Exercises
Exercises
Exercises
John Peloquin
Fall 2007
Abstract
Notes on Abstract Algebra by Dummit and Foote.
Exercises
Chapter 0
Section 0.3
We first note that the set on the right is well defined. For suppose a, b ∈ Z and a ≡ b
mod n, so a = b + nk for some k ∈ Z. If (b, n) = d > 1, write b = d k 1 and n = d k 2 ,
where k 1 , k 2 ∈ Z. Then
a = b + nk = d k 1 + (d k 2 )k = d (k 1 + k 2 k)
so d |a, and (a, n) > 1. Hence if (a, n) = 1, then (b, n) = 1. Since b was arbitrary, this
shows that the set is well defined. Now for the proof.
k 2 = 1 · k 2 = (a · b) · k 2 = b · (a · k 2 ) = ·k 2 = b · 0 = 0
It follows from the above proof that Z/nZ is partitioned into two subsets, one
consisting of all elements with multiplicative inverses and the other consisting of all
elements which are zero divisors. (Recall that a ∈ Z/nZ is called a zero divisor iff
a 6= 0 and there exists b ∈ Z/nZ, b 6= 0 such that ab = 0.)
1
Chapter 1
Section 1
Proof. Note that 0, 1 ∈ Z/nZ and 0 6= 1. If Z/nZ were a group under multiplication,
then there would exist a ∈ Z/nZ such that
1 = 0·a = 0·a = 0
—a contradiction.
E XERCISE 25. Let G be a group and suppose x 2 = 1 for all x ∈ G. Then G is abelian.
E XERCISE 31. Let G be a group of even order. Then G contains an element of order 2.
A = { x ∈ G | x = x −1 } and B = { x ∈ G | x 6= x −1 }
Then G = A ∪ B , where this union is disjoint, so |G| = |A| + |B |. Now |B | is even since
the elements of B come in pairs. Since |G| is even by hypothesis, |A| must also be
even. We know 1 ∈ A, hence |A| ≥ 2 and thus there exists x ∈ A, x 6= 1 as desired.
E XERCISE 32. Let G be a group and x ∈ G with |x| = n. Then 1, x, . . . , x n−1 are distinct.
This implies that if |x| = n, G contains at least n distinct elements, so |x| ≤ |G|.
Let H = 〈x〉 (see Exercise 26). If x m ∈ H , then write m = nk + r where 0 ≤ r < n.
We have
x m = x nk+r = (x n )k · x r = 1 · x r = x r
Hence H = {1, x, . . . , x n−1 }. In addition, these elements are distinct by this exercise.
Hence |H | = n. Since x was arbitrary, we have: the order of an element in a group is
the order of the cyclic subgroup generated by that element.
2
E XERCISE 36. Let G = {1, a, b, c} be a group with order 4 and no elements of order 4.
Then the group table for G is uniquely determined. In particular G is abelian.
Proof. Note that 1, a, b, c are all distinct. By Exercise 31, we may assume a 2 = 1. We
compute ab by considering possible cases and using the cancellation laws:
c 2 = (ba)(ab) = b(a 2 )b = b 2
Thus we must have b 2 = 1. This also implies c 2 = 1. Finally this gives bc = a = cb.
The group table for G is thus:
1 a b c
1 1 a b c
a a 1 c b
b b c 1 a
c c b a 1
The symmetry about the diagonal shows that G is abelian. (This also follows from
Exercise 25 and the fact that 12 = a 2 = b 2 = c 2 = 1.)
Section 2
D 6 = { 1, r, r 2 , s, sr, sr 2 }
so |sr 2 | = 2.
3
E XERCISE 2. Let x ∈ D 2n and suppose x is not a power of r . Then r x = xr −1 .
4
E XERCISE 15. Z/nZ = 〈 1 | n · 1 = 1 · · + 1} = 0 〉
| + ·{z
n times
Proof. Recall that Z/nZ = {0, . . . , n − 1}. Thus it is clear that 1 generates Z/nZ.
Suppose r , s ∈ Z/nZ and r = s. We claim that this relation can be derived from
n · 1 = 0. Indeed, write r − s = nk for k ∈ Z. Then we have
r − s = r − s = nk = k(n · 1) = k · 0 = 0
Section 3
σ: 1 7→ 3 2 7→ 4 3 7→ 5 4 7→ 2 5 7→ 1
τ: 1 7→ 5 2 7→ 3 3 7→ 2 4 7→ 4 5 7→ 1
σ = (135)(24) σ2 = (153)
τ = (15)(23) τ2 = 1
στ = (2534) τσ = (1243)
τ2 σ = σ
σ = (12345678) σ2 = (1357)(2468)
σ3 = (14725836) σ4 = (15)(26)(37)(48)
σ5 = (16385274) σ6 = (1753)(2864)
σ7 = (18765432) σ8 = 1
5
E XERCISE 10. Let σ = (a 0 a 2 · · · a m−1 ) be an m-cycle. Then |σ| = m.
Proof. Note that σi (a j ) = a j +i mod m (by induction on i ). Since all of the a j are
distinct in the cycle notation, it follows that σi (a j ) 6= a j for 0 < i < m, so σi 6= 1
for 0 < i < m. But σm (a j ) = a j for all j —that is, σm = 1. Hence |σ| = m.
σi (k) = k + i mod m (0 ≤ k ≤ m − 1)
E XERCISE 15. Let σ ∈ S n and let m be the least common multiple of the lengths of
the cycles in the cycle decomposition of σ. Then |σ| = m.
σm = (σ1 · · · σk )m = σm m
1 · · · σk
by Exercise 1.24. Now |σi | is just the length of σi , for all 1 ≤ i ≤ k, by Exercise 10
above. Hence since m is a common multiple of these lengths by hypothesis, we
have σm i
= 1 for 1 ≤ i ≤ k. Thus σm = 1.
On the other hand, since m is the least common multiple, if 0 < j < m, then there
j
exists some cycle, say σi , such that j is not a multiple of |σi |. We claim that σi 6= 1.
Indeed, let p = |σi | and write j = pq + r where 0 ≤ r < p. Then
j pq+r p
σi = σi = (σi )q σri = 1q σri = σri
j
If σi = 1, then σri = 1, so r = 0 and j is multiple of |σi |—contradicting our hypothesis.
j
Hence σi 6= 1 as claimed.
Now by the pairwise disjointness of the cycles in σ, this implies that σ 6= 1. Hence
we have shown that m is the least positive integer with σm = 1, that is, |σ| = m.
6
One might conjecture, based on the above exercise, that for any group G and any
arbitrary elements x, y ∈ G with |x| = m and |y| = n, that |x y| = lcm(m, n). But this is
false. For example, let G = S 3 , and set x = (13) and y = (12). Then |x| = 2 and |y| = 2,
so lcm(|x|, |y|) = 2. But x y = (123), so |x y| = 3 6= 2.
The key in the case of cycle decompositions is that the cycles are disjoint, so the
action of one factor cannot ‘alter’ the action of another factor.
E XERCISE 18. We find all numbers n such that S 5 contains an element of order n.
Note that every permutation in S 5 must take one of the following forms:
S 3 = 〈 σ, τ | σ3 = 1, τ2 = 1, στ = τσ−1 〉
Section 4
This allows for six possible distinct configurations, hence |GL 2 (F2 )| = 6.
7
E XERCISE 5. Let F be a field. Then GL n (F ) is finite iff F is finite.
Proof. If F is finite, then there are only finitely many n × n matrices over F , hence in
particular only finitely many such matrices with inverses, so GL n (F ) is finite.
If F is infinite, then there exist infinitely many matrices in GL n (F ) of the form
a
.. = a · I n×n
.
a
with a ∈ F .
AB (1,1) = 1 · 0 + · · · + 1 · 0 + 1 · 1 = 1 6= 0 = 0 · 1 + 0 · 0 + · · · + 0 · 0 + 1 · 0 = B A (1,1)
Section 6
Thus ϕ(x −1 ) = ϕ(x)−1 by uniqueness of inverses, so the result holds for n = −1.
For n > 0, proceed by induction. If the result holds for n ≥ 0, then
8
And hence the result holds for n + 1. Finally, if the result holds for n < 0, then
by Exercise 1. On the other hand, if 0 < m < n, then x m 6= 1G , and hence ϕ(x)n 6= 1H
by injectivity of ϕ. So |ϕ(x)| = n as desired.
For each n > 0, let
Proof. Suppose G is abelian. Let a 0 , b 0 ∈ H and set a = ϕ−1 (a 0 ) and b = ϕ−1 (b 0 ). Then
Hence H is abelian.
Interchanging the roles of G and H establishes the converse.
E XERCISE 4. R× ∼
6 C×
=
E XERCISE 5. R ∼
6 Q
=
9
E XERCISE 6. Q ∼
6 Z
=
Proof. Suppose ϕ : Q ∼
= Z. Set m = ϕ(1) and fix n > 0. Then
E XERCISE 9. D 24 ∼
6 S4
=
Proof. We know that D 24 has an element of order 12. However, by looking at cycle
decompositions (cf. Exercise 3.18), we see that S 4 only has elements of orders m =
1, 2, 3, 4. Thus the result follows from Exercise 4.
Φ : σ 7→ π−1 ◦ σ ◦ π
Note that Φ is well defined since if σ ∈ S ∆ , then Φ(σ) ∈ S Ω (clearly Φ(σ) : Ω → Ω, and
Φ(σ) is a bijection since the composite of bijections is a bijection).
Note that Φ is a homomorphism since if σ, τ ∈ S ∆ , then
ϕ(ab −1 ) = ϕ(a)ϕ(b)−1 = 1
10
E XERCISE 19. Let G = { z ∈ C | (∃n > 0)[ z n = 1 ] } and fix k > 1. Then the map z 7→ z k is
a surjective homomorphism from G to G, but is not an isomorphism.
E XERCISE 25. Let n > 0 and set θ = 2π/n. Recall from linear algebra that the matrices
cos θ − sin θ
µ ¶ µ ¶
0 1
R= S=
sin θ cos θ 1 0
r k1 s k2 · · · r km s km+1 7→ R k1 S k2 · · · R km S km+1
Note that the map is not surjective since there are more elements in GL 2 (R) than
those generated by R, S. We note that the map ϕ provides a matrix representation of
the dihedral group (cf. permutation representations).
Section 7
Proof. For x, y ∈ R, 0 · (x, y) = (x + 0y, y) = (x, y), so the identity property holds. Also,
if r, s ∈ R,
(r + s) · (x, y) = (x + (r + s)y, y)
= (x + r y + s y, y)
= (x + s y + r y, y)
= r · (x + s y, y) = r · (s · (x, y))
Note that this action is just a horizontal shear. The geometric fact that horizontal
shears accumulate additively is reflected in the fact that the action is given by R. (To
compare, see that the action is not given by R× , for example.)
11
E XERCISE 8( A ). Let A be a nonempty set. Fix k ∈ Z, 0 < k ≤ |A|, and let A be the set
of all subsets of A of cardinality k. Then S A acts on A by
σ · {a 1 , . . . , a k } 7→ {σ(a 1 ), . . . , σ(a k )}
E XERCISE 11. Let D 8 act on the vertices of a square. Label the vertices consecutively
1, 2, 3, 4, starting from a fixed vertex and moving in a clockwise manner. We write
cycle decompositions of the elements in S 4 corresponding to the elements of D 8
under this action.
1 7→ 1
r 7→ (1234)
r 2 7→ (13)(24)
r 3 7→ (1423)
s 7→ (24)
sr 7→ (24)(1234) = (14)(23)
sr 2 7→ (24)(13)(24) = (13)
sr 3 7→ (24)(1423) = (12)(34)
Note that this map is injective, hence we have obtained a faithful representation
of D 8 in S 4 . In other words, D 8 is isomorphic to a subgroup of S 4 (though note this
must be a proper subgroup, since |D 8 | = 8 and |S 4 | = 24). Compare this with the
faithful (matrix) representation obtained in Exercise 6.25.
E XERCISE 12. Let n ∈ Z+ be even, and label the vertices of a regular n-gon as above
with labels from V = {1, . . . , n}. Set
Then D 2n acts on P .
Proof. For convenience we identify V with the set of vertices of the n-gon, and
P with the set of pairs of opposite vertices.
12
Let φ : D 2n → S V be the representation of D 2n given by the usual action of D 2n
on V . Define ψ : D 2n → S P as follows: for α ∈ D 2n , set
ψ(α) : P → P
{a, b} 7→ {φ(α)(a), φ(α)(b)}
Note that ψ(α) is a well defined permutation of P , hence ψ is well defined. Also ψ is
a homomorphism since φ is. Thus ψ defines an action of D 2n on P .
1. If n = 2, then direct computation reveals that the action is trivial (that is, the
kernel is all of D 4 ).
ker ψ = {1, r 2 , s, sr 2 }
where r is the usual clockwise rotation through θ = π/2, and s the reflection in
the line passing through opposing vertices 1 and 3.
g · x = g xg −1
for all g , x ∈ G.
(g 0 g 2 ) · x = (g 0 g 2 )x(g 0 g 2 )−1
= (g 0 g 2 )x(g 2−1 g 0−1 )
= g 0 (g 2 xg 2−1 )g 0−1
= g 0 (g 2 · x)g 0−1 = g 0 · (g 2 · x)
13
E XERCISE 17. Let G be a group and fix g ∈ G. Then the operation of conjugation by g
is an automorphism of G.
The previous two exercises show that conjugation in a group G is a very special group
action, since from each group element we obtain not merely a permutation but an
automorphism of G.
As an example, consider G = GL 2 (R), which for convenience we identify with the
group of invertible linear transformations on R2 . Each element g ∈ G gives rise to
a transformation of R2 . Now if x ∈ G, and we view x as an operation on the pre-
transformed space, then the transformation naturally corresponding to x in the g -
transformed space is simply g xg −1 . Thus the operation of g on all of G produces
a ‘translated version’ of G, where every x ∈ G is ‘relativized’ to g . This ‘translated
version’ of G is structurally identical to G.
b = 1 · b = (h −1 h)b = h −1 (hb) = h −1 a
a = hb = h(g c) = (hg )c
hence a ∼ c. Since a, b, c were arbitrary, this shows that reflexivity, symmetry, and
transitivity hold for ∼, hence ∼ is an equivalence relation on A.
Proof. Let H act on G by the left regular action. By the previous exercise, and the
assumption that G is finite, we may partition G into finitely many orbits under H :
G = O(x 1 ) ∪ · · · ∪ O(x n )
But note that O(x i ) is merely the image of H under the permutation (in G) of right
multiplication by x i . Hence |O(x i )| = |H | for all 1 ≤ i ≤ n. Since G is partitioned,
|G| = n · |H |, so the result holds.
14
E XERCISE 23. Let G be the group of rotations of a cube, and let G act on the set
A = {p 1 , p 2 , p 3 } of three pairs of opposite faces of the cube. Then the kernel of this
action is the subgroup H of G generated by {1, r 1 , r 2 , r 3 }, where r i is the rotation by π
radians about the line passing through the centers of the two faces in p i .
Proof. It is immediate that the kernel contains all of H , since it contains {1, r 1 , r 2 , r 3 }.
To prove the converse, let ρ be any rotation in the kernel, and consider the effect
of ρ on p 1 . Either ρ swaps the two faces in p 1 or it does not. In the non-swapping
case, either ρ = 1 or ρ = r 1 . In the swapping case, it is easy to see that either ρ = r 2 or
ρ = r 3 . Hence H contains the kernel.
This shows that H is the kernel.
In fact we see from the above proof that H = {1, r 1 , r 2 , r 3 }. In addition we see that the
action’s representation of G is not faithful.
Note that the above proof could be simplified by just arguing directly that any
element in the kernel is one of 1, r 1 , r 2 , or r 3 . We leave the above proof as it is to
illustrate the technique of using a subgroup generated by particular elements; in
more complex cases, it might not be as easy to obtain an explicit list of the elements
in the kernel.
Chapter 2
Section 1
E XERCISE 1.
E XERCISE 2.
(a) Fix n ≥ 3, set G = S n and let S be the set of 2-cycles in G. Then S 6≤ G. Proof:
Note that (12), (13) ∈ S, but (13)(12) = (123) 6∈ S.
p p
(e) Let G = R and let S = { r ∈ G | r 2 ∈ Q }. Then S 6≤ G. Proof: Note 2, 3 ∈ S, but
p p p p p
( 2 + 3)2 = 2 + 2 2 3 + 3 = 5 + 6
p p p
which is not rational since 6 is not rational. Hence ( 2 + 3) 6∈ S.
15
E XERCISE 4. Let G = Z and let N be the subset of nonnegative integers. Then N is
infinite and closed under the group operation, but N is not a subgroup since it does
not contain additive inverses.
Alternately let G = Q× . Then Z is an infinite subset and is closed under the group
operation, but it is not a subgroup since it does not contain multiplicative inverses.
E XERCISE 5. Let G be a group suppose n = |G| > 2. Then G does not have a subgroup
of order n − 1.
E XERCISE 7. Fix n ∈ Z, n > 1 and let G = Z × (Z/nZ). Then G is an abelian group, and
it is clear that the torsion subgroup of G is { (0, a) | a ∈ Z/nZ }.
The subset of G consisting of the elements of infinite order (together with 0) is
not a subgroup. Indeed, (−1, 0) and (1, 1) have infinite order, but (−1, 0)+(1, 1) = (0, 1)
has finite order (and is not 0).
16
E XERCISE 10( B ). Let G be a group and H be an arbitrary nonempty collection of
subgroups of G. Then H ≤ G.
T
E XERCISE 13. Let 0 < H ≤ Q and suppose 1/x ∈ H for all nonzero x ∈ H . Then H = Q.
This result shows, in particular, that Q is the smallest additive subgroup containing Z
and closed under multiplicative inverses.
x −1 ∈ H ⊆ H . It follows that H ≤ G.
S S
Section 2
17
In the above proof, we see an instance of the general result that if G is a group, S ⊆ G,
T
and G = 〈S〉, then Z (G) = {CG (s) | s ∈ S }. In particular, |Z (G)| divides |CG (s)| for all
s ∈ S (by Lagrange’s Theorem).
The following result is also useful in the computation of centralizers: if G is a
group and g ∈ G, then CG (g ) = CG (〈g 〉). In particular, if |g | = n, and (m, n) = 1, then
CG (g m ) = CG (g ) since 〈g m 〉 = 〈g 〉.
E XERCISE 13. Fix n ∈ Z, n > 0 and let R be the set (ring) of all polynomials in x 1 , . . . , x n
over Z. For each σ ∈ S n , define the map σ : R → R by
From this exercise we see that the permutations in S n naturally induce permutations
on polynomials in n variables.
Section 3
It follows from these results that for G = 〈x〉, if |G| = n, then 〈x b 〉 = G iff (b, n) = 1;
and if |G| = ∞, then 〈x b 〉 = G iff b = ±1 (cf. Proposition 6).
We also obtain the following specific case (cf. p. 59): if n ≥ 1 and 1 ≤ a, b ≤ n,
then 〈a〉 ≤ 〈b〉 in Z/nZ iff (b, n)|(a, n).
18
E XERCISE 1. We determine the subgroup structure (lattice) of Z45 = 〈x〉.
Note that 45 = 3 · 3 · 5, with positive divisors 1, 3, 5, 9, 15, 45. Hence by Theorem 7,
the subgroups of Z45 are
〈1〉 〈x 15 〉 〈x 9 〉 〈x 5 〉 〈x 3 〉 〈x〉
By the result above, we see that 〈x a 〉 ≤ 〈x b 〉 iff (b, 45)|(a, 45) for 1 ≤ a, b ≤ 45.
Proof. Note that each element in Z2 ×Z2 has order at most 2, hence the group cannot
be cyclic (for this requires an element of order 4).
If (a, b) generates Z2 × Z, then we must have b = ±1 by Proposition 6. Clearly
we cannot have a = 1, for 〈(1, b)〉 = 1 × Z < Z2 × Z. But we also cannot have a = x,
because for example (1, 1) 6∈ 〈(x, b)〉. Since these possibilities are exhaustive, there is
no generator for the group.
Let (a, b) ∈ Z × Z. Define
(
(1, −1) if a and b have the same sign
x=
(1, 1) if not
Then x 6∈ 〈(a, b)〉. Since (a, b) was arbitrary, there is no generator for the group.
—a contradiction.
19
E XERCISE 16. Let G be a group and suppose x, y ∈ G commute. Set m = |x| and
n = |y|. Then |x y| divides lcm(m, n).
Proof. Indeed, write α = lcm(m, n). Since m|α and n|α, x α = 1 and y α = 1. Then
since x and y commute, (x y)α = x α · y α = 1. The result follows from Proposition 3.
Note that |x y| need not equal lcm(m, n). Indeed, let G = S 3 and set x = y = (12).
Then m = n = 2, so lcm(m, n) = 2, but x y = 1 so |x y| = 1 < 2.
Also the result need not hold if x, y do not commute. Again in S 3 , set x = (12) and
y = (13). Then lcm(m, n) = 2. But x y = (132), so |x y| = 3, which does not divide 2.
Proof. Fix n ≥ 3 and let G = (Z/2n Z)× . We claim that x = 2n−1 −1 and y = 2n−1 +1 are
two distinct elements of order 2 in G, from which it follows that G cannot be cyclic
by Theorem 7(3).
Note that 0 < x < y < 2n . Thus if x = 1 (in G), then 2n−1 − 1 = 1 (in Z), which
implies 2n−2 = 1—a contradiction since n ≥ 3. On the other hand,
E XERCISE 25. Let G be a cyclic group of order n and let k be relatively prime to n.
Then the map ϕ : g 7→ g k is an automorphism of G.
20
However, in general the map is still surjective. Indeed, for G arbitrary and g ∈ G, let
H = 〈g 〉 and let m = |g |. Then by Lagrange’s Theorem, m|n, hence (k, m) = 1. Note
that ϕ|H : H → H . Hence by the above exercise there exists h ∈ H such that ϕ(h) = g .
On the other hand, if (k, n) > 1, then by Cauchy’s Theorem (Section 3.2) there
exists x ∈ G with |x| = d > 1 and d |k. Then x 6= 1 but x k = 1, so the the map g 7→ g k
is not injective, and hence not surjective since G is finite. Thus we obtain: for G of
order n, the map g 7→ g k is surjective iff (k, n) = 1.
σa : Z n → Z n by σa (x) = x a (x ∈ Zn )
Thus σ = σa .
(d) (Z/nZ)× ∼
= Aut(Zn )
Note that this result shows that Aut(Zn ) is an abelian group of order ϕ(n). By
Theorem 4, this result actually gives the structure of the automorphism group
for any finite cyclic group of order n.
Section 4
21
E XERCISE 3. Suppose H ≤ G is abelian. Then 〈H , Z (G)〉 is abelian.
Proof. Let x, y ∈ 〈H , Z (G)〉. Then by Proposition 9, x and y are both finite products
of elements all of which commute with each other. Hence x y = y x.
Note this result does not hold in general for the centralizer of an abelian subgroup.
Let G = S 5 and H = 〈(12)〉. Then H is abelian, and (34), (45) ⊆ C (H ), but 〈H ,C (H )〉 is
nonabelian since (34)(45) = (345) 6= (354) = (45)(34).
Proof. Note that {1, x, y, x y} ⊆ 〈x, y〉. It is easily verified using cancellation that these
elements are distinct, hence |〈x, y〉| ≥ 4. But then by Lagrange’s Theorem, we must
have |〈x, y〉| = 6, that is, 〈x, y〉 = S 3 .
E XERCISE 6. Let x = (12) and y = (12)(34) (in S 4 ). Then 〈x, y〉 is noncyclic of order 4.
Proof. It is easily verified that 〈x, y〉 = {1, x, y, x y}, where these elements are distinct.
Hence |〈x, y〉| = 4. But there are no elements in this group of order 4, hence it is
noncyclic (and isomorphic to the Klein Four group by Exercise 1.1.36).
E XERCISE 14.
(c) Let H ≤ Q be finitely generated. Then H is cyclic.
Proof. This follows from (c) and the fact that Q is not cyclic: for q = m/n > 0
arbitrary, note that
m m
−kq < 0 < < ≤ kq (k ∈ Z, k > 0)
n +1 n
Hence 〈q〉 < Q. Since q was arbitrary, Q is noncyclic.
22
E XERCISE 16.
(a) Let H be a proper subgroup of a finite group G. Then there exists a maximal
subgroup of G containing H .
Proof. Since G is finitely generated, this follows from the proof of Exercise 17
by setting H0 = H .
(c) Let G = 〈x〉 be cyclic of order n. Then H ≤ G is maximal iff H = 〈x p 〉 for some
prime p dividing n.
E XERCISE 17. Let G be a nontrivial finitely generated group. Then G has a maximal
subgroup.
H R g ⇐⇒ H < 〈H ∪ {g }〉 < G
Let F be a choice function for R. Note that for H ∈ G , H 6∈ dom(F ) iff either H = G or
H is maximal in G.
Let Ω denote the class of all ordinals and construct a recursion H : Ω → G as
follows:
1 if α = 0
〈H ∪ {F (H )}〉 if α = β+ and H ∈ dom(F )
β β β
Hα =
Hβ if α = β and Hβ 6∈ dom(F )
+
if α is a limit ordinal
S
H
β<α β
It is easily verified by induction that α < β implies Hα ⊆ Hβ for all α, β ∈ Ω, and hence
by induction again that Hα ∈ G for all α ∈ Ω.
23
Let γ be the Hartogs number of G . Since H |γ : γ → G cannot be injective, there
must exist ordinals δ < ε < γ such that Hδ = Hε . Note that δ < δ+ ≤ ε, hence
H δ ⊆ H δ+ ⊆ H ε ⊆ H δ
so Hδ = Hδ+ . By the definition of our recursion above, this implies that Hδ 6∈ dom(F ),
which means either Hδ = G or Hδ is maximal in G.
To rule out the first case, we prove (by induction) that Hα < G for all α ∈ Ω. Fix α
and suppose the result holds for all β < α. If α = 0, the result holds since H0 = 1
and 1 < G by hypothesis. If α = β+ , then both possible cases in our recursion yield
Hα < G. Finally, if α is a limit ordinal, suppose towards a contradiction that
[
G = Hα = Hβ
β<α
Note that we present this direct proof instead of using Zorn’s Lemma because it well
illustrates both the principle behind Zorn’s Lemma as well as the natural way in
which the ordinals can be used to extend the counting process beyond the finite
(cf. the proof of this result for a finite group G).
E XERCISE 19.
Section 5
24
E XERCISE 6. We compute the centralizers of each element in the following groups G:
C (1) = C (r 2 ) = D 8
C (r ) = C (r 3 ) = 〈r 〉
C (s) = C (r 2 s) = 〈s, r 2 〉
C (r s) = C (r 3 s) = 〈r s, r 2 〉
(c) Let G = S 3 = 〈(12), (123)〉 = {1, (12), (13), (23), (123), (132)}. We have
C (1) = S 3
C ((12)) = 〈(12)〉 C ((13)) = 〈(13)〉 C ((23)) = 〈(23)〉
C ((123)) = C ((132)) = 〈(123)〉
N (1) = N (〈123〉) = N (S 3 ) = S 3
N (〈(12)〉) = 〈(12)〉 N (〈(13)〉) = 〈(13)〉 N (〈(23)〉) = 〈(23)〉
Chapter 3
Section 1
25
E XERCISE 1. Let ϕ : G → H be a homomorphism and let E ≤ H . Then ϕ−1 (E ) ≤ G. If
E is normal, so is ϕ−1 (E ).
ϕ(x y −1 ) = ϕ(x)ϕ(y −1 )
= ϕ(x)ϕ(y)−1
= x 0 y 0−1 ∈ E
It follows that F ≤ G.
Suppose now that E E H . Let g ∈ G be arbitrary and x ∈ F . Then
g α N = ϕ(g α ) = ϕ(g )α = (g N )α
Proof. Trivial.
26
E XERCISE 14. Let G = Q/Z.
r + Z = (q + m) + Z
= q + (m + Z)
= q +Z
(b) All elements in G have finite order, but there exist elements of arbitrarily large
order.
n · (q + Z) = nq + Z = m + Z = Z
Hence |q + Z| ≤ n.
Note if (m, n) = 1, then |q + Z| = n. Thus for m = 1, |q + Z| → ∞ as n → ∞.
Proof. By (b), every element in Q/Z has finite order. Suppose α + Z ∈ R/Z and
|α + Z| = n. Then
n · (α + Z) = nα + Z = Z
In particular, nα = m for some m ∈ Z, so α = m/n ∈ Q. Thus α + Z ∈ Q/Z. This
establishes the claim.
Proof. Define ϕ : G → C× by
m 2mπi
ϕ: + Z 7→ e n
n
It is verified that this a well defined, injective, surjective homomorphism.
27
E XERCISE 15. Let G be a divisible abelian group and N < G. Then G/N is divisible.
Proof. Let g ∈ G be arbitrary. Then since g ∈ 〈S〉, there exist s 1 , . . . , s n ∈ S such that
ε ε
g = s 11 · · · s nn (εi = ±1)
ε ε
g = s 11 · · · s nn
= s 1 ε1 · · · s n ε n ∈ 〈 S 〉
28
E XERCISE 25.
(a) Let G be a group and N ≤ G. Then N E G iff g N g −1 ⊆ N for all g ∈ G.
Proof. The forward direction is trivial. If g N g −1 ⊆ N for all g ∈ G, then for all
g ∈ G we have
N = (g g −1 )N (g g −1 ) = g (g −1 N g )g −1 ⊆ g N g −1
Hence g N g −1 ⊆ N . But note that g does not normalize N , since not every
element in N has an even integer in entry (1, 2).
E XERCISE 26.
(a),(b) Recall that conjugation is an automorphism.
(c) Let G be a group, S ⊆ G, and N = 〈S〉. Then N E G iff g Sg −1 ⊆ N for all g ∈ G.
g xg −1 = (g s 1 g −1 )ε1 · · · (g s n g −1 )εn ∈ N
(d) It follows immediately that if N = 〈x〉 is cyclic in G, then N is normal iff for all
g ∈ G, g xg −1 = x k for some k ∈ Z.
(e) Let G be a group, n ∈ Z, n > 0, and let N be the subgroup generated by all
elements in G of order n. Then N is normal.
29
E XERCISE 29. Let G = 〈T 〉. Suppose S ⊆ G, N = 〈S〉, and N is finite. Then N E G iff
t St −1 ⊆ N for all t ∈ T .
D 8 /1 ∼= D8
2 ∼
D 8 /〈r 〉 = V4 (or D 4 )
D 8 /〈r 〉 ∼
= Z2 (or D 2 )
D 8 /〈s, r 2 〉 ∼
= Z2
2 ∼
D 8 /〈r s, r 〉 = Z2
ϕ : D 2n → D 2k
s α r β 7→ σα ρ β (α, β ∈ Z)
Proof. Write Z = Z (G). Suppose G/Z is cyclic and fix g ∈ G such that G/Z = 〈g Z 〉.
Then for all x ∈ G, there exists α ∈ Z such that x ∈ (g Z )α = g α Z . In other words, for
all x ∈ G there exists α ∈ Z and z ∈ Z such that x = g α z.
Let x, y ∈ G be arbitrary, and write x = g α w and y = g β z where α, β ∈ Z and
w, z ∈ Z . Then we have
x y = (g α w)(g β z) = g α g β w z = g β g α zw = (g β z)(g α w) = y x
since w, z commute with all elements, and powers of g commute with each other.
Since x, y were arbitrary, it follows that G is abelian.
Another way to state this result is that G/Z (G) cannot be nontrivially cyclic (since,
when G is abelian, Z (G) = G so G/Z (G) ∼= 1).
30
E XERCISE 40. Let G be a group, N E G, and write G = G/N . For x, y ∈ G, x, y com-
mute in G iff x −1 y −1 x y ∈ N . (Note x −1 y −1 x y is called the commutator of x and y.)
Proof.
x y = y x ⇐⇒ x −1 y −1 x y = 1
⇐⇒ x −1 y −1 x y = 1
⇐⇒ x −1 y −1 x y ∈ N
It follows from Exercise 26(c) that N E G. Since N contains all commutator elements
in G, it follows from Exercise 40 that G/N is abelian.
This exercise shows that the quotient construction is in a certain sense ‘unique’. It is
the only construction on a partition of a group whose operation may be induced by
the operation of the group elements in the above manner.
Section 2
E XERCISE 2. We prove that the lattice of subgroups for S 3 in Section 2.5 is correct.
Proof. We first claim that the lattice is exhaustive. Indeed, if H is any subgroup of S 3 ,
then by Lagrange’s Theorem we know the order of H is one of 1, 2, 3, or 6. Cases
|H | = 1 and |H | = 6 are trivial. If |H | = 2, then H is cyclic and generated by one of the
three elements of order 2, that is, H is one of
Similarly if |H | = 3, then H must be cyclic and equal to 〈(123)〉. This shows that the
lattice is exhaustive.
31
Clearly any two of the distinct subgroups of order 2 have trivial intersection. Also,
by Lagrange’s Theorem, each of these subgroups have trivial intersection with the
subgroup of order 3 since (2, 3) = 1. This shows that the intersections on the lattice
are correct.
Finally, it is easily verified that any two distinct nontrivial subgroups join to all
of S 3 . Hence the joins on the lattice are correct.
Proof. Let H = G/Z (G) and suppose Z (G) > 1. If Z (G) = G, then trivially G is abelian.
If Z (G) 6= G, then by Lagrange’s Theorem, |Z (G)| is either p or q, and thus the same
is true of |H |. But then H is cyclic, so again G is abelian by Exercise 3.1.36.
E XERCISE 6. Suppose H ≤ G, g ∈ G, and the right coset H g is equal to some left coset
of H . Then g H = H g .
E XERCISE 8. Suppose H and K are finite subgroups of G whose orders are relatively
prime. Then H ∩ K = 1
P = { (x 1 , . . . , x p ) | x i ∈ G ∧ x 1 · · · x p = 1 }
32
(e) We claim that each equivalence class has order 1 or p. To see this, let E be
an arbitrary equivalence class and note that the set of cyclic permutations of
elements of E forms a group H of order p under composition. Moreover, this
group acts on E . Now fix a tuple t ∈ E and consider the stabilizer S of t in H .
By Lagrange’s Theorem, since p is prime, either S = 1 or S = H . If S = H , then
E has order 1. If S = 1, then distinct cyclic permutations give rise to distinct
images of t . So H is in bijective correspondence with E , and E has order p.
It follows now that |G|p−1 = k +pd , where k is the number of 1-element classes
and d is the number of p-element classes.
(f) Since p divides |G|p−1 , p divides k + pd , and so p divides k. We know k > 0
since {(1, . . . , 1)} is a 1-element equivalence class. But this means k > 1, so there
must exist a nonidentity element x ∈ G with x p = 1, and |x| = p.
Note that this theorem is a generalization of Exercise 1.1.31, and the proof is in some
ways a generalization of the proof used in that exercise.
π : k H 7→ g k H (k ∈ K )
|G : H |
|G : K | =
|K : H |
this suggests that the quotient G/K might be isomorphic to the quotient (G/H )/(K /H ).
Indeed, this is so (see the Third Isomorphism Theorem).
33
E XERCISE 15. Let G = S n , fix i ∈ {1, . . . , n}, and let G i be the stabilizer of i . Then
Gi ∼
= S n−1 .
Gi ∼
= σG i σ = G n ∼
−1
= S n−1
where the first isomorphism is obtained by conjugation and the second is obtained
by restriction.
E XERCISE 21. Neither Q nor Q/Z have proper subgroups of finite index.
Proof. Recall that if G is divisible, then G cannot be finite (Exercise 2.4.19(b)). Also
if G is divisible and H is a proper normal subgroup of G, then G/H is also divisible
(Exercise 3.1.15). Thus a divisible group cannot have proper normal subgroups of
finite index.
The result now follows from the fact that Q is abelian and divisible (Exercise
2.4.19(a)).
Proof. Consider G = (Z/nZ)× . Recall a ∈ G since (a, n) = 1, and |G| = ϕ(n). Hence by
Lagrange’s Theorem, |a| divides ϕ(n), so a ϕ(n) = 1—that is, a ϕ(n) ≡ 1 mod n.
100
E XERCISE 23. We compute the last two digits of 33 using Euler’s Theorem. Note
100
that this is equivalent to computing 33 mod 100. By Euler’s Theorem, note that
100
3ϕ(100) mod 100 = 1, hence we need only compute 33 mod ϕ(100) mod 100.
100
Write r = 3 mod ϕ(100). Note ϕ(100) = ϕ(2 · 52 ) = 2(1) · 5(4) = 40. Thus to
2
compute r , we may use Euler’s Theorem again to note r = 3100 mod ϕ(40) mod 40.
Now ϕ(40) = ϕ(23 · 5) = 22 (1) · 1(4) = 16, and 100 mod 16 = 4. So r = 34 mod 40 = 1.
Thus our answer is 31 mod 100 = 3–that is, the last two digits are 03.
Section 3
A = A/N = { aN | a ∈ A }
Then the map A 7→ A forms a bijective correspondence between the set of subgroups
of G containing N and the set of subgroups of G = G/N . Moreover, this bijection
preserves lattice structure: for A, B ≤ G with N ≤ A ∩ B ,
(1) A ≤ B iff A ≤ B
(2) if A ≤ B then |B : A| = |B : A|
(3) 〈A, B 〉 = 〈 A, B 〉
(4) A ∩ B = A ∩ B
(5) A E G iff A E G
34
Proof. Let π : G → G be the natural projection. Then for N ≤ A ≤ G,
A = π[A] ≤ π[G] = G
f : B /A → (B /N )/(A/N )
b A 7→ bN (A/N )
By definition
bN (A/N ) = { (bN )(aN ) | a ∈ A } = { baN | a ∈ A }
Note f is well-defined. Indeed, if b A = b 0 A, then b = b 0 a 0 for some a 0 ∈ A, hence
baN = b 0 (a 0 a)N for all a ∈ A, and so bN (A/N ) ⊆ b 0 N (A/N ). The converse follows by
symmetry.
Also f is injective. If bN (A/N ) = b 0 N (A/N ), then bN = b 0 aN ⊆ b 0 A for some
a ∈ A. In particular b ∈ b A ∩ b 0 A, hence b A = b 0 A.
Clearly f is surjective. Hence f is a bijection and |B : A| = |B : A|.
p = |G : H | = |G : H K ||H K : H |
(A × B )/(C × D) ∼
= (A/C ) × (B /D)
ϕ : (A × B ) → (A/C ) × (B /D)
(a, b) 7→ (aC , bD)
35
E XERCISE 7. Suppose G = M N where M E G and N E G. Then
G/(M ∩ N ) ∼
= (G/M ) × (G/N )
Proof. Define the map
ϕ : G → (G/M ) × (G/N )
g 7→ (g M , g N )
Note that this result can be seen visually by looking at the (partial) subgroup lattice.
The ‘diamond’ formed by G/(M ∩ N ) is reobtained by taking the cross product of the
‘lines’ G/M and G/N .
Section 4
Proof. By definition |G| > 1. Fix x ∈ G, x 6= 1 and set N = 〈x〉. Since G is abelian,
any subgroup is normal and so 1 < N E G. But this implies N = G since G is simple,
so G is cyclic. By Theorem 2.3.7, G cannot have infinite or finite composite order
because it does not have any nontrivial proper subgroups. Therefore G ∼ = Z p for
some prime p.
36
E XERCISE 4. Let G be a finite abelian group and suppose n > 0 divides the order of G.
Then G has a subgroup of order n.
|H | = |N ||H : N | = |N ||H | = pm = n
1 = H0 E · · · E Hn = H
Hi +1 /Hi = A/(A ∩ B ) ∼
= AB /B ≤ G i +1 /G i
1 = G0 E · · · E Gn = G
E XERCISE 6,7 ( J ORDAN -H ÖLDER T HEOREM PART I). Let G be a nontrivial finite group.
Then G has a composition series.
37
Isomorphism Theorem, G i +1 /G i ∼
= G i +1 /G i , and hence is simple, for 0 ≤ i < k. Thus
we have a composition series for G given by
1 = N0 E · · · E N j = N = G 0 E · · · E G k = G
Note that this proof also shows that if H E G, there exists a composition series for G
containing H as a term. Indeed, if H = 1 or H = G, this is trivial. Otherwise, combine
a composition series for H with the preimage of a composition series for G/H under
the natural projection to obtain a composition series for G which includes H .
An analogous note also holds for solvable groups.
Section 5
(a 1 · · · a m ) = (a 1 a m ) · · · (a 1 a 2 )
Indeed, this follows by induction. For m = 1 the result is trivial. If the result holds for
fixed m ≥ 1, note that
Thus the result holds for m + 1. By induction the result holds for all m.
Now for any σ ∈ S n , σ may be written as a product of cycles (cycle decomposition
algorithm). Since each of these cycles may be written as a product of transpositions
by the above, the result holds for σ.
E XERCISE 3. S n = 〈 (i i + 1) | 1 ≤ i < n 〉
Proof. Let S denote the set on the right. Claim that for any 1 ≤ i < j ≤ n, (i j ) ∈ S.
Indeed, this follows by induction on j . For j = 1 the result is trivial. If the result
holds for j < n, note that (i j + 1) = (i j )( j j + 1)(i j ) ∈ S.
Since any element of S n is a product of transpositions, the result follows.
Proof. Let S denote the set on the right. By assumption (12) ∈ S. Suppose 1 < i < n.
Set ρ = (1 · · · n)i −1 . Then (i i + 1) = ρ(12)ρ −1 ∈ S.
The result now follows from Exercise 3.
38
E XERCISE 7. The group of rigid motions (rotations) of a regular tetrahedron in 3-
space is isomorphic to A 4 .
Proof. Let G be the group of rigid motions. By letting G act on numbered faces of the
tetrahedron, we obtain a permutation representation ϕ : G → S 4 . Note |ϕ[G]| ≤ 12
since any rigid motion must take face 1 to one of four possible faces, and at any one
of these must put it in one of three possible positions. And |ϕ[G]| ≥ 12 since each
one of these configurations is distinct and is realized by a rigid motion in G. Hence
we must have |ϕ[G]| = 12.
We claim that ϕ[G] ⊆ A 4 . First note that G is generated by two rotations ρ 1 and ρ 2
through 2π/3 radians about axes passing through any two distinct vertices and the
centers of their respective opposing faces. Therefore since ϕ is a homomorphism, it
suffices to observe that ϕ(ρ 1 ) and ϕ(ρ 2 ) can be written as even permutations.
Since ϕ[G] ⊆ A 4 and |ϕ[G]| = 12 = |A 4 |, ϕ[G] = A 4 and so ϕ : G ∼
= A4.
N OTE .Recall that groups of rigid motions of 2-dimensional regular n-gons in 3-space
are not in general isomorphic to subgroups of A n (e.g. D 6 ∼
= S 3 ). Intuitively it makes
sense that the groups of rigid motions of a 3-dimensional object in 3-space (acting
on a set of n elements) might be, since it does not contain reflections.
39
By the lattice, the factors in this chain are cyclic of orders 2 or 3, and hence are
abelian simple. It follows that this is a composition series for A 4 and moreover that
A 4 is solvable.
Chapter 4
Section 1
hence g xg −1 ∈ G b . Thus gG a g −1 ⊆ G b .
Conversely, if x ∈ G b , then by the above (with g −1 ), g −1 xg ∈ G a , hence
x = g (g −1 xg )g −1 ∈ gG a g −1
40
Proof. Each set in B is nonempty since B is nonempty. Fix b ∈ B . Since G is
transitive, for any a ∈ A there exists σ ∈ G such that a = σ(b) ∈ σ[B ]. Hence
A = B. We claim that σ[B ] 6= τ[B ] implies σ[B ] ∩ τ[B ] = ;. Indeed, suppose
S
(c) S 4 is primitive on A = {1, 2, 3, 4}, but D 8 is not primitive on the vertices of the
square.
41
E XERCISE 9 (O RBIT-S TABILIZER ). Let G act transitively on the finite set A and H E G.
Let O 1 , . . . ,O r denote the distinct orbits of H on A.
(a) G acts transitively on the orbits where g ·O i = { g · a | a ∈ O i }, and all the orbits
have the same cardinality.
Proof. The first claim follows from the Orbit-Stabilizer Theorem (Proposition 2)
since H a = H ∩ G a . The second claim follows similarly by noting that HG a is
precisely the stabilizer of O 1 in the action of G on the orbits.
Section 2
Proof. Let G act on the coset space A = G/H , let π : G → S A be the permutation
representation and set K = ker π. Then K ≤ H , K E G, and by the First Isomorphism
Theorem,
|G : K | = |G/K | = |ϕ[G]| ≤ |S A | = n!
42
E XERCISE 11. Let G be a finite group and πG : G → SG denote the left regular action
of G. If x ∈ G with |x| = n and |G| = mn, then π(x) is a product of m n-cycles. Hence
π(x) is an odd permutation iff |x| is even and |G|/|x| is odd.
E XERCISE 12. Let G be a finite group and π : G → SG be the left regular action. If
π[G] contains an odd permutation, then G has a subgroup of index 2.
E XERCISE 13. Let G be a group with |G| = 2k where k is odd. Then G has a subgroup
of index 2.
Proof. Since G has even order, there exists x ∈ G of order 2. But then by Exercise 11,
πG (x) is an odd permutation (where πG denotes the left regular representation of G).
Hence G has a subgroup of index 2 by Exercise 12.
Note that the previous three exercises illustrate a recurring use of group actions:
pulling information about elements and subgroups of S n back into arbitrary groups.
In this case we have used information about the existence of A n in S n to obtain
information about more general groups.
E XERCISE 14. Let |G| = n with n composite. Suppose for each positive k|n there
exists a subgroup of G of order k. Then G is not simple.
Proof. Note n > 1. Let p be the smallest prime dividing n and write k = n/p. Then
1 < k < n. By assumption there exists a subgroup K with |K | = k, and 1 < K < G.
By Lagrange’s Theorem, |G : K | = p, so K E G by Corollary 5. It follows that G is not
simple.
Section 3
43
E XERCISE 2. We find conjugacy class decompositions.
N OTE .For the following exercise we first observe the following result: If A and B are
groups, then the conjugacy classes in A × B are precisely the direct products of the
conjugacy classes of A with the conjugacy classes of B .
(a) Let G = Z2 ×S 3 . Write Z2 = {±1}. Then the conjugacy classes of Z2 are {1} and {−1}.
The conjugacy classes in S 3 are {1}, {(12), (13), (23)}, and {(123), (132)}. Hence
by the above note, the conjugacy classes in G are
(b) Let G = S 3 × S 3 . The conjugacy classes in G are just the direct products of the
conjugacy classes in S 3 with each other.
(c) Let G = Z3 × A 4 . The conjugacy classes in G are just the direct products of the
conjugacy classes in Z3 ({1}, {x}, and {x 2 }) with the conjugacy classes in A 4 .
44
E XERCISE 5. Let G be finite and suppose |G : Z (G)| = n. Then every conjugacy class
in G has at most n elements.
E XERCISE 6. Let G be a nonabelian group of order 15. Then there is only one possible
class equation for G.
Proof. We must have Z (G) = 1, since otherwise G/Z (G) is of prime order and hence
cyclic, contradicting that G is nonabelian. Thus there is only the trivial singleton
conjugacy class. The other conjugacy classes must have orders 3 or 5. It is easy to
check that the only possibility is 15 = 1 + 3 + 3 + 3 + 5.
E XERCISE 13. Let G be a finite group with two conjugacy classes. Then G ∼
= Z2 .
Proof. Let n = |G|. Then we must have n = 1+k where k is the order of the nontrivial
conjugacy class. But k|n by the Orbit-Stabilizer Theorem. Hence there exists m with
mk = n = k + 1. Now m ≥ 1, but since k ≥ 1, m 6= 1. Also 2k > 1, so 3k > k + 1, and
hence we have m < 3. This leaves m = 2, in which case k = 1, n = 2, and G ∼
= Z2 .
F (X ) = { a ∈ A | (∀σ ∈ X )(σ(a) = a) }
Proof. Trivially 1 ∈ D. Suppose σ, τ ∈ D, so that M (σ) and M (τ) are finite. Note that
M (τ−1 ) = τ[M (τ)] = M (τ), and M (στ−1 ) ⊆ M (σ) ∪ M (τ−1 ), hence M (στ−1 ) is finite
and στ−1 ∈ D. This shows that D ≤ S A .
Suppose now σ ∈ D and τ ∈ S A . Note that M (τστ−1 ) = τ[M (σ)], hence
E XERCISE 18. Let A be a nonempty set, H ≤ S A , and F (H ) the fixed set of H . Then
NS A (H ) stabilizes F (H ) and its complement M (H ).
45
E XERCISE 19. Let G be a group, H E G, and K be a conjugacy class of G such that
K ⊆ H . Then K is a union of k conjugacy classes of equal size in H , where for any
x ∈ K , k = |G : HCG (x)|.
Proof. This follows from the proof of Exercise 1.9. Note that under conjugation,
G acts transitively on K , hence K is partitioned into H -orbits of equal cardinality.
Now G acts transitively on the set of H -orbits. Hence k is just the size of the one
orbit under this action. The stabilizer of this action is precisely HCG (x), hence by
the Orbit-Stabilizer Theorem we have k = |G : HCG (x)| as desired.
Recall A n E S n and |S n | = 2·|A n | (hence A n is maximal for n ≥ 2). It follows from this
result that any conjugacy class in S n which is contained in A n consists either of one
conjugacy class in A n or else two conjugacy classes in A n of equal size.
E XERCISE 22. Suppose n ≥ 3 is odd. Then the set of all n-cycles forms two conjugacy
classes of equal size in A n .
Proof. Since n is odd, the n-cycles are even and hence are contained in A n . Define
σ = (123 · · · n), τ = (12), and σ∗ = τστ−1 = (213 · · · n). We claim that σ and σ∗ are
not conjugate in A n . Indeed, if ρ ∈ S n and ρσρ −1 = σ∗ , then by Proposition 10 and
the fact that the only alternate representations of the cycle for σ∗ are obtained by
cyclically permuting its elements, we must have ρ = (σ∗ )k τ for some k, which is an
odd permutation. Hence there is not one conjugacy class of n-cycles in A n , and the
desired result follows from Exercise 19.
E XERCISE 25. Let G = GL 2 (C) and let H be the subgroup of upper triangular matrices
in G. Then every element in G is conjugate to some element in H .
E XERCISE 29. Let G be a group of order p α . Then G has a subgroup of order p β for
all 0 ≤ β ≤ α.
|H | = |H ||Z | = p ρ p δ = p β−δ p δ = p β
46
Thus in either case there exists a subgroup of G of order p β , as desired.
E XERCISE 30. Let G be a group of odd order. Then for all nonidentity elements x ∈ G,
x is not conjugate to x −1 .
E XERCISE 33. Let σ ∈ S n , let m 1 , . . . , m s be the distinct cycle lengths in the (complete)
cycle decomposition of σ, and for each 1 ≤ i ≤ s let k i be the number of m i -cycles in
the decomposition (so n = is=1 k i m i ). Then the number of conjugates of σ is
P
s
Y 1
n! k
i =1 ki ! mi i
Proof. Recall that the number of conjugates of σ is precisely the number of elements
in S n with the same cycle type as σ, hence we need only count the number of ways
to form a permutation in S n with this cycle type.
We give a combinatorial argument. To obtain any such permutation we may start
with any linear arrangement of the numbers 1, . . . , n and then partition the numbers
into cycles, by drawing parentheses, according to the cycle type of σ. Now there
are n! distinct linear arrangements of the numbers 1, . . . , n. However, for any such
permutation formed, for each 1 ≤ i ≤ s, the m i -cycles may be permuted amongst
each other in any of k i ! ways, and the elements in each m i -cycle may be cyclically
permuted in m i ways, while still preserving identity and form. Thus each linear
k
arrangement of 1, . . . , n is in an equivalence class of is=1 k i ! m i i (including itself),
Q
all of which give the same permutation. Thus the number of distinct permutations
with the cycle type of σ is as above.
(a) C SG (π(G)) = λ(G) and C SG (λ(G)) = π(G). In particular, σg and τh commute for
all g , h ∈ G.
1 Note that under our definitions, the right regular representation λ of G is not associated with a right
action of G, but a left action h · x = xh −1 , for otherwise λ would fail to be a homomorphism under our
usual left-based functional notation.
47
Proof. We prove the first equality; the other follows by symmetry. Note that
λ(G) ⊆ C SG (π(G)) since for g , h ∈ G,
(τh σg )(x) = τh (σg (x)) = (g x)h −1 = g (xh −1 ) = σg (τh (x)) = (σg τh )(x)
Note this exercise shows in particular that the centralizer of the left or right regular
permutation representation of a group is isomorphic to the group itself.
Section 4
ϕ∗ (x) = (σϕg σ−1 )(x) = σ(g σ−1 (x)g −1 ) = σ(g )xσ(g )−1
48
E XERCISE 2. Let G be an abelian group, |G| = pq where p 6= q are prime. Then G is
cyclic.
Proof. By Cauchy’s Theorem, there exist x, y ∈ G with |x| = p and |y| = q. Set z = x y.
Note that
z pq = (x y)pq = x pq y pq = (x p )q (y q )p = 1 · 1 = 1
Hence |z| divides pq. We claim |z| = pq.
If |z| divides p, then 1 = (x y)p = x p · y p = y p . But this implies q|p, which is a
contradiction since p 6= q. Similarly |z| does not divide q since p does not divide q.
Hence |z| = pq as claimed, and G is cyclic.
Recall from the text that if |G| = pq with p ≤ q and p does not divide q − 1, then G is
abelian. Combining this with the above result we obtain the following classification
theorem: if |G| = pq with p < q and p does not divide q − 1, then G is cyclic.
As an example, if |G| = 15 = 3 · 5, then G is cyclic.
E XERCISE 3. |Aut(D 8 )| ≤ 8
E XERCISE 5. Aut(D 8 ) ∼
= D8
This exercise illustrates that if inner automorphisms do not exhaust Aut(G), then it
is useful to look at larger normalizers of G.
Proof. Trivial since automorphisms preserve order and map subgroups to subgroups.
49
E XERCISE 8. Suppose H ≤ K ≤ G.
Proof. If σ ∈ Aut(G), then σ(K ) = K since K char G, hence σ|K ∈ Aut(K ). But
then σ(H ) = σ|K (H ) = H since H char K . Since σ was arbitrary, H char G.
Proof. Let H ≤ K . Note K E G (by checking generators). Recall that Aut(K ) consists
of maps r 7→ r a where 0 < a < n and (a, n) = 1. Hence H is mapped to itself under any
automorphism of K , that is, H char K . It follows from Exercise 8(a) that H E G.
E XERCISE 13. If |G| = 203 and G has a normal subgroup of order 7, then G is abelian
(and hence cyclic).
Proof. Note |G| = 7 · 29, where 7 and 29 are both prime. Let H E G with |H | = 7. We
know G/C (H ) is isomorphic to a subgroup of Aut(H ), hence |G|/|C (H )| must divide
|Aut(H )| = ϕ(7) = 6. This is only possible if C (H ) = G, that is, H ≤ Z (G). This implies
G/Z is of prime order and hence cyclic, so G is abelian. It follows that G is cyclic by
Exercise 2.
E XERCISE 15. The groups (Z/5Z)× , (Z/9Z)× , and (Z/18Z)× are each cyclic.
Proof. The groups are each abelian. The latter two are of orders ϕ(9) = 3(3 − 1) = 6
and ϕ(18) = (2 − 1) · 3(3 − 1) = 6, respectively, hence cyclic by Exercise 2. The first
group is cyclic and generated by the element 2.
Section 5
50
N OTE .The proof of Sylow’s Theorem illustrates several recurring techniques:
1. It uses induction to prove (part of) an existence claim (cf. Cauchy’s Theorem
for abelian groups, Exercise 3.29, etc.).
2. It makes use of the Class Equation, combining information about the order
of G and an assumption about the order of Z (G) to obtain information about
subgroups (centralizers) in G (cf. an example in Section 4.4).
3. To count Sylow subgroups, it restricts the source of a group action to a sub-
group about which we have more information, which provides additional in-
formation about the original orbits (cf. Exercise 1.9).
4. In using Sylow p-subgroups computationally, it utilizes their maximality in
order among p-subgroups.
These techniques are important to keep in mind.
E XERCISE 5. Let G = D 2n . If p is an odd prime and P ∈ Sylp (G), then P is cyclic and
normal.
Proof. Let P ∈ Sylp (G). Let r, s be the usual generators of G and set H = 〈r 〉. Recall
|H | = n, hence (since p 6= 2) any Sylow p-subgroup of H is also a Sylow p-subgroup
of G. By Sylow’s Theorem then, all Sylow p-subgroups of G are contained in H , so
in particular P ≤ H . It follows that P is cyclic, and P char H E G (see Exercise 4.7),
hence P E G.
E XERCISE 13. Let G be a group with |G| = 56. Then G has a normal Sylow p-subgroup
for some prime p dividing |G|.
Proof. Note |G| = 23 ·7. By Sylow’s Theorem, n 7 ≡ 1 (7) and n 7 |8, thus n 7 = 1 or n 7 = 8.
If n 7 = 1, then there is a unique normal Sylow 7-subgroup of G. If n 7 = 8, then there
are 8 · (7 − 1) = 48 elements of order 7 in G (since distinct Sylow 7-subgroups must
have trivial intersection), leaving 8 remaining elements. These 8 elements must be
precisely those in a unique, and hence normal, Sylow 2-subgroup of G.
51
E XERCISE 16. Let G be a group with |G| = pqr , where p, q, r are prime and p < q < r .
Then G has a normal Sylow subgroup for one of p, q, r .
E XERCISE 17. Let G be a group with |G| = 105. Then there exists a normal Sylow
5-subgroup and a normal Sylow 7-subgroup in G.
Note this proof utilizes the technique of constructing an intermediate cyclic normal
subgroup (cf. the example on groups of order 30).
E XERCISE 22. Let G be a group and |G| = 132. Then G is not simple.
52
E XERCISE 24. Let G be a group, |G| = 231. Then Z (G) contains a Sylow 11-subgroup
and there exists a normal Sylow 7-subgroup.
Proof. Note |G| = 3 · 7 · 11. By Sylow’s Theorem, n 7 ≡ 1 (7) and n 7 |33, leaving only the
possibility n 7 = 1 and a normal Sylow 7-subgroup.
Similarly n 11 = 1. Let P be the unique Sylow 11-subgroup of G. Since P is normal,
N (P ) = G, and hence G/C (P ) is isomorphic to a subgroup of Aut(P ). In particular,
|G|/|C (G)| divides |Aut(P )| = ϕ(11) = 10. But this is only possible if |C (P )| = |G|, that
is, C (P ) = G so P ≤ Z (G).
E XERCISE 26. Let G be a group with |G| = 105 and suppose G has a normal Sylow
3-subgroup. Then G is abelian.
E XERCISE 30. A simple group of order 168 must have 48 elements of order 7.
Proof. Suppose |G| = 168 = 23 · 3 · 7. By Sylow’s Theorem, n 7 ≡ 1 (7) and n 7 |24, hence
n 7 = 1 or n 7 = 8. But since G is simple, n 7 6= 1. Hence since every element of order 7
lies in a Sylow 7-subgroup, and distinct Sylow 7-subgroups have trivial intersection,
there are 8 · (7 − 1) = 48 elements of order 7 in G.
53
E XERCISE 34. Let G be a group, P a Sylow p-subgroup of G, and N E G. Then P ∩ N
is a Sylow p-subgroup of N .
Proof. Let P ∗ ∈ Sylp (N ). Then there exists g ∈ G such that P ∗ ≤ g P g −1 , and hence
g −1 P ∗ g ≤ g −1 (g P g −1 )g = P
Note that
|P N | |P ||N | |P |
|P N /N | = = =
|N | |P ∩ N ||N | |P ∩ N |
which is the highest power of p dividing |G|/|N |. Hence P N /N ∈ Sylp (G/N ). (See
also Exercise 3.3.9.)
E XERCISE 35. Note that for P ∈ Sylp (G) and H ≤ G, it is not true in general that
P ∩H ∈ Sylp (H ). For example, P = 〈(12)〉 is a Sylow 2-subgroup of S 3 , but if H = 〈(23)〉,
P ∩ H = 1, which is not a Sylow 2-subgroup of H .
Proof. We know from a counting argument that |G| = 60 (see Exercise 1.2.12). Hence
by Proposition 23, it suffices to prove that G is simple, and hence by Proposition 21
it suffices to prove that G has more than one Sylow 5-subgroup.
Since 60 = 22 · 3 · 5, any subgroup of G of order 5 is a Sylow 5-subgroup of G.
Recall that G acts on the set of 12 faces of the dodecahedron and each such face has
a stabilizer of order 5. Clearly there exists more than one distinct such stabilizer. By
the above remarks this completes the proof.
E XERCISE 44. Let G be a group and p be the smallest prime dividing |G|. Then if
P ∈ Sylp (G) and P is cyclic, N (P ) = C (P ).
E XERCISE 49. Let G be a group of order 2n m, where m is odd, and suppose G has a
cyclic Sylow 2-subgroup. Then G has a normal subgroup of order m.
Section 6
54
E XERCISE 1. For n ≥ 5, A n has no proper subgroups with index strictly less than n.
Proof. Suppose towards a contradiction G is not simple, so there exists 1 < H < G
with H E G. For all i , set Hi = G i ∩ H . Then H1 ≤ H2 ≤ · · · ≤ H = Hi and Hi E G i
S
for all i . Since H 6= 1, there exists a least j with H j 6= 1. If Hk = G k for all k ≥ j , then
[ [ [
H= Hi = Hk = Gk = G
k≥ j k≥ j
contradicting that H < G. Hence there exists k ≥ j such that 1 < H j ≤ Hk < G k —
in other words, Hk is a nontrivial proper normal subgroup of G k , contradicting our
assumption that G k is simple.
Chapter 5
Section 1
Q
E XERCISE 4. Let G 1 , · · · ,G n be finite groups, G = G i , and p a prime. Then
©Y ª
Sylp (G) = P i | P i ∈ Sylp (G i )
Q
Hence n p (G) = n p (G i ).
Q Q Q
Proof. The reverse inclusion is immediate since |G| = |G i | and | P i | = |P i | for
any subgroups P i ≤ G i .
To prove the forward inclusion, suppose P ∈ Sylp (G), and set P i = P ∩ G i for all i
Q
so P = P i (here we are identifying G i with the naturally corresponding subgroup
of G). Then by Lagrange’s Theorem, P i is a p-subgroup of G i for all i , so there exist
subgroups P i∗ ∈ Sylp (G i ) with P i ≤ P i∗ . Note P ∗ = P i∗ ∈ Sylp (G) by the above, and
Q
∗ ∗ ∗ ∗
P ≤ P . But P = P since |P | = |P |, hence P i = P i and P i ∈ Sylp (G i ) for all i .
The value of n p (G) follows immediately.
55
E XERCISE 5. We exhibit a nonnormal subgroup of G = Q 8 × Z4 . Recall
j (i x) j −1 = j (i x)(− j ) = j i (− j )x = j (−k)x = −i x 6∈ H
ϕπ : G i → G π−1 (i )
Y Y
(g i ) 7→ (g π−1 (i ) )
Then ϕπ is an isomorphism.
ϕπ ((g i )(h i )) = ϕπ (g i h i )
= (g π−1 (i ) h π−1 (i ) )
= (g π−1 (i ) )(h π−1 (i ) )
= ϕπ (g i )ϕπ (h i )
(Note the third equality follows since if we write h i = g τ−1 (i ) , then ϕσ (h i ) = (h σ−1 (i ) )
by definition and h σ−1 (i ) = g τ−1 (σ−1 (i )) .) Therefore π 7→ ϕπ is a homomorphism. It is
clear that the kernel of this homomorphism is trivial, so it is injective.
56
E XERCISE 9. Let F be a field, and let H be the set of n × n matrices over F with only a
single 1 in each row and column (and 0’s elsewhere). Then H ≤ GL n (F ) and H ∼ = Sn .
Proof. Note H acts faithfully on the standard basis vectors e 1 , . . . , e n in the vector
space F n . Hence H ≤ GL n (F ) (since each matrix in H has a natural inverse), and
H∼= Sn .
The matrices in H are called permutation matrices for the obvious reason.
E XERCISE 11. Let n be a positive integer and p a prime. The number of subgroups
of order p in the elementary abelian group E p n is (p n − 1)/(p − 1).
Proof. Each of the p n − 1 nonidentity elements in E p n generates, and hence lives in,
a subgroup of order p. By Lagrange’s Theorem, distinct subgroups of order p have
trivial intersection, hence the p n − 1 nonidentity elements are partitioned into sets
of p − 1 elements, namely the subgroups of order p minus the identity element. It
follows that there are (p n − 1)/(p − 1) such subgroups.
E XERCISE 12,13 (C ENTRAL P RODUCTS ). Let A and B be groups with Z1 ≤ Z (A) and
Z2 ≤ Z (B ) such that Z1 ∼
= Z2 by the mapping x i 7→ y i . Set Z = { (x i , y i−1 ) | x i ∈ Z1 } and
define A ∗ B = (A × B )/Z .
Proof. Let π project A onto A/Z . Then ker π = {(a, 1) ∈ Z }. But if (a, 1) ∈ Z ,
(a, 1) = (x i , y i−1 ) for some i , that is, y i−1 = 1 so y i = 1 and hence a = x i = 1.
Thus ker π = 1 and A ∼ = A by the First Isomorphism Theorem. Similarly B ∼ = B.
If x ∈ A ∩ B , then x = a = b for some a ∈ A, b ∈ B . In particular, (a, 1) ∈ (1, b)Z ,
so there exists i with
A ∩ B = { x i | x i ∈ Z1 } = Z1 = Z2 = { y i | y i ∈ Z2 }
(b) Let Z4 = 〈x〉 and D 8 = 〈r, s〉 (r, s as usual). Then we may form Z4 ∗D 8 under the
identification x 2 7→ r 2 . We then have
Z4 ∗ D 8 = 〈x, r , s | x 4 = 1 = r 4 = s 2 , r s = s r −1 , x 2 = r 2 〉
57
The central product illustrates well how quotient groups can be used to ‘identify’
elements in existing groups. In this case we desire to identify each element x i in
a central subgroup of A with a corresponding element y i in an isomorphic central
subgroup of B . In other words, we desire x i = y i . We achieve this by taking a quotient
to force x i = y i . Since
x i = y i ⇐⇒ x i y i −1 = 1 ⇐⇒ x i y i−1 = 1
we divide out the elements x i y i−1 = (x i , y i−1 )—that is, the elements in Z .
In the quotient A ∗ B , we retain the structures of A and B , but the two distinct
central subgroups of A and B have collapsed into one central subgroup of both.
Note that even if every subgroup of each A i is normal, it is not necessarily true that
Q
every subgroup of A i is normal (see Exercise 5).
(a) G = ∞
Q
n=1 Z2 is an infinite group, but every element has order 1 or 2.
(b) G = Q/Z is an infinite group in which every element has finite order, but there
is an element of order n for each positive integer n (cf Exercise 3.1.14).
(c) G = Z2 × Z has an element of order 2 and an element of infinite order.
(d) G = ∞
Q
n=1 S n is such that any finite group is isomorphic to a subgroup of G.
(e) G =
Q∞
Z2 is nontrivial and G =∼ G ×G.
n=1
Section 2
E XERCISE 3. For each n, we find all possible elementary divisors and invariant fac-
tors for abelian groups of order n.
(a) n = 270: Note 270 = 2 · 33 · 5, hence the possible lists of elementary divisors are
{2, 33 , 5}, {2, 32 , 3, 5}, and {2, 3, 3, 3, 5}. Using the algorithm to obtain invariant
factors from elementary divisors, we obtain the corresponding lists {2 · 33 · 5},
{2 · 32 · 5, 3}, and {2 · 3 · 5, 3, 3}, respectively, of invariant factors.
58
E XERCISE 4. Let [a 1 , · · · , a k ] denote the group Z a1 × · · · × Z ak . We determine which
pairs of groups given are isomorphic.
(a) For groups [4, 9], [6, 6], [8, 3], [9, 4], [6, 4], and [64], we have the following lists of
elementary divisors:
[4, 9] −→ {22 , 32 }
[6, 6] −→ {2, 2, 3, 3}
[8, 3] −→ {23 , 3}
[9, 4] −→ {22 , 32 }
[6, 4] −→ {2, 22 , 3}
[64] −→ {26 }
Hence [4, 9] ∼
= [9, 4], but no other pairs of distinct groups above are isomorphic.
E XERCISE 5. Let G be a finite abelian group of type (n 1 , . . . , n t ). Then G contains an
element of order m iff m|n 1 . In particular, G has exponent n 1 .
E XERCISE 6. A finite group has finite exponent. An infinite group may have finite
exponent. A group with exponent m need not contain an element of order m.
Proof. If |G| = n, then for all x ∈ G, |x| divides n (Lagrange), hence x n = 1. Let m be
the least integer satisfying this property. Then m is the exponent of G.
Let G = ∞ Z2 . Then G is an infinite group with finite exponent 2.
Q
m m m
Proof. Write [x i i ] = (x 1 1 , · · · , x n n ). Then
© m ª © m ª © p ª Y p
imϕ = [x i i ]p = [(x i i )p ] = [(x i )mi ] = 〈x i 〉
¡Q ¢p Q p
If we let A p denote imϕ on A, then this shows: A i = A i . Also
© m m
ker ϕ = [x i i ] | [x i i ]p = 1
ª
59
mi p
= 1 for all i , which is true iff p αi |m i p for all i , which is
m
Now [x i i ]p = 1 iff x i
Q p αi −1
true iff p αi −1 |m i . Hence ker ϕ = 〈x i 〉 as desired.
(c) A/imϕ ∼
= Epn ∼
= ker ϕ, hence in particular A/imϕ and ker ϕ both have rank n.
Proof. By part (b), |ker ϕ| = ni=1 p = p n , and x p = 1 for all x ∈ ker ϕ, hence
Q
ker ϕ ∼
= E p n . Also
¡Y ¢ ¡Y p ¢ Y p
Ai ∼= (A i /A i ) ∼
Y
A/imϕ = Ai / = Zp = E p n
p
where the second isomorphism follows since |A i | = p αi −1 .
E XERCISE 8. Let p be a prime and G be a finite abelian group. Let G p and G p denote
the image and kernel of the p-th power map, respectively.
(a) G/G p ∼
= Gp .
Proof. Let n = |G|. If p does not divide n, the result is trivial since in this case
α α α
G p = G and G p = 1. Hence we assume p divides n. Write n = p 1 1 p 2 2 · · · p k k ,
α1
with pairwise distinct primes and p 1 = p (so p 1 > 1). By Theorem 5, we may
α Q p
write G = A i where |A i | = p i i for 1 ≤ i ≤ k. Now G p = A i , and it is easily
Q
p
seen that A i = A i and (A i )p = 1 for i > 1, hence we have
¡Y ¢ ¡Y p ¢
G/G p = Ai / Ai
∼
Y p
= (A i /A ) i
p
= (A 1 /A 1 ) × 1 × · · · × 1
∼
= (A 1 )p × 1 × · · · × 1 by Exercise 7
Y
= (A i )p
= Gp
E XERCISE 13. Let A = ri=1 〈x i 〉 where |x i | = n i for 1 ≤ i ≤ r . Let G be any group with
Q
n
commuting elements g 1 , . . . , g r such that g i i = 1 for 1 ≤ i ≤ r . Then there is a unique
homomorphism from A to G which maps x i 7→ g i .
Q α Q α
Proof. Define ϕ : A → G by x i i 7→ g i i . Note ϕ is well defined since if x = y
for x, y ∈ A, then both x and y (as products of powers of the x i ) can be reduced,
using commutativity and the relations |x i | = n i , to the same expression of the form
Qr ki
i =1 x i where 0 ≤ k i < n i for 1 ≤ i ≤ r . Since the g i also commute and satisfy the
n
relations g i i = 1, the elements ϕ(x) and ϕ(y) can both be reduced to the product
Qr ki
i =1 g i , hence ϕ(x) = ϕ(y).
Clearly ϕ : x i 7→ g i for 1 ≤ i ≤ r , and trivially ϕ is a homomorphism. Also ϕ is
unique since any homomorphism sending x i 7→ g i for 1 ≤ i ≤ r must map according
to ϕ. This completes the proof.
60
Note this proof uses some of the same reasoning used in the proof that two groups
with the same presentations are isomorphic. Also, this proof captures the essence of
the proof of the Fundamental Theorem for Finitely Generated Abelian Groups.
Section 4
Proof. Since P is a nontrivial p-group, Z (P ) 6= 1. Also p 2 does not divide |Z (P )|, lest
P /Z (P ) be cyclic, contradicting that P is nonabelian. Hence |Z (P )| = p.
It follows that |P /Z (P )| = p 2 , hence P /Z (P ) is abelian and P 0 ≤ Z (P ). Since P is
nonabelian, P 0 6= 1, thus we must have P 0 = Z (P ).
E XERCISE 10. A finite abelian group is the direct product of its Sylow subgroups.
α α
Proof. Let A be a finite abelian group. Set n = |A| and write n = p 1 1 · · · p k k (where
the p i are pairwise distinct primes). Let A i be the Sylow subgroup of A with |A i | =
α
p i i . Then A i E A, and by Lagrange’s Theorem, A i ∩ A j = 1 for i 6= j . Thus we have
A = A1 · · · Ak ∼
Y
= Ai
61
E XERCISE 11. Let G be a group and suppose G = H K for characteristic subgroups
H , K with H ∩ K = 1. Then Aut(G) ∼
= Aut(H ) × Aut(K ).
This result shows that if a group can be factored into disjoint subgroups, each of
which is preserved under automorphisms of the group, then any automorphism of
the group can be naturally factored into automorphisms of those subgroups.
Section 5
For the following exercises, unless otherwise noted, H and K denote arbitrary groups,
ϕ : K → Aut(H ), and G = H oϕ K (with H and K treated as subgroups of G).
62
E XERCISE 1. C K (H ) = ker ϕ
ker ϕ = { k ∈ K | ϕ(k) = 1 }
= { k ∈ K | (∀h ∈ H )[ϕ(k)(h) = h] }
= { k ∈ K | (∀h ∈ H )(khk −1 = h) }
= { k ∈ K | (∀h ∈ H )(kh = hk) }
= { k ∈ K | k ∈ CG (H ) }
= K ∩CG (H ) = C K (H )
E XERCISE 2. C H (K ) = N H (K )
(b) G ∼
= S4.
63
E XERCISE 6 (I SOMORPHIC S EMIDIRECT P RODUCTS ). Suppose H is arbitrary, K is
cyclic, and ϕ1 and ϕ2 are homomorphisms from K to Aut(H ) (assumed injective
if K is infinite) such that ϕ1 (K ) is conjugate to ϕ2 (K ). Then H oϕ1 K ∼
= H oϕ2 K .
Proof. Suppose σϕ1 (K )σ−1 = ϕ2 (K ). Since K is cyclic, there exists some n such that
σϕ1 (k)σ−1 = ϕ2 (k)n for all k ∈ K .
Recall that for h ∈ H and k ∈ K , k acts on h in H oϕ1 K by conjugation with
khk −1 = ϕ1 (k)(h). From our intuitive understanding of conjugation, we know that
σϕ1 (k)σ−1 = ϕ2 (k)n = ϕ2 (k n ) acts ‘in the same way’ on σ(h), that is, k n acts ‘in the
same way’ on σ(h) in H oϕ2 K . This suggests the isomorphism
ψ : (h, k) 7→ (σ(h), k n )
= σ(h 1 ϕ1 (k 1 )(h 2 )) , (k 1 k 2 )n
¡ ¢
= (σ(h 1 ), k 1n )(σ(h 2 ), k 2n )
= ψ(h 1 , k 1 )ψ(h 2 , k 2 )
|K | |K |
|ker ϕ1 | = = = |ker ϕ2 |
|ϕ1 (K )| |ϕ2 (K )|
Therefore ker ϕ1 = ker ϕ2 since K is cyclic (and hence has a unique subgroup of this
order). Now set N = ker ϕi , K = K /N , and let ϕi : K → ϕi (K ) be the naturally induced
isomorphisms. Also let c σ denote conjugation by σ. Then ϕ = ϕ2 −1 c σ ϕ1 ∈ Aut(K ).
2 For this case I required some assistance from Professor Google.
64
But there is a natural surjective homomorphism Aut(K ) → Aut(K ), hence there is a
map k 7→ k r ∈ Aut(K ) whose image is ϕ. It follows from computation that c σ (ϕ1 (k)) =
ϕ2 (k)r for all k ∈ K , so n = r . Thus there exists m such that (k n )m = k for all k ∈ K ,
and hence the map ψ∗ above, with this m, is a two-sided inverse of ψ.
Thus in either case ψ is an isomorphism, completing the proof.
Z 8 × Z 7 (∼
= Z56 ) Z2 × Z4 × Z7 Z2 × Z2 × Z2 × Z7
Z8 Z2 × Z4 Z2 × Z2 × Z2 D 8 Q8
We consider these cases in turn and, for each such K , find all possible maps ϕ
(up to a possible choice of generators in K , which does not change the iso-
morphism type by Exercise 6). In the first three cases we do not consider case
ϕ = 1 explicitly, since the trivial map gives rise to abelian groups which we
have already classified above. Note since (8, 6) = 2, if ϕ 6= 1, then ϕ(K ) must be
the unique subgroup of Aut(H ) of order 2.
If K ∼
= Z8 , then we must have ker ϕ = Z4 .
65
If K ∼
= Z2 × Z4 , then we must have either ker ϕ = Z4 or ker ϕ = Z2 × Z2 .
If K ∼
= Z2 × Z2 × Z2 , then we must have ker ϕ ∼
= Z2 × Z2 .
If K ∼
= D 8 , then ker ϕ must be one of D 8 , Z4 , or Z2 × Z2 .
If K ∼
= Q 8 , then ker ϕ must be one of Q 8 or Z4 (recall there are no subgroups
of Q 8 isomorphic to Z2 × Z2 ).
In addition, each one of these cases is realized, and gives rise to a distinct
semidirect product type (since for each distinct type of K , the kernels listed
are of different types).
Thus there are nine possible isomorphism types for nonabelian G with normal
Sylow 7-subgroup.
(d),(e) We classify (nonabelian) G with a nonnormal Sylow 7-subgroup.
We know that G must have a normal Sylow 2-subgroup, so G ∼= H oϕ K where
|H | = 8, K ∼
= Z7 , and ϕ : K → Aut(H ).
We claim H ∼ = Z2 ×Z2 ×Z2 . Indeed, if k ∈ Z7 , k 6= 1, then k acts nontrivially on H
by conjugation. By the Orbit-Stabilizer Theorem then, the orbit of any element
in H moved by k must have order 7. Hence the 7 nonidentity elements in H
must all be in a single orbit, and so have the same order. The only possibility
is the one stated by the claim.
We know Aut(H ) ∼ = GL 3 (F2 ), so |Aut(H )| = 168 = 23 ·3·7. Thus in particular there
exists an element in Aut(H ) of order 7, so a semidirect product of the current
form does exist. In addition, all subgroups of order 7 in Aut(H ) are conjugate
by Sylow’s Theorem, so by Exercise 6, this gives only one isomorphism type.
It follows from this exercise that there are exactly thirteen distinct isomorphism
types for groups of order 56.
Z3 × Z25 (∼
= Z75 ) Z3 × Z5 × Z5
66
E XERCISE 12 (G ROUPS OF ORDER 20).
Suppose G is a group with |G| = 20 = 22 · 5. By Sylow’s Theorem, n 5 ≡ 1(5) and
n 5 |4, hence n 5 = 1 and G ∼= H oϕ K where H ∼ = Z5 , |K | = 4 (so K ∼
= Z4 or K ∼ = Z2 × Z2 ),
∼
and ϕ : K → Aut(H ) = Z4 .
If ϕ = 1, G is one of the abelian groups Z5 × Z4 (∼ = Z20 ) or Z5 × Z2 × Z2 .
If ϕ 6= 1, then if K ∼
= Z4 , either ker ϕ = 1 and G ∼ = Hol(Z5 ) or else ker ϕ ∼ = Z2 . If
∼ ∼
K = Z2 × Z2 , then we must have ker ϕ = Z2 .
Since all of these cases are possible and nonisomorphic, there are exactly five
isomorphism types for groups of order 20.
Note that this is equivalent to proving that for any A ∈ GL 2 (Fp ) with |A| = 2, there is
a change of basis under which A becomes one of the above. (Note that since p 6= 2,
and hence 2 6= 0, the two matrices above are not conjugate, since the one at right has
no eigenvectors but the one at left does.)
Suppose A ∈ GL 2 (Fp ) with |A| = 2. Note3 that for any v ∈ Z p2 ,
A(v + Av) = Av + A 2 v = Av + v = v + Av
A(v − Av) = Av − A 2 v = Av − v = −(v − Av)
Let (v, w) be a basis for the space. Note it cannot be true that both v + Av and v − Av
are in the span of w, lest
1
v = ((v + Av) + (v − Av))
2
is also in the span of w, contradicting the linear independence of (v, w). (Note the
above equation relies on the fact that p 6= 2, so 2 6= 0 and 1/2 is defined.) Let v ∗ be
whichever is not in the span of w. Then (v ∗ , w) is a basis and Av ∗ = ±v ∗ . Similarly
there is w ∗ with Aw ∗ = ±w ∗ such that (v ∗ , w ∗ ) is a basis of the space. This proves
the conjugacy claim above.
It follows that there are at most two (and at least one, taking say p = 3) possible
isomorphism types for G in this case.
3 For the following observations I required some assistance from Professor Google.
67
E XERCISE 18. Let H be any group. There exists a group G containing H as a normal
subgroup such that any automorphism of H is witnessed as an inner automorphism
in G (restricted to H ).
(a) The elements of H are coset representatives in G/K , and if we identify H and G/K ,
π|H is just the left regular representation of H .
(b) NS n (π(H )) = π(G). Therefore in general the normalizer of the (left) regular
representation of a group H is isomorphic to Hol(H ).
Proof. Note π(H ) E π(G) since H E G, hence π(G) ≤ N (π(H )). Because H is
finite, it is sufficient to prove that |π(G)| = |N (π(H ))|.
We need |N (π(H ))| ≤ |π(G)|.
Recall N (π(H ))/C (π(H )) is isomorphic to a subgroup of Aut(π(H )). Since π is
injective, Aut(π(H )) ∼
= Aut(H ) = K . By Exercise 4.3.36, C (π(H )) ∼
= H . Hence
|N (π(H ))|
divides |K |
|H |
Thus |N (π(H ))| divides |H ||K | = |G|, so |N (π(H ))| ≤ |π(G)| as desired.
Thus N (π(H )) = π(G) ∼= G = Hol(H ). The general result follows from (a).
Proof. If σ is the n-cycle and Zn = 〈x〉, the map x 7→ σ realizes 〈σ〉 as a (left)
regular representation of Zn under some numbering of Zn . Thus by part (b),
NS n (σ) ∼
= Hol(Zn ) = Zn oAut(Zn ), and since |Aut(Zn )| = ϕ(n), |NS n (σ)| = nϕ(n).
Note this exercise shows how holomorphs arise naturally from permutation groups.
Also, this and Exercise 4.3.36 together give us characterizations of the centralizers
and normalizers of regular permutation representations.
68
E XERCISE 23 ( W REATH P RODUCT ). Let K and L be groups with ρ : K → S n for some
positive integer n.
Let H = n L and ϕ : K → Aut(H ) be naturally induced from ρ by letting the
Q
elements of K permute the factors of H (see Exercise 1.8). Then the wreath product
of L by K is defined by
L o K = H oϕ K
If ρ is not mentioned, it is assumed to be the left regular representation of (finite) K .
(a) Suppose K and L are finite. Then |L o K | = |L||K | |K |. Proof: Immediate.
(b) Let p be a prime. Then Z p o Z p is a nonabelian group of order p p+1 isomorphic
to a subgroup of S p 2 .
Note examples like (b) might shed some light on the name ‘wreath product’. In that
example, each of the factors in the direct product can be seen as a little circle, and
the factors themselves are cyclically permuted in a larger circle.
Chapter 6
Section 1
Letting g vary we see σ(x) ∈ Z (G), then letting x vary we see σ[Z (G)] ⊆ Z (G). Now
substituting σ−1 for σ, we have σ−1 [Z (G)] ⊆ Z (G), so Z (G) = σ[σ−1 [Z (G)]] ⊆ σ[Z (G)],
so σ[Z (G)] = Z (G). Since σ was arbitrary, Z (G) char G as claimed.
Now if Zi (G) char G for some i ≥ 1, set G = G/Zi (G). If σ ∈ Aut(G), σ naturally
induces a map
σ:G →G
g 7→ σ(g )
69
This map is well-defined since if x = y, that is, x Zi (G) = y Zi (G), then
σ(x) = σ(x)Zi (G) = σ[x Zi (G)] = σ[y Zi (G)] = σ(y)Zi (G) = σ(y)
where the second and fourth equalities hold since Zi (G) char G. It is immediate that
σ is a homomorphism, and σ−1 is a two-sided inverse of σ, hence σ ∈ Aut(G).
Now it follows from the reasoning above that σ[Z (G)] = Z (G). We claim that
σ[Zi +1 (G)] = Zi +1 (G). Indeed, recall by definition Zi +1 (G) = Z (G). Thus if g ∈ Zi +1 ,
then g ∈ Z (G), hence σ(g ) = σ(g ) ∈ Z (G), that is, σ(g ) ∈ Zi +1 . Since g was arbitrary,
σ[Zi +1 (G)] ⊆ Zi +1 (G), and as above it follows that σ[Zi +1 (G)] = Zi +1 (G) as claimed.
Thus Zi +1 char G. By induction, the result is true for all i .
Proof. If H ≤ Z (G), the result holds, so assume that H 6≤ Z (G). Then we have
Z (G) < H Z (G) E G, so by the Lattice Theorem 1 < H E G where G = G/Z (G).
By induction on nilpotence class, we may assume H ∩ Z (G) 6= 1. Now it follows
from the Lattice Theorem again that H Z (G) ∩ Z2 (G) 6≤ Z (G), which implies
H ∩ Z2 (G) 6= 1. Since H E G, [G, H ] ≤ H (Proposition 5.7(2)), and by definition
of Z2 (G), [G, Z2 (G)] ≤ Z (G). Therefore we have [G, H ∩ Z2 (G)] ≤ H ∩ Z (G). If
[G, H ∩ Z2 (G)] 6= 1, the result holds, otherwise that means H ∩ Z2 (G) ≤ Z (G), so
again H ∩ Z (G) 6= 1 and the result holds.
Together with the fact that all p-groups are nilpotent (Proposition 2), these results
exhibit strong connections between p-groups and nilpotent groups in general. For
this reason it is often convenient to think about (simpler) p-groups when working
with nilpotent groups.
4 For this problem, I required some assistance from Professor Google.
70
E XERCISE 3. Let G be a finite group. Then G is nilpotent iff G has a normal subgroup
of every order dividing |G|, and G is cyclic iff G has a unique subgroup of every order
dividing |G|.
Proof. If G has a normal subgroup of every order dividing |G|, then in particular
G has a normal (and hence unique) Sylow p-subgroup for all primes p dividing |G|.
Thus all Sylow p-subgroups of G are normal, so G is nilpotent by Theorem 3.
α α
Conversely, if G is nilpotent, write |G| = p 1 1 · · · p k k (p i prime), so by Theorem 3
β β
G = P 1 × · · · × P k where P i ∈ Sylp i (G). If n divides |G|, write n = p 1 1 · · · p k k . Then by
β
Theorem 1(3), we know each P i has a normal subgroup Ni with |Ni | = p i i . Now set
N = N1 × · · · × Nk . Then N E G and |N | = n, as desired.
If G is cyclic, we know G has a unique subgroup of each order dividing |G| by
Theorem 2.7(3). To prove the converse, we exhibit for any noncyclic G two distinct
subgroups having the same order. So consider an arbitrary noncyclic group G. If G is
abelian, then by the fundamental theorem and Proposition 5.6(2), we must be able
to write G as a direct product of cyclic groups where, for some prime p, at least two
of the factors are nontrivial p-groups. From these p-groups we can select elements
of order p which generate distinct subgroups of order p.
If G is nilpotent but not abelian, proceed by induction. Note G/Z (G) is also
nilpotent but not cyclic, lest G be abelian. Hence we may assume G/Z (G) contains
distinct subgroups of the same order. The complete preimages of these subgroups
in G are distinct subgroups of G of the same order.
Finally if G is not nilpotent, then by Theorem 3 at least one of its Sylow subgroups
is nonnormal and hence not unique. Therefore in all cases, G has distinct subgroups
of the same order.
Proof. Set G = G/Z (G). We first establish a natural relationship between the upper
central series in G and the upper central series in G:
Zi +1 (G) = Zi (G) (i ≥ 0)
On the other hand, Zi +2 (G)/N = Zi +2 (G)/Zi +1 (G). Therefore we must prove that
Zi +2 (G)/Zi +1 (G) = Z (G/Zi +1 (G)).
71
We know that Zi +2 (G)/Zi +1 (G) = Z (G/Zi +1 (G)) by definition of Zi +2 (G). And by
the Third Isomorphism Theorem, G/Zi +1 (G) ∼ = G/Zi +1 (G). Using this isomorphism,
and the fact that centers are preserved under isomorphism, it can be seen that the
required equality holds.
Therefore our claim holds for i + 1. By induction, our result holds for all i ≥ 0.
Returning to the problem, note for i ≥ 0 that Zi +1 (G) = G iff Zi (G) = G by our
result and the Lattice Theorem. Hence G is nilpotent iff G is nilpotent, as desired.
Note the proof gives us more information: for n > 0, G is of nilpotence class n + 1 iff
G/Z (G) is of nilpotence class n.
H i +1 = [H , H i ] ≤ [G,G i ] = G i +1
Thus the claim holds for all i ≥ 0 by induction. Since G is nilpotent, G n = 1 for
some n, so H n = 1 and H is nilpotent.
Suppose now ϕ : G → K is a surjective homomorphism. We claim ϕ[G i ] = K i for
i ≥ 0. Note ϕ[G 0 ] = ϕ[G] = H = H 0 since ϕ is surjective. If i ≥ 0 and ϕ[G i ] = H i , then
The second equality follows since if x ∈ G and y ∈ G i , then ϕ([x, y]) = [ϕ(x), ϕ(y)] by
the definition of the commutator. Hence the claim holds for all i ≥ 0 by induction.
Since G is nilpotent, G n = 1 for some n, hence K n = 1 and K is nilpotent. Thus we
see that the homomorphic image of a nilpotent group is nilpotent.
If H E G, then G/H is the homomorphic image of G under a natural projection,
so G/H is nilpotent.
72
E XERCISE 9. Let G be finite. Then G is nilpotent iff for all x, y ∈ G if (|x|, |y|) = 1 then
x y = y x.
E XERCISE 12. We find the upper and lower central series for A 4 and S 4 .
Note by this exercise and Exercise 3.5.10, A 4 is solvable but not nilpotent.
73
E XERCISE 13. We find the upper and lower central series for A n and S n , for n ≥ 5.
Proof. Fix n ≥ 5. Recall A n is simple (Theorem 4.24), that is, A n has no nontrivial
proper normal subgroups. We know that Z0 (A n ) = 1. By Exercise 14, we must have
Z1 (A n ) = 1 or Z1 (A n ) = A n . But A n is not abelian, hence Z1 (A n ) = 1 and it follows
that Zk (A n ) = 1 for all k ≥ 0. It is trivial that Zk (S n ) = 1 for all k ≥ 0.
We know A 0n = A n , and since A n is simple, A 1n = A n or A 1n = 1. Again since A n is
not abelian, we must have A 1n = A n , and it follows that A kn = A n for all k ≥ 0. Now
S n0 = S n , but A n E S n and S n /A n ∼ = Z2 is abelian, hence S n1 ≤ A n . Since S n is not
abelian, the simplicity of A n forces S n1 = A n . Now S n2 = [S n , A n ] ≤ A n . Again by the
simplicity of A n , and the fact that A n 6≤ Z (S n ) = 1, we must have S n2 = A n . It follows
that S nk = A n for all n ≥ 1.
In particular, these results show that neither A n nor S n is nilpotent.
Similarly σ[G (i +1) ] = G (i +1) . Since σ was arbitrary, this shows that G i +1 char G and
G (i +1) char G. By induction, the result holds for all i ≥ 0.
H , q + H , . . . , (µ − 1)q + H
r = βq + h 0 = (ρ + δµ)q + h 0 = ρq + (δµq + h 0 ) = ρq + h 00
74
E XERCISE 20. Let G be finite, P ∈ Sylp (G), and N E G with (|N |, p) = 1.
(a) NG (P ) = NG (P )
NG (N P ) = NG (N P ) = NG (P )
We claim N NG (P ) = NG (N P ), so that NG (P ) = N NG (P ) = NG (N P ) = NG (P ),
establishing the desired result.
Indeed, note that N P E N (N P ) and P ∈ Sylp (N P ) since p does not divide |N |.
Therefore by Frattini’s Argument, N NG (P ) = (N P )NG (P ) = NG (N P ), which is
precisely the claim.
For a group G, the Frattini subgroup Φ(G) of G is defined to be the intersection of all
maximal subgroups of G (if G has no maximal subgroups, Φ(G) = G). The following
exercises establish some basic properties of the Frattini subgroup.
where M varies over maximal subgroups of G. Since σ was arbitrary, Φ(G) char G.
E XERCISE 24. Let G be a group. Call an element x ∈ G a nongenerator if for all H < G,
〈H , x〉 < G. Then Φ(G) = { x ∈ G | x is a nongenerator }.
Proof. If x 6∈ Φ(G), then there exists some maximal M < G such that x 6∈ M . But then
M < 〈M , x〉 = G by maximality of M , hence x is not a nongenerator.
Conversely, if x is not a nongenerator, there exists H < G with 〈H , x〉 = G. Now
if H ≤ M < G with M maximal, then x 6∈ M lest G = 〈H , x〉 ≤ M < G, a contradiction.
Therefore we assume H is not contained in a maximal subgroup, so for all K with
H ≤ K < G, we may choose C (K ) with K < C (K ) < G. Set C (G) = G. Now define
recursively on the ordinals
H0 = H
Hα+ = C (Hα )
[
Hλ = Hα (λ a limit ordinal)
α<λ
75
By induction, this forms an ascending chain of subgroups in G, and there must exist
a limit ordinal µ with G = α<µ Hα and H ≤ Hα < G for all α < µ. Now x ∈ α<µ Hα ,
S S
Proof. Let P ∈ Sylp (Φ(G)). We claim that P E Φ(G). Indeed, by Frattini’s argument
(Proposition 6), G = Φ(G)NG (P ). Now if NG (P ) < G, then NG (P ) ≤ M < G for some
maximal subgroup M . But then since Φ(G) ≤ M , we have G = Φ(G)NG (P ) ≤ M < G—
a contradiction. Therefore P E G, so in particular P E Φ(G) as claimed. Since P was
arbitrary, Φ(G) is nilpotent by Theorem 3(3).
Φ(P f = Φ(Pe) = e
\ \
) = M≤ M 1
M M
(c) Burnside’s Basis Theorem: The set {y 1 , . . . , y k } is a minimal generating set for P
iff {y 1 , . . . , y k } is a basis of P (as a vector space over Fp ).
76
Proof. If P is cyclic, then P is spanned by a single element, so by Burnside’s
Basis Theorem, P is generated by a single element, that is, P is cyclic.
If P /P 0 is cyclic, then P 0 = Φ(P ) by parts (a) and (b), hence P is cyclic, and P is
cyclic by the previous case.
E XERCISE 31. Let G be a finite solvable group and suppose M is a minimal normal
subgroup of G. Then M is an elementary abelian p-group for some prime p.
The previous two exercises show that the ‘top’ and ‘bottom’ of a finite solvable group
have a relatively simple structure.
Section 2
In the following exercises, for a group G, µ(G) denotes the minimal possible index of
a proper subgroup in G (cf. p. 203).
E XERCISE 1. Let G be a group, P ∈ Sylp (G), and suppose for all Q ∈ Sylp (G) with
P 6= Q, P ∩Q = 1. Then for all P 1 , P 2 ∈ Sylp (G) with P 1 6= P 2 , P 1 ∩P 2 = 1. Therefore the
number of nonidentity elements of p-power order in G is (|P | − 1)|G : NG (P )|.
This contradicts that P has trivial intersection with distinct Sylow p-subgroups. Thus
P 1 ∩ P 2 = 1 as desired. The computational result follows trivially.
77
E XERCISE 4. There are no simple groups of order 80.
E XERCISE 5. Let G be a solvable group with |G| = pm, p prime and (p, m) = 1. If
P ∈ Sylp (G) and NG (P ) = P , then G has a normal subgroup of order m.
Proof. Suppose G is simple with |G| = 4125 = 3·53 ·11. Note µ(G) ≥ 15. Now n 5 ≡ 1(5),
n 5 |3 · 11, and n 5 > 1. But then n 5 = 11 < 15, a contradiction.
Proof. Suppose G is simple with |G| = 792 = 23 · 32 · 11. Then n 11 ≡ 1(11), n 11 |23 · 32 ,
and n 11 > 1, hence n 11 = 12. By simplicity then we may assume G ≤ S 12 . Now let
P ∈ Syl11 (G). Note (p. 204) |NS 12 (P )| = 11 · (11 − 1) = 2 · 5 · 11. But |NG (P )| = 2 · 3 · 11,
contradicting that NG (P ) ≤ NS 12 (P ) by Lagrange’s Theorem.
Section 3
78
E XERCISE 1. Let F (R) and F (S) be free groups. Then F (R) ∼
= F (S) iff |S| = |R|.
Proof. If |S| = 1, then F (S) is cyclic and hence abelian. If |S| > 1, choose a, b ∈ S with
a 6= b. Then by definition of products in F (S), [a, b] = a −1 b −1 ab 6= 1, hence a and b
do not commute, so F (S) is nonabelian.
E XERCISE 5. We give a presentation for A 4 using two generators. Set σ = (123) and
τ = (124). We know from the lattice that A 4 = 〈σ, τ〉. Note σ3 = τ3 = (στ)2 = 1. The
latter relation implies that τσ = σ−1 τ−1 , so (as in the dihedral group) these relations
allow us to reduce any element of A 4 to a specific form σi τ j where 0 ≤ i , j < 3. Thus
any relations in A 4 may be derived from these three relations. It follows that
A 4 = 〈 σ, τ | σ3 = τ3 = (στ)2 = 1 〉
E XERCISE 11. Let S be a set and R = 〈 [s, t ] | s, t ∈ S 〉. Then A(S) = 〈S | R〉 is called the
free abelian group on S. If G is abelian and ϕ : S → G is any set map, then there exists
a unique homomorphism Φ : A(S) → G such that Φ|S = ϕ.
In particular, if |S| = n, then A(S) ∼
= n Z.
Q
Proof. Note we could simply define the homomorphism directly as in Theorem 17.
Alternately, note by the universal property of free groups that there exists a unique
Φ : F (S) → G with Φ|S = ϕ. Since G is abelian, R ≤ ker Φ, hence if N is the normal
closure of R, N ≤ ker Φ. Therefore we obtain a natural sequence
79