You are on page 1of 25

Solutions to Abstract Algebra (Dummit and Foote 3e)

Chapter 1 : Group Theory


Jason Rosendale
jason.rosendale@gmail.com
February 11, 2012
This work was done as an undergraduate student: if you really dont understand something in one of these
proofs, it is very possible that it doesnt make sense because its wrong. Any questions or corrections can be
directed to jason.rosendale@gmail.com.
Exercise 1.1.1
(a) is not associate: 3 (2 1) ,= (3 2) 1. (b),(c),and (d) can be shown to be associative by grinding out the
algebra and showing that a (b c) = (a b) c. (e) is not associative, as can be shown by letting a = b = c = 2.
Exercise 1.1.2
(a) and (e) can be shown to not be commutative by letting a = 1, b = 2. (b),(c) and (d) can be shown to be
commutative by grinding out the algebra.
Exercise 1.1.3
a + b+ c
= a+ b+ c def. of modular addition
= a+ b + c def. of modular addition
Exercise 1.1.4
ab c
= a b c def. of modular multiplication
= a bc def. of modular multiplication
Exercise 1.1.5
To be a group, each element a would need a multiplicative inverse b such that ab = 1; but 0 has no such inverse.
Exercise 1.1.6
(b) is not closed: 1/14 + 1/14 = 1/7. (c) is not closed: 2/3 + 2/3 = 4/3. (d) is not closed: 2 + (3/2) = 1/2.
(f) is not closed: 1/2 + 1/3 = 5/6. (a) and (e) are groups under addition.
Exercise 1.1.7
That xy is well-dened follows directly from the fact that +,-, and the greatest integer function are well dened
on R. To prove that G is a group: G has an identity element (0 G and 0 x = x 0 = x [x] = 0), each
1
element a G has an inverse (1 a G and a (1 a) = 1 [1] = 0 = (1 a) a), and associativity can be
shown with a little tedious algebra. That G is abelian follows from the commutativity of addition:
x y = (x + y) [x + y] = (y + x) [y + x] = y x
Exercise 1.1.8a
G has an identity element (1 G and z1 = z = 1z) and an inverse (z
n
= 1 implies (z
1
)
n
= 1 so z
1
G) and
complex multiplication is associative. To show closure:
x, y G assumed
(m, n Z
+
)x
m
= 1 y
n
= 1 def. of membership in G
(mn Z
+
)x
mn
= 1
n
= 1 y
mn
= 1
m
= 1
(mn Z
+
)x
mn
y
mn
= 1
(mn Z
+
)(xy)
mn
= 1
xy G def. of membership in G
Exercise 1.1.8b
The group is not closed under addition. 1 and 1 are both elements of G, but 1 + (1) = 0 is not.
Exercise 1.1.9a
The identity is (0, 0), the inverse of (a, b) is (a, b), associativity and closure are trivial.
Exercise 1.1.9b
The identity is (1, 0), associativity and closure are trivial. The inverse of (a, b) can be found by solving (a +
b

2)x = 1 and is given by:


(a, b)
1
=
_
a
a
2
2b
2
,
b
a
2
2b
2
_
Exercise 1.1.10
The denition of a symmetric matrix is that for all i, j we have a
ij
= a
ji
. In the group table, this is the case i
for all i, j we have a
i
a
j
= a
j
a
i
which is the denition of abelian.
Exercise 1.1.11
General formula is: [a[ = lcm(a, 12)/a.
[0[ = 1
[1[ = 12
[2[ = 6
[3[ = 4
[4[ = 3
[5[ = 12
[6[ = 2
[7[ = 12
[8[ = 3
[9[ = 4
[10[ = 6
[11[ = 12
2
Exercise 1.1.12
[1[ = 1
[ 1[ = 2
[7[ = 2
[ 7[ = 2
[13[ = 2
Exercise 1.1.13
General formula is: [a[ = lcm(a, 36)/a.
[1[ = 38
[2[ = 18
[6[ = 6
[9[ = 4
[10[ = 18
[12[ = 3
[ 1[ = 36
[ 10[ = 18
[ 18[ = 2
Exercise 1.1.14
This would be tedious to type out. For each a we need to use the methods of exercises 0.3.15 to nd y such that
36x + ay = 1
and this y would be the order of a.
Exercise 1.1.15
Proof by induction. For the n = 1 case, its clear that a
1
= a
1
1
. Supposing the equality holds for the n = k
case, we have:
(a
1
a
2
. . . a
k
a
k+1
)(a
1
k+1
a
1
k
. . . a
1
1
)
= (a
1
a
2
. . . a
k
)a
k+1
a
1
k+1
(a
1
k
. . . a
1
1
) associativity
= (a
1
a
2
. . . a
k
)1(a
1
k
. . . a
1
1
) denition of inverses
= (a
1
a
2
. . . a
k
)(a
1
k
. . . a
1
1
) denition of identity
= e equality holds for n = k
and thus (a
1
k+1
a
1
k
. . . a
1
1
) is shown to be the right inverse of (a
1
a
2
. . . a
k
a
k+1
). The proof for left inverse is
similar. Thus equality holds for the n = k + 1 case. By induction, equality holds for all n Z
+
.
Exercise 1.1.16
This follows directly from the denition of [x[. The order of x is the smallest integer n such that x
n
= 1; if
x
2
= 1, then this smallest integer must be less than or equal to 2. Thus it can be only 1 or 2.
Exercise 1.1.17
If [x[ = n, then x
n
= 1, and thus x x
n1
= 1, which makes x
n1
the inverse of x.
3
Exercise 1.1.18
xy = yx assumed
y
1
xy = y
1
yx left multiplication
y
1
xy = 1x = x denition of inverse
x
1
y
1
xy = x
1
x left multiplication
x
1
y
1
xy = 1 denition of inverse
(x
1
y
1
)
1
= xy denition of inverse
xy = yx consequence of exercise 1.1.15
Exercise 1.1.19a
Part (a) is a direct consequence of associativity.
Exercise 1.1.19a
Part (b) is a special case of exercise 1.1.15.
Exercise 1.1.19c
Parts (a) and (b) trivially hold when a = 0 or b = 0. Parts (a) and (b) can be shown to hold for a, b 0
by taking the inverses of each side of each equation. The only thing left to show is that part (a) holds when
a < 0 < b or b < 0 < a:
Assume that a < 0 < b or b < 0 < a. If [a[ = [b[ (absolute value, not order) then clearly
x
a+b
= x
0
= x
a
x
a
= x
a
x
b
ow assume, without loss of generality, that [a[ < [b[ Then b has the same sign as a and b has the same sign as
a + b. Therefore a has the same sign as a + b and we can deduce:
x
a
x
b
= x
a
x
a+b+a
from b = b + a a
= x
a
x
a
x
b+a
from part (a), since a and b + a have same sign
= (x
b+a
left cancellation
Exercise 1.1.20
From 1.1.19(b), we have
x
n
= 1 (x
n
)
1
= 1
1
= 1 x
n
= 1 (x
1
)
n
= 1
Thus the least n such that x
n
= 1 must also be the least n such that (x
1
)
n
= 1.
Exercise 1.1.21
x
2k+1
= 1 assumed, since n = 2k + 1
x
2k+1
x = 1x = x right multiplication
x
2k+2
= x 1.1.19(a)
(x
2
)
k+1
= x 1.1.19(a)
4
Exercise 1.1.22
x
n
= 1 assumed
(xgg
1
)
n
= 1 gg
1
= 1
xg(g
1
xg)
n1
g
1
= 1 tricky associativity
g
1
xg(g
1
xg)
n1
g
1
g = g
1
1g = 1 left and right multiplication
g
1
xg(g
1
xg)
n1
= 1 right cancellation
(g
1
xg)
n
= 1 1.1.19(a)
This shows that the least n such that x
n
= 1 must also be the least n such that (g
1
xg)
n
= 1. To prove that
[ab[ = [ba[, let x = ab and g = b
1
in the preceeding proof and conclude that the least n such that (ab)
n
= 1
must also be the least n such that (babb
1
)
n
= (ba)
n
= 1.
Exercise 1.1.23
(x
s
)
t
= x
st
= x
n
= 1, so [x
s
[ t. But if there were some k such that 0 k < t such that (x
s
)
k
= 1, then we
would have 0 sk < st = n with x
sk
= 1 contradicting the assumption that [x[ = n.
Exercise 1.1.24
Proof by induction. The n = 0 and n = 1 cases are trivial. Assume the equality holds for n = k:
(ab)
k+1
= a(ba)
k
b tricky associativity
(= a(ab)
k
b a, b commute
(= aa
k
b
k
b equality holds for n = k
(= a
k+1
b
k+1
Thus equality holds for all n N. To prove that this holds for negative n, just take the inverse of each side of
the equation to yield (ab)
n
= b
n
a
n
, then apply commutativity to conclude (ba)
n
= b
n
a
n
.
Exercise 1.1.25
Given that x
2
= 1, we see that (ab)(ab) = b(aa)b = 1. Right-multiplying by b
1
a
1
yields ab = ba.
Exercise 1.1.26
Were told that H is nonempty so it contains some element h; its closed under inverses and the binary operation,
it contains the identity element hh
1
= 1. It inherits associativity from G. Thats sucient for H to be a group.
Exercise 1.1.27
Let H = x
n
: n Z. The identity x
0
= 1 exists in n; if x
n
H then x
n
H so H is closed under inverses; its
trivially closed under the binary operation; and (x
i
x
j
)x
k
= x
i+j+k
= x
i
(x
j
x
k
) so it is associative.
Exercise 1.1.28
All parts of this exercise can be proven through simple but tedious algebra.
Exercise 1.1.29
(a, b) (c, d) = (c, d) (a, b)
(ac, bd) = (ca, db) def. of
ac = ca and bd = db def. of equality in AB
5
Exercise 1.1.30
To prove commutativity:
(a, 1)(1, b) = (a1, 1b) = (a, b) = (1a, b1) = (1, b)(a, 1)
Now let p = [a[, q = [b[, and x = [p, q]. Then:
(a, b)
x
= (a
x
, b
x
) = ((a
p
)
x/p
, (b
q
)
x/q
) = (1
x/p
, 1
x/q
) = (1, 1) = 1
Thus the order of (a, b) divides [p, q] = x. Now let n = [(a, b)[. From exercise 1.1.24:
(1, 1) = (a, b)
n
= [(a, 1)(1, b)]
n
= (a, 1)
n
(1, b)
n
Right-multiplying gives us:
(1, b)
n
= (a, 1)
n
which means
(1, b
n
) = (a
n
, 1)
thus a
n
= b
n
= 1. Thus p and q both divide n, which means that [p, q] divides n. Weve now shown that
n = [(a, b)[ divides x = [p, q] and that x divides n: therefore x = n.
Exercise 1.1.31
To prove that t(G) has an even number of elements, establish an equivalence relation a b : (ab = 1ora = b)
on t(G). This is clearly reexive,symmetric, and easily shown to be transitive. There are two elements per
equivalence class on t(G): one element and its inverse. Thus t(G) has an even number of elements.
Thus the set Gt(G) (that is, g G[g = g
1
) must also have an even number of elements. It contains the
identity, so it must contain at least one nonidentity element. Each element in Gt(G) has g
2
= 1, and therefore
must have order 1 or 2 by exercise 1.1.16. Only the identity has order 1, so this nonidentity element must have
order 2.
Exercise 1.1.32
Suppose x
i
= x
j
for 0 i < j n 1. Then 1 = x
j
x
i
= x
ji
, which would contradict the claim that [x[ = n.
Exercise 1.1.33a
Part (a) is the contrapositive of exercise 1.1.31.
Exercise 1.1.33b
Its trivial to show that i = k implies x
i
x
i
= x
2k
1 and therefore x
i
= x
1
. To prove in the other direction:
x
i
= x
i
assumed
x
i
x
i
= x
i
x
i
x
2i
= 1
This shows that x
i
= x
i
i n (the order of x) divides 2i. But 2i < n by the previous exercise, so n[2i implies
2i = 0 or 2i = n. And were asked to assume that i > 0, so we must have 2i = n = 2k and so i = k.
Exercise 1.1.34
Were it the case that x
m
= x
n
for some m ,= n, then we would have x
|mn|
= 1 which would contradict the
assumption that x was of innite order.
Exercise 1.1.35
Let x G have order n. Let k be an arbitrary power of x. By the division algorithm, we can nd unique values
of a, b with 0 b < n such that k = an + b. Thus we have
x
k
= x
an+b
= x
a
nx
b
= x
b
6
Exercise 1.1.36
We know that at least one element must be its own inverse from exercise 1.1.31. WLOG, assume that this
element is a. From 1.1.31 we know that b is its own inverse i c is its own inverse: if both are their own inverses,
then the group is abelian by 1.1.25 and we are done.
If neither b nor c are their own inverses, then they must have order 3 (b
3
= c
3
= 1). We cannot have
b
2
= b (for we would have o(b) = 1) or b
2
= 1 (for we would have o(b) = 2) or b
2
= a (for we would have
b = b, b
2
= a, b
3
= ab, b
4
= a
2
= 1 so that o(b) = 4) so we must have b
2
= c. And this gives us a contradiction,
since it would imply that o(b
2
) = o(c) = o(b).
Exercise 1.2.2
We know that D
2n
has order 2n with distinct elements r
0
, r
1
, . . . r
n1
, r
0
s, r
1
s, . . . r
n1
s. So any element that is
not equivalent to r
k
for some k is equivalent to sr
k
for some k. We can now use a proof by induction: the case
for x = r
0
s and x = r
1
s is trivial. Now assume that the equality holds for x = r
k
s:
r(r
k+1
)s = rr(r
k
)s = r(r
k
)sr
1
= r
k+1
sr
1
and thus equality holds for x = r
k+1
s.
Exercise 1.2.3
As above, every element of D
2n
which is not a power of r has the form r
k
s = sr
k
. Thus:
(r
k
s)
2
= (r
k
s)(r
k
s) = sr
k
r
k
s = ss = 1
This proves only that the order of (r
k
s) divides 2; we must show that it is 2. Proof by contradiction: If the
order were 1, then we would have r
k
s = 1 which would imply s = r
k
which would imply that all 2n elements
could be written uniquely as a power of r; but r
n
= 1 so there could be at most n such elements. Thus the
order of (r
k
s) cannot be 1.
To show that s, sr are generators is suces to show that ssr = r, so all powers of r can be generated.
Exercise 1.2.4
z
2
= r
2k
= r
n
= 1, so o(z) = 2 (k > 1, so z is not the identity element). r
2
k trivially commutes with elements
of D
2n
that are powers of r; since o(z) = 2 it commutes with all other elements of order 2 (cf 1.1.33); that is,
all elements of D
2n
that are not a power of r (previous exercise).
To show that this is the only element other than the identity that commutes with all elemenets of D
2n
, we
use a very unsatisfying proof by exhaustion of cases. Let a be an element such that ab = ba for all b D
2n
.
Either a = r
i
or a = sr
i
for some i 0 . . . 2k 1.
case i) Suppose a = r
i
and let b be an arbitrary element of D
2n
. The element a obviously commutes for all b
of the form b = r
j
. If b = sr
j
, then ab = ba when:
ab = ba assumed
r
i
sr
j
= sr
j
r
i
denition of a, b
sr
i
r
j
= sr
j
r
i
from relation rs = sr
1
r
i+j
= r
j+i
left cancellation
r
i+j
r
ji
= 1 right multiplication by r
(j+i)
r
2i
= 1
n[2i o(r) = n
k[i n = 2k
This shows that, for this choice of a = r
i
, ab = ba for all b i i is a multiple of k. But weve assumed that
i 0, . . . , 2k 1 so either i = 0 or i = k. Thus a = r
i
commutes with all b i a = r
0
(the identity) or
a = r
k
.
7
case ii) Suppose a = sr
i
. Then this a does not commute with all b: let b = r. Then we have
ab = ba assumed
sr
i
r = rsr
i
denition of a, b
sr
i+1
= sr
1+i
from relation rs = sr
1
r
i+1
= r
i1
left cancellation
r
i+1
r
1i
= 1 right multiplication by r
(j+i)
r
2
= 1
But n is dened by the order of r, so r
2
= 1 implies n = 2, contradicting the requirement that n 4. Thus
we have shown that the only nonidentity element that commutes with all elements of D
2n
(n = 2k, n 4)
is r
k
.
Exercise 1.2.5
As shown in the previous exercise, the only element that commutes with element of D
2n
for n > 2 is r
i
where
n[2i. If n is odd then the requirements that i 0, . . . , n and i = kn together imply i = 0 and thus the only
commuting element is r
0
= 1.
Exercise 1.2.6
Since o(x) = o(y) = 2, we have x = x
1
, y = y
1
. Thus we have
tx = xyx right-multiply t = xy by x
tx = xy
1
x
1
y = y
1
, x = x
1
tx = x(xy)
1
tx = xt
1
Exercise 1.2.7
Let a = s and b = sr so that ab = r. The relations follow from each other: a
2
= s
2
= 1, (ab)
n
= r
n
= 1, and
b
2
= (sr)
2
= ssr
1
r = 1.
Exercise 1.2.8
n is, by denition, the smallest r such that r
n
= 1 so o(r) = n.
Exercise 1.2.9 through 1.2.13
A tetrahedron has four three-sided faces; a cube has six four-sided faces; an octahedron has eight three-sided
faces; a dodecahedron has twelve ve-sided faces; an icosahedron has twenty three-sided faces. Each side (pair
of vertices) can be ipped, so the total number of positions for a side is : The number of rigid motions for a
given solid is given by:
2 (number of sides) =
(number of faces) (number of sides per face)
number of faces connected to each side
Each solid has two faces connected to each side, so this reduces to (number of faces) * (number of sides per
face).
Exercise 1.2.14
For addition, 1. For multiplication, p : p is prime (by the fundamental theorem of arithmetic).
Exercise 1.2.15
For addition, p : (p, n) = 1. For multiplication, p < n : p is prime.
8
Exercise 1.2.16
Letting n = 2, we have r
2
= s
2
= 1. So let x = r, y = s. The relations follow: x
2
= r
2
= 1,y
2
= s
2
= 1, and
xy = y
1
x :
xy = (xy)
1
from o(xy) = 2
= y
1
x
1
= yx
1
from o(y) = 2
Exercise 1.2.17a
The presentation in 1.2 is x, y[x
n
= y
2
= 1, xy = yx
2
. The text demonstrated that xy = yx
2
implies x
3
= 1.
Thus, if n = 3k then x
3
= 1 but x
0
, x
1
, x
2
are distinct elements. Letting x = r, y = s gives us x
3
= r
3
= 1,
y
2
= s
2
= 1, and rs = sr
1
:
rs = xy from o(xy) = 2
= yx
2
= yx
1
xx
2
= 1 x
2
= x
1
= sr
1
Exercise 1.2.17b
We know that x
3
= 1 so x
3k
= 1. By denition, x
n
=1. If (3, n) = 1 then either n = 3k + 1 or n = 3k + 2 =
3(k + 1) 1 for some k. In either case, we have
1 = x
n
= x
3k1
= x
3k
x
1
= x
1
But x
1
= 1 i x = 1, so in either case we have x = 1 and X
2n
is generated uniquely by s, which has order 2.
Exercise 1.2.18
The presentation in 1.3 is u, v[u
4
= v
3
= 1, uv = v
2
u
2
.
a) If v
3
= vv
2
= 1, then v
2
= v
1
. Similarly, u
2
= u
2
and u
3
= u
1
.
b) u
3
= u(v
3
)u
2
from v
3
= 1
= (uv)(v
2
u
2
) associative property
= (v
2
u
2
)(uv) from uv = v
2
u
2
= v
2
u
3
v associative property
= v
1
u
3
v from part (a)
Left-multiplying both sides by v gives us vu
3
= u
3
v; right-multiplying by v
1
gives us u
3
v
1
= v
1
u
3
.
c) From part (b), we have u
3
v
1
= v
1
u
3
. Using part (a) this becomes u
1
v
1
= v
1
u
1
. Taking inverses of
each side yields vu = uv.
d) Were told uv = v
2
u
2
; by commutativity of u, v this becomes uv = (uv)
2
; by left- or right-cancellation this
becomes uv = 1.
e) 1 = (uv)
4
= u
4
v
3
v = v, and uv = 1 implies u = v
1
= 1.
Exercise 1.3.8
Dene the function f : N S

to be f(n) = (1n). The function is clearly injective (albeit not surjective) so


[S

[ [N[.
Exercise 1.3.9
See exercise 1.3.11.
9
Exercise 1.3.10
Trivial proof by induction on i. Since
m
maps a
k
a
k
, it must be the identity function; since m is the least
integer such that
m
= I, we have o() = m.
Exercise 1.3.11
Let k be the order of
i
. It must be the case that m[ik or, by the division algorithm, we could nd ik such that
ik = xm + b with 0 < b < m which would give us
(
i
)
k
=
ik
=
xm+b
= (
m
)
x

b
=
b
which would contradict the previous exercises conclusion that o() = m.
Having proven that ik must be a multiple of m, we see that the least such multiple of ik is ik = lcm(m, ik). If
we want the o(
i
) = m, we must have im = lcm(m, im) = lcm(m, i) which occurs only when gcd(m, i) = 1:
ab = lcm(a, b) gcd(a, b) lcm(a, b) =
ab
gcd(a, b)
Exercise 1.3.12a
yes: let = (1 3 5 7 9 2 4 6 8 10) and k = 5.
Exercise 1.3.12b
no. Let k be the smallest k such that
k
= (1 2)(3 4 5). Then
2k
= (3 5 4). Thus
2k
(1) = 1 (implying that
m[2k from exercise 10) but
2k
(3) = 5 (implying m 2k by exercise 10). This establishes a contradiction so
there can be no k such that
k
= .
Exercise 1.3.13
Assuming that can be written has a product of commuting 2-cycles, we have

2
= [(a
1
b
1
) . . . (a
k
b
k
)]
2
= (a
1
b
1
)
2
. . . [(a
k
b
k
)
2
= (1)
2
. . . (1)
2
= (1)
so that o() = 2.
If we assume that o() = 2 then maps a (a) and maps (a) ((a)) = a (with the possiblity
that (a) = a). Consider the set of 2-cycles:
(a (a)) : a 1, . . . , m and a < (a)
We can dene as the product of every element of this set:
= (a
1
(a
1
))(a
2
(a
2
)) . . . (a
k
(a
k
))
Each a 1, . . . , m appears in at most one of these disjoint 2-cycles, and appears i a ,= (a).
Exercise 1.3.14
Follow the preceeding proof. Assuming that can be written has a product of commuting p-cycles, we have

p
= [(a
1
b
1
) . . . (a
k
b
k
)]
p
= (a
1
b
1
)
p
. . . [(a
k
b
k
)
p
= (1)
p
. . . (1)
p
= (1)
And we can write as the product of elements of the disjoint collection of p-cycles:
(a
1
(a)
2
(a) . . .
p1
(a)) : a 1, . . . , m and a = min(a, (a), . . . ,
m1
(a))
If p is not prime then we can nd a, b > 1 such that p = ab = lcm(a, b) and the element
(1 2 . . . a)(a + 1 a + 2 . . . a + b)
has order p despite not being a product of disjoint p-cycles. A more explicit example: In S
6
we have
(1 2 3)(4 5)
which has order 6.
10
Exercise 1.3.15
Choose S
n
and let k be the least common multiple of the lengths of the cycles in its cycle decomposition.
From exercise 1.1.24 we know that
k
= (1) means that (a
1
. . . a
i
)
k
= (1) for each cycle in the decomposition
of , so k must be a common multiple of the order (length) of each cycle in the decomposition of ; thus the
order of must be the least such k, which is the least common multiple of the lengths of the cycles of .
Exercise 1.3.16
From n elements, there are
_
n
m
_
ways of selecting m elements and m! ways of writing an mcycle with them.
Each distinct m-cycle can be written in m dierent ways. Thus the number of distinct m-cycles in S
n
is given
by
_
n
m
_
m!
m
=
n!m!
m!(n m)!m
=
n(n 1)(n 2) . . . (n m + 1)
m
Exercise 1.3.17
There are
_
n
4
_
ways of selecting the elements of two disjoint 2-cycles, and 3 unique ways to construct the two
2-cycles with them. Thus the number of distinct products of two disjoitn 2-cycles is:
_
n
4
_
1
3
=
n!
(n 4)!4!
1
3
=
n(n 1)(n 2)(n 3)
8
Exercise 1.3.18
For each S
n
we can choose i, k > 0 such that i + k n and construct following disjoint product S
n
:
(1 2 . . . i)(n n 1 . . . n (k 1))
By exercise 15, this element has order of lcm(i, k). Thus for S
5
we can construct elements of orders 1,2,3,4,5
(trivially) and 6. The order 6 element is given by
(1 2 3)(4 5)
Exercise 1.3.19
With the explanation from the previous exercise we see that S
7
has elements of order 1, 2, 3, 4, 5, 6, 7 (trivially),
lcm(2, 5) = 10, and lcm(3, 4) = 12.
Exercise 1.3.20
Let a = (1 2), b = (2 3) and dene the generator to be a, b[a
2
= b
2
= 1, abab = ba. All other elements of S
3
can be written with these two elements:
(1) = a
2
= b
2
(1 3) = aba = bab
(1 2 3) = ab
(1 3 2) = ba
The fact that a
2
= b
2
= 1 means that every element of S
3
can be reduced to an alternating string of a, b. The
choice of the relation abab = ba (and its equivalents, ab = baba and aba = bab) means that any such alternating
string of length n 4 can be reduced to a string of length n 2. Thus S
3
can be represented as string of 3 or
fewer alternating elements a, b. There are only 6 such strings: 1, a, b, ab, ba, and aba = bab.
Exercise 1.4.1
Proof by enumeration. Consider all 16 possible 2 2 matrices with entries in 0, 1 and show that exactly 6 of
them have nonzero determinants.
11
Exercise 1.4.2
_
0 1
1 0
_
2
=
_
1 1
0 1
_
2
=
_
1 0
1 1
_
2
= I
_
1 0
0 1
_
1
_
0 1
1 1
_
3
=
_
1 1
1 0
_
3
= I
Exercise 1.4.3
_
1 0
1 1
_ _
1 1
0 1
_
=
_
1 1
1 0
_
,=
_
0 1
1 1
_
=
_
1 1
0 1
_ _
1 0
1 1
_
Exercise 1.4.4
Suppose n is not prime, and let a(1 < a < n) be a divisor of n. Then a has no multiplicative inverse. Proof
by contradiction: assume that ak = 1. Then we would have ak + mn = 1 (by denition of equivalence mod n)
which means that gcd(a, n) = 1 which contradicts our assumption that a(1 < a < n) is a divisor of n.
Exercise 1.4.5
If [F[ = q is nite, then there are at most q
n
2
possible n n matrices; GL
n
(F) is a subset of these matrices
and at least one such matrix has a zero determinant, so [GL
n
(F)[ q
n
2
and thus GL
n
(F) is nite. If [F[ is
innite, then fI GL
n
(F) for each f F (where I is the identity matrix) and thus [GL
n
(F)[ [F[ which
means [GL
n
(F)[ is innite.
Exercise 1.4.6
see previous exercise
Exercise 1.4.7
The determinant of a 2 2 matrix is given by the formula
det
__
a b
c d
__
= ad bc
So we see that the deterimant is zero if ad = bc. Using basic combinatorics, its easy to show that:
We can choose a, d such that ad = 0 in p + (p 1) distinct ways
We can choose a, d such that ad = 1 in p 1 distinct ways
We can choose a, d such that ad = 2 in p 1 distinct ways
. . .
We can choose a, d such that ad = p 1 in p 1 distinct ways
The same holds true for the number of ways we can choose b, c. Which means
We can choose a, b, c, d such that ad = bc = 0 in (p + (p 1))
2
distinct ways
We can choose a, b, c, d such that ad = bc = 1 in (p 1)
2
distinct ways
We can choose a, b, c, d such that ad = bc = 2 in (p 1)
2
distinct ways
. . .
We can choose a, b, c, d such that ad = bc = p 1 in (p 1)
2
distinct ways
12
Thus there are (2p 1)
2
ways to have ad and bc equal to 0, and (p 1)
2
ways to have them equal each of the
(p 1) other values. Thus the total number of ways we can construct a 2 2 matrix with ad = bc is
(2p 1)
2
+ (p 1)(p 1)
2
= p
3
+ p
2
p
And since there are p
4
possible 2 2 matrices over F
p
, the total number of such matrices with nonzero deter-
minants is
p
4
p
3
p
2
+ p
Exercise 1.4.8
Let A be the matrix with a
1,2
= 1 as the only nonzero entry, and let B be the matrix with a
2,1
as the only
nonzero entry. Then AB has a
1,1
= 1 as the only nonzero entry while BA has a
2,2
as the only nonzero entry.
Exercise 1.4.9
We want to show that
__
a
1
b
1
c
1
d
1
_ _
a
2
b
2
c
2
d
2
___
a
3
b
3
c
3
d
3
_
=
_
a
1
b
1
c
1
d
1
_ __
a
2
b
2
c
2
d
2
_ _
a
3
b
3
c
3
d
3
__
This can be done tediously through algebra.
Exercise 1.4.10a
_
a
1
b
1
0 c
1
_ _
a
2
b
2
0 c
2
_
=
_
a
1
a
2
a
1
b
2
+ b
1
c
2
0 c
1
c
2
_
Exercise 1.4.10b
We want to nd values of a
2
, b
2
, and c
2
such that the product in part (a) is the identity matrix. Its immediately
clear that we need to have a
2
= a
1
1
and c
2
= c
1
1
. With these substitutions, we have
a
1
b
2
+ b
1
c
2
= a
1
b
2
+ b
1
c
1
1
which equals 1 exactly when b
2
= a
1
1
(1 b
1
c
1
1
). So the inverse is
_
a
1
1
a
1
1
(1 b
1
c
1
1
)
0 c
1
1
_
This is an element of G since a
1
1
,= 0, c
1
1
,= 0.
Exercise 1.4.10c
G is a group: weve shown closure under the operation in part (a), closure of inverses in part (b), associativity
in exercise 9, and its clear that the identity matrix is an element of G. Were told that a ,= 0 and c ,= 0, so all
elements of G have a nonzero determinant ac 0b = ac.
Exercise 1.4.10d
Parts (a) through (c) are still valid after adding the further restriction that a
1
= c
1
. The proof changes very
little.
13
Exercise 1.4.11a
XY =
_
_
1 a b
0 1 c
0 0 1
_
_
_
_
1 d e
0 1 f
0 0 1
_
_
=
_
_
1 d + a e + af + b
0 1 f + c
0 0 1
_
_
To prove non-abelianism, we see calculate Y X:
Y X =
_
_
1 d e
0 1 f
0 0 1
_
_
_
_
1 a b
0 1 c
0 0 1
_
_
=
_
_
1 a + d b + dc + e
0 1 c + f
0 0 1
_
_
So we have XY ,= Y X whenever af ,= cd. An explicit example can be given by letting a = b = f = 0, c = d =
e = 1.
Exercise 1.4.11b
We want to nd values of d, e, and f such that both of the products in part (a) are the identity matrix. This
immediately yields a system of equations:
d + a = 0
e + af + b = 0
b + dc + e = 0
c + f = 0
Whose solution is d = a, f = c, e = ac b. So the inverse matrix is
X
1
=
_
_
1 a ac b
0 1 c
0 0 1
_
_
Exercise 1.4.11c
Associativity can be proven with tedious algebra. The previous parts of this exercise show that H(F) is a group.
The fact that each of the 3 entries can take [F[ possible values implies that o(H(F)) = [F[
3
.
Exercise 1.4.11d
Too tedious to typeset.
Exercise 1.4.11e
Let X be an arbitrary element of H(R). We prove by induction that for all n N, the matrix X
n
has the form
X
n
=
_
_
1 na nb +
n(n1)
2
ac
0 1 nc
0 0 1
_
_
The case for k = 1 is trivial. For k = 2 we have
X
2
=
_
_
1 a b
0 1 c
0 0 1
_
_
2
=
_
_
1 2a 2b + ac
0 1 2c
0 0 1
_
_
Now assume that we have established the form of X
k
. For X
k+1
:
X
k+1
= XX
k
=
_
_
1 a b
0 1 c
0 0 1
_
_
=
_
_
1 ka kb +
k(k1)
2
ac
0 1 kc
0 0 1
_
_
=
_
_
1 (k + 1)a (k + 1)b +
k(k+1)
2
ac
0 1 (k + 1)c
0 0 1
_
_
So that the proof by induction is complete. From this, we see that X
n
= I only if na = 0, nc = 0, and
nb + n(n + 1)ac = 0. This occurs only when a = b = c = 0: that is, when X is the identity matrix. Thus every
nonidentity element has innite order.
14
Exercise 1.5.1
o(1) = 1, o(1) = 2, o(i) = o(j) = o(k) = 4.
Exercise 1.5.3
All the given relations of Q
8
can be derived from 1, i, j, k[i
2
= j
2
= k
2
= 1, ij = k:
ij = k iij = ik j = ik, i = kj
ij = k kij = 1 ij = k j = ki, i = kj
Exercise 1.6.1
Trivial proof by induction on n. Its trivial for n = 1, and true by denition of homomorphism for n = 2.
Assuming it holds for n = k, we have
(x
k+1
) = (x
k
x) = (x
k
)(x) = (x)
k
(x) = (x)
k+1
From this, we have
1 = (x
0
) = (x
n
x
n
) = (x
n
)(x
n
) = phi(x
n
)(x)
n
so that, by the denition of inverses,
(x
n
) = ((x)
n
)
1
= (x)
n
Exercise 1.6.2 lemma
From the previous exercise, we know that (x
0
) = (x)
0
so that (1
G
) = 1
H
. Now, choose an arbitrary x G
and suppose o(x) = n is nite.
x
n
= 1
G
assumed
(x
n
) = (1
G
) is well-dened
(x)
n
= 1
H
from previous exercise
o((x))[n
o((x))[o(x) denition of n
Note that this last step also implies that o((x)) is nite. Now assume that o((x)) = m is nite:
(x)
m
= 1
H
assumed
(x
m
) = (1
G
) is well-dened
x
m
= 1
G
is an isomorphism, thus 1-to-1
o(x)[m
o(x)[o((x)) denition of m
This last step implies that o(x) is nite. Thus we have shown that o(x) is nite i o(x) is nite, and if either is
nite then o(x) = o((x)). The result is not true if is only assumed to be a homormorphism (the step requir-
ing isomorphism is clearly labeled). As a counter example, consider the following homormorphism : f : Z Z
dened as f(n) = 1. f(3)f(2) = 1 1 = 1 = f(6), but clearly o(3) = while o(f(3)) = 1.
Exercise 1.6.3
G is abelian assumed
(a, b G)ab = ba def. of abelianism
(a, b G)(ab) = (ba) isomorphisms are well-dened and bijective
(a, b G)(a)(b) = (b)(a) def. of homomorphisms
(G) is abelian def. of abelianism
H is abelian isomorphisms are surjective
Each step in this proof is bidirectional, so weve proven that when is isomorphic, then G is abelian i H is
15
abelian. If is only a homomorphism, then we have the unidirectional proof:
G is abelian assumed
(a, b G)ab = ba def. of abelianism
(a, b G)(ab) = (ba) isomorphisms are well-dened
(a, b G)(a)(b) = (b)(a) def. of homomorphisms
(G) is abelian def. of abelianism
This shows that when is a homormorphism, then G is abelian implies (G) is abelian.
Note that nothing can be assumed from the abelianism of H or (G). Consider : G 1
H
dened as (g) = 1
H
.
(G) = H is trivially abelian, but G can be any group whatsoever.
Exercise 1.6.4
C 0 has an element of order 4 (i), but no such element exists in R. This contradicts exercise 1.6.2.
Exercise 1.6.5
Proof 1: There can be no bijective function between the two sets, as proven by Cantors Theorem.
Proof 2: suppose : Q R were an isomorphism. Let x R be the element such that (2) = x; let a Q be
the element such that (a) =

k. From this, we have


(a
2
) = (a)
2
= (

k)
2
= k = (2)
Which, since is a bijection, means that a
2
= 2 and thus a =

2. But this cant be true of any a Q.


Exercise 1.6.6
Let Q has an element of order 3 (
1
3
), but there is no such element in Z. This contradicts exercise 1.6.2.
Exercise 1.6.7
D
8
has only one element of order 4 (o(r) = 4) while Q
8
has three such elements (i, j, k). This contradicts exercise
1.6.2.
Exercise 1.6.8
Assuming m, n > 0, S
m
contains m! elements while S
n
contains n!, so there can be no bijection between them
unless n = m.
Exercise 1.6.9
D
24
has an element of order 12 (r) while S
4
can have no such element (exercise 1.3.15)
Exercise 1.6.10a
a = b

1
(a) =
1
(b) is bijective

1
(a) =
1
(b) permutations are well-dened

1
(a) =
1
(b) permutations are well-dened
Exercise 1.6.10b
The same logic used in part (a) can show that
1

1
is well-dened, and this clearly acts as an inverse
to .
16
Exercise 1.6.10c
( ) = ( )
1
= (
1
)
1
= () ()
Exercise 1.6.11
The function f : AB B A dened as f(a, b) = (b, a) is easily shown to be an isomorphism.
Exercise 1.6.12
The function f : (A B) C A (B C) dened as f( ((a, b), c) ) = (a, (b, c)) is easily shown to be an
isomorphism.
Exercise 1.6.13
Verifying the group properties is tedious but easy. We just need to show that 1
H
(G) and that (G) is
closed under its operation and inverses. If is injective then the homomorphism is bijective, since is clearly
surjective onto (G), and bijective homomorphisms are isomorphisms.
Exercise 1.6.14
Verifying the group properties of K is tedious but easy. Its clear that cant be injective if more than one
element maps to 1 H, so is injective only if K = 1
G
. Proof by contrapositive that is injective if
K = 1
G
: Assume that K = 1
G
.
(a) = (b)
(a)(b)
1
= 1
H
(G) is a group by previous exercise
(a)(b
1
) = 1
H
(ab
1
) = 1
H
ab
1
= 1
G
K = 1
G

a = b
Exercise 1.6.15
(x, y) is in the kernel of if (x, y) = x = 1, so the kernel is
K = (x, y) R
2
[x = 1 = 1 R
Exercise 1.6.16
Following the logic above, we see that the kernel of
1
is
K = (a, b) AB [ a = 1
A
= 1
A
B
and the kernel of
2
is
K = (a, b) AB [ b = 1
B
= A1
B

Exercise 1.6.17
Let : G G be dened by (g) = g
1
. This function is clearly onto, so (G) = G. So we are asked to prove
that is a homomorphism i G is abelian. First assume that is a homomorphism:
(a
1
b
1
) = (a
1
)(b
1
) assumed
((ba)
1
) = (a
1
)(b
1
) properties of inverses
ba = ab denition of
Thus (G) = G is abelian. Now assume that G is abelian:
17
b
1
a
1
= a
1
b
1
assumed
(ab)
1
= a
1
b
1
property of inverses
(ab) = (a)(b) denition of
Thus is a homomorphism.
Exercise 1.6.18
Let : G G be dened by (g) = g
2
.
ba = ab
a(ba)b = a(ab)b
a(ba)b = a(ab)b left- and right-multiplication
(ab)(ab) = (aa)(bb) associativity
(ab) = (a)(b) denition of
Exercise 1.6.19
As denied here, G is the set of nite roots of unity in C. From complex analysis, we know that for each
k N there are k distinct elements of order k. Let k be xed and dene f
k
as f
k
(z) = z
k
. This is clearly a
homomorphism, and is surjective since
z G z
1/k
G f
k
(z
1/k
) = z
But by exercise 14, f
k
cannot be injective: the k roots of unity of order k mean that the kernel of f
k
is of size k.
Exercise 1.6.20
The identity element is the identity mapping; isomorphisms are invertible and therefore have inverses in Aut(G).
Associativity and closure is inherited from the properties of function composition.
Exercise 1.6.21
Dene f
k
: Q Q to be f
k
(q) = kq. This function is clearly injective and a homomorphism. To prove
surjectivity:
q Q
q
k
Q f
k
_
q
k
_
= q
Exercise 1.6.22
Dene f
k
: A A to be f
k
(a) = a
k
. Since A is abelian, we have
f
k
(ab) = (ab)
k
= a
k
b
k
= f
k
(a)f
k
(b)
If k = 1, then the function is injective (a = b a
1
= b
1
f(a) = f(b)) and surjective (a A
f(a
1
) = a).
Exercise 1.6.23
Let be an automorphism such that
2
is the identity map and (g) = g i g = 1. Choose arbitrary elements
a, b G. is bijective, so there are x, y G such that (x) = a, (y) = b. From this we have
(xy)
2
= xy = (x)
2
(y)
2

2
is the identity map
(x)(y)(x)(y) = (x)(x)(y)(y) is a homomorphism
abab = aabb denition of a, b
ba = ab left- and right- cancellation of previous step
and since a, b were arbitrary this suces to prove that G is abelian.
18
Exercise 1.6.24
We need to show that mapping preserves the properties of each generator and relation. Were told that x
2
=
r
2
= 1 and y
2
= s
2
= 1. From the exercise 1.2.6, the fact that x
2
= y
2
= 1 is sucient to conclude that
xy = yx
1
.
Exercise 1.6.25a
Let v = [v[ cos() +[v[ sin() be an arbitrary vector. The given matrix transforms v as follows:
_
cos() sin()
sin() cos()
_ _
[v[ sin()
[v[ cos()
_
=
_
[v[[cos() cos() sin() sin()]
[v[[sin() cos() + cos() sin()]
_
which, via the angle addition formulas from the trigonmetric identities, is equivalent to
_
[v[ cos( + )
[v[ sin( + )
_
which is, of course, the original vector with its endpoint rotated by an additional radians counterclockwise
about the origin. Our vector v was arbitrary, so every vector endpoint (and thus every point in R
2
) is also
rotated in the same way.
Exercise 1.6.25b
We need to show that mapping preserves the properties of each generator and relation. represents a rotation
of 2/n radians, so clearly o((r)) = n. And (s)
2
= I, so o((s)) = 2. We can show that the relationship
rs = sr
1
has an associated relationship (r)(s) = (s)(r)
1
by some tedious algebraic verication. The
text (bottom of p38) assures us that this is sucient to guarantee an isomorphism between G and D
2n
.
Exercise 1.6.26
Further dene as follows:
(1) = I, (k) =
_
0

1 0
_
We need to show that mapping preserves the properties of each generator and relation. There are a lot of
relations that need to be algebraically veried (e.g., (i)
2
= (1), (i)(j) = k) but they are trivial (albeit
tedious). The identity is the only element of Q
8
that maps to I GL
2
(C), so by exercise 1.6.14, this is sucient
to prove that is injective.
Exercise 1.7.1
Let F be a eld and dene a group action of G = F

on A = F by g a = ga. The element 1 F

satises
property (i) of group actions (1 a = a for all a F). To prove property (ii), we note that (F, ) is not a group
(0 has no inverse), but it is still a semigroup (its associative and closed under its operation). Property (ii) is
then justied as follows:
(g
1
g
2
) a
= g
1
g
2
a denition of the group action; F is closed under multiplication
= g
1
(g
2
a) multiplication in F is associative
= g
1
(g
2
a) denition of group action
= g
1
(g
2
a) denition of group action
19
Exercise 1.7.2
The element 0 Z satises property (i) of group actions (0 a = a for all a Z). To justify property (ii), let
g
1
, g
2
, a be arbitrary elements of Z:
(g
1
+ g
2
) a
= (g
1
+ g
2
) + a denition of the group action; Z is closed under addition
= g
1
+ (g
2
+ a) addition in Z is associative
= g
1
+ (g
2
a) denition of group action
= g
1
(g
2
a) denition of group action
Exercise 1.7.3
The element 0 R satises property (i) of group actions. To justify property (ii), let r
1
, r
2
be arbitrary elements
of R and let (x, y) be an arbitrary point in R R:
(r
1
r
2
) (x, y)
= (r
1
+ r
2
) (x, y) operation on R is addition
= (x + (r
1
+ r
2
)y, y) denition of group action
= (x + (r
1
+ r
2
)y, y) operation on R is addition
= ((x + r
1
y) + r
2
y, y associativity, distributive property of eld R R
= r
2
(x + r
1
y, y) denition of group action
= r
2
(r
1
(x, y)) denition of group action
= g
1
+ (g
2
+ a) addition in Z is associative
= g
1
+ (g
2
a) denition of group action
= g
1
(g
2
a) denition of group action
Exercise 1.7.4a
Let K represent the kernel of the action (g G : g a = a for all a A). By the subgroup criterion (proven in
chapter 2), we need show that a left identity exists and that a, b K ab
1
K. Its clear that 1 K, so an
identity exists. Now assume that g
1
, g
2
K:
1, g
1
, g
2
K assumed
(a A)g
1
a = a g
2
a = a 1 a = a denition of K
(a A)g
1
(g
2
a) = a algebraic substitution
(a A)(g
1
g
2
) a = a property (ii) of group actions
g
1
g
2
K denition of K
Exercise 1.7.4b
Fix some a A and let S represent the stabilizer of a in G. By the subgroup criterion (proven in chapter 2), we
need show that a left identity exists and that a, b S ab
1
S. Its clear that 1 S, so an identity exists.
Now assume that g
1
, g
2
S:
20
1, g
1
, g
2
S assumed
g
1
a = a g
2
a = a 1 a = a denition of S
g
1
(g
2
a) = a algebraic substitution
(g
1
g
2
) a = a property (ii) of group actions
g
1
g
2
S denition of S
Exercise 1.7.5
Each step in the following proof is bidrectional (i):
g K assumed
(a A)ga = a denition of K
(a A)
g
(a) = a denition of
g

g
is the identity permutation on G denition of the identity function

g
is the identity element of S
A
denition of the group S
A
g is in the kernel of : G S
A
denition of kernel,
Exercise 1.7.6
Proof by contradiction. Assume that G is not faithful: then there are distinct nonidentity elements g
1
, g
2
such
that g
1
a = g
2
a for all a A. From this, we obtain
(g
1
1
g
2
) a = g
1
1
(g
2
a) = g
1
1
(g
1
a) = (g
1
1
g
1
) a = 1 a = a
and thus g
1
1
g
2
K. And this element cannot be the identity since g
1
,= g
2
. Thus the kernel contains a
nonidentity element. By contrapositive, if K = 1 then G is faithful.
Exercise 1.7.7
The kernel of the given action is 1; by the previous exercise, this suces to prove that the action is faithful.
Exercise 1.7.8a
Let G = S
A
and let B = T(A). The identity permutation
1
satises property (i) of group actions (
1
(b) = b for
all b B). To show that property (ii) is satised, let
g
,
h
be arbitrary elements of G and let b be an arbitrary
element of B:
(
g

h
) b
= (
g

h
)(b) denition of
=
g
(
h
(b)) denition of
=
g
(
h
b) denition of
Exercise 1.7.8b
The element (1 2) acts on each subset by replacing 1 (if it exists) with 2 and vice-versa. For example, (1 2)1, 4 =
2, 4. The element (1 2 3) replaces each 1 with 2, each 2 with 3, and each 3 with 1. For example, (1 2 3)2, 3, 4 =
3, 1, 4.
Exercise 1.7.9
The proof in 1.7.8(a) and the description in 1.7.8(b) still apply when subsets are replaced with ordered k-tuples.
21
Exercise 1.7.10a
We prove that the action is faithful for k < [A[ or k [Z[.
case 1) Suppose [A[ is nite and k < [A[. We show that the action of S
n
on k-element subsets of A is faithful.
Let
x
,
y
S
n
be any two distinct permutations of A. Since these permutations are distinct, there is
some a A such that
x
(a) ,=
y
(a). From the invertibility of permutations this gives us the inequality

x
(a) ,=
y
(a)
1
y

x
(a) ,=
1
y
(
y
(a)) = a
Since 1 k < n, we can choose a k-element subset B A such that a B but
1
y
(
x
(a)) , B. We
can now demonstrate that the action is faithful, since
x
(B) contains
x
(a), but
y
(B) does not contain

y
(
1
y
(
x
(a))) =
x
(a). Thus
x
and
y
do not perform the same action on B. But these were arbitrary
elements of S
A
, thus no two elements of S
A
perform the same action. By denition, this means that S
A
is faithful on the set of k-element subsets of A.
case 2) Suppose [A[ is nite and k = [A[. There is only one distinct subset of size k: A itself. And every
permutation of A is still the same set (just rearranged). Thus S
n
is the trivial action on k-element subsets
of [A[.
case 3) Suppose [A[ is innite. For all nite k, we can follow the logic of case (1) and conclude that the action
is faithful. If k is also innite, then we can still choose an arbitrary a A and let B = Aa, and then
follow the logic of case (1) and conclude that the action is still faithful.
Exercise 1.7.10b
Choose two arbitrary permutations
x
,
y
S
n
. For these to be distinct, there must be some a A such that

x
(a) ,=
y
(a). Having chosen such an a, let B be the k-tuple consisting of the element a repeated k times. Its
clear that
x
(B) ,=
y
(B), so the action of S
A
is faithful. But the value of k was never specied, so this proof
holds for all values of k (nite and innite).
Exercise 1.7.12
Note that a regular n-gon is a two-dimensional shape; the three-dimensional version is a regular n-hedron. Let n
be even, and label the vertices of the n-gon clockwise as 0, 1, 2, . . . , n 1. Let the n/2 pairs of opposite vertices
be represented by the set of ordered pairs a
i
dened as
a
i
=
__
i,
n
2
+ i
_
: 0 i <
n
2
_
We dene the action of D
2n
on the elements of a
i
so that the elements of D
2n
permute the set as follows:
r
k
a
i
= a
(ik) mod n
s
k
a
i
= a
(ikn/2) mod n
To prove this is an action, we note that r
0
a
i
= a
i
for all a
i
, so property (i) of group actions is satised. To
verify property (ii), let r
a
s
b
and r
x
s
y
be two arbitrary elements of D
2n
:
(r
a
s
b
r
x
s
y
) a
i
= a
j
, with j = ((((i ny/2) x) nb/2) a) mod n denition of , associativity of modular arith-
metic
= r
a
s
b
a
k
, with k = ((i ny/2) x) mod n denition of , associativity of modular arith-
metic
= r
a
s
b
(r
x
s
y
a
i
) denition of
The kernel of this action is r
0
= 1.
22
Exercise 1.7.13
We prove that the identity element of G is the only element in the kernel K. Property (i) of group actions
guarantees that 1
G
K. But let g G be an arbitrary non-identity element of G and choose 1
G
A = G.
Then g1
G
,= 1
G
and so g , K.
Exercise 1.7.14
We prove by contradiction that property (ii) of group actions is not satised. Assume A = G is non-abelian and
dene the action g a = ag. Let g
1
, g
2
be arbitrary elements of G. Hypothesis to be contradicted: suppose that
property (ii) of group actions were satised.
(g
1
g
2
) a
= a(g
1
g
2
) denition of this group action
= (ag
1
)g
2
) associativity of operation on G
= g
2
(ag
1
)) denition of this group action
= g
2
(g
1
a)) denition of this group action
= (g
2
g
1
) a)) property (ii) of group actions
This demonstrates that g
1
commutes with g
2
. But these were arbitrary elements of G so all elements of G
commute, which means that G is abelian. This contradicts our initial assumption that property (ii) of group
actions is satised.
Exercise 1.7.15
Let A = G and dene the action g a = ag
1
. The identity element 1
G
satises the condition 1 a = a for all
a A. To verify property (ii):
(g
1
g
2
) a
= a(g
1
g
2
)
1
denition of this group action
= a(g
1
2
g
1
1
) property of inverses
= (ag
1
2
)g
1
1
associativity of operation on G
= g
1
(ag
1
2
) denition of this group action
= g
1
(g
2
a)) denition of this group action
Exercise 1.7.16
Let A = G and dene the action g a = gag
1
. The identity element 1
G
satises the condition 1 a = a for all
a A. To verify property (ii):
(g
1
g
2
) a
= (g
1
g
2
)a(g
1
g
2
)
1
denition of this group action
= (g
1
g
2
)a(g
1
2
g
1
1
) property of inverses
= g
1
(g
2
(ag
1
2
)g
1
1
associativity of operation on G
= g
1
(g
2
ag
1
2
) denition of this group action
= g
1
(g
2
a)) denition of this group action
Exercise 1.7.17
Let G be a group and x a value for g G. Dene a mapping f : G G as f(x) gxg
1
. We want to prove
that the given function is an isomorphism. We do so by showing that its a bijective homomorphism.
23
f is a homomorphism
f(xy) = g(xy)g
1
= g(xg
1
gy)g
1
= (gxg
1
)(gyg
1
) = f(x)f(y)
f is surjective
Choose h G. Then g
1
hg G and f(g
1
hg) = h. Thus f is surjective
f is injective:
f(x) = 1 i gxg
1
= 1 i x = g
1
g = 1. The kernel of f consists only of the identity element; by exercise 1.6.14,
this is sucient to prove that f is injective.
Exercise 1.7.18
reexive:
1 H and a = 1a so a a.
transitive:
If a b then a = hb, which implies b = h
1
a, which implies b a.
symmetric:
If a b and b c, then a = bh
1
and b = ch
2
, which implies a = ch
2
h
1
, which implies a c.
Exercise 1.7.19
Let H be a subgroup of nite group G, and let H act on G by left multiplication. Fix an element x G and
let O
x
be the orbit of x under the action of H. Dene the map f : H O
x
as h hx. We prove that f is a
bijection.
f is surjective:
O
x
is dened as the set hx : h G which is exactly the image of f(H).
f is injective:
h(a) = h(b) i ha = hb i a = b (by left-cancellation).
Weve shown that [H[ = [O
x
[. But H was an arbitrary subgroup and x was an arbitrary element of G: thus
[O
x
[ = [O
y
[ for all x, y G. Applying the preceeding exercise, we conclude that G can be partitioned into
disjoint sets of size [O
x
[ = [H[; this can only occur if [G[ is an integer multiple of [H[.
Exercise 1.7.20
Label the 4 vertices of the tetrahedron as 1, 2, 3, 4. Let G be the group of rigid motions of the tetrahedron.
Each rigid motion corresponds to some permutation of the 4 vertices, so dene the function : G S
4
so that (g) is the permutation corresponding to the rigid motion g. Its clear that is a homomorphism:
(g
1
g
2
) = (g
1
)(g
2
), since each side of the equation represents the rigid motion g
1
followed by the rigid motion
g
2
. is injective: the kernel of consists only of the identity element of G, so is injective by exercise 1.6.14.
Thus G is isomorphic to (G), which is a subgroup by exercise 1.6.13.
24
Exercise 1.7.21
Choose one face of the cube to call the front and label its vertices a
1
, a
2
, a
3
, a
4
. On the opposite face, label
the diagonally opposite points to be b
1
, b
2
, b
3
, b
4
. This labels all eight vertices of the cube. No matter how we
rotate this cube, we know several things:
exactly four of these vertices will be on the front of the cube
for each i exactly one of a
i
, b
i
will be on the front of the cube (because they are diagonally opposite)
no rigid motion consists only of swapping one or more points with the points diagonally opposite (i.e.,
swap one or more a
i
with its corresponding b
i
)
This last point is particuarly important: it means that if there were a set of rigid motions that would give us
front vertices of (arbitrary example) (a
1
, b
2
, b
3
, a
4
), then none of the 15 other ordered 4-tuples we could form
from replacing a
i
with b
i
(or vice-versa) would represent a rigid motion. This means that each rigid motion of
the cube can be uniquely expressed as a permutation of (1, 2, 3, 4). We now follow the logic of the preceeding
exercise to prove an isomorphism.
Let G be the group of rigid motions of the cube. Each rigid motion corresponds to some permutation of the
4 pairs of opposite vertices, so dene the function : G S
4
so that (g) is the permutation corresponding to
the rigid motion g. Its clear that is a homomorphism: (g
1
g
2
) = (g
1
)(g
2
), since each side of the equation
represents the rigid motion g
1
followed by the rigid motion g
2
. is injective: the kernel of consists only of the
identity element of G, so is injective by exercise 1.6.14. Thus G is isomorphic to (G), which is a subgroup
by exercise 1.6.13.
Exercise 1.7.22
We can apply the preceeding exercise with opposite faces taking the place of opposite vertices. This is intuitively
true: we could place a cube inside an octohedron by placing one vertex of the cube in the center of each face
of the octohedron, and the rigid motions of the octohedron would correspond perfectly with the rigid motions
of the cube. So the rigid motions of an octahedron are isomorphic to S
4
itself (the errata for this textbook
eliminates the subgroup qualication, although S
4
is technically a subgroup of S
4
).
Exercise 1.7.23
Label the front, top, and one side face of the cube as (respectively) A, B, C. Label the respective opposite faces
D, E, F. The set of rigid motions of the cube can be made up from various combinations of the three basic
rotations:
r
1
= (A B D E), r
2
= (A C D F), r
3
= (C B F E)
When these rotations act on the set of opposite vertices, the action faithful. But when they act on the set of
opposite faces, they are equivalent to
r
1
= (A, D B, E), r
2
= (A, D C, F), r
3
= (C, F B, E)
so that r
2
1
, r
2
2
, and r
2
3
are all equal to the identity motion. The kernel of the action becomes
K = i, r
2
1
, r
2
2
, r
2
3

(Any product of elements of K can be simplied to one of these elements). Note that 4[24, in concordance with
exercise 19.
25

You might also like