You are on page 1of 34

Jonathan Bergkno

Herstein Solutions
Chapters 1 and 2
Throughout, G is a group and p is a prime integer unless otherwise stated. A B denotes that A is a
subgroup of B while AB denotes that A is a normal subgroup of B.
H 1.3.14* (Fermats Little Theorem) Prove that if a Z then a
p
a mod p.
Proceed by induction on (positive) integer a. The theorem holds for a = 1 because 1
p
= 1 1 mod p.
Suppose that k
p
k mod p and then compute, by the binomial theorem,
(k + 1)
p
=
p

i=0

p
i

k
i
1
pi
= k
p
+ 1 +p(. . .) k + 1 mod p.
In the last step, we used the induction hypothesis. This proves the result for positive integer a. The expansion
(k + 1)
p
= k
p
+ 1 +p(. . .) is justied by lemma 2 (below) which states that p [

p
n

for n 1, . . . , p 1.
The case a = 0 is trivial: 0
p
= 0 0 mod p. Let a Z
+
as before. As was just proven, there exists c Z
such that a
p
a = cp. Now, if p > 2 (odd), we have (a)
p
(a) = a
p
+a = (c)p so (a)
p
a mod p.
On the other hand, if p = 2, we just need to see that modulo 2 picks out even/odd parity. Regardless of
being positive or negative, an even integer squared is even and an odd integer squared is odd. Therefore
a
2
a mod 2 and the theorem is proven in its entirety.
Lemma 1 Let n Z
+
and r 0, . . . , n. The binomial coecient

n
r

=
n!
r!(nr)!
is an integer.
Proof: Proceed by induction on n. For n = 1,

1
0

= 1 and

1
1

= 1 are both integers, as claimed. Suppose


that

n1
r

is an integer for all r 0, . . . , n 2. Now

n 1
r

n 1
r 1

=
(n 1)!
r!(n r 1)!
+
(n 1)!
(r 1)!(n r)!
= (n 1)!

(n r) +r
r!(n r)!

n
r

.
Hence, when the above computation goes through,

n
r

is a sum of integers and thus is, itself, an integer.


However, the computation fails from the outset for r = 0 and r = n (were interested in things like (r 1)!
and (n r 1)!), so those cases must be considered independently. We have

n
0

=
n!
0!n!
= 1

n
n

=
n!
n!0!
= 1
and the claim is proven.
Lemma 2 Let n 1, . . . , p 1. Then p [

p
n

.
Proof: Intuitively, this lemma is true because the numerator has a factor of p and the denominator has no
factors that cancel it (relying crucially on the primality of p). By the fundamental theorem of arithmetic (Z
is a UFD), we can write the denominator as n!(p n)! =

q
a
i
i
with q
i
the unique ascending list of prime
divisors of n!(pn)! and a
i
their respective powers. As every factor of n(n1) 21(pn)(pn1) 21
is smaller than p (this fails if n = 0 or n = p), p divides none of them and hence, as a prime, does not divide
their product. Then none of the q
i
is p. The factor of p in the numerator is then preserved upon taking the
quotient, and p [

p
n

. Note that it makes no sense to talk about p dividing

p
n

unless

p
n

Z (lemma 1).
H 1.3.15 Let m, n, a, b Z with (m, n) = 1. Prove there exists x with x a mod m and x b mod n.
As m and n are coprime, there exist integers c, d such that cm + dn = 1. We see that (ac)m + (ad)n = a,
so that adn = a acm a mod m. Similarly, (bc)m + (bd)n = b, so bcm = b bdn b mod n. Set
x = adn +bcm. Then we have
x adn mod m a mod m x bcm mod n b mod n.
H 2.3.3 Let (ab)
2
= a
2
b
2
for all a, b G. Prove that G is abelian.
The statement is that, for all a, b G, we have abab = a
2
b
2
. Multiply both sides of the equation on the left
by a
1
and on the right by b
1
. Then we have ba = ab and hence G is abelian.
H 2.3.4* Let G be such that, for three consecutive integers i, (ab)
i
= a
i
b
i
for all a, b G. Prove that G
is abelian.
Let N be the smallest of the three consecutive integers. Then we have that (ab)
N
= a
N
b
N
, (ab)
N+1
=
a
N+1
b
N+1
and (ab)
N+2
= a
N+2
b
N+2
for all a, b G. Inverting the rst equation and right multiplying it to
the second equation implies
ab = a
N+1
b
N+1
b
N
a
N
= a
N+1
ba
N
hence ba
N
= a
N
b.
Inverting the second equation and right multiplying it to the third equation gives
ab = a
N+2
b
N+2
b
(N+1)
a
(N+1)
= a
N+2
ba
(N+1)
hence ba
N+1
= a
N+1
b.
Therefore a
N+1
b = ba
N+1
= ba
N
a = a
N
ba and left multiplying by a
N
yields ab = ba for arbitrary a, b G.
Hence G is abelian.
H 2.3.8 Let G be nite. Prove the existence of an N Z such that a
N
= e for all a G.
Let a G. As G is nite and closed under multiplication, the set a
0
, a
1
, a
2
, . . . is nite. Hence there exist
m, n 0, 1, . . . such that a
m
= a
n
(without loss of generality, take m > n). By the division algorithm, we
can write m = kn+r with k, r Z and 0 r < n. Now we have the statement a
kn+r
= a
n
or a
(k1)n+r
= e.
Enumerate G as a
1
, . . . , a
n
. For each a
i
, there exists an N
i
= (k
i
1)n
i
+ r
i
, computed by the above
method, such that a
N
i
i
= e. Let N =

N
i
. Then a
N
i
= (a
N
i
i
)
N/N
i
= e
N/N
i
= e for each a
i
G.
H 2.3.9a Let G have three elements. Prove that G is abelian.
Let G = e, a, b. In order to show that G is abelian, we only need to show that ab = ba because the identity
commutes with everything. Suppose ab ,= ba. We have only three choices: ab = e, ab = a or ab = b.
(1) If ab = e, then a = b
1
so ba = e = ab, which is a contradiction. Hence ab ,= e.
(2) If ab = a, then b = e so ba = a = ab, which is a contradiction. Hence ab ,= a.
(3) If ab = b, then a = e so ba = b = ab, which is a contradiction. Hence ab ,= b.
Therefore ab = ba necessarily, and hence G is abelian.
H 2.3.10 Let G be such that every element is its own inverse. Prove that G is abelian.
Let a, b G. Then ab = a
1
b
1
= (ba)
1
= ba, so G is abelian.
H 2.3.11 Let G have even order. Prove there exists a non-identity element a G with a
2
= e.
If an element a of a group doesnt satisfy a
2
= e, then there exists a unique inverse element a
1
,= a in the
group. Elements of this type can be counted in pairs a, a
1
. There are therefore an even number 2k of
elements with a
2
,= e. The identity satises e
2
= e, so there are [G[ 2k 1 0 non-identity elements a
satisfying a
2
= e. If [G[ is even, then [G[ 2k 1 is non-zero and hence there exists a non-identity element
a G with a
2
= e.
H 2.3.12 Let G be a non-empty set closed under an associative product with an e G such that ae = a
for all a G as well as the property that, for each a G, there exists y(a) G with ay(a) = e. Prove that
G is a group under this product.
In order for this set to be a group, right inverses must also be left inverses and it must hold that ea = a for all
a G. Multiply the equation ay(a) = e by y(y(a)), the right inverse of y(a), on the right: by associativity,
y(y(a)) = [ay(a)]y(y(a)) = a[y(a)y(y(a))] = a.
Hence y(y(a)) = a is the right inverse of y(a). Now we see that, in addition to ay(a) = e, we have y(a)a = e
so that y(a) is the inverse (both left and right) of a. Now we can trivially show the property ea = a. Multiply
the equation ae = a on the left by y(a) and on the right by a (notice that y(a)aa = [y(a)a]a = a):
a = y(a)[ae]a = [y(a)a]ea = ea.
Therefore G is a group.
H 2.5.1 Let H and K be subgroups of the group G. Prove that H K G.
H K is closed: let a, b H K. Then a, b H, K so ab H, K because both are subgroups. Hence
ab H K. H K is closed under inverses: let a H K. Then a
1
H, K, and hence a
1
H K. By
lemma 2.3, H K G.
H 2.5.2 Let G have a subgroup H. Dene a left coset of H by aH = ah [ h H. Show there is a
bijection between left and right cosets of H in G.
Dene f : aH [ a G Ha [ a G by f(aH) = Ha
1
. This map is well-dened: suppose a
1
, a
2
G
are such that a
1
H = a
2
H. Then there exists h H such that a
1
= a
2
h and hence f(a
1
H) = Ha
1
1
=
Hh
1
a
1
2
= Ha
1
2
= f(a
2
H). The map is trivially surjective: for any a G, Ha is the image of a
1
H. The
map is injective: suppose a
1
H, a
2
H are such that f(a
1
H) = f(a
2
H). Then Ha
1
1
= Ha
1
2
which implies the
existence of h H such that a
1
1
= ha
1
2
. Inverting, we nd that a
1
= a
2
h
1
, i.e. that a
1
H = a
2
H which
proves injectivity.
H 2.5.3 Let G have no proper subgroups. Prove that [G[ is prime.
Let g G with g ,= e. g) is a subgroup of G. Because g ,= e, g) is not the trivial subgroup e. Hence,
because G has no proper subgroups, it must be that g) = G which gives [g[ = [G[. Suppose [G[ = mn with
m, n > 1 (so m, n < [G[). Then (g
m
)
n
= e which implies that [g
m
[ = n < [G[. The subgroup g
m
) is proper,
with 1 < n < [G[ elements, which is a contradiction. Therefore [G[ must be prime.
H 2.5.6* Let H, K G have nite indices in G. Give an upper bound for the index of H K.
Missing.
H 2.5.7 With a, b R, let
ab
: R R be given by
ab
= ax +b. Let G =
ab
[ a ,= 0. Prove that G is a
group under composition. What is
ab

cd
?
(
ab

cd
)(x) = a(cx +d) +b = acx +ad +b =
ac,ad+b
(x). As R is a eld and a, c ,= 0,
ac,ad+b
G so that
G is closed under composition. The operation is associative:
(
ab

cd
)
ef
=
ac,ad+b

ef
=
ace,acf+ad+b
=
ab

ce,cf+d
=
ab
(
cd

ef
).
The set is closed under inverses:
1
ab
= 1
a
,
b
a
G. Finally, there is an identity element e =
1,0
. Therefore
G is a group. Note that G is non-abelian because
ab

cd
=
ac,ad+b
,=
ac,cb+d
=
cd

ab
.
H 2.5.8 Taking the group of 2.5.7, let H =
ab
G [ a Q. Prove that H G and list the right cosets
of H in G.
Because Q is a eld, we have that, for
ab
,
cd
H,
ab

cd
=
ac,ad+b
H and
1
ab
= 1
a
,
b
a
H. By
lemma 2.3, H G.
Now let
ab
,
cd
G be such that H
ab
= H
cd
. Then
ab

1
cd
= a
c
,b
ad
c
H. Therefore
a
c
Q is the
condition on this situation. In particular,
ab
and
ac
are in the same right coset regardless of b and c, which
tells us that the right cosets are indexed by just one real parameter.
H
ab
=
cd
G [ c = qa for some q Q0 [ a R0 .
H 2.5.9 (a) In the context of 2.5.8, prove that every left coset of H in G is also a right coset of H in G.
(b) Give an example of a group G and a subgroup H of G such that the above is not true.
(a) Let
ab
,
cd
G be such that
ab
H =
cd
H. Then
1
ab

cd
= c
a
,
db
a
H. Again, this is the statement
that
c
a
Q, so two elements of G are equivalent left-modulo H if the ratio of their rst parameters is rational.
Therefore consider the cosets
ab
H and H
ab
. If
cd

ab
H then
c
a
Q. Hence
a
c
Q which gives that

cd
H
ab
. Therefore
ab
H H
ab
. The reverse inclusion is identical, so
ab
H = H
ab
.
(b) Consider G = S
3
= e, (12), (13), (23), (123), (213) and the subgroup H = e, (12). The left cosets are:
eH = e, (12) (213)H = (213), (213)(12) = (23) (123)H = (123), (123)(12) = (13).
On the other hand, the right cosets are:
He = e, (12) H(213) = (213), (12)(213) = (13) H(123) = (123), (12)(123) = (23).
For this choice of G and H, there exist left cosets that are not right cosets and vice versa.
H 2.5.11 Let G have subgroups of orders n and m. Prove that G has a subgroup of order lcm(m, n).
Missing.
H 2.5.12 Let a G. Prove that N(a) = g G [ ga = ag is a subgroup of G. (N(a) is the normalizer
of a in G)
Let g, h N(a). Then (gh)a = gha = gah = agh = a(gh), so gh N(a). Additionally, multiply the
equation ga = ag on the left and right by g
1
to nd that g
1
gag
1
= g
1
agg
1
, so ag
1
= g
1
a. Hence
g
1
N(a). By lemma 2.3, N(a) G.
H 2.5.13 Prove that Z(G) = g G [ gx = xg for all x G G. (Z is the center of G)
The proof is identical to that of 2.5.12. Alternately, notice that Z(G) =

aG
N(a) G.
H 2.5.14 Let G = g) be cyclic and let H G. Prove that H is cyclic.
If H is trivial, the result holds. Therefore let H be non-trivial. Then there exists a smallest positive n
0
such
that g
n
0
H (remember H is closed under inverses, so it cant have just negative powers of g). The claim is
that H = g
n
0
). It is clear, because H is a group containing g
n
0
, that g
n
0
) H. Let g
k
H be arbitrary.
By the division algorithm, we can write k = qn
0
+r with q Z and r 0, 1, . . . , n
0
1. Now, by closure,
g
k
g
qn
0
= g
r
H. By our assumption that n
0
is the smallest positive exponent in H, we must conclude
that r = 0. Therefore n
0
[ k and we see that every element of H is a power of g
n
0
, whence H g
n
0
). This
proves the claim, so H = g
n
0
) is cyclic and the result is shown.
H 2.5.15 Let G be cyclic with [G[ = n. How many generators does G have?
Let g be a generator of G, i.e. G = g). The claim is that g
m
is also a generator if and only if m is
relatively prime to n. If m is relatively prime to n, then there exist a, b Z such that am + bn = 1. Now
g = g
am+bn
= (g
m
)
a
(g
n
)
b
= (g
m
)
a
which tells us that g g
m
) so G = g) g
m
) G. Hence m, n
relatively prime implies that g
m
generates G. On the other hand, if G = g
m
), then g g
m
), so that for
some a Z we have (g
m
)
a
= g. This is impossible if am + bn > 1 for all a, b Z, as would be the case if
m and n were not relatively prime. Hence G = g
m
) implies that m is coprime to n. This proves the claim,
and therefore the number of generators of G is (n), with the Euler totient function.
H 2.5.16 Let a G. If a
m
= e, prove that [a[ divides m.
By the division algorithm, we may write m = q[a[ + r with q Z and r 0, 1, . . . , [a[ 1. Now
e = a
m
= a
q|a|+r
= (a
|a|
)
q
a
r
= a
r
. As r < [a[ and [a[ is the minimal exponent taking a to the identity, we
must conclude that r = 0 and hence [a[ [ m.
H 2.5.17 Let a, b G be such that a
5
= e and aba
1
= b
2
. What is [b[?
We have that a
2
ba
2
= a(aba
1
)a
1
= ab
2
a
1
. It follows by induction that a
n
ba
n
= b
2
n
: suppose this is
true for n and compute
a
n+1
ba
(n+1)
= a(a
n
ba
n
)a
1
= ab
2
n
a
1
= b
2
n+1
.
The last equality follows from raising the condition aba
1
= b
2
to powers: (aba
1
)
k
= ab
k
a
1
= b
2k
.
Therefore a
5
ba
5
= b
32
, but, because a
5
= a
5
= e, the left hand side is simply b. We have, nally, b = b
32
,
or b
31
= e. By 2.5.16, the order of b divides 31, so [b[ = 1 or [b[ = 31.
H 2.5.18* Let G be nite, abelian and such that x
n
= e has at most n solutions for every n Z
+
. Prove
that G is cyclic.
Missing.
H 2.6.1* Let H G be such that the product (Ha)(Hb) is again a right coset of H for a, b G. Prove
that H G.
Consider the product (Ha)(Ha
1
) for arbitrary a G. As sets, it is true that (Ha)(Ha
1
) = H(aHa
1
):
(h
1
a)(h
2
a
1
) (Ha)(Ha
1
) is h
1
(ah
2
a
1
) H(aHa
1
) so (Ha)(Ha
1
) H(aHa
1
); conversely,
h
1
(ah
2
a
1
) H(aHa
1
) is (h
1
a)(h
2
a
1
) (Ha)(Ha
1
) so H(aHa
1
) (Ha)(Ha
1
). Then, by the con-
dition of the problem, (Ha)(Ha
1
) = H(aHa
1
) is a right coset of H. The set aHa
1
contains e = aea
1
,
so in fact H(aHa
1
) = He = H which implies that aHa
1
H, i.e. that H G.
H 2.6.2 Let H G have index 2. Prove that H G.
Let g GH. The right cosets of H in G may be enumerated as H, Hg. Because distinct cosets are
disjoint, Hg is exactly the set GH of elements in G not belonging to H. It is trivial that H = He = eH is
a left coset in addition to being a right coset. Furthermore, gH is again GH (The only alternative would
be gH = H, but that isnt the case: ge = g , H while ge gH.), so that gH = Hg. By lemma 2.10, H G
because every one of its right cosets in G is also a left coset in G.
H 2.6.3 Let H G and N G. Prove that NH G.
Let n
1
, n
2
N and h
1
, h
2
H. Compute (n
1
h
1
)(n
2
h
2
) = n
1
h
1
n
2
h
1
1
h
1
h
2
= n
1
(h
1
n
2
h
1
1
)h
1
h
2
. Because N
is normal, there exists n
3
N such that h
1
n
2
h
1
1
= n
3
. Then we have (n
1
h
1
)(n
2
h
2
) = (n
1
n
3
)(h
1
h
2
) NH
so NH is closed under the group product. Furthermore, NH is closed under inverses: Let n N and h H.
We have (nh)
1
= h
1
n
1
= h
1
n
1
hh
1
= n

h
1
NH, again using the existence of n

N such that
n

= h
1
n
1
h. Therefore, by lemma 2.3, NH is a group.
H 2.6.4 Let M, N G. Prove that M N G.
Let x M N (so x M, x N) and g G. Because M and N are normal, we see that gxg
1
M and
gxg
1
N so that gxg
1
M N. Therefore g(M N)g
1
M N and hence M N G.
H 2.6.5 Let H G and N G. Prove that H N H.
Let x H N and h H. We have hxh
1
H because it is a product of three elements of H. We also
have that hxh
1
N because N is normal and x N. Therefore hxh
1
H N so H N H.
H 2.6.6 Let G be abelian. Prove that every subgroup of G is normal.
Let H G and let h H, g G. Then ghg
1
= gg
1
h = h H, so H G.
H 2.6.7* If every subgroup of a group G is normal, is G necessarily abelian?
No. Consider the order 8 quaternion group Q
8
= 1, i j k with (1)
2
= 1, 1 commuting with
everything, and i
2
= j
2
= k
2
= ijk = 1. First observe that Q
8
is non-abelian: ij = (ij)(k
2
) = (ijk)k =
k while ji = j
1
i
1
= (ij)
1
= k
1
= k.
We can enumerate all the subgroups of Q
8
by considering subgroups generated by combinations of elements.
First of all, we have the trivial subgroups 1) = 1 and i, j, k) = Q
8
of orders 1 and 8 respectively. Both are
trivially normal. Next, the subgroup 1) = 1, 1 is of order 2. Finally, the subgroups i), j), k) of order
4 round out the list. We can see that to be the case by noting that 1 is redundant as a generator if we include
any of i, j, k because each already squares to 1. Furthermore, any subgroup generated by 2 or more of
i, j, k is all of Q
8
: for instance, i, j) = 1 = i
4
, 1 = i
2
, i, i = i
3
, j, j = j
3
, k = ij, k = ij
3
= Q
8
.
It is easy to see that 1) is normal because its elements commute with everything: i1) = i, i = 1)i,
and so forth (in fact, its the center of Q
8
. See 2.5.13). The rest of the subgroups are of order 4, so of index
2. By 2.6.2, i), j), k) are all normal. Therefore all subgroups of Q
8
are normal, but, as displayed above,
Q
8
is non-abelian.
H 2.6.8 Let H G. For g G, prove that gHg
1
G.
Let h
1
, h
2
H. We have gh
1
g
1
gh
2
g
1
= gh
1
h
2
g
1
gHg
1
because h
1
h
2
H. Furthermore,
(gh
1
g
1
)
1
= gh
1
1
g
1
gHg
1
because h
1
H. By lemma 2.3, gHg
1
G.
H 2.6.9 Let G be nite and let H G be the only subgroup of G of order [H[. Prove that H G.
The map f : H gHg
1
given by f(h) = ghg
1
is a bijection. Injectivity: f(h
1
) = f(h
2
) implies
gh
1
g
1
= gh
2
g
1
so h
1
= h
2
. Surjectivity: any ghg
1
gHg
1
is the image of h under f. Therefore
[gHg
1
[ = [H[. Now conjugation brings H into another subgroup, according to 2.6.8, and that subgroup
has the same cardinality as H, by assumption. If H is the only subgroup of order [H[, then conjugation xes
H, i.e. gHg
1
= H for all g G. This is the condition that H G. The restriction to nite G is likely just
to allow Hersteins use of the word order.
H 2.6.11 Let M, N G. Prove that NM G.
By 2.6.3, NM is a subgroup. Let g G and compute gNMg
1
= gNg
1
gMg
1
= NM. Hence NM G.
H 2.6.12* Let M, N G be such that M N = e. Prove that mn = nm for all m M and n N.
Consider the commutator [n, m] = nmn
1
m
1
. As M G and n G, there exists m

M such that
m

= nmn
1
. Hence nmn
1
m
1
= m

m
1
M. On the other hand, because N G and m G,
there exists n

N such that n

= mn
1
m
1
. Then nmn
1
m
1
= nn

N. Now we must have that


nmn
1
m
1
M N which is assumed to be trivial. Hence nmn
1
m
1
= e or, multiplying on the right by
m and then n, nm = mn.
H 2.6.13 Let T G be cyclic. Prove that every subgroup of T is normal in G.
Say T = t). By 2.5.14, a subgroup S of T is again cyclic, so let S = t
m
). Because T is normal, there
exists a k

such that gt
k
g
1
= t
k

T (k

depends on k and g). Then consider an element (t


m
)
k
of S.
g(t
m
)
k
g
1
= (gt
k
g
1
)
m
= (t
k

)
m
= (t
m
)
k

S. Therefore gSg
1
= S and S G.
H 2.6.14* Give an example of groups E F G with E F and F G but E not normal in G.
Missing.
H 2.6.15 Let N G. For a G, prove that [Na[ in G/N divides [a[ in G.
We have (Na)
|a|
= N(a
|a|
) = Ne = N, the identity in G/N. By 2.5.16, [Na[ divides [a[.
H 2.6.16 Let G be nite and let N G be such that [G : N] and [N[ are coprime. Prove that any element
x G satisfying x
|N|
= e must be in N.
First notice that x
[G:N]
N because the order of G/N is [G : N] so N = (Nx)
[G:N]
= Nx
[G:N]
. There exist
a, b Z such that a[G : N] +b[N[ = 1. Let x G be such that x
|N|
= e. Now
x = x
a[G:N]+b|N|
= (x
[G:N]
)
a
(x
|N|
)
b
= (x
[G:N]
)
a
N.
H 2.7.1 Are the following maps homomorphisms? If yes, what are their kernels? (a) G = R0 under
multiplication, : G G given by (x) = x
2
. (b) G as in (a), : G G given by (x) = 2
x
. (c) G = R
under addition, : G G given by (x) = x + 1. (d) G as in (c), : G G given by (x) = 13x. (e) G
any abelian group, : G G given by (x) = x
5
.
(a) Yes: (xy) = (xy)
2
= x
2
y
2
= (x)(y). The identity in G is 1, so the kernel is the set of those x G
such that (x) = x
2
= 1. Hence ker = 1, 1.
(b) No: (2 3) = 2
6
= 64 while (2)(3) = 2
2
2
3
= 32.
(c) No: (x +y) = x +y + 1 while (x) +(y) = (x + 1) + (y + 1) = x +y + 2.
(d) Yes: (x +y) = 13(x +y) = 13x + 13y = (x) +(y). The identity in G is 0, so the kernel is the set of
those x G such that (x) = 13x = 0. Hence ker = 0.
(e) Yes: (xy) = (xy)
5
= x
5
y
5
= (x)(y). ker = x G [ x
5
= e.
H 2.7.2 Let : G G be given by (x) = gxg
1
for xed g G. Prove is an isomorphism.
is a homomorphism: (xy) = gxyg
1
= gx(g
1
g)yg
1
= (x)(y). The kernel of is trivial: if x G is
such that gxg
1
= e, then x = e. Hence ker = e and therefore is an isomorphism.
H 2.7.3 Let G be nite and let n Z be relatively prime to [G[. Prove that every x G may be written
as g = x
n
with x G.
There exist a, b Z such that an +b[G[ = 1. For g G, we have g = g
an+b|G|
= (g
a
)
n
.
Herstein assumes, additionally, that G is abelian and hints to consider the map : G G given by
(y) = y
n
. This map is a homomorphism when G is abelian, but is not in general. The map, however, is
always surjective, as shown above (g = (g
a
)). Therefore , the n-th power map, is a bijection for any nite
G and choice of n coprime to [G[. In that context, it makes sense to talk about a well-dened, unique nth
root of a group element.
H 2.7.4 Let U G. Let U) G be the smallest subgroup of G containing U. (a) Prove that such a U)
exists. (b) If gug
1
U for all g G and u U, prove that U) G.
(a) Let / be the collection of all subgroups V of G which contain U. G /, so / is not empty. U) =

V A
V
is a subgroup of Gwhich contains U and is a subset of every element of /. Hence any subgroup of Gcontaining
U also contains U). In this sense, U) ts the criterion.
(b) By lemma 3 (below), U) is exactly the set of all nite products of elements of U and their inverses.
Note that if gug
1
= u

U then taking the inverse gives gu


1
g
1
= u
1
U). Now an arbitrary element
of U) may be written as u
k
1
1
u
k
n
n
(with n 0 and k
i
= 1) so conjugation gives
gu
k
1
1
u
k
n
n
g
1
= gu
k
1
1
g
1
g g
1
gu
k
n
n
g
1
= u
k
1
1
u
k
n
n
U),
where u

i
= gu
i
g
1
. Therefore U) G.
Lemma 3 Dene V = u
k
1
1
u
k
2
2
u
k
n
n
G [ n 0, 1, , u
i
U, k
i
= 1, the set of all nite (or
empty, in which case the result is e) products of elements from U or their inverses. Then V = U).
Proof: As U) is a subgroup containing U, it also contains all inverses of elements of U. Furthermore,
it is closed under multiplication, so V U) immediately. Notice that V is both trivially closed under
multiplication and non-empty (because when n = 0 we see that e V ). Furthermore, the inverse of
u
k
1
1
u
k
n
n
is u
k
n
n
u
k
1
1
which is again in V . By lemma 2.3, V G. Now U V , so, by (a), U) V .
Therefore V = U) and the claim is proven.
H 2.7.5 Let U = xyx
1
y
1
[ x, y G. Dene G

= U) (the commutator subgroup of G). (a) Prove


that G

G. (b) Prove that G/G

is abelian. (c) If G/N is abelian, prove that G

N. (d) Prove that, if


H G

then H G.
(a) Let g, x, y G and compute the conjugate of an element xyx
1
y
1
U:
gxyx
1
y
1
g
1
= gxg
1
gyg
1
gx
1
g
1
gy
1
g
1
= (gxg
1
)(gyg
1
)(gxg
1
)
1
(gyg
1
)
1
U.
By 2.7.4b, G

= U) G.
(b) We have, for x, y G, that (G

x)(G

y)(G

x
1
)(G

y
1
) = G

(xyx
1
y
1
) = G

. Right multiplying the


equation by G

y and then by G

x gives that (G

x)(G

y) = (G

y)(G

x), so G/G

is abelian.
(c) If G/N is abelian, then, for x, y G we have (Nx)(Ny)(Nx
1
)(Ny
1
) = (Nx)(Nx
1
)(Ny)(Ny
1
) =
N(xx
1
yy
1
) = N. Hence xyx
1
y
1
N. Then U N and therefore G

N by virtue of being the


smallest subgroup containing U.
(d) Let g G and h H. Because G

H, we have that ghg


1
h
1
H. Therefore ghg
1
H, so H G.
H 2.7.6 Let M, N G. Prove that NM/M

= N/(N M).
By 2.6.5, N M N. It is trivial that M NM because M G and NM G. Therefore, with the
knowledge that all quantities are meaningful, dene : NM/M N/(N M) by (nmM) = n(N M)
for n N, m M. The map is well-dened: suppose n
1
m
1
M = n
2
m
2
M for n
1
, n
2
N and m
1
, m
2
M.
In fact, this also means that n
1
M = n
2
M, so that n
1
= n
2
m
3
for some m
3
M. However, also notice
that m
3
= n
1
2
n
1
N, so m
3
M N. Then n
1
(N M) = n
2
m
3
(N M) = n
2
(N M) and the map is
well-dened as claimed.
is a homomorphism: let n
1
, n
2
N and m
1
, m
2
M. Then ((n
1
m
1
M)(n
2
m
2
M)) = (n
1
Mn
2
M) =
(n
1
n
2
M) = n
1
n
2
(N M) = n
1
(N M)n
2
(N M) = (n
1
m
1
M)(n
2
m
2
M). Finally, the kernel of is
trivial: suppose (nmM) = N M for n N and m M. This means that n N M M, i.e. we have
nmM = M. Therefore is an isomorphism and the result is proven.
H 2.7.7 For a, b R, let
ab
: R R be given by
ab
(x) = ax +b. Let G =
ab
[ a, b R, a ,= 0 and let
N =
1b
G. Prove that N G and that G/N

= R0 under multiplication.
By 2.5.7, G is a group under composition,
ab

cd
=
ac,ad+b
and
1
ab
= 1
a
,
b
a
. Then, with
ab
G,

ab

1c

1
ab
=
a,ac+b
1
a
,
b
a
=
1,ac
N
so that N G. Dene : G/N R0 by (
ab
N) = a. is well-dened: suppose
ab
N =
cd
N so that

cd
=
ab

1e
for some e R. Then
cd
=
a,ae+b
so that c = a and (
cd
N) = c = a as required. is a
homomorphism: (
ab
N
cd
N) = (
ab

cd
N) = (
ac,ad+b
N) = ac = (
ab
)(
cd
). The kernel of is trivial:
if (
ab
N) = 1, then a = 1 so that
ab
N, i.e.
ab
N = N. Therefore is an isomorphism and the result is
proven.
H 2.7.8 Let D
2n
be the dihedral group with generators x, y such that x
2
= y
n
= e and xy = y
1
x. (a)
Prove that y) D
2n
. (b) Prove that D
2n
/y)

= Z
2
.
(a) y) has order n, and D
2n
has order 2n, so y) has index 2n/n = 2 in D
2n
. By 2.6.2, y) D
2n
.
(b) The order of D
2n
/y) is 2, and the unique group of order 2, up to isomorphism, is Z
2
. More constructively,
: D
2n
/y) Z
2
given by (x
i
y
j
Y )) = i is an isomorphism.
H 2.7.9 Prove that Z(G) G. (see 2.5.13)
For g G, z Z(G), we have gzg
1
= zgg
1
= z Z(G), so Z(G) G.
H 2.7.10 Prove that a group of order 9 is necessarily abelian.
Missing.
H 2.7.11 Let G be non-abelian with order 6. Prove that G

= S
3
.
The non-identity elements of G have order 2, 3 or 6 by Lagranges theorem. If G contained an element of
order 6, then G would be cyclic and hence abelian. Therefore if G is non-abelian, its non-identity elements
have orders 2 and/or 3. Lets recall S
3
= e, (12), (13), (23), (123), (213). It contains 2 elements of order 3
and 3 elements of order 2. Its multiplication table is:
S
3
e (123) (213) (12) (13) (23)
e e (123) (213) (12) (13) (23)
(123) (123) (213) e (13) (23) (12)
(213) (213) e (123) (23) (12) (13)
(12) (12) (23) (13) e (213) (123)
(13) (13) (12) (23) (123) e (213)
(23) (23) (13) (12) (213) (123) e
Suppose G has no elements of order 3. Then all of its elements have order 2, but this forces G to be abelian
by 2.3.10. Therefore there must exist an element a G with order 3. Let b G be distinct from e, a, a
2
.
If the order of b is 3, then e, a, a
2
, b, b
2
contains no duplicates: if b
2
= a
2
then multiplying on the left by
a and on the right by b gives a = b, which is a contradiction; if b
2
= a then squaring gives a
2
= b
4
= b,
which is a contradiction. Hence we have 5 elements of the group in the case that b has order 3. Consider
ab: the cases ab = e, ab = a, ab = a
2
, ab = b and ab = b
2
all yield immediate contradictions, so we must
conclude that ab is the sixth element of the group. Furthermore, the same argument has us conclude that
ba is none of e, a, a
2
, b, b
2
, so ab = ba. For elements u, v of an arbitrary group, uv = vu easily implies that
u
m
v
n
= v
n
u
m
. In our case, the fact that ab = ba gives us that G is abelian, so this is not the desired group.
Therefore a and a
2
are the only elements of order 3, and the remaining elements of G all have order 2:
G = e, a, a
2
, b, c, d with b
2
= c
2
= d
2
= e. Consider the product ab. Each of the possibilities ab = e,
ab = a, ab = a
2
, and ab = b gives an immediate contradiction, so ab c, d, and, by a similar argument,
ac b, d and ad b, c. There are two distinct situations possible: ab = c, ac = d, ad = b and
ab = d, ac = b, ad = c. From those congurations, its easy to construct the entire multiplication table for
the two Gs:
G
1
e a a
2
b c d
e e a a
2
b c d
a a a
2
e c d b
a
2
a
2
e a d b c
b b d c e a
2
a
c c b d a e a
2
d d c b a
2
a e
G
2
e a a
2
b c d
e e a a
2
b c d
a a a
2
e d b c
a
2
a
2
e a c d b
b b c d e a a
2
c c d b a
2
e a
d d b c a a
2
e
Comparing to the group table for S
3
, we see that : G
1
S
3
given by (e) = e, (a) = (123), (a
2
) = (213),
(b) = (12), (c) = (13), (d) = (23) is a homomorphism. It is also a bijection, so is an isomorphism,
proving the equivalence of G
1
to S
3
. On the other hand, : G
2
S
3
given by (e) = e, (a) = (213),
(a
2
) = (123), (b) = (12), (c) = (13), (d) = (23) is again a bijective homomorphism, giving the
equivalence of G
2
to S
3
. These were the only two possible non-abelian groups of order 6, and both are
isomorphic to S
3
, so the claim is proven.
H 2.7.12 Let G be abelian and let N G. Prove that G/N is abelian.
Because G is abelian, N is normal in G and G/N is a group. Let g
1
, g
2
G. Then (g
1
N)(g
2
N) = (g
1
g
2
)N =
(g
2
g
1
)N = (g
2
N)(g
1
N) so that G/N is abelian.
H 2.8.1 Are the following maps automorphisms? (a) G = Z under addition, T : x x. (b) G = R
+
under multiplication, T : x x
2
. (c) G cyclic of order 12, T : x x
3
. (d) G = S
3
, T : x x
1
.
(a) Yes. Let a, b Z. T is a homomorphism: we have T(a+b) = (a+b) = (a) +(b) = T(a) +T(b). T is
injective: if T(a) = T(b), then a = b so a = b. T is surjective: a = T(a). Hence T is an automorphism.
(b) Yes. Let x, y R
+
. T is a homomorphism: we have T(xy) = (xy)
2
= x
2
y
2
= T(x)T(y). T is injective: if
T(x) = T(y), then x
2
= y
2
so x = y because we restrict to positive reals. T is surjective: x = T(

x) where

x is the unique, positive square root which exists for all positive x.
(c) No. Say G = g). Then g
4
,= e, but T(g
4
) = g
12
= e, so the kernel of T is non-trivial. Hence T is not
injective and thus not an isomorphism.
(d) No. T[(12)(13)] = T[(213)] = (123), while T[(12)]T[(13)] = (12)(13) = (213).
H 2.8.2 Let H G, and let T be an automorphism of G. Prove T(H) G.
For h
1
, h
2
H, we have T(h
1
)T(h
2
) = T(h
1
h
2
) T(H) and T(h
1
)
1
= T(h
1
1
) T(H). By lemma 2.3,
T(H) G.
H 2.8.3 Let N G, and let T be an automorphism of G. Prove T(N) G.
By 2.8.2, T(N) G. Let g G and n N. Then gT(n)g
1
= T(T
1
(g)nT
1
(g
1
)) = T(n

) T(N) for
some n

= T
1
(g)nT
1
(g
1
) N because N is normal.
H 2.8.4 Prove that Inn(S
3
)

= S
3
.
The center of S
3
is trivial, as we can easily check: (12)(13) = (213) while (13)(12) = (123), so (12) , Z(S
3
)
and (13) , Z(S
3
). (23)(123) = (13) while (123)(23) = (12) so (13) , Z(S
3
) and (123) , Z(S
3
). Finally,
(12)(213) = (13) while (213)(12) = (23), so (213) , Z(S
3
). Therefore Z(S
3
) = e and, by lemma 2.19,
Inn(S
3
)

= S
3
/Z(S
3
)

= S
3
.
H 2.8.5 Prove that Inn(G) Aut(G).
Let g G, T
w
: x wxw
1
Inn(G) and Aut(G). Then (T
w

1
)(g) = (T
w
(
1
(g))) =
(w
1
(g)w
1
) = (w)g(w
1
) = (w)g(w)
1
= T
(w)
(g), where T
(w)
: x (w)x(w)
1
. Hence
T
w

1
= T
(w)
Inn(G), which proves that Inn(G) Aut(G).
H 2.8.6 Let G = e, a, b, ab be a group of order 4 with a
2
= b
2
= e and ab = ba. Determine Aut(G).
An automorphism of G xes the identity, and permutes the three elements of order 2. Then it is clear that
Aut(G) is isomorphic to a subgroup of S
3
. Furthermore, we can exhibit elements of order 2 and 3 in Aut(G)
which proves that Aut(G)

= S
3
in its entirety.
For example, : G G given by (a) = b, (b) = a and (ab) = ab is an automorphism of order
2. To see that is a homomorphism, we just need to check that (ab) = ab = (b)(a) = (a)(b),
(aab) = a = bab = (a)(ab) and (bab) = b = aab = (b)(ab). All other possible products automatically
work because the group is abelian. Therefore is a homomorphism, and we see easily that
2
(a) = (b) = a
and
2
(b) = (a) = b, so
2
= id and hence is an order 2 automorphism.
Additionally, : G G given by (a) = b, (b) = ab and (ab) = a is an automorphism of order 3. We
can check that is a homomorphism: (ab) = a = bab = (a)(b), (aab) = ab = (ab)(a) = (a)(ab),
and (bab) = b = aba = (b)(ab). To check the order of , we raise it to powers:
2
(a) = (b) = ab,

2
(b) = (ab) = a and
2
(ab) = (a) = b. Therefore
2
,= id, but
3
(a) = (ab) = a,
3
(b) = (a) = b
and
3
(ab) = (b) = ab so, in fact,
3
= id.
As a result, we know that Aut(G) must be isomorphic to a subgroup of S
3
which can accomodate elements
of orders 2 and 3. This forces the subgroup to have order at least 6 = lcm(2, 3), but thats the order of S
3
itself. Hence Aut(G)

= S
3
.
H 2.8.7 Let C G. C is characteristic if (C) C for all Aut(G). (a) Prove that a characteristic
subgroup is normal. (b) Prove that the converse of (a) is false.
(a) Suppose C is characteristic and let g G. gCg
1
is the image of C under the inner automorphism
T
g
: x gxg
1
. Because T
g
Aut(G), we have that gCg
1
= T
g
(C) C. This holds for arbitrary g G,
so C G.
(b) A normal subgroup, N G, is xed by all inner automorphisms, by denition. In order for the converse
to fail to hold, there need to be automorphisms outside of Inn(G) which dont x N. Consider G = V
4
, the
Klein group of 2.8.6, and N = a) = e, a which is normal because it is of index 2. Then take the order 2
automorphism of G, with (a) = b, (b) = a and (ab) = ab. We have (a)) = e, b , = a). Hence N is
normal, but not characteristic because doesnt x it.
H 2.8.8 Prove that the commutator subgroup G

= aba
1
b
1
[ a, b G) (see 2.7.5) is characteristic.
Let Aut(G) and let a
1
b
1
a
1
1
b
1
1
a
n
b
n
a
1
n
b
1
n
G

. Because is a homomorphism, we have


(a
1
b
1
a
1
1
b
1
1
a
n
b
n
a
1
n
b
1
n
) = (a
1
)(b
1
)(a
1
)
1
(b
1
)
1
(a
n
)(b
n
)(a
n
)
1
(b
n
)
1
G

.
Therefore (G

) G

, so G

is characteristic.
H 2.8.9 Let N G and let M be a characteristic subgroup of N. Prove M G.
Every inner automorphism xes N but, a priori, could move M N around. Notice that if T
g
Inn(G),
it restricts to an automorphism T
g
[
N
of N because T
g
(N) = N. Now M is characteristic in N, so we have
T
g
[
N
(M) = M. This is simply a restriction of a map to the domain N, so, in fact, T
g
(M) = M as well.
Hence M is xed by every inner automorphism, i.e. M G.
Some remarks: If N had not been normal, then inner automorphisms T
g
wouldnt restrict to automorphisms
of N (because T
g
might throw some elements of N aeld), and the argument breaks down. Also, if N were
normal but M merely normal (not characteristic) in N, the argument would also break down: T
g
restricts
to an automorphism of N which is not, in general, an inner automorphism of N (not unless g N, in fact).
Thus the restriction T
g
[
N
would not x M if it were only normal but not characteristic. Therefore this
exercise represents a strengthening of the conditions of 2.6.14, where we see that normality of subgroups is
not a transitive property.
H 2.8.10 Let G be nite and let Aut(G) x only the identity. Prove that every g G may be written
as g = x
1
(x) for some x G.
Consider the map : G G given by (x) = x
1
(x). This map is injective: let x, y G be such that
(x) = (y). Then x
1
(x) = y
1
(y). Multiplying on the left by y, on the right by (x
1
), and using
the fact that is a homomorphism, we have yx
1
= (yx
1
). By our assumption on , we must conclude
that yx
1
= e, so that x = y. This proves injectivity. Because G is nite, is also necessarily surjective
(however, is not a homomorphism in general, and this is immaterial). Therefore for every g G, there
exists an x G such that g = (x) = x
1
(x).
H 2.8.11 Let G be nite, and let Aut(G) x only the identity. Suppose additionally that
2
= id.
Prove that G is abelian.
Let g G. By 2.8.10, there exists x G such that g = x
1
(x). Now (g) = (x
1
)
2
(x) = (x)
1
x =
(x
1
(x))
1
= g
1
. Suppose G were not abelian, with a, b G such that ab ,= ba. Then (ab) = b
1
a
1
while (a)(b) = a
1
b
1
. If we were to have that (ab) = (a)(b), then inverting b
1
a
1
= a
1
b
1
would
produce the contradiction ab = ba. Therefore, because is assumed to be a homomorphism, G must be
abelian.
Remark: In fact, if G is a group then G is abelian if and only if : g g
1
is an automorphism of G.
is trivially a bijection (uniqueness of inverses gives injectivity, and g = (g
1
) gives surjectivity). If G is
abelian, then is a homomorphism because (gh) = h
1
g
1
= g
1
h
1
= (g)(h). If G is not abelian,
then is not a homomorphism by the above argument. The restriction to nite G in the problem statement
enables us to use the obscure (in my opinion) result 2.8.10.
H 2.8.12* Let G be nite, and let Aut(G) be such that (x) = x
1
for at least three quarters of the
elements of G. Prove that (x) = x
1
for all x in G and that G is abelian.
Missing.
H 2.8.13 Give an example of a non-abelian nite group G with an automorphism that maps exactly three
quarters of the elements of G to their inverses.
Missing.
H 2.8.14* Let G be nite with [G[ > 2. Prove Aut(G) is non-trivial.
If G is abelian, then inversion : g g
1
is an automorphism (see 2.8.11). If = id, then g = g
1
for
all g G, i.e. all non-identity elements have order 2. In that case, construct the map : G G which
transposes two non-identity elements and xes everything else. is an automorphism (this needs to be
checked!). Therefore, if G is abelian, either or is a non-trivial automorphism of G.
If G is non-abelian, take an element a , Z(G). Then there exists g G with ag ,= ga, and therefore
aga
1
,= g. Thus the inner automorphism T
a
: g aga
1
of conjugation by a does not x g, so T
a
,= id but
T
a
Aut(G).
H 2.8.15* Let G have even order 2n. Suppose that exactly half of the elements of G have order 2 and the
rest form a subgroup H of order n. Prove that [H[ is odd and that H is abelian.
If H were of even order, then by 2.3.11 or Cauchys theorem, it would contain an element of order 2. It is
assumed that H is the collection of elements with order dierent from 2, so [H[ must be odd.
Let x , H. Then xh , H for any h H. Furthermore, xh has order 2, because it is outside of H. The map
: G G given by (g) = xgx
1
= xgx is an inner automorphism of G. Because H is normal (index 2),
xes H and hence restricts to an automorphism [
H
of H. Now we see that h(h) = hxhx = (hx)
2
= e, so
that [
H
(h) = h
1
. Therefore inversion is an automorphism on H, and, as seen in 2.8.11, this gives that H
is abelian.
H 2.8.16* Let (n) be the Euler -function. Let a Z, with a > 1. Prove that n [ (a
n
1).
Consider Z

a
n
1
, the multiplicative (modulo a
n
1) group of integers coprime to (a
n
1). The number of
elements in Z

a
n
1
is (a
n
1). Furthermore, a is coprime to a
n
1 because (a
n1
)a + (1)(a
n
1) = 1 so
that a Z

a
n
1
. The order of a is n because a
n
1 mod (a
n
1) while, for 1 < k < n, 1 < a
k
< a
n
1.
By Lagrange, this order must divide the order of the group, and therefore n [ (a
n
1).
H 2.9.1 Let g G. Dene
g
: G G by
g
(x) = gx. Prove that
g
is a bijection and that
g

h
=
gh
.

g
is surjective: if y G, then y = gg
1
y =
g
(g
1
y).
g
is injective: if x, y G, then
g
(x) =
g
(y) implies
that gx = gy so x = y. Finally, if g, h G, then
gh
(x) = ghx while
g
(
h
(x)) =
g
(hx) = ghx. Therefore

gh
=
g

h
. Hence each
g
is a permutation of G and the
g
form a subgroup (under composition) of the
set of bijections G G.
H 2.9.2 Let
g
be dened as in 2.9.1. Dene
g
: G G by
g
(x) = xg. Prove that, for g, h G, we
have
g

h
=
h

g
.
Let x G. We have
g
(
h
(x)) =
g
(xh) = gxh while
h
(
g
(x)) =
h
(gx) = gxh. Hence
g

h
=
h

g
.
H 2.9.3 Let
g
and
g
be dened as in 2.9.2. If : G G is a bijection such that
g
=
g
for all
g G, prove that =
h
for some h G.
For x G,
g
((x)) = g(x) while (
g
(x)) = (gx). Note that we dont assume to be a homomorphism.
However, we see that (gx) = g(x) for all x, g G. If we pick g = x
1
, then we nd that (e) = x
1
(x) for
all x G. Solving for (x) yields (x) = x(e), which is the desired result. Specically, we have =
(e)
.
H 2.9.4 Let H G. (a) Show that gHg
1
G for every g G. (b) Prove that W =

gG
gHg
1
is a
normal subgroup of G.
(a) Let h
1
, h
2
H and let g G. We have gh
1
g
1
gh
2
g
1
= gh
1
h
2
g
1
gHg
1
, so gHg
1
is closed under
multiplication. Additionally, the inverse of gh
1
g
1
is gh
1
1
g
1
gHg
1
. By lemma 2.3, gHg
1
G.
(b) W G because it is the intersection of subgroups. Suppose w W so that, for every g G, there exists
h H such that w = ghg
1
. Let x G, and consider xwx
1
. For any g G, there exists h H such
that w may be written as w = (x
1
g)h(x
1
g)
1
= x
1
ghg
1
x. Consequently, xwx
1
= xx
1
ghg
1
xx
1
=
ghg
1
gHg
1
. This can be done for arbitrary x, g G, so xwx
1
W and W G.
H 2.9.5 Let [G[ = p
2
. Prove that G has a normal subgroup of order p.
Lemma 2.21 states that if G is a nite group, and H G is a proper subgroup such that [G[ [G : H]!, then
H must contain a non-trivial normal subgroup of G. The lemma is proven by considering the action of left
multiplication by G on the set G/H of left cosets of H in G.
As p
2
is not prime, we know by 2.5.3 that G has a proper subgroup H. By Lagrange, this subgroup must
have order p. By the fundamental theorem of arithmetic, [G[ = p
2
does not divide [G : H]! = p!, so the
conditions of lemma 2.21 are satised. Therefore H must contain a non-trivial normal subgroup K of G. By
Lagrange, this subgroup must have order p, i.e. K = H. Thus H G and [H[ = p.
H 2.9.6* Let [G[ = p
2
and let H G with [H[ = p. Prove that H Z(G).
If G contains an element of order p
2
, then G is cyclic and hence abelian. In that case, Z(G) = G so
H Z(G). Otherwise, all non-identity elements of G have order p. Let x H. In fact, because H has
prime order, it is cyclic and H = x). Let y G. By lemma 2.21 (as used in 2.9.5), y) G. Lagranges
theorem applied to x) y) (as a subgroup of x) or y)) tells us that it has order 1 or p. In other words,
x) = y) = x) y) or x) y) = e. In the former case, x) = y) gives that x and y commute. In the
latter case, x) y) = e invites the application of 2.6.12, which again gives that xy = yx. As x H and
y G were arbitrary, this shows that H Z(G).
H 2.9.7* Let [G[ = p
2
. Prove that G is abelian.
If G contains an element of order p
2
, then G is cyclic and hence abelian. Suppose G contains no element
of order p
2
. Then, by Lagrange, every element g G has order p. By lemma 2.21 (as used in 2.9.5), g) is
normal and, by 2.9.6, g) Z(G). In particular, g Z(G). This argument applies to arbitrary g G, so
Z(G) = G, i.e. G is abelian.
H 2.9.8 Let [G[ = 2p. Prove G has a subgroup H of order p and that H G.
If x G has order 2p, then x
2
has order p, so H = x
2
) is a subgroup of order p. Otherwise, suppose there
is no element in G of order 2p. If x G has order p, then H = x) is a subgroup of order p. Otherwise,
suppose G has no elements of order p or 2p. Then all non-identity elements of G have order 2. Therefore G
is abelian by 2.3.10, and we reach a contradiction: if G is abelian, then Cauchys theorem for abelian groups
implies the existence of an element x G of order p. Therefore G necessarily has a subgroup H of order p.
The index of H is 2, and so it is normal by 2.6.2.
H 2.9.9 Let [G[ = pq, where p ,= q are both prime. Suppose H, K G with [H[ = p and [K[ = q. Prove
that G is cyclic.
Let H = h) and K = k) so h
p
= e and k
q
= e. If x h) k) then [x[ [ p and [x[ [ q, so [x[ = 1 and
consequently h) k) = e. By 2.6.12, hk = kh. Now consider the element hk. Its order must be 1, p,
q or pq. As k , h), we cannot have k = h
1
, so hk ,= e. Compute (hk)
p
= h
p
k
p
= k
p
,= e because q p.
Similarly, (hk)
q
= h
q
k
q
= h
q
,= e because p q. Therefore [hk[ = pq, and G = hk) is cyclic.
H 2.9.10* Let [G[ = pq, where p > q are both prime. (a) Prove that G has a subgroup of order p and a
subgroup of order q. (b) Prove that if q (p 1) then G is cyclic. (c) Prove that, given two primes p, q with
q (p 1), there exists a non-abelian group of order pq. (d) Prove that any two non-abelian groups of order
pq are isomorphic.
Missing.
H 2.10.1, 2.10.2 Decompose into products of disjoint cycles:
(a)

1 2 3 4 5 6 7 8 9
2 3 4 5 1 6 7 9 8

(b)

1 2 3 4 5 6
6 5 4 3 1 2

(a) (12345)(6)(7)(89).
(b) (1625)(34).
H 2.10.3 Express as products of disjoint cycles: (a) (15)(16789)(45)(123). (b) (12)(123)(12).
Herstein composes permutation from left to right, while I compose permutations from right to left. This
problem statement has been modied accordingly. Disjoint cycles commute, so this is often irrelevant.
(a) Consider the action of each cycle in order. Starting with 123456789, (123) sends this into 312456789.
(45) sends this into 312546789. (16789) sends this into 912543678. (15) sends this into 412593678. Hence
(15)(16789)(45)(123) = (123678954).
(b) Starting with 123, (12) sends this into 213. (123) sends this into 321. (12) sends this into 231. Therefore
(12)(123)(12) = (132).
H 2.10.4 Prove that (12 n)
1
= (n, n 1, n 2, . . . , 2, 1)
First note that a transposition (2-cycle) is its own inverse. Its straightforward to check by hand Hersteins
assertion, pg. 67, that (a
1
, a
2
, . . . , a
m
) = (a
1
, a
m
)(a
1
, a
m1
) (a
1
, a
2
). Therefore
(12 n)
1
= [(1n)(1, n 1) (12)]
1
= (12) (1n) = (1, n, n 1, . . . , 2) = (n, n 1, n 2, . . . , 2, 1).
H 2.10.5 Find the cycle structure of all the powers of (12 8)
Let = (12 8). Well compute what
i
does to 12345678. The sequence goes
12345678

81234567

78123456

67812345

56781234

45678123

34567812

23456781
Then we have = (12345678),
2
= (1357)(2468),
3
= (14725836),
4
= (15)(26)(37)(48),
5
= (16385274),

6
= (1753)(2864),
7
= (18765432) and
8
= e.
H 2.10.6 (a) What is the order of an n-cycle? (b) What is the order of the product of disjoint cycles of
lengths m
1
, m
2
, . . . , m
k
? (c) How do you nd the order of a given permutation?
(a) Let be an n-cycle. If
k
(i) = i for some k < n, then can be decomposed into smaller cycles because
i has an orbit which is a cardinality k subset of 1, 2, . . . , n. Furthermore, by the pigeonhole principle,

n
(i) = i. Hence the order of an n-cycle is n.
(b) Let
1
, . . . ,
k
be disjoint cycles. As disjoint cycles commute, (

i
)
m
=

m
i
. The order of

i
is
therefore N = lcm
i
(m
i
), the least common multiple of the orders (lengths, by (a)) of all the
i
. It is clear
that (

i
)
N
= e because m
i
[ N for all i. For any smaller exponent K, there is at least one
i
with m
i
K
so that (

i
)
K
,= e.
(c) Decompose the permutation into disjoint cycles. The order of the permutation is the least common
multiple of the lengths of the disjoint cycles.
H 2.10.7 Compute a
1
ba where (a) a = (12)(135), b = (1579). (b) a = (579), b = (123).
(a) a = (1352) and a
1
= a
3
= (1253). Compute a
1
ba by applying each permutation in order.
a(123456789) = 251436789, b(251436789) = 951426387, and a
1
(951426387) = 192456387. Therefore
a
1
ba = (2379).
(b) a
1
= a
2
= (597). a(123456789) = 123496587, b(123496587) = 312496587, and a
1
(312496587) =
312456789. Therefore a
1
ba = (123) = b. We could have seen this immediately because a and b are disjoint.
H 2.10.8 (a) For x = (12)(34) and y = (56)(13), nd a permutation a such that a
1
xa = y. (b) Prove
there is no a such that a
1
(123)a = (13)(578). (c) Prove there is no a such that a
1
(12)a = (15)(34)
(a) We must have xa = ay. Then x(a(1)) = a(y(1)) = a(3), x(a(3)) = a(y(3)) = a(1), x(a(2)) = a(y(2)) =
a(2), and so on. We see that a(2) and a(4) must be xed by x, so a(2), a(4) 5, 6. On the other hand,
x should transpose a(3) with a(1) and a(5) with a(6). There is freedom in creating a, but the following
choice works: a(1) = 1, a(3) = 2, a(5) = 3, a(6) = 4, a(2) = 5, and a(4) = 5. This may also be written as
a = (253)(46). This a works, though it is not unique. For example, a = (1263)(45) is another solution.
(b) Let x = (123) and y = (13)(578). Applying the same analysis as above, we see that x(x(a(1))) = a(1)
while x(a(1)) = a(3) ,= a(1). Hence a(1) must be brought on an orbit of length 2 by x, but it is clear that x
does this to no symbol. The cycles of x are all of length 1 or 3. Therefore no such a exists. A more direct
way to see this is that [a
1
xa[ = [x[ = 3 while [y[ = 6 by 2.10.6.
(c) Let x = (12) and y = (15)(34). We have x(a(1)) = a(5), x(a(2)) = a(2), x(a(3)) = a(4), x(a(4)) = a(3),
and x(a(5)) = a(1). Only a(2) is xed here, but x must x three of its ve inputs. This is a contradiction,
and so it is impossible to construct such an a.
H 2.10.9 For what m is an m-cycle an even permutation?
The m-cycle = (a
1
a
m
) = (a
1
a
m
) (a
1
a
3
)(a
1
a
2
) may be written as the product of m 1 transposi-
tions. While this decomposition into transpositions is not unique, the parity (even/odd) of the number of
transpositions is (see pg. 67). Then is an even permutation if and only if m1 is even, i.e. if m is odd.
H 2.10.10 What are the parities of the following permutations? (a) (12)(123). (b) (45)(123)(12345). (c)
(25)(14)(13)(12).
(a) (12)(123) = (12)(13)(12) is odd, as it is the product of three transpositions. Alternately, because sgn is
multiplicative, sgn((12)(123)) = (1)(+1) = 1.
(b) (45)(123)(12345) = (45)(13)(12)(15)(14)(13)(12) is the product of seven transpositions, so it is odd.
Again, we can also compute sgn((45)(123)(12345)) = (1)(+1)(+1) = 1.
(c) (25)(14)(13)(12) is the product of four transpositions, so it is even.
H 2.10.11 Prove that S
n
= (12 n), (12)).
Let = (12 n), = (12) and U = , ). The result of applying to n symbols is to rotate them all to
the right by one spot. Similarly, applying
1
eects a left rotation by one. Compute
1
by considering
its action on 12 n:
123 n

1
234 n1

324 n1

1324 n.
We see that
1
= (23). Computing
2

2
, a pattern becomes clear which suggests that
k

k
=
(k + 1, k + 2). The base case k = 1 was just shown, so assume this result to be true. Then
k+1

(k+1)
=
(k + 1, k + 2)
1
. Observe the action of this permutation:
1 2 3 k + 1 k + 2 k + 3 n 1 n

1
2 3 4 k + 2 k + 3 k + 4 n 1
(k + 1, k + 2) 2 3 4 k + 3 k + 2 k + 4 n 1
1 2 3 k + 1 k + 3 k + 2 n 1 n
from which we see that (k + 1, k + 2)
1
= (k + 2, k + 3). Therefore the claim is proven that
k

k
=
(k + 1, k + 2). In particular, (k, k + 1) U for k 1, 2, . . . , n 1.
Lets investigate the product (ab)(bc)(ab). This permutation sends abc bac bca cba, so (ab)(bc)(ab) =
(ac). This is useful because we can write (1k) = (1, k 1)(k 1, k)(1, k 1) where we use the just-discovered
fact that (k1, k) U. Iterating this procedure, starting with k = 3 (because (1, k1) = (12) is given to be
in U), we nd that (1k) U for k 2, 3, . . . , n. For instance, (13) = (12)(23)(12) and (14) = (13)(34)(13).
Now the arbitrary transposition (ab) we see to be in U because (ab) = (1a)(1b)(1a). Therefore, as every
permutation may be decomposed into transpositions, every permutation is contained in U. Now S
n
U
S
n
, so this nalizes the proof that U = S
n
.
H 2.10.12* Prove that, for n 3, the subgroup U
n
generated by the 3-cycles is A
n
.
Because (i) 3-cycles are even permutations, (ii) inverses of 3-cycles are again 3-cycles, and (iii) products of
even permutations are again even, we have that U
n
A
n
. Every even permutation is the product of an
even number of transpositions. Let a, b, c, d 1, 2, . . . , n be distinct. We have (ab)(cd) = (adc)(abc) and
(ab)(ad) = (adb). Every product of two transpositions is in one of these two forms (the rst case has no index
shared, the second case has one index shared), and both forms may be rewritten as a product of 3-cycles.
Hence any even permutation may be written as the product of 3-cycles, i.e. U
n
. Therefore A
n
U
n
,
and so A
n
= U
n
.
H 2.10.13* Let N A
n
contain a 3-cycle. Prove that N = A
n
.
The spirit of this proof is to show that if N contains a 3-cycle then it contains all other 3-cycles. Then by
2.10.12, the result would be proven. Let (abc) N. Note that (abc)
2
= (abc)
1
= (acb) N. Using the
normality of N, we can compute things like (cde)
1
(abc)(cde) = (abe) N. This allows us to swap out an
index, with the ultimate goal of generating an arbitrary 3-cycle from our original (abc). Continuing in this
fashion, we see that, if a, b, c, d, e 1, 2, . . . , n are distinct and f 1, 2, . . . , na, b, d, e, then
(abf)
1
(bcd)
1
(cde)
1
(abc)(cde)(bcd)(abf) = (def) N.
This conguration of indices is only possible for n 5. Thankfully, smaller cases are almost trivial. A
1
and A
2
dont accomodate 3-cycles, so they are irrelevant. A
3
= (123)) has order 3 so it contains no non-
trivial subgroups. A
4
is actually susceptible to the above argument, although the long formula breaks down
because there arent 5 symbols to choose from. Instead, we see that to go from any one 3-cycle in A
4
to
another, the only steps which might be required are swapping up to two indices and squaring (as noted
above, (abc)
2
= (acb)). Swapping the indices is done by conjugation, e.g. (cde)
1
(abc)(cde) = (abe), as
explored above. Therefore in A
4
it is also the case that a normal subgroup containing a 3-cycle must contain
every 3-cycle. By 2.10.12, A
n
N, so N = A
n
.
H 2.10.14* Prove that A
5
has no non-trivial normal subgroups. (i.e. it is simple)
Missing, although a proof may be found in 2.11.6c.
H 2.10.15 Assume that A
5
is simple. Prove that if H A
5
is proper, then [H[ 12.
A
5
has order 60, so, by Lagranges theorem, [H[ 2, 3, 4, 5, 6, 10, 12, 15, 20, 30. A subgroup of order
30 has index 2, so it would be normal by 2.6.2. Because A
5
is simple, this is disallowed, so [H[ , = 30.
Suppose [H[ = 20 so [A
5
: H] = 3. By lemma 2.21, because [A
5
[ = 60 3! = [A
5
: H]!, H contains a
non-trivial normal subgroup of A
5
. However, as A
5
contains no non-trivial normal subgroups, we come to a
contradiction. Therefore [H[ , = 20. The same argument rules out [H[ = 15 because 60 4!. Therefore [H[
must be 12 or smaller. Lemma 2.21 says nothing about an order 12 subgroup because 60 [ 5! = 60.
H 2.11.1 (a) In S
n
, prove that there are
n!
r(nr)!
distinct r-cycles. (b) Find the number of conjugates
of (12 r) in S
n
. (c) Prove that, if S
n
commutes with (12 r), then = (12 r)
i
with i
0, 1, . . . , r 1 and S
n
leaving all of 1, 2, . . . , r xed.
(a) An r-cycle permutes r out of n symbols, and there are

n
r

=
n!
r!(nr)!
ways of choosing those symbols.
Beyond this choice, there are (r 1)! ways of permuting all but one of the symbols in the r-cycle. Thus, for
instance, we count the distinct 3-cycles (123) and (132) but, ignoring movement of the rth symbol, we dont
overcount (312) = (123) as unique. Thus the number of distinct r-cycles in S
n
is (r 1)!

n
r

=
n!
r(nr)!
.
(b) All permutations of the same cycle decomposition are conjugate to one another. In particular, all r-cycles
in S
n
have the same cycle decomposition 1, 1, . . . , 1, r, with n r ones. Hence (12 r) is conjugate to
every one of the
n!
r(nr)!
r-cycles in S
n
.
(c) It is clear that (12 r) commutes with its powers as well as with , which is disjoint from it. Therefore
if A = (12 r)
i
[ i 0, 1, . . . , r 1, xing 1, 2, . . . , r is the set of interest, then A N((12 r)),
where N(a) is the normalizer of a, i.e. the set of elements commuting with a. The cardinality of A is
r(n r)! because there are (n r)! ways of permuting the n r elements r + 1, r + 2, . . . , n. We also
know that the normalizer satises (# conjugates of a) = [S
n
[/[N(a)[. In the case of a = (12 r), this gives
n!
r(nr)!
= n!/[N(a)[, so [N(a)[ = r(n r)!. Therefore A = N(a).
H 2.11.2 (a) Find the number of conjugates of (12)(34) in S
n
for n 4. (b) Determine N((12)(34)).
(a) For S
n
, we have (12)(34)
1
= ((1)(2))((3)(4)). If were to commute with (12)(34), then we
would need ((1)(2)) = (12), ((3)(4)) = (34) or ((1)(2)) = (34), ((3)(4)) = (12). The action of on
the n4 other symbols is irrelevant. Enumerate these by considering rst (1) 1, 2, 3, 4 with 4 choices.
Once (1) is picked, (2) is immediately determined and two choices remain for (3). After picking (3),
(4) is also determined. Hence there are 4 2 (n 4)! = 8(n 4)! distinct which commute with (12)(34).
Then the number of conjugates of (12)(34) is [S
n
[/[N((12)(34))[ = n!/(8(n4)!) =
1
8
n(n1)(n2)(n3).
Alternately, count in the following way: the conjugates of (12)(34) are those permutations with the same
cycle structure (ab)(cd). We need to pick 2 elements from n to constitute a, b, and 2 elements from n 2 to
constitute c, d. Then we will have overcounted by a factor of 2, getting, for instance, (12)(34) and (34)(12)
which are equal. Therefore the number of conjugates is

n
2

n2
2

/2 =
1
8
n(n 1)(n 2)(n 3).
(b) All commuting elements are described in (a).
H 2.11.3 Prove that S
p
contains (p 1)! + 1 elements satisfying x
p
= e.
The identity element satises e
p
= e. By 2.11.1, the number of p-cycles in S
p
is p!/(p 0!) = (p 1)! which,
by 2.10.6a, have order p. Therefore we have exhibited (p 1)! + 1 elements satisfying x
p
= e. Are there
others? No. Consider a non-identity element S
p
. The order of is equal to the least common multiple
of the lengths of the cycles in its disjoint cycle decomposition by 2.10.6b. If it were the case that
p
= e,
then the order of must be p because p is prime, so that the cycles in its decomposition must have length 1
or p. Because there are only p symbols to choose from in S
p
, one p-cycle already exhausts them all and we
therefore exclude possibilities such as being the product of two or more p-cycles. Thus for
p
= e to hold,
we must have that is a p-cycle, and so our original count is acceptable.
H 2.11.4 Let G be nite and let a G have exactly two conjugates. Prove that G is not simple.
We have that [G[/[N(a)[ = (# conjugates of a) = 2, so that N(a) has index 2. By 2.6.2, N(a) is normal.
The only oddball case to consider is if N(a) = e, so that G is of order 2 and is therefore abelian. If this
were the case, then N(a) = G, which is a contradiction.
H 2.11.5 (a) Find , A
5
such that and are conjugate in S
5
but not in A
5
. (b) Find all conjugacy
classes in A
5
and the cardinalities of each.
(a) Let = (12345) and = (21345). They are both in A
5
because they are 5-cycles which, by 2.10.9, are
even permutations. Furthermore, we easily see that (12)(12) = so that conjugation by (12) , A
5
brings
into . Now suppose , are such that both
1
= and
1
= . Then we have
1

1
= ,
i.e. that
1
N(). For this particular = (12345), we are very familiar with its normalizer. From
2.11.1c, we see that N() = (12345)
i
[ i 0, 1, 2, 3, 4. Notice that N() A
5
. Now put = (12) and
let S
5
be any other solution to
1
= . We have that
1
(12) A
5
. If A
5
, then we would have
(12) A
5
, a contradiction. Therefore , A
5
so we have proved that this and are conjugate in S
5
but
not in A
5
.
Notice that the same cant be done for 3-cycles. That is, all 3-cycles are conjugate in A
5
. Take, for example,
= (123) and = (234). We have (1234)(123)(4321) = (234), but if is another element of N((123)),
then, by the above comments,
1
(1234) = (123)
i
(45)
j
with i 0, 1, 2 and j 0, 1. Solving for
yields = (1234)(123)
i
(45)
j
= (14)(13)(12)(123)
i
(45)
j
which may be in A
5
if i = 0 and j = 1. Then we
have = (14)(13)(12)(45) = (12345) A
5
. Indeed, (12345)(123)(54321) = (234), so the two 3-cycles are
conjugate in A
5
. More generally, = (abc) is conjugate to = (bcd) by (abcdf) A
5
and = (abc) is
conjugate to = (cdf) by (acfbd) A
5
. In summary, all 3-cycles are conjugate in A
5
, with
e(abc)e
1
= (abc) (abcdf)(abc)(abcdf)
1
= (bcd) (acfbd)(abc)(acfbd)
1
= (cdf).
(b) The possible cycle structures for elements of A
5
are 1, 1, 1, 1, 1, 1, 2, 2, 1, 1, 3, and 5. The rst
cycle structure has only one element: the identity. The conjugacy class of the identity is trivially C(e) = e
with only one member.
Pick the element (12)(34) with structure 1, 2, 2. Lets investigate its conjugacy class: in 2.11.2a, we
considered the elements which commute with (12)(34), however we now restrict our attention to those
commuting elements which live in A
5
. The full normalizer in S
5
can be seen by the remarks of 2.11.2a to
be N
S
5
((12)(34)) = e, (12), (34), (12)(34), (13)(24), (14)(23), (1324), (1423). Half of these elements are odd
permutations, so the normalizer of interest is N
A
5
((12)(34)) = e, (12)(34), (13)(24), (14)(23). Now the
conjugacy class of (12)(34) has 60/4 = 15 elements, i.e. [C
A
5
((12)(34))[ = 15. At this point, its natural
to wonder if all elements of structure 1, 2, 2 share this conjugacy class. In S
5
, they certainly do, but here
we only allow conjugation by even permutations. As seen in (a), this is a nontrivial restriction. In fact, it
is true that all elements of structure 1, 2, 2 are conjugate to one another under A
5
. This may be proven
by a direct argument of the type given for 3-cycles at the end of the discussion of (a). However, we will
momentarily prove it indirectly by ignoring the issue for now.
Elements with structure 1, 1, 3 are 3-cycles. As per 2.11.1c, N
S
5
((123)) = (123)
i
(45)
j
[ i 0, 1, 2, j
0, 1. We only consider N
A
5
= N
S
5
A
5
= (123)
i
[ i 0, 1, 2. From this we see that [C((123))[ =
[A
5
[/[N
A
5
((123))[ = 60/3 = 20 is the size of the conjugacy class of (123). Furthermore, by the discussion at
the end of (a), we know that, for any 3-cycle S
5
, C
A
5
() = C
A
5
((123)) because all 3-cycles are conjugate
in A
5
.
The only remaining cycle structure is 5, the 5-cycles. We saw above that (12345) and (21345) are not
conjugate in A
5
, so they spawn distinct conjugacy classes. Specically, N
A
5
((12345)) = (12345)
i
[ i
0, 1, 2, 3, 4 and N
A
5
((21345)) = (21345)
i
[ i 0, 1, 2, 3, 4 as discussed in (a), so that [C
A
5
((12345))[ =
[C
A
5
((21345))[ = 60/5 = 12.
Now we have demonstrated 5 distinct conjugacy classes of total size 1 + 15 + 20 + 12 + 12 = 60 = [A
5
[.
Therefore this constitutes a complete collection of conjugacy classes. As promised, this indirectly proves
that all permutations of cycle type 1, 2, 2 are conjugate to one another in A
5
.
H 2.11.6 Let N G. (a) Let a N and prove that every conjugate of a in G is also in N. (b) Prove that
[N[ =

[G : N(a)] for some choices of a N. (c) Prove that A


5
is simple.
(a) Let g G. Then gag
1
N because N is normal.
(b) For a, b N, a b if a = gbg
1
for some g G is trivially an equivalence relation. These equivalence
classes (conjugacy classes) partition N. The conjugacy class of element a N has order [G : N(a)] by
theorem 2.h, where N(a) still refers to the normalizer of a with respect to G. If we take a single representative
a from each conjugacy class and add up [G : N(a)], we will nd [N[.
Alternately, consider the conjugacy classes C(a) = gag
1
[ g G of G. If C(a) N ,= e, then there
exist g G and n N such that gag
1
= n N. Now a = g
1
ng N, so in fact gag
1
N for all g G
because N is normal. Therefore C(a) N or C(a) N = e. Apply the class equation to G N = N to
nd [N[ =

a
[C(a) N[. Those a producing C(a) disjoint from N may be excluded from the sum because
they contribute 0.
(c) Let NA
5
be non-trivial. The order of N must be one of 2, 3, 4, 5, 6, 10, 12, 15, 20, 30. The ve distinct
conjugacy classes of A
5
have sizes 1, 12, 12, 15, 20 as per 2.11.5b. By (b), the possible orders under 30 for
N are then 1, 13, 16, 21, 25, 28 (notice that N must contain the identity, so the conjugacy class of size 1
must always be included). This list does not intersect with the list of orders allowed by Lagranges theorem.
Therefore A
5
has no non-trivial normal subgroups.
H 2.11.7 Let n Z
+
. Prove that if [G[ = p
n
then G has a subgroup of order m for all 0 m < n.
This statement is trivial for n = 1. Assume the result to be true for n 1. Because G is a p-group, it has a
non-trivial center Z. Therefore p [ [Z[ so, by Cauchys theorem, there exists an element a Z with order p.
The subgroup A = a) is normal in G because it is a subgroup of the center, so consider the quotient group

G = G/a) of order p
n1
. By induction,

G contains a subgroup

H of order p
n2
.
Dene the quotient map : G

Gby (g) = gA) and let g
1
, g
2
G. is well-dened: if g
1
A) = g
2
A) then
g
1
2
g
1
A, so (g
1
) = g
1
A) = g
2
g
1
2
g
1
A) = g
2
A) = (g
2
). is a homomorphism: (g
1
g
2
) = g
1
g
2
A) =
(g
1
A))(g
2
A)) = (g
1
)(g
2
). Now dene H = g G [ (g)

H. Because is a homomorphism, H is a
subgroup of G. restricts to a homomorphism = [
H
: H

G. We see that ker() = A and im() =

H,
so by theorem 2.d,

H

= H/A. Now it is clear that [H[ = [

H[ [A[ = p
n1
so that G has a subgroup of
order p
n1
. By induction, this subgroup contains subgroups of orders p
m
for all 0 m < n, so the result is
proven.
As a tangential result to be used in 2.11.8, we can prove that H G, i.e. that a group of order p
n
has a
normal subgroup of order p
n1
. We will proceed again by inducting on n. We can use n = 1 as the trivial
base case (though 2.9.5 shows explicitly that the claim holds for n = 2). Assume the result for n1 and again
consider the subgroup H = g G [ (g)

H. Let g G and h H. Then (ghg
1
) = (g)(h)(g)
1
is
in

H by induction (because

H is a normal subgroup of order n 2 in order n 1

G)
Lemma 4 Let G be nite and let p be the smallest prime dividing [G[. Let H G be of index p. Prove
that H G.
Suppose that H is not normal, so that N(H) ,= G. We still have H N(H), so we must have N(H) = H
by order considerations. Let G act by conjugation on G/H, i.e. dene : G S(G/H) by ((g))(H) =
gg
1
H for G. is a homomorphism: ((g)(h))(H) = ghh
1
g
1
H = (gh). Consider its kernel:
ker = k G [ kk
1
H = H for all G. We see that if k ker then kk
1

1
H for all G.
Then let H and it must be that kk
1
H, so k N(H). However, we assumed that N(H) = H, so
this implies that k H. Therefore ker H.
Theorem 2.d tells us that im()

= G/ ker so that [G : ker ] = [im()[ divides p! because im() is a subgroup
of S(G/H), a group of order p!. Counting in another way, we have that [G : ker ] [ ker [ = [G[, which is the
statement that [G : ker ] [ [G[. Now there exist a, b, c, d Z such that a[G : ker ] = p!, b[G : ker ] = [G[
and cp!+d[G[ = (p!, [G[). Plugging the rst two into the third yields (ac+bd)[G : ker ] = (p!, [G[). Therefore
[G : ker ] [ (p!, [G[). Because p[p! and p[[G[, and no smaller (non-unit) integer divides [G[, it must be that
(p!, [G[) = p. Finally, we can conclude our argument: we have that [G : ker ] [ p, and, because ker H,
[G : ker ] = [G : H][H : ker ] = p[H : ker ]. Then [G : ker ] = 1 is clearly forbidden, while [G : ker ] = p
implies that H = ker which is normal as it is the kernel of a homomorphism. This is a contradiction of
our assumption that H isnt normal. Thus it must be the case that H is normal.
H 2.11.8 Let n Z
+
. Prove that if [G[ = p
n
then there exists r Z and subgroups N
i
, i 0, 1, . . . , r
such that e = N
0
N
1
N
r1
N
r
= G where N
i
N
i+1
and N
i+1
/N
i
is abelian. (i.e. p-groups
are solvable)
Put r = n. By 2.11.7, N
n
= G has a subgroup N
n1
of order p
n1
. As this subgroup has index p, lemma 4
gives that it is normal. Furthermore, N
n
/N
n1
has order p, so it is cyclic and therefore abelian. Iterating,
we construct N
i1
from N
i
by taking a normal subgroup of order p
i1
from the order p
i
group N
i
. This
generates the desired tower of subgroups. Note that this is not the only such tower.
H 2.11.9 Let n Z
+
. Let [G[ = p
n
and let H G be proper. Prove that N(H) ,= H.
As H N(H), this is the statement that a proper subgroup of a p-group is properly contained in its
normalizer. Induct on the exponent n. For n = 1, the result holds because the only non-trivial subgroup is
e, which is normalized by all of G. Therefore assume that, in group of orders p
m
, m < n, all subgroups
are properly contained in their normalizers.
In the group G of order p
n
, consider a proper subgroup H and let Z = Z(G) be the center of G, which is
non-trivial because G is a p-group. If Z is not a subgroup of H, then, because Z N(H) (regardless of
H), we have H < Z, H) properly while Z, H) N(H). Therefore we may restrict our attention to the
case that Z H. As the center is always a normal subgroup, consider the map : G G/Z given by
(g) = gZ. This map is a well-dened homomorphism (see 2.11.7, for instance). Because [Z[ > 1, the group
G/Z is a p-group of order strictly less than p
n
. Note that (H) < G/H properly, for if gZ = hZ for some
h H, g G, then gh
1
Z H implies that g H. Therefore gZ , (H) for any g GH. By
induction, (H) is properly contained in its normalizer N((H)) G/Z. Therefore there exists x GH
such that (xZ)(hZ)(xZ)
1
(H) for all h H. But then for each h H there exists h
1
H such that
xhx
1
Z = h
1
Z, implying that xhx
1
h
1
1
Z H, i.e. that xhx
1
H. Thus we have an x N(H) while
x , H, so the claim is proven.
H 2.11.10 Let n Z
+
. Let [G[ = p
n
and let H G have order p
n1
. Prove that H G.
H has index p, so by lemma 4, H G.
H 2.11.11* Let n Z
+
. Let N G be non-trivial. Prove that N Z ,= e.
Let a
1
, . . . , a
k
be representatives of the conjugacy classes of G, ordered such that a
1
, . . . , a
m
N and
a
m+1
, . . . , a
k
, N. Recall, as was argued in 2.11.6b, that the conjugacy classes C(a
i
) have either C(a
i
) N
or C(a
i
) N = e. First arrange the a
1
, . . . , a
m
so that the rst r represent conjugacy classes of size 1
(i.e. elements in N Z) and the latter mr represent classes of size larger than 1. Then we can write the
class equation for G N = N as
[N[ =
m

i=1
[C(a
i
) N[ = [N Z[ +
m

i=r
[C(a
i
)[ = [N Z[ +
m

i=r
[G[
[N(a
i
)[
.
As [N[ < p
n
, every term in the sum is divisible by p, hence [N Z[ = [N[

|G|
|N(a
i
)|
is divisible by p,
implying that N Z is non-trivial, as desired.
As an aside, the equivalence [C(a)[ = [G[/[N(a)[ may be proven by considering the map f : G/N(a) C(a)
given by f(xN(a)) = xax
1
. f is well-dened: if xN(a) = yN(a) for x, y G, then x
1
y N(a) whence
x
1
yay
1
x = a. This gives that yay
1
= xax
1
, so f(xN(a)) = f(yN(a)). f is injective: by almost the
same argument, if f(xN(a)) = f(yN(a)) for x, y G then xax
1
= yay
1
so that x
1
yay
1
x = a, or
xN(a) = yN(a). Because f is a map between nite sets, this proves that f is a bijection, so [C(a)[ =
[G[/[N(a)[.
H 2.11.12 If G/Z is cyclic, prove that G is abelian.
By lemma 2.19, G/Z

= Inn(G). Say Inn(G) =
x
) where
x
(g) = xgx
1
. Now let a, b G. For
some r, s Z, we have
a
=
r
x
and
b
=
s
x
. Consider the commutator aba
1
b
1
. We can cast this as

a
(b)b
1
=
r
x
(b)b
1
= x
r
bx
r
b
1
. Once again, rewrite this as x
r

b
(x
r
) = x
r

s
x
(x
r
) = x
r
x
s
x
r
x
s
= e.
Therefore aba
1
b
1
= e for arbitrary a, b G so G is abelian.
Alternately, let G/Z = xZ) and let a, b G. Then there exist r, s Z with aZ = x
r
Z and bZ = x
s
Z.
Therefore a = x
r
z
1
and b = x
s
z
2
for some z
1
, z
2
Z. Now ab = x
r
z
1
x
s
z
2
= x
(r+s)
z
1
z
2
while ba =
x
s
z
2
x
r
z
1
= x
(r+s)
z
1
z
2
by virtue of z
1
and z
2
being in the center. Therefore ab = ba for arbitrary a, b G,
so G is abelian.
H 2.11.13 If [G[ = 15, prove that G is cyclic.
By Cauchys theorem, there exists f G with order 5. By lemma 4, N = f) G because [G : N] = 3 is
the smallest prime dividing 15. Consider the action of G on N by conjugation, : G Aut(N) given by
(g) =
g
where
g
(x) = gxg
1
. Notice that N must be normal so that G may act on it by conjugation.
is a homomorphism: (gh)(x) =
gh
= ghxh
1
g
1
= ((g) (h))(x). Because N is cyclic, we can easily
count the order of Aut(N). If N = x) for some x N, then every automorphism Aut(N) sends x
to another generator, and that choice of (x) uniquely determines the entire map because (x
n
) = (x)
n
suces to compute the image of any other element x
n
N. By 2.5.15, there are (5) = 4 (Euler totient
function) generators of N, so 4 choices of where to send x and hence [Aut(N)[ = 4.
Now, by Lagranges theorem, [(G)[ divides [Aut(N)[ = 4. Also, by theorem 2.d, [(G)[ divides [G[ = 15 (in
particular, [(G)[ [ ker [ = [G[). Therefore [(G)[ = 1, i.e. ker = G so that the inner automorphism for
every g G xes every element of N. Let t G be an element of order 3 (guaranteed to exist by Cauchys
theorem). Then tft
1
= f because (t) = id, so tf = ft. Now we see that tf has order lcm(3, 5) = 15, so,
in fact, G = tf) is cyclic.
H 2.11.14 If [G[ = 28, prove that G has a normal subgroup of order 7.
By Cauchys theorem, there exists x G with order 7. By lemma 2.21, a subgroup of order 7 in G must be
normal because [G[ = 28 24 = 4!. Therefore x) is a normal subgroup of G with order 7.
H 2.11.15 If [G[ = 28 and G contains a normal subgroup F of order 4, prove that G is abelian.
By 2.11.14, there exists S G with [S[ = 7. By Lagranges theorem, S F = e, so SF has order
[S[[F[/[S F[ = 28 and therefore SF = G. Because F is normal by assumption, 2.6.12 gives that sf = fs
for all s S, f F. Furthermore, groups of order 4 and 7 are both necessarily abelian. Then consider
g
1
= s
1
f
1
G and g
2
= s
2
f
2
G. We have g
1
g
2
= s
1
f
1
s
2
f
2
= s
2
f
2
s
1
f
1
= g
2
g
1
and hence G is abelian.
H 2.12.1 Let G = S
4
. Exhibit a 2-Sylow subgroup and a 3-Sylow subgroup of G.
[G[ = 24 = 2
3
3, so a 3-Sylow subgroup has order 3 while a 2-Sylow subgroup has order 8. A 3-Sylow
subgroup is (123)) = e, (123), (213). A 2-Sylow subgroup is (1234), (14)(23))

= D
4
. As dened in 2.7.8,
a presentation of D
4
is x, y [ x
2
= y
4
= e and xy = y
1
x). We see that ((14)(23))
2
= (1234)
4
= e and
(14)(23)(1234) = (123) = (4321)(14)(23) so that x = (14)(23) and y = (1234) does provide a set of generators
for an order 8 subgroup of S
4
isomorphic to D
4
.
As an aside, coming up with the 3-Sylow subgroup is somewhat non-trivial. Some things worth noticing at
the outset are that S
4
has no elements of order 8, so well need a non-cyclic group generated by at least 2
elements, and that 8 12 = [A
4
[, so the 3-Sylow subgroup must contain an odd permutation. There is a very
limited list of possibilities for a group of order 8, and the dihedral group has come up in Hersteins problems
already. Then trying to nd generators for D
4
is fairly easy. An order 4 pick for y might as well be (1234),
and then one looks for an order 2 (i.e. transposition or product of two disjoint transpositions) x such that
xyx = y
1
= (4321). Because this is conjugation by x, we immediately see that x has to be (14)(23) for this
choice of y. Other choices of x and y can be made.
H 2.12.2 If [G[ = 108 = 2
2
3
3
, prove that G has a normal subgroup of order 9 or 27.
By Sylows theorem, G contains a 3-Sylow subgroup H of order 27. Let G act by left multiplication on
G/H, with : G S(G/H) given by ((g))(aH) = gaH. This map is a homomorphism into S(G/H)

= S
4
:
((g
1
g
2
))(aH) = g
1
g
2
aH = g
1
(g
2
aH) = ((g
1
) (g
2
))(aH) so (g
1
g
2
) = (g
1
)(g
2
). We must have, then,
that [(G)[ divides 24 = [S
4
[ by Lagrange and that [(G)[ divides 108 = [G[ by theorem 2.d ((G)

=
G/ ker ). Therefore [(G)[ divides their greatest common divisor (24, 108) = 12.
Consider the kernel of the action . By virtue of being a kernel, we have ker G. Additionally, we see
that ker H, because if k ker then kaH = aH for all a G. If we put a H, then we see that
kH = H, which is the statement that k H. Now return to the specics of this problem. We may compute
[ ker [ = [G[/[(G)[ by theorem 2.d. Now, as the only possible values of [(G)[ are 1, 2, 3, 4, 6, 12, the
only possible values of [ ker [ are 108, 54, 36, 27, 18, 9. Recalling that ker H, we must have that [ ker [
divides 27 = [H[. Therefore the only acceptable cases are [ ker [ = 27 and [ ker [ = 9. These subgroups will
always be normal in G, so the result is proven.
Supplementary Problems
H 2.S.1 Let G be nite and abelian with G = g
1
, . . . , g
n
. (a) Prove that g
1
g
2
g
n
has order 1 or 2.
(c) If G has exactly one element y of order 2, prove that y = g
1
g
2
g
n
. (d) If p Z is prime, prove that
(p 1)! 1 mod p. (Wilsons theorem)
(a) Number the group elements so that, for 1 i r, they satisfy g
1
i
= g
i
, and, for r < i n, g
1
i
,= g
i
.
Because e
1
= e, r 1, and we are free to always put g
1
= e. Now for every element g
i
with r < i n, its
inverse g
1
i
is another g
j
with r < j n. Hence the product g
r+1
g
n
reduces to the identity e. Now we
have
g
1
g
r
g
r+1
g
n
= (g
1
g
r
)(g
r+1
g
n
) = g
1
g
r
and (g
1
g
n
)
2
= (g
1
g
r
)
2
= e because, for each 1 i r, we have g
2
i
= e. Therefore g
1
g
n
has order
1 or 2.
(c) If G contains one element g
2
= y of order 2, then g
1
g
n
= ey(g
3
g
n
) = y, where every one of g
i
(3 i n) is paired o with its inverse.
(d) Consider the multiplicative group Z

p
modulo p. We note that this is a nite, abelian group and the
product of all of its elements is 1 2 (p 1) = (p 1)!. Then we seek elements of order 2, that is, a Z

p
such that a
2
1 mod p. Equivalently, there must exist k Z such that a
2
1 = (a + 1)(a 1) = kp. As
p is a prime dividing the right hand side, it must divide at least one factor of the left hand side. Because a
is restricted to the range 1 a p 1, we see that this can only be satised if a + 1 = p, i.e. if a = p 1.
Therefore a = p 1 is the unique element in Z

p
of order 2. By (c), it follows that (p 1)! (p 1)
mod p 1 mod p.
H 2.S.2 Let p Z be an odd prime such that 1 +
1
2
+
1
3
+ +
1
p1
= a/b with a, b Z. Prove that p [ a.
If p > 3, prove that p
2
[ a.
Missing.

You might also like