2 , sr, sr 3 }.
Putting this into a lattice diagram, we obtain the following subgroup structure for groups
of D4 containing K:
D4
H1 H2 H3
K
♦
31
2.4. The Correspondence Theorem
Before we return to our new favourite kind of groups (metabelian groups), we need a
lemma.
Thu 14 Mar Lemma 2.17. If H and K are normal subgroups of a group G, then HK = KH and
HK is a normal subgroup of G.
Proof. By the Second Isomorphism Theorem, since K ◁G, we have that HK is a subgroup
of G. By Lemma 2.8, KH ⊆ HK so now we need to show that HK ⊆ KH. Take
hk ∈ HK. Since K ◁ G, we have hkh−1 ∈ K. Thus hk = (hkh−1 )h ∈ KH. So
HK = KH.
To show that HK ◁ G, take hk ∈ HK and g ∈ G. We will show that ghkg −1 ∈ HK.
We have
Example 2.18. Using the Correspondence Theorem, we can return to the concept of
metabelian groups (Example 2.10) and look at whether the factor groups of a metabelian
group are metabelian.
Let G be a metabelian group where K ◁ G such that K and G/K are abelian. For
an arbitrary H ◁ G, we will prove that G/H is metabelian. To do so, we will find some
S ◁ G/H such that S and (G/H)/S are abelian. By the Correspondence Theorem, S
necessarily has the form N /H for some subgroup N of G with H ◁ N . So, let’s look for
this subgroup of G that contains H as a normal subgroup.
From the Second Isomorphism Theorem, H ◁ KH (here we have swapped the tradi-
tional roles of H and K, but that’s fine because H is normal now too). Moreover,
K/(K ∩ H) ∼
= KH/H.
By Lemma 2.17, KH ◁G. Then Correspondence Theorem (iv) tells us that KH/H ◁G/H.
We claim that KH/H is our special normal abelian subgroup S.
So, we need to show that KH/H and (G/H)/(KH/H) are abelian. By assumption,
K is a abelian, thus the factor group K/(K ∩ H) is also abelian and so is KH/H, by
the Second Isomorphism Theorem. Using two applications of the Third Isomorphism
Theorem, we have
(G/H)/(KH/H) ∼
= G/KH ∼
= (G/K)/(KH/K).
32
2.4. The Correspondence Theorem
Again, by assumption G/K is abelian and thus so are any of its factor groups. Therefore
(G/H)/(KH/H) is also abelian. Finally, we conclude that any factor group G/H of a
metabelian group G is also metabelian. ♦
We end this chapter with an illustration depicting the First Isomorphism Theorem:
33
p-Groups 3
Recall that one of our goals for this course is to find a partial converse to Lagrange’s
Theorem. p-Groups will help us do so. These are groups whose orders are prime powers.
However, before we look at p-groups, we need to look at a new kind of algebraic structure
which will allow us to examine properties of p-groups.
Throughout this chapter, we will be considering X to be a nonempty set.
Note: The set X is not necessarily a group itself. Also note the convention of using infix
notation g ∗ x instead of ∗(g, x) when apply a group action.
Example 3.2. The following are some simple examples of group actions:
34
3.1. Group actions
1. For any group G and set X, there is the trivial group action defined by g ∗ x = x
for all g ∈ G and x ∈ X.
Example 3.3. A group G can act on itself by the group action defined by g ∗ x = gxg −1 .
This action is called conjugation. Moreover, if H is a normal subgroup of G, then G
acts on H by conjugation. ♦
Exercise 3.4. Prove that the maps defined in Examples 3.2 and 3.3 are group actions.
S(x) = {g ∈ G : g ∗ x = x}. #
Note: You can think of a stable subset as a set that is closed under the group action.
Note that under this action, S itself becomes a G-set. Also note that subsets of X can be
called stable, whereas a stabiliser is a subgroup of G.
2. Symmetry: x ∼ y implies y ∼ x,
35
3.1. Group actions
The set [x] = {y ∈ X : x ∼ y} is called the equivalence class of x. All of the equivalence
classes of a set X under a given relation ∼ form a partition of X. An example of an
equivalence relation is the relation of congruency modulo some integer n on the set of
integers, Z. That is, the relation ∼ defined by a ∼ b if and only if a ≡ b mod n. Recall
that from this equivalence relation, we obtain the group Zn of integers modulo n, whose
elements are the equivalence classes under that relation.
Proof. We must show that ∼ satisfies the three equivalence relation conditions: reflexivity,
symmetry and transitivity. Take x, y, z ∈ X. We have 1 ∗ x = x, therefore x ∼ x and
so reflexivity holds. Suppose that x ∼ y, then y = g ∗ x for some g ∈ G. Consider
g −1 ∗ (g ∗ x) = g −1 ∗ y. However, we also have g −1 ∗ (g ∗ x) = (g −1 g) ∗ x = x. Thus,
x = g −1 ∗ y and so symmetry holds. Now suppose that x ∼ y and y ∼ z. Thus, y = g ∗ x
and z = g ′ ∗ y for some g, g ′ ∈ G. Thus z = g ′ ∗ (g ∗ x) = (g ′ g) ∗ x and so x ∼ z. Hence ∼
is an equivalence relation on X.
We say that G acts transitively on X if X has one unique orbit. Meaning for all
x, x′ ∈ X, we can find some g ∈ G such that g ∗ x = x′ . We call the action transitive.#
36
3.1. Group actions
g5 ∗ x5
x5
·
g5 ∗ x4 ·
·
g∗x
x4 x1 ∀g ∈ G
g3 ∗ x3 g1 ∗ x1 x
x3 x2
g2 ∗ x2
On the left-hand side, we have the orbit of x1 , which is G ∗ x1 = {x1 , x2 , x3 , . . .}, where
each xi is obtained by acting upon some other xj by some g ∈ G. On the right-hand side,
we have the orbit of an element x. This orbit consists of only one element, G ∗ x = {x}.
A singleton orbit arises when g ∗ x = x for all g ∈ G. Meaning no matter which g acts on
x, we still get x after doing the group action. We will later see that these singleton orbits
can be quite useful.
Example 3.8. Consider the subgroup G = {1, r2 } of the dihedral group D4 , which rep-
resents the symmetries of a square. Recall that r2 represents a rotation by 180 degrees.
We can label the vertices of the square as a, b, c and d, as shown in the following diagram:
1 = R0 r2 = R180
a b c d
d c b a
Then G acts on the set X = {a, b, c, d} by the action defined by g ∗ x = the vertex where
x has moved to under the rotation g. For example, when we apply the rotation r2 = R180
to the vertex a, it moves into the original position of vertex c. Thus r2 ∗ a = c. (As an
exercise, prove that this is indeed a group action).
Let’s consider the orbit of each vertex. To do so, let’s calculate g ∗x for each g ∈ G and
x ∈ X. By the definition of group action, we do not need to perform these calculations
for g = 1, since 1 ∗ x = x for all x ∈ X. We have r2 ∗ a = c, r2 ∗ b = d, r2 ∗ c = a and
37
3.1. Group actions
G ∗ a = G ∗ c = {a, c},
G ∗ b = G ∗ d = {b, d}.
Proof. To prove this, let us construct a bijection between elements in the orbit G ∗ x and
the left cosets of S(x). Take y ∈ G ∗ x, then there is some g ∈ G such that y = g ∗ x. We
define a map φ from G ∗ x to the set of left cosets of S(x) by φ(y) = gS(x). To show that
this is well-defined, suppose y = g1 ∗ x = g2 ∗ x for some g1 , g2 ∈ G. We need to show that
g1 S(x) = g2 S(x), that is, g1−1 g2 ∈ S(x). We have
Thus g1−1 g2 ∈ S(x) and so g1 S(x) = g2 S(x). Therefore the image of y under φ is in fact
unique and independent of the choice of representative g.
Now we will show that φ is injective. Let y, y ′ ∈ G ∗ x. Then y = g ∗ x and y ′ = g ′ ∗ x
for some g, g ′ ∈ G. Suppose that φ(y) = φ(y ′ ). Then gS(x) = g ′ S(x), meaning that
g −1 g ′ ∈ S(x) and so g −1 g ′ ∗ x = x. This simplifies to g −1 ∗ y ′ = x. We have y = g ∗ x and
so y = g ∗ (g −1 ∗ y ′ ) = y ′ . Thus, φ is injective.
To show that φ is surjective, let gS(x) be a left coset of S(x). Let y = g ∗ x ∈ G ∗ x.
Then φ(y) = gS(x). So, φ is surjective. Therefore, we have shown that there is a bijection
between the orbit G ∗ x and the set of left cosets of S(x). Thus, |G ∗ x| = |G : S(x)|.
The Orbit-Stabiliser Theorem actually gives us a way to relate the size of the set X
to the indices of its stabilisers.
Corollary 3.10 (Orbit Equation). Let G be a group acting on a finite set X. Let T
denote a set of representatives of the distinct equivalence classes of ∼ (equivalently, the
representatives of all distinct orbits). Then,
X X
|X| = |G ∗ x| = |G : S(x)|.
x∈T x∈T
Proof. The first equality holds since the equivalence classes of ∼ (or, orbits) partition X.
The second equality holds by the Orbit-Stabiliser Theorem.
38
3.1. Group actions
X g = {x ∈ X : g ∗ x = x}.
On the other hand, the fixed subset of G (or just fixed set) is the set of all x ∈ X that
are fixed by all elements of G. We write this as
Note: Fixed sets X g play a similar role to stabilisers S(x). One looks at the elements
of X fixed by a given g, the other looks at the elements of G which fix a given x.
Returning to the concept of orbits, we can divide the set of orbits of X into orbits
which are singletons and orbit which are not singletons. If an orbit of some x ∈ X is a
singleton, it means that the only element y ∈ X that satisfies y = g ∗ x, for some g ∈ G,
is the element y = x. That is, g ∗ x = x for all g ∈ G. This means that x belongs to the
fixed set Xf .
Because the orbits of a set X partition that set, we have another version of the Orbit
Equation, where the sum is split between orbits which are singletons and orbits which are
not singletons.
39
3.2. Conjugation
3.2 Conjugation
In Example 3.3, we saw that a group G can act on itself by conjugation, defined by
g ∗ x = gxg −1 . This is an important example of a group action. In later sections, we
will apply our results about group actions to the group action of conjugation to obtain
a partial converse to Lagrange’s Theorem. For now, let’s look at how our results about
group actions work for conjugation in particular.
Definition 3.14. Let G be a group and a, b ∈ G. We say that a and b are conjugate in
G if b = gag −1 for some g ∈ G. That is, a and b belong to the same orbit.
The conjugates of an element a are all elements of the form gag −1 for some g ∈ G.
That is, all elements in G ∗ a, the orbit of a. We call the orbit of a under conjugation the
conjugacy class of a. We write
Note: Since the conjugacy classes of G are precisely the orbits of G, they also form a
partition of G.
Example 3.15.
2. In any group G, class 1 = {1}. Moreover, class a = {a} if and only if a ∈ Z(G),
since gag −1 = a for all g ∈ G if and only if a ∈ Z(G). ♦
So, the stabiliser consists of all elements that commute with x. This has a special name,
the centraliser of x. It is a subgroup of G denoted by
CG (x) = {g ∈ G : gx = xg}.
40
3.2. Conjugation
Question: Why is the centraliser of x a subgroup of G? Why is the centre equal to the
intersection of the centralisers?
Theorem 3.17. Let G be a group and take a, b ∈ G. If a and b are conjugate, then their
orders are equal, that is, |a| = |b|.
Proof. Since a and b are conjugate, b = gag −1 for some g ∈ G. Recall that for each g ∈ G,
we have an inner automorphism σg : G → G defined by σg (x) = gxg −1 (Example 2.4).
Since a and b are conjugate, b = σg (a). Since σg is an automorphism, orders are preserved
and |a| = |b|.
This theorem means that we can determine conjugacy classes by looking at the orders
of the elements. Namely, the conjugacy class of a will be a subset of the set of elements
of G which have order |a|.
Example 3.18. Consider the dihedral group D3 = {1, r, r2 , s, sr, sr 2 } where |r| = 3,
|s| = 2 and rsr = s. The only elements of order 3 are r and r2 . By Theorem 3.17,
the only possible conjugate of r is r2 , that is class r ⊆ {r, r2 }. We confirm that r2 is
a conjugate of r by looking at srs−1 = r2 . Thus class r = {r, r2 }. Through similar
reasoning (now considering elements of order 2), we obtain class s ⊆ {s, sr, sr 2 }. Notice
that sr = rsr−1 and sr2 = r2 sr−2 . Thus class s = {s, sr, sr 2 }. Therefore, the distinct
conjugacy classes of D3 are class 1, class r and class s. ♦
Note: Using inner automorphisms, we have another way to express the conjugacy class
of an element a, namely,
41
3.2. Conjugation
In Remark 3.16, we saw that the stabiliser of x under conjugation is also known as
the centraliser CG (x) of x. We similarly have a special name for the stabiliser of a set H
under conjugation.
Remark 3.22. The stabiliser of a set H under conjugation is called the normaliser of
H in G, written as
Now take H to be a subgroup of G, not just a subset. Notice how NG (H) consists of all
elements g for which gHg −1 = H, one of the Normality Test conditions. It follows that
H ◁ NG (H). In fact, NG (H) is the largest subgroup of G that contains H as a normal
subgroup. Therefore, H ◁ G if and only if NG (H) = G. ✠
By the above remark, we can see how the conjugation group action on a subgroup
gives us a way to determine whether a group is normal by examining the stabiliser.
Now let’s return to conjugation by G on G and look at the fixed set of the action. We
have
So, the fixed set of G is the centre of G. Using this, we obtain a special version of the
Orbit Equation for the group action of conjugation called the Class Equation.
Theorem 3.23 (Class Equation). Let G be a finite group and class a1 , class a2 , . . .,
class an be the distinct nonsingleton conjugacy classes. Then
X
n
|G| = |Z(G)| + |G : CG (ai )|.
i=1
42
3.3. Cauchy’s Theorem
Notice how each of the terms on the right-hand side are divisors of |G|. The centre is a
subgroup of G thus its order must divide the order of G by Lagrange’s Theorem. Moreover,
each centraliser CG (ai ) is a subgroup of G, thus its index |G : CG (ai )| = |G|/|CG (ai )| also
divides |G|.
Example 3.24. In Example 3.18, we saw that the conjugacy classes of D3 were class 1,
class r and class s. The only singleton conjugacy class is class 1 = {1}. Therefore,
|Z(D3 )| = 1. On the other hand, |D3 : CD3 (r)| = |class r| = 2, by the Orbit-Stabiliser
Theorem. Similarly, |D3 : CD3 (s)| = |class s| = 3. Thus,
c ∗ (g1 , g2 , . . . , gp ) = (g2 , g3 , . . . , gp , g1 ).
Notice how we have only defined the action for the element c ∈ Cp . However, all other
elements of Cp can be written as cn where 1 ≤ n ≤ p. Thus, cn ∗ x = c ∗ (c ∗ (· · · ∗ x)) for
all x ∈ X, where we are applying the group action n times. Basically, this group action
shifts all entries in a p-tuple by the order of the group element.
43
3.3. Cauchy’s Theorem
Our goal is to use the Orbit Equation. For this, we need to understand the fixed set
and the nonsingleton orbits of the group action.
Note that Xf = {(g, g, . . . , g) : g ∈ G}, since the p-tuples where all entries are equal
will not change under this group action. Furthermore, we have X 6= Xf . This is because,
for a nonidentity element g ∈ G, the p-tuple (g, g −1 , 1, . . . , 1) belongs to X but not Xf .
Take x ∈ X with x ∈
/ Xf , that is, some x has a nonsingleton orbit. Then there is some
cm ∈ Cp such that cm ∗ x 6= x. Thus S(x) is a proper subgroup of Cp since cm ∈
/ S(x). By
Lagrange’s Theorem, |S(x)| must divide |Cp | = p. Since S(x) is a proper subgroup, we
must have |S(x)| = 1. Thus, |Cp : S(x)| = p/1 = p whenever x ∈ X has a nonsingleton
orbit.
Suppose that there are s nonsingleton orbits in X. Substituting |X| = |G|p−1 , n = s
and |Cp : S(xi )| = p into the Orbit Equation, we obtain
Notice that p divides both |G|p−1 and sp. Thus p must divide |Xf |. We know that |Xf | ≥ 1
since (1, 1, . . . , 1) ∈ X. Since p divides |Xf | and |Xf | ≥ 1, there are at least p elements in
|Xf |, including at least p − 1 tuples consisting of nonidentity elements. Thus, there are
at least p − 1 nonidentity elements g ∈ G such that g p = 1. Therefore, there are at least
p − 1 elements of G of order p. For any prime p, we have p − 1 ≥ 1. Thus G has at least
one element of order p.
Note: In the final step of the proof, the order of g must be p since g p = 1 implies |g| | p
so either |g| = 1 or p. However, g is not the identity, so we must have |g| = p.
The original proof of Cauchy’s Theorem used the Class Equation and induction on
the order of G. Give it a try if you’re interested!
Corollary 3.26. Let G be a finite group. If p is a prime that divides the order of G, then
G has a subgroup of order p.
Proof. By Cauchy’s Theorem, there exists some g ∈ G with order p. Then the cyclic
subgroup hgi of G has order p.
44
3.4. p-Groups
Proof. Since 2 and 3 are primes which divide |G|, Cauchy’s Theorem says that there
exist a, b ∈ G such that |a| = 2 and |b| = 3. Consider the cosets hbi and ahbi. We have
hbi = {1, b, b2 } and ahbi = {a, ab, ab2 }. This makes for a total of 6 distinct elements. Thus
G = {1, b, b2 , a, ab, ab2 }, since G has order 6.
Consider the product ba, which doesn’t appear in the above list of elements of G. So
ba must equal one of the previously listed elements. The only possibilities are that ba = ab
or ba = ab2 .
If ba = ab then (ab)2 = a2 b2 = b2 and (ab)3 = a3 b3 = a. So the order of ab must
instead be 6. This means that habi = G and so G is isomorphic to the cyclic group Z6 .
On the other hand, if ba = ab2 , then bab = a. From this, we see that the map
φ : G → D3 , defined by φ(a) = s and φ(b) = r, is an isomorphism. Thus G ∼
= D3 .
Question: Why are the elements of ahbi distinct? Why are the only two possibilities
ba = ab or ba = ab2 ? Why is φ an isomorphism?
The following corollary generalises the previous corollary to groups of order 2p for odd
prime p. The proof is not included, but it is very similar to the above proof. You won’t
be expected to know this result though.
Corollary 3.28. If p is an odd prime, then every group of order 2p is either cyclic or
dihedral.
3.4 p-Groups
Following Cauchy’s Theorem, we go a step further to look at groups whose orders are
only divisible by a single prime p.
Definition 3.29. Let p be a prime. A group G is called a p-group if |G| = pn for some
n ≥ 1. A subgroup H of a group G is called a p-subgroup if H is a p-group. #
Example 3.30.
1. For each prime p and positive integer k, the groups Zpk and Zp ×Zp ×· · ·×Zp = (Zp )k
are p-groups.
2. The dihedral group D4 is a 2-group since it has order 8 = 23 . On the other hand,
D3 is not a p-group since it has order 6, which is not a power of a prime.
45
3.4. p-Groups
Theorem 3.31. A finite nontrivial group G is a p-group if and only if the order of every
element of G is a power of p.
Proof. By Lagrange’s Theorem, if |G| is a power of p, the only possible orders of elements
are powers of p. On the other hand, suppose the order of every element of G is a power
6 p that also divides |G|. Then q divides |G|. By
of p. Suppose there is a prime q =
Cauchy’s Theorem, there is an element of order q. This contradicts the assumption that
every element’s order is a power of p. Thus |G| must be a power of p since p is the only
prime divisor of |G|.
Exercise 3.33. Prove the above theorem using the Class Equation.
Corollary 3.34. If G is a p-group with order pn , then for each k with 1 ≤ k ≤ n, G has
a normal subgroup of order pk .
46
3.4. p-Groups
Proof. Note that G is a p-group. Therefore, the centre of G is nontrivial and so has order
p or p2 . If |Z(G)| = p then Z(G) 6= G. However, the factor group G/Z(G) has prime
order p and so it is cyclic. By Theorem 1.45, G must be abelian and so Z(G) = G. This
is a contradiction so we must have |Z(G)| = p2 . Therefore G = Z(G) and G is abelian.
Now, if G has an element of order p2 , then G is cyclic and is thus isomorphic to Zp2 .
On the other hand, if there are no elements of order p2 , all nonidentity elements must have
order p. Take some nonidentity elements a, b ∈ G such that b ∈ / hai. Since |a| = p, we
have hai ∼
= Zp . Moreover, hai is a subgroup of ha, bi. Since b ∈
/ hai, this must be a strict
subgroup and so the only possibility for the order of ha, bi is p2 . That is, G = ha, bi. The
map φ(an , bm ) = an bm is an isomorphism from hai × hbi to G. However, hai ∼ = Zp ∼ = hbi
and so G ∼
= Zp × Zp .
Note: The theorem does not extend to groups of order p3 . For example, the group D4
has order 23 = 8 and is not abelian.
Wed 26 Apr Theorem 3.36. Let G be a p-group. Suppose G acts on a set X, then p divides |X|−|Xf |.
Equivalently, |X| ≡ |Xf | mod p.
Exercise 3.37. Prove the above theorem. Hint: Use the Orbit Equation.
Proof. Let X denote the set of left cosets of H in G. We define a group action on X by H
using left multiplication: h∗gH = hgH. Note that |Xf | ≥ 1 since h∗H = H for all h ∈ H
and so H ∈ Xf . By Theorem 3.36, p divides |X|−|Xf |. However, |X| = |G : H| = |G|/|H|.
So p also divides |X|. Thus |Xf | > 1. Notice that gH ∈ Xf if and only if hgH = gH for
all h ∈ H. This is equivalent to saying g −1 hg ∈ H. That is, g −1 Hg = H. However, this
is precisely the condition for g to be in NG (H). Thus gH ∈ Xf if and only if g ∈ NG (H).
Since |Xf | > 1, there is some element gH ∈ Xf , with g ∈ G, such that gH 6= H, that is,
g∈
/ H. Meaning that there exists some g ∈ G such that g ∈ NG (H) and g ∈
/ H. Thus we
conclude that NG (H) 6= H.
47
3.4. p-Groups
Proof. In the proof of the previous theorem, we see that g ∈ NG (H) if and only if
gH ∈ Xf . So Xf consists of left cosets with representatives from NG (H). This means
|Xf | = |NG (H) : H|. Recall that |X| = |G : H|. Furthermore, from Theorem 3.36, we
have |G : H| ≡ |NG (H) : H| mod p.
48
3.4. p-Groups
Coming soon...
Sylow Theorems
Finitely Generated Abelian Groups
49