Sei sulla pagina 1di 295

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
don't
understand
something
in
one
of
these
proofs,
it
is
very
possible
that
it
doesn't
make
sense
because
it's
wrong.
Any
questions
or
corrections

can
be
directed
to
jason.rosendale@gmail.com.
Exercise
1.1.1
(a)
is
not
associate:
3
(2
1)
6
=
(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

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
x?y
is
well-de

ned
follows
directly
from
the
fact
that
+,-,
and
the
greatest
integer
function
are
well
de

ned
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
(zn
=
1
implies
(z..1)n
=1
so
z..1
.
G)
and
complex
multiplication
is
associative.
To
show
closure:
7
.
x,
y
.
G
assumed
.
(9m,
n
.
Z+)xm
=1
.
yn
=
1

def.
of
membership
in
G
.
(9mn
.
Z+)xmn
=1n
=1
.
ymn
=1m
=1
mn
.
(9mn
.
Z+)xymn
=1
.
(9mn
.
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
+
v
b
2)x
=
1
and
is

given
by:
(a,
b)..1
=
a,
b
a2
2b2
a2
2b2
Exercise
1.1.10
The
de

nition
of
a
symmetric
matrix
is
that
for
all
i,
j
we
have
aij
=
aji.
In
the
group
table,
this
is
the
case
iff
for
all
i,
j
we
have
aiaj
=
aj
ai
which
is
the
de

nition
of
abelian.
Exercise
1.1.11
General
formula
is:
ja|
=
lcm(a,
12)=a.
j0|
=1
j1|
=
12
j2|
=6
j3|
=4
j4|
=3
j5|
=
12
j6|
=2
j7|
=
12
j8|
=3
j9|
=4
j10|
=6
j11|
=
12
2

Exercise
1.1.12
j1|
=1
j1|
=2
j7|
=2
j7|
=2
j13|
=2
Exercise
1.1.13
General
formula
is:
ja|
=
lcm(a,
36)=a.
j1|
=
38
j2|
=
18
j6|
=6
j9|
=4
j10|
=
18
j12|
=3
j1|
=
36
j10|
=
18
j18|
=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,
it's
clear
that
a1
=
a
..1
.
Supposing
the
equality
holds
for
the
n
=
k
1
case,
we
have:

..1
..1
..1
(a1a2
:::akak+1)(a
:::a
)
k+1ak
1
..1
..1
..1
=(a1a2
:::ak)ak+1a
(a
:::a
)
associativity
k+1k
1
..1
..1
=(a1a2
:::ak)1(ak
:::a
1
)
de

nition
of
inverses
..1
..1
=(a1a2
:::ak)(ak
:::a
1
)
de

nition
of
identity
=
e
equality
holds
for
n
=
k
..1
..1
..1
and
thus
(a
:::a
)
is
shown
to
be
the
right
inverse
of
(a1a2
:::akak+1).
The
proof
for
left
inverse
is
k+1ak
1
similar.
Thus
equality
holds
for
the
n
=
k
+
1
case.
By
induction,
equality
holds

for
all
n
.
Z+
.
Exercise
1.1.16
n
This
follows
directly
from
the
de

nition
of
jxj.
The
order
of
x
is
the
smallest
integer
n
such
that
x=
1;
if
x2
=
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
jx|
=
n,
then
xn
=
1,
and
thus
x

xn..1
=

1,
which
makes
xn..1
the
inverse
of
x.
3

Exercise
1.1.18
7
.
xy
=
yx
assumed
..1..1
.
yxy
=
yyx
left
multiplication
..1
.
yxy
=1x
=
x
de

nition
of
inverse
..1..1..1
.
xyxy
=
xx
left
multiplication
..1..1
.
xyxy
=
1
de

nition
of
inverse
..1..1)..1
.
(xy=
xy
de

nition
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
ja|
=
jb|
(absolute
value,
not
order)

then
clearly
a+b
0
a
..a
ab
x
=
x
=
xx
=
xx
ow
assume,
without
loss
of
generality,
that
ja|
<
jb|
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:
ab
a..a+b+a
xx=
xxfrom
b
=
b
+
a
a
a..ab+a
=
xxxfrom
part
(a),
since
..a
and
b
+
a
have
same
sign
b+a
=(xleft
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
xn
=
1
must
also
be
the
least
n
such
that
(x..1)n
=
1.
Exercise
1.1.21
7
.
x2k+1
=
1
assumed,
since
n
=2k
+1
2k+1
.
xx
=1x
=
x
right
multiplication
2k+2
.
x=
x
1.1.19(a)
2)k+1

.
(x=
x
1.1.19(a)
4

Exercise
1.1.22
7
.
xn
=
1
assumed
.
(xgg..1)n
=1
gg..1
=1
..1
.
xg(gxg)n..1g..1
=
1
tricky
associativity
..1..1..1
.
gxg(gxg)n..1gg
=
g..11g
=
1
left
and
right
multiplication
..1..1
.
gxg(gxg)n..1
=
1
right
cancellation
..1
.
(gxg)n
=
1
1.1.19(a)
n
..1
This
shows
that
the
least
n
such
that
x=
1

must
also
be
the
least
n
such
that
(gxg)n
=
1.
To
prove
that
b..1
jab|
=
jbaj,
let
x
=
ab
and
g
=
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
st
n

(xs)t
=
x=
x=
1,
so
jxsj=
t.
But
if
there
were
some
k
such
that
0
=
k<t
such
that
(xs)k
=
1,
then
we
would
have
0
=
sk
<
st
=
n
with
xsk
=
1
contradicting
the
assumption
that
jx|
=
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)kb
tricky
associativity
(=
a(ab)kb
a;b
commute
(=
aakbkb
equality
holds
for
n
=
k
k+1bk+1
(=
a
Thus
equality
holds
for
all
n
.
N.
To
prove
that
this
holds
for
negative
n,
just
take
the

inverse
of
each
side
of
=
b..n=
b..n..n
the
equation
to
yield
(ab)..n
a..n,
then
apply
commutativity
to
conclude
(ba)..n
a.
Exercise
1.1.25
Given
that
x2
=
1,
we
see
that
(ab)(ab)=
b(aa)b
=
1.
Right-multiplying
by
b..1a..1
yields
ab
=
ba.
Exercise
1.1.26
We're
told
that
H
is
nonempty
so

it
contains
some
element
h;
it's
closed
under
inverses
and
the
binary
operation,
it
contains
the
identity
element
hh..1
=
1.
It
inherits
associativity
from
G.
That's
sucient
for
H
to
be
a
group.
Exercise
1.1.27
n
Let
H
=
x:
n
.
Z.
The
identity
x0
=
1
exists
in
n;
if
xn
.

H
then
x..n
.
H
so
H
is
closed
under
inverses;
it's
iki+j+kj
trivially
closed
under
the
binary
operation;
and
(xxj
)x=
x=
xi(xxk)
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
7
.
(a,
b)


(c,
d)=(c,
d)

(a,
b)
.
(ac,
bd)=(ca,
db)
def.
of

.
ac
=
ca
and
bd
=
db
def.
of
equality
in
A

B
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
=
jaj;q
=
jbj,
and
x
=[p,
q].
Then:
p)x=p
(a,
b)x
=(a
x;bx)
=
((a,
(bq)x=q)
=
(1x=p,
1x=q)
=
(1,
1)
=
1
Thus
the
order
of
(a,
b)
divides
[p,
q]=
x.
Now

let
n
=
j(a,
b)j.
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
n
(1;b..n)=(a,
1)
n
b..n
thus
a=
=
1.
Thus
p
and
q
both
divide
n,
which
means
that
[p,
q]
divides
n.

We've
now
shown
that
n
=
j(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
ab
:(ab
=1ora
=
b)
on
t(G).
This
is
clearly
re
exive,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
G
t(G)
(that
is,
g
.
Gjg
=
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
G

t(G)
has
g2
=
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
ij
..i
Suppose
x=
xj
for
0
=
i<j
=
n
1.
Then1=
xx=
xj..i,
which
would
contradict
the
claim
that

jx|
=
n.
Exercise
1.1.33a
Part
(a)
is
the
contrapositive
of
exercise
1.1.31.
Exercise
1.1.33b
ii
i
..1
It's
trivial
to
show
that
i
=
k
implies
xx=
x2k1
and
therefore
x=
x.
To
prove
in
the
other
direction:
i
..i
7
.
x=
xassumed
ii
..ii

.
xx=
xx
.
x2i
=1
i
This
shows
that
x=
x..i
iff
n
(the
order
of
x)
divides
2i.
But
2i<n
by
the
previous
exercise,
so
nj2i
implies
2i
=0
or
2i
=
n.
And
we're
asked
to
assume
that
i>
0,
so
we
must
have
2i
=
n
=2k
and
so
i
=
k.
Exercise

1.1.34
mn
jm..nj
Were
it
the
case
that
x=
xfor
some
m
6
=
n,
then
we
would
have
x=
1
which
would
contradict
the
assumption
that
x
was
of
in

nite
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
k
an+b
ab
b
x
=
x
=
x
nx
=
x
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
iff
c
is
its
own
inverse:
if
both
are
their
own
inverses,
then
the
group
is
abelian
by
1.1.25
and
we
are
done.

3
If
neither
b
nor
c
are
their
own
inverses,
then
they
must
have
order
3
(b3
=
c=
1).
We
cannot
have
b2
=
b
(for
we
would
have
o(b)
=
1)
or
b2
=
1
(for
we
would
have
o(b)
=
2)
or
b2
=
a
(for
we
would
have
b
=
b,
b2
=

a,
b3
=
ab,
b4
=
a2
=
1
so
that
o(b)
=
4)
so
we
must
have
b2
=
c.
And
this
gives
us
a
contradiction,
since
it
would
imply
that
o(b2)=
o(c)=
o(b).
Exercise
1.2.2
01n..101n..1
We
know
that
D2n
has
order
2n
with
distinct
elements
r;r;:::r;rs;rs;:::rs.
So
any
element
that
is
not

equivalent
to
rk
for
some
k
is
equivalent
to
srk
for
some
k.
We
can
now
use
a
proof
by
induction:
the
case
01k
for
x
=
rs
and
x
=
rs
is
trivial.
Now
assume
that
the
equality
holds
for
x
=
rs:
..1
k+1
..1
r(r
k+1)s
=
rr(r
k)s
=

r(r
k)sr
=
r
sr
k+1
and
thus
equality
holds
for
x
=
rs.
Exercise
1.2.3
As
above,
every
element
of
D2n
which
is
not
a
power
of
r
has
the
form
rks
=
sr..k
.
Thus:
k
kk
..kk
(rs)2
=(rs)(rs)=
sr
rs
=
ss
=1
k
This
proves
only

that
the
order
of
(rs)
divides
2;
we
must
show
that
it
is
2.
Proof
by
contradiction:
If
the
order
were
1,
then
we
would
have
rks
=
1
which
would
imply
s
=
r..k
which
would
imply
that
all
2n
elements
n
could
be
written
uniquely
as
a
power
of
r;
but
r=
1
so
there

could
be
at
most
n
such
elements.
Thus
the
k
order
of
(rs)
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
22kn
z=
r=
r=
1,
so
o(z)=2(k>
1,
so

z
is
not
the
identity
element).
r2k
trivially
commutes
with
elements
of
D2n
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
D2n
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
D2n,
we
use
a
very
unsatisfying
proof
by
exhaustion
of
cases.
Let
a
be
an
element
such
that
ab
=
ba
for
all
b
.
D2n.
Either
a
=
ri
or
a
=
sri
for
some
i
.
0
::.
2k
1.
case
i)
Suppose
a
=
ri
and

let
b
be
an
arbitrary
element
of
D2n.
The
element
a
obviously
commutes
for
all
b
of
the
form
b
=
rj.
If
b
=
srj,
then
ab
=
ba
when:
7
.
ab
=
ba
assumed
ij
ji
.
rsr=
srrde

nition
of
a,
b
..ij
ji
..1
.
srr=
srrfrom
relation
rs
=
sr
..i+jj+i
.
r=
rleft
cancellation
..i+j
..(j+i)
.
rr..j..i
=
1
right
multiplication
by
r
.
r..2i
=1
.
nj2io(r)=
n
.
kjin
=2k
i
This
shows
that,
for
this
choice
of
a
=
r,
ab
=
ba
for
all
b
iff

i
is
a
multiple
of
k.
But
we've
assumed
that
i
0
i
.
0;:::,
2k
1
so
either
i
=0
or
i
=
k.
Thus
a
=
rcommutes
with
all
b
iff
a
=
r(the
identity)
or
k
a
=
r.
7

i
case
ii)
Suppose
a
=
sr.
Then
this
a
does
not
commute
with
all
b:
let
b
=
r.
Then
we
have
7
.
ab
=
ba
assumed
.
srir
=
rsri
de

nition
of
a,
b
i+1
..1+i
..1
.
sr=
srfrom
relation
rs
=
sr
i+1
i..1
.
r=
rleft
cancellation
i+1..(..j+i)
.
rr1..i
=
1
right
multiplication
by
r
.
r2
=1
But
n
is
de

ned
by
the
order
of
r,
so
r2
=
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
D2n(n
=2k,
n
=
4)
k
is
r.
Exercise
1.2.5
As
shown
in
the
previous
exercise,
the

only
element
that
commutes
with
element
of
D2n
for
n>
2
is
ri
where
nj2i.
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
r0
=
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
..1..1
..1..1
tx
=
xyxy
=
y;x
=
x
tx
=
x(xy)..1
tx
=
xt..1
Exercise
1.2.7
22
n
Let
a
=
s
and
b
=
sr
so
that
ab
=

r.
The
relations
follow
from
each
other:
a=
s=
1,
(ab)n
=
r=
1,
and
..1
b2
=(sr)2
=
ssrr
=
1.
Exercise
1.2.8
n
is,
by
de

nition,
the
smallest
r
such
that
rn
=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:
(number
of
faces)
*
(number
of
sides
per
face)
2
*
(number
of
sides)
=

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
22
2222
Letting
n
=
2,
we
have
r=
s=
1.
So
let
x
=
r,
y
=
s.
The
relations
follow:
x=
r=
1,y=
s=
1,
and
xy
=
y..1x
:
xy
=(xy)..1
from
o(xy)=2
..1..1
=
yx
=
yx..1
from
o(y)=2
Exercise
1.2.17a
n

2
The
presentation
in
1.2
is
hx,
yjx=
y2
=1,
xy
=
yx2i.
The
text
demonstrated
that
xy
=
yximplies
x3
=
1.
3
012
33
Thus,
if
n
=3k
then
x=
1
but
x;x;xare
distinct
elements.
Letting
x
=
r,
y
=
s
gives
us
x=
r=
1,
2
..1
y=
s2

=
1,
and
rs
=
sr:
rs
=
xy
from
o(xy)=2
2
=
yx
..12
..1
=
yxxx2
=1
.
x=
x
..1
=
sr
Exercise
1.2.17b
33k
We
know
that
x=
1
so
x=
1.
By
de

nition,
xn=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
n
3k13k
1
1
1=
x
=
x
=
xx
=
x
But
x..1
=
1
iff
x
=
1,
so
in
either
case
we
have

x
=
1
and
X2n
is
generated
uniquely
by
s,
which
has
order
2.
Exercise
1.2.18
4
22i.
The
presentation
in
1.3
is
hu,
vju=
v3
=1,
uv
=
vu
32
..12
3
..1
a)
If
v=
vv2
=
1,
then
v=
v.
Similarly,
u=
u..2
and
u=
u.
32

b)
u=
u(v3)ufrom
v3
=1
2
=(uv)(vu2)
associative
property
222
=(vu2)(uv)
from
uv
=
vu
23
=
vuv
associative
property
=
v..1u3v
from
part
(a)
333..1
..13
Left-multiplying
both
sides
by
v
gives
us
vu=
uv;
right-multiplying
by
v..1
gives
us
uv=
vu.
3..1
..13
..1..1
..1..1
c)
From
part
(b),
we
have

uv=
vu.
Using
part
(a)
this
becomes
uv=
vu.
Taking
inverses
of
each
side
yields
vu
=
uv.
2
d)
We're
told
uv
=
vu2;
by
commutativity
of
u,
v
this
becomes
uv
=(uv)2;
by
left-or
right-cancellation
this
becomes
uv
=
1.
43
e)
1=(uv)4
=
uvv
=
v,
and
uv
=
1
implies

u
=
v..1
=
1.
Exercise
1.3.8
De

ne
the
function
f
:
N
.
S.
to
be
f(n)
=
(1n).
The
function
is
clearly
injective
(albeit
not
surjective)
so
jS!jjNj.
Exercise
1.3.9
See
exercise
1.3.11.
9

Exercise
1.3.10
Trivial
proof
by
induction
on
i.
Since
maps
ak
.
ak,
it
must
be
the
identity
function;
since
m
is
the
least
integer
such
that
=
I,
we
have
o()=
m.
Exercise
1.3.11
Let
k
be
the
order
of
.
It
must
be
the
case
that
mjik
or,

by
the
division
algorithm,
we
could

nd
ik
such
that
ik
=
xm
+
b
with
0
<b<m
which
would
give
us
=
(i)k
=
=()xb
=
which
would
contradict
the
previous
exercise's
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
ab
=
lcm(a,
b)

gcd(a,
b)
.
lcm(a,
b)=
gcd(a,
b)
Exercise
1.3.12a
yes:
let
s
=(13579246810)
and
k
=
5.

Exercise
1.3.12b
no.
Let
k
be
the
smallest
k
such
that
=
(1
2)(3
4
5).
Then
=
(3
5
4).
Thus
=
1
(implying
that
mj2k
from
exercise
10)
but
=
5
(implying
m
.
2k
by
exercise
10).
This
establishes
a
contradiction
so
there
can
be
no
k
such
that

=
.
Exercise
1.3.13
Assuming
that
s
can
be
written
has
a
product
of
commuting
2-cycles,
we
have

=
[(a1
b1)
::.
(ak
bk)]2
=(a1
b1)2
::.
[(ak
bk)2
=
(1)2
::.
(1)2
=
(1)
so
that
o()
=
2.
If
we
assume
that
o()
=
2
then
s
maps

a
.
and
maps
.
=
a
(with
the
possiblity
that
a).
Consider
the
set
of
2-cycles:
f(a(a))
:
a
.
1;:::;m
and
a<(a)}
We
can
de

ne
s
as
the
product
of
every
element
of
this
set:
s
=(a1
::.
(ak
Each
a
.
1;:::;m
appears
in
at
most
one
of
these
disjoint
2-cycles,
and
appears
iff
a
6
=

Exercise
1.3.14
Follow
the
preceeding
proof.
Assuming
that
s
can
be
written
has
a
product

of
commuting
p-cycles,
we
have
=
[(a1
b1)
::.
(ak
bk)]p
=(a1
b1)p
::.
[(ak
bk)p
=
(1)p
::.
(1)p
=
(1)
And
we
can
write
s
as
the
product
of
elements
of
the
disjoint
collection
of
p-cycles:
f(a1(a)
:::p..1(a))
:
a
.
1;:::;m
and
a
=
min(a,
..1(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
S6
we
have
(1
2
3)(4
5)
which
has
order
6.
10

Exercise
1.3.15
Choose
s
.
Sn
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
=
(1)
means
that
(a1
::.
ai)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
s
must
be
the
least
such
k,
which
is
the
least
common
multiple
of
the
lengths
of
the
cycles
of

Exercise
1.3.16
n
From
n
elements,
there
are
ways
of
selecting
m
elements
and
m!
ways
of
writing
an
m..cycle

with
them.
m
Each
distinct
m-cycle
can
be
written
in
m
dierent
ways.
Thus
the
number
of
distinct
m-cycles
in
Sn
is
given
by
nm!
n!m!
n(n
1)(n
2)
::.
(n
m
+
1)
==
mm
m!(n
m)!mm
Exercise
1.3.17
n

There
are
ways
of
selecting
the
elements
of
two
disjoint
2-cycles,
and
3
unique
ways
to
construct
the
two
4
2-cycles
with
them.
Thus
the
number
of
distinct
products
of
two
disjoitn
2-cycles
is:
n
1
n!1
n(n
1)(n
2)(n
3)
==
43
(n
4)!4!
3

8
Exercise
1.3.18
For
each
Sn
we
can
choose
i;k
>
0
such
that
i
+
k
=
n
and
construct
following
disjoint
product
Sn:
(1
2
::.
i)(nn
1
::.
n
(k
1))
By
exercise
15,
this
element
has
order
of
lcm(i,
k).
Thus
for
S5
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
S7
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
de

ne
the
generator
to
be
ha,
bja2
=
b2
=1,
abab
=
bai.
All
other
elements
of
S3
can
be
written
with
these
two
elements:
2
(1)
=
a
=
b2
(1
3)
=
aba
=
bab
(1
2
3)
=
ab
(1
3
2)
=
ba
2
The
fact
that
a=
b2
=
1
means

that
every
element
of
S3
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
S3
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
f0,
1}
and
show
that
exactly
6
of
them
have
nonzero
determinants.
11

Exercise
1.4.2
2
2
2
01
11
10
===
I
10
01
11
1
3
3
1001
11
==
I
0111
10
Exercise
1.4.3
.
.
.
.
.
.
.
.
.
.
.
.
1
1
0
1
1

0
1
1
=
1
1
1
0
6
=
0
1
1
1
=
1
0
1
1
1
1
0
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
de

nition
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
n
If
jF
|
=
q
is

nite,
then
there
are
at
most
q2
possible
n

n
matrices;
GLn(F
)
is
a
subset
of
these
matrices
n
and
at
least
one
such
matrix
has
a
zero
determinant,
so
jGLn(F
)j=
q2
and
thus
GLn(F
)
is

nite.
If
jF
|
is
in

nite,
then
fI
.
GLn(F
)
for
each
f
.
F
(where
I
is
the
identity
matrix)
and
thus
jGLn(F
)jjF
|
which
means
jGLn(F
)|
is
in

nite.
Exercise
1.4.6
see
previous
exercise
Exercise
1.4.7
The
determinant
of
a
2

2
matrix
is
given
by
the
formula
ab
det
=
ad
bc
cd
So
we
see
that
the
deterimant
is
zero
if
ad
=
bc.
Using
basic
combinatorics,
it's
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
3
(2p
1)2
+(p

1)(p
1)2
=
p
+
p
2
p
4
And
since
there
are
ppossible
2

2
matrices
over
Fp,
the
total
number
of
such
matrices
with
nonzero
determ
inants
is
2
p
4
p
3
p
+
p
Exercise
1.4.8
Let
A
be

the
matrix
with
a1;2
=
1
as
the
only
nonzero
entry,
and
let
B
be
the
matrix
with
a2;1
as
the
only
nonzero
entry.
Then
AB
has
a1;1
=
1
as
the
only
nonzero
entry
while
BA
has
a2;2
as
the
only
nonzero
entry.
Exercise
1.4.9
We
want
to
show
that
.
a1
b1
.
.

a2
b2
.
.
a3
b3
.
=
.
a1
b1
.
.
a2
b2
.
.
a3
b3
.
c1
d1
c2
d2
c3
d3
c1
d1
c2
d2
c3
d3
This
can
be
done
tediously
through
algebra.
Exercise
1.4.10a
.
.
.
a1
b1
a2
b2
a1a2
a1b2
+
b1c2

=
0
c1
0
c2
0
c1c2
Exercise
1.4.10b
We
want
to

nd
values
of
a2;b2,
and
c2
such
that
the
product
in
part
(a)
is
the
identity
matrix.
It's
immediately
..1
..1
clear
that
we
need
to
have
a2
=
a1
and
c2
=
c1
.
With
these
substitutions,
we
have
a1b2
+
b1c2
=
a1b2
+
b1c
..1
1
..1
..1
which
equals
1

exactly
when
b2
=
a
(1
b1c
).
So
the
inverse
is
11
..1
..1
..1
aa
(1
b1c
)
11
1
..1
0
c
1
..1
..1
This
is
an
element
of
G
since
a
66
1
=0;c
1
=
0.

Exercise
1.4.10c
G
is
a
group:
we've
shown
closure
under
the
operation
in
part
(a),
closure
of
inverses
in
part
(b),
associativity
in
exercise
9,
and
it's
clear
that
the
identity
matrix
is
an
element
of
G.
We're
told
that
a
6
=
0
and
c
6
=
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
a1
=
c1.
The
proof
changes
very
little.
13

Exercise
1.4.11a
2323.
.
1
ab
1
de
1
d
+
ae
+
af
+
b
XY
=
01
c
01
f
=
01
f
+
c
4545.
.
001001
00
1
To
prove
non-abelianism,
we
see
calculate
YX:
2323.
.
1
de
1
ab
1
a
+

db
+
dc
+
e
YX
=
01
f
01
c
=
01
c
+
f
4545.
.
001001
00
1
So
we
have
XY
.
6An
explicit
example
can
be
given
by
letting
a
=
b
=
f
=0;c
=
d
=
=
YX
whenever
af
=
cd.
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
.
.
1
..a
ac
b
X..1
=
.
01
..c
.
00
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
jF
|
possible
values
implies
that
o(H(F
))
=
jF
j3
.
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
Xn
has
the
form

.
.
n(n..1)
1
na
nb
+
ac
2
Xn
.
5
=
01
nc
00
1
The
case
for
k
=
1
is
trivial.
For
k
=
2
we
have
23.
3
2
1
ab
12a
2b
+
ac
X2
=
.
01
c

.
=
.
01
2c
.
001
001
Now
assume
that
we
have
established
the
form
of
Xk
.
For
Xk+1:
2323.
.
k(k..1)
k(k+1)
1
ab
1
ka
kb
+
ac
1(k
+
1)a
(k
+
1)b
+
ac
2
2
Xk+1
=
XXk
=
.
01
c
.

=
.
01
kc
.
=
.
01
(k
+
1)c
.
001001
00
1
So
that
the
proof
by
induction
is
complete.
From
this,
we
see
that
Xn
=
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
in

nite
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
Q8
can
be
derived
from
h1,
i,
j,
kji2
=
j2
=
k2
=
..1,
ij
=
ki:
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.
It's
trivial
for
n
=
1,
and
true
by
de

nition
of
homomorphism
for
n
=
2.
Assuming
it
holds
for
n
=
k,
we
have
k
'(x
k+1)=
'(xx)=
'(x
k)'(x)=
'(x)k'(x)=
'(x)k+1
From
this,
we
have
..n
1=
'(x
0)=
'(xx
n)=
'(x
..n)'(x
n)=
phi(x
..n)'(x)n
so
that,
by
the
de

nition
of
inverses,
..1
'(x
..n)=('(x)n)=
'(x)..n
Exercise
1.6.2
lemma
From
the
previous
exercise,
we
know
that
'(x0)=
'(x)0
so
that
'(1G)=1H
.
Now,
choose
an
arbitrary
x
.
G
and
suppose
o(x)=
n
is

nite.
7
.
xn
=1G
assumed
.
'(xn)=
'(1G)
.
is
well-de

ned
.
'(x)n
=1H
from
previous
exercise
.
o('(x))jn
.
o('(x))jo(x)
de

nition
of
n
Note
that
this
last
step
also
implies
that
o('(x))
is

nite.
Now
assume
that
o('(x))
=
m
is

nite:
7
.
'(x)m
=1H
assumed
.
'(xm)=
'(1G)
.
is
well-de

ned
.
xm
=1G
.
is
an
isomorphism,
thus
1-to-1
.
o(x)jm
.
o(x)jo('(x))
de

nition
of
m
This
last
step
implies
that
o(x)
is

nite.
Thus
we
have
shown
that
o(x)
is

nite
iff
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
requiri
ng
isomorphism
is
clearly
labeled).
As
a
counter
example,
consider
the
following
homormorphism
:
f
:
Z
.
Z
de

ned
as
f(n)
=
1.
f(3)f(2)
=
1

1=1=
f(6),
but
clearly
o(3)
=
8
while
o(f(3))
=
1.
Exercise
1.6.3
7
.
G
is
abelian
assumed
.
(8a,
b
.
G)ab
=
ba
def.
of
abelianism
.
(8a,
b
.
G)'(ab)=
'(ba)
isomorphisms
are
well-de

ned
and
bijective
.
(8a,
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
we've
proven
that
when
.
is
isomorphic,
then
G
is
abelian
iff
H
is
15

abelian.
If
.
is
only
a
homomorphism,
then
we
have
the
unidirectional
proof:
7
.
G
is
abelian
assumed
.
(8a,
b
.
G)ab
=
ba
def.
of
abelianism
.
(8a,
b
.
G)'(ab)
=
'(ba)
isomorphisms
are
well-de

ned
.
(8a,
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
.
1H
de

ned
as
'(g)=1H
.
'(G)=
H
is
trivially
abelian,
but
G
can
be
any
group
whatsoever.
Exercise
1.6.4
C..f0}
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
Cantor's
Theorem.
Proof
2:
suppose
.
:
Q
.
R
were
an
isomorphism.
Let
x
.
R
be
the
element
such
that
'(2)
=
x;
let
a
.
Q
be
v
the
element
such
that
'(a)=
k.
From
this,
we
have
v
'(a
2)=
'(a)2
=(
k)2
=
k
=

'(2)
v
Which,
since
.
is
a
bijection,
means
that
a2
=
2
and
thus
a
=
2.
But
this
can't
be
true
of
any
a
.
Q.
Exercise
1.6.6
Let
Q
has
an
element
of
order
3
(
1
),
but
there
is
no
such
element
in
Z.
This
contradicts
exercise

1.6.2.
3
Exercise
1.6.7
D8
has
only
one
element
of
order
4
(o(r)
=
4)
while
Q8
has
three
such
elements
(i,
j,
k).
This
contradicts
exercise
1.6.2.
Exercise
1.6.8
Assuming
m;n
>
0,
Sm
contains
m!
elements
while
Sn
contains
n!,
so
there
can
be
no
bijection
between
them

unless
n
=
m.
Exercise
1.6.9
D24
has
an
element
of
order
12
(r)
while
S4
can
have
no
such
element
(exercise
1.3.15)
Exercise
1.6.10a
7
.
a
=
b
.
. .1(a)=
..1(b)
.
is
bijective
.
s
.
..1(a)=
s
.
..1(b)
permutations
are
well-de

ned
.
.
.
s
.
..1(a)=
.
.
s
.
..1(b)
permutations
are
well-de

ned
Exercise
1.6.10b
The
same
logic
used
in
part
(a)
can
show
that
.
.
.
. .1
is
well-de

ned,
and
this
clearly
acts
as
an
inverse
to
'.
16

Exercise
1.6.10c
'(s
.
)=
.
.
(s
.
t
)
.
..1
=
.
.
(s
.
..1
.
.
.
t
)
.
..1
=
'()
.
'()
Exercise
1.6.11
The
function
f
:
A

B
.
B

A
de

ned
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)
de

ned
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
1H
.
'(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.
It's
clear
that
.
can't
be
injective
if
more
than
one
element
maps
to
1
.
H,
so
.
is
injective
only
if
K
=
f1Gg.
Proof
by
contrapositive
that
.
is
injective
if

K
=
f1Gg:
Assume
that
K
=
f1Gg.
7
.
'(a)=
'(b)
.
'(a)'(b)..1
=1H
'(G)
is
a
group
by
previous
exercise
.
'(a)'(b..1)=1H
.
'(ab..1)=1H
.
ab..1
=1G
K
=
f1Gg
.
a
=
b
Exercise
1.6.15
(x,
y)
is
in
the
kernel
of
p
if
(x,
y)=
x
=
1,
so
the
kernel
is

K
=
f(x,
y)
.
R2jx
=1}
=
f1g
R
Exercise
1.6.16
Following
the
logic
above,
we
see
that
the
kernel
of
1
is
K
=
f(a,
b)
.
A

B
|
a
=1A}
=
f1Ag
B
and
the
kernel
of
2
is
K
=
f(a,
b)
.
A


B
|
b
=1B}
=
A
f1B}
Exercise
1.6.17
..1
Let
.
:
G
.
G
be
de

ned
by
'(g)=
g.
This
function
is
clearly
onto,
so
'(G)=
G.
So
we
are
asked
to
prove
that
.
is
a
homomorphism
iff
G
is
abelian.
First
assume
that
.
is
a
homomorphism:
7
.
'(a..1b..1)=
'(a..1)'(b..1)
assumed
.
'((ba)..1)=
'(a..1)'(b..1)
properties
of
inverses
.
ba
=
ab
de

nition
of
.
Thus
'(G)=
G
is
abelian.
Now
assume
that
G
is
abelian:
17

..1
..1b..1
7
.
b..1a=
aassumed
..1b..1
.
(ab)..1
=
aproperty
of
inverses
.
'(ab)=
'(a)'(b)
de

nition
of
.
Thus
.
is
a
homomorphism.
Exercise
1.6.18
2
Let
.
:
G
.
G
be
de

ned
by
'(g)=
g.
7
.
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)
de

nition
of
.
Exercise
1.6.19
As
de

nied
here,
G
is
the
set
of

nite
roots
of
unity
in
C.
From
complex
analysis,
we
know
that
for
each
k
k
.
N
there
are
k
distinct
elements
of
order
k.
Let
k
be

xed
and
de

ne
fk
as
fk(z)=
z.
This
is
clearly
a
homomorphism,
and
is
surjective
since
1=k
.
G
z
.
G
.
z
.
fk(z
1=k)=
z
But
by
exercise
14,
fk
cannot
be
injective:
the
k
roots
of
unity
of
order
k
mean
that
the
kernel
of
fk
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
De

ne
fk
:
Q
.
Q
to
be
fk(q)
=
kq.
This
function
is
clearly
injective
and
a
homomorphism.
To
prove
surjectivity:
q
.
Q
.
q
k
.
Q
.
fk
.
q
k
.
=
q
Exercise
1.6.22
k
De

ne
fk
:
A
.
A
to
be
fk(a)=
a.
Since
A
is
abelian,
we
have
kbk
fk(ab)=(ab)k
=
a
=
fk(a)fk(b)
..1
b..1
If
k
=
..1,
then
the
function
is
injective
(a
=
b
.
a=
.
f(a)=
f(b))
and
surjective
(a
.
A
.
f(a..1)=
a).
Exercise
1.6.23
Let
s
be

an
automorphism
such
that
is
the
identity
map
and
g
iff
g
=
1.
Choose
arbitrary
elements
a,
b
.
G.
s
is
bijective,
so
there
are
x,
y
.
G
such
that
a,
b.
From
this
we
have
7
.
=
xy
=
is
the
identity
map
.

s
is
a
homomorphism
.
abab
=
aabb
de

nition
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
2
We
need
to
show
that
mapping
preserves
the
properties
of
each
generator
and
relation.
We're
told
that
x=
22
2
r=
1
and
y=
s2
=
1.
From
the
exercise
1.2.6,
the
fact
that
x=
y2
=
1
is
sucient
to
conclude
that
..1
xy
=
yx.

Exercise
1.6.25a
Let
~v
=
jv|
cos()+
jv|
sin()
be
an
arbitrary
vector.
The
given
matrix
transforms
~v
as
follows:
.
.
.
.
cos()
sin()
jv|
sin()
jvj[cos()
cos()
sin()
sin()]
=
sin()
cos()
jv|
cos()
jvj[sin()
cos()
+
cos()
sin()]
which,
via
the

angle
addition
formulas
from
the
trigonmetric
identities,
is
equivalent
to
jv|
cos(f
+
)
jv|
sin(f
+
)
which
is,
of
course,
the
original
vector
with
its
endpoint
rotated
by
an
additional
f
radians
counterclockwise
about
the
origin.
Our
vector
~v
was
arbitrary,
so
every
vector
endpoint
(and
thus
every
point
in
R2)
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
..1
rs
=
srhas
an

associated
relationship
'(r)'(s)=
'(s)'(r)..1
by
some
tedious
algebraic
veri

cation.
The
text
(bottom
of
p38)
assures
us
that
this
is
sucient
to
guarantee
an
isomorphism
between
G
and
D2n.
Exercise
1.6.26
Further
de

ne
.
as
follows:
.
v
.
0
....1
'(1)
=
I,
'(k)=
v
....10
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
veri

ed
(e.g.,
'(i)2
=
..'(1);'(i)'(j)=
k)
but
they
are
trivial
(albeit
tedious).
The
identity
is
the
only
element
of
Q8
that
maps
to
I
.
GL2(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
de

ne
a
group
action
of
G
=
F

on
A
=
F
by
g

a
=
ga.
The
element
1
.
F

satis

es
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
(it's
associative
and
closed
under
its
operation).
Property
(ii)
is
then
justi

ed
as
follows:
(g1g2)

a
=
g1g2a
de

nition
of
the
group
action;
F
is
closed
under
multiplication
=
g1(g2a)
multiplication
in
F
is
associative
=
g1(g2

a)
de

nition
of
group
action
=
g1

(g2

a)
de

nition
of
group
action
19

Exercise
1.7.2
The
element
0
.
Z
satis

es
property
(i)
of
group
actions
(0

a
=
a
for
all
a
.
Z).
To
justify
property
(ii),
let
g1;g2;a
be
arbitrary
elements
of
Z:
(g1
+
g2)

a
=(g1
+
g2)+
a
de

nition
of
the
group
action;
Z
is
closed
under
addition
=
g1
+(g2
+
a)
addition
in
Z
is
associative
=
g1
+(g2

a)
de

nition
of
group
action
=
g1

(g2

a)
de

nition
of
group
action
Exercise
1.7.3
The
element
0
.
R
satis

es
property
(i)
of
group
actions.
To
justify
property
(ii),
let
r1;r2
be
arbitrary
elements
of
R
and
let
(x,
y)
be
an
arbitrary
point
in
R
R:
(r1r2)

(x,
y)
=(r1
+
r2)

(x,
y)
operation
on
R
is
addition
=(x
+(r1
+
r2)y,
y)
de

nition
of
group
action
=(x
+(r1
+
r2)y,
y)
operation
on
R
is
addition
=
((x
+
r1y)+
r2y,
y
associativity,
distributive
property
of

eld
R
R
=
r2

(x
+
r1y,
y)
de

nition
of
group
action
=
r2

(r1

(x,
y))
de

nition
of
group
action
=
g1
+(g2
+
a)
addition
in
Z
is
associative
=
g1
+(g2

a)
de

nition
of
group
action
=
g1

(g2

a)
de

nition
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.
It's
clear

that
1
.
K,
so
an
identity
exists.
Now
assume
that
g1;g2
.
K:
7
.
1;g1;g2
.
K
assumed
.
(8a
.
A)g1

a
=
a
.
g2

a
=
a
.
1

a
=
a
de

nition
of
K
.
(8a
.
A)g1

(g2

a)=
a
algebraic
substitution
.
(8a
.
A)(g1g2)

a
=
a
property
(ii)
of
group
actions
.
g1g2
.
K
de

nition
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.
It's
clear
that
1
.
S,
so
an
identity
exists.
Now
assume

that
g1;g2
.
S:
20

7
.
1;g1;g2
.
S
assumed
.
g1

a
=
a
.
g2

a
=
a
.
1

a
=
a
de

nition
of
S
.
g1

(g2

a)=
a
algebraic
substitution
.
(g1g2)

a
=
a
property
(ii)
of
group
actions
.
g1g2
.
S
de

nition
of
S
Exercise
1.7.5
Each
step
in
the
following
proof
is
bidrectional
(i):
7
.
g
.
K
assumed
.
(8a
.
A)ga
=
a
de

nition
of
K
.
(8a
.
A)g(a)=
a
de

nition
of
.
is
the
identity
permutation
on
G
de

nition
of
the
identity
function
.
is
the
identity
element
of
SA
de

nition
of
the
group
SA
.
g
is
in
the
kernel
of
.
:
G
.
SA
de

nition
of
kernel,
.
Exercise
1.7.6
Proof
by
contradiction.
Assume
that
G
is
not
faithful:
then
there
are
distinct
nonidentity
elements
g1;g2
such
that
g1

a
=
g2

a
for
all
a
.
A.
From
this,
we
obtain
..1
..1
..1
..1
(gg2)

a
=
g

(g2

a)=
g

(g1

a)=(gg1)

a
=1

a
=
a
111
1
..1
and
thus
gg2
.
K.
And
this
element
cannot
be
the
identity
since
g1
6g2.
Thus
the
kernel
contains
=a
1
nonidentity
element.
By
contrapositive,
if
K
=
f1}
then
G
is
faithful.
Exercise
1.7.7

The
kernel
of
the
given
action
is
f1g;
by
the
previous
exercise,
this
suces
to
prove
that
the
action
is
faithful.
Exercise
1.7.8a
Let
G
=
SA
and
let
B
=
P(A).
The
identity
permutation
satis

es
property
(i)
of
group
actions
(1(b)=
b
for
all
b
.
B).
To
show
that
property
(ii)
is
satis

ed,
let
be
arbitrary
elements
of
G
and
let
b
be
an
arbitrary
element
of
B:
(g
.

b
=(g
.
de

nition
of

=
de

nition
of
.
=

(h

b)
de

nition
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)f1,
4}
=
f2,
4g.
The
element
(1
2
3)
replaces
each
1
with
2,
each
2
with
3,
and
each
3
with
1.
For
example,
(1
2
3)f2,
3,

4}
=
f3,
1,
4g.
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<
jA|
or
k
jZj.
case
1)
Suppose
jA|
is

nite
and
k<
jAj.
We
show
that
the
action
of
Sn
on
k-element
subsets
of
A
is
faithful.
Let
.
Sn
be
any
two
distinct
permutations
of
A.
Since
these
permutations
are
distinct,
there
is
some
a
.
A
such
that
6
=
From
the
invertibility
of
permutations
this
gives
us
the
inequality

.
6(y(a))
=
a
yy
Since
1
=
k<n,
we
can
choose
a
k-element
subset
B
.
A
such
that
a
.
B
but
6
.
B.
We
y
can
now
demonstrate
that
the
action
is
faithful,
since
contains
but
does
not
contain
..1(x(a)))

=
Thus
and
do
not
perform
the
same
action
on
B.
But
these
were
arbitrary
elements
of
SA,
thus
no
two
elements
of
SA
perform
the
same
action.
By
de

nition,
this
means
that
SA
is
faithful
on
the
set
of
k-element
subsets
of
A.
case
2)
Suppose
jA|
is

nite
and
k
=
jAj.
There
is
only
one
distinct
subset
of
size
k:
A
itself.
And
every
permutation
of
A
is
still
the
same
set
(just
rearranged).
Thus
Sn
is
the
trivial
action
on
k-element
subsets
of
jAj.
case
3)
Suppose
jA|
is
in

nite.
For
all

nite
k,
we
can
follow
the
logic
of
case
(1)
and
conclude
that
the
action
is
faithful.
If
k
is
also
in

nite,
then
we
can
still
choose
an
arbitrary
a
.
A
and
let
B
=
A
..fag,
and
then
follow
the
logic
of
case
(1)
and
conclude
that
the
action
is
still
faithful.
Exercise
1.7.10b
Choose
two
arbitrary
permutations
.
Sn.
For
these
to
be
distinct,
there
must
be
some
a
.
A
such
that

6(a).
Having
chosen
such
an
a,
let
B
be
the
k-tuple
consisting
of
the
element
a
repeated
k
times.
It's
=
clear
that
6
=
so
the
action
of
SA
is
faithful.
But
the
value
of
k
was
never
speci

ed,
so
this
proof
holds
for
all
values
of
k
(

nite
and
in

nite).
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
fai}
de

ned
as
nn
fai}
=
i,
+
i
:0
=
i<
22
We
de

ne
the
action
of
D2n
on
the
elements
of
fai}
so
that
the
elements
of
D2n
permute
the
set
as
follows:
k
rai
=
a(i..k)
mod
n
k
sai
=
a(i..kn=2)
mod
n
0
To
prove
this
is
an
action,
we
note
that
rai
=
ai
for
all
ai,
so
property

(i)
of
group
actions
is
satis

ed.
To
ax
verify
property
(ii),
let
rsb
and
rsy
be
two
arbitrary
elements
of
D2n:
abx
(rsrsy)

ai
=
aj
,
with
j
=
((((i
ny=2)
x)
nb=2)
a)
mod
n
de

nition
of
,
associativity
of
modular
arithm
etic
ab
=
rs
ak,
with
k
=
((i
ny=2)
x)
mod
n
de

nition
of
,
associativity
of
modular
arithm
etic
ab
xy
=
rs
(rs
ai)
de

nition
of

The
kernel
of
this
action
is
fr0}
=
f1g.
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
1G
.
K.
But
let
g
.
G
be
an
arbitrary
non-identity
element
of
G
and
choose
1G
.
A
=
G.
Then
g1G
6
=1G
and
so
g
6
.

K.
Exercise
1.7.14
We
prove
by
contradiction
that
property
(ii)
of
group
actions
is
not
satis

ed.
Assume
A
=
G
is
non-abelian
and
de

ne
the
action
g

a
=
ag.
Let
g1;g2
be
arbitrary
elements
of
G.
Hypothesis
to
be
contradicted:
suppose
that
property
(ii)
of
group
actions
were
satis

ed.
(g1g2)

a
=
a(g1g2)
de

nition
of
this
group
action
=(ag1)g2)
associativity
of
operation
on
G
=
g2

(ag1))
de

nition
of
this
group
action
=
g2

(g1

a))
de

nition
of
this
group
action
=(g2g1)

a))
property
(ii)
of
group
actions
This
demonstrates
that
g1
commutes
with
g2.
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
satis

ed.
Exercise
1.7.15
Let
A
=
G
and
de

ne
the
action
g

a
=
ag..1
.
The
identity
element
1G
satis

es
the
condition
1

a
=
a
for
all
a
.
A.
To
verify
property
(ii):
(g1g2)

a
=
a(g1g2)..1
de

nition
of
this
group
action
..1
..1
=
a(gg
)
property
of
inverses
21
..1
..1
=(ag
)g
associativity
of
operation
on
G
21
=
g1

(ag
..1)
de

nition
of
this
group
action
2
=
g1

(g2

a))
de

nition
of
this
group
action
Exercise
1.7.16
Let
A
=
G
and
de

ne
the
action
g

a
=
gag..1
.
The
identity
element
1G
satis

es
the
condition
1

a
=
a
for
all
a
.
A.
To
verify
property
(ii):
(g1g2)

a
=(g1g2)a(g1g2)..1
de

nition
of
this
group
action
..1
..1
=(g1g2)a(gg
)
property
of
inverses
21
..1
..1
=
g1(g2(ag
)g
associativity
of
operation
on
G
21
=
g1

(g2ag
..1)
de

nition
of
this
group
action
2
=
g1

(g2

a))
de

nition
of
this
group
action
Exercise
1.7.17
..1
Let
G
be
a
group
and

x
a
value
for
g
.
G.
De

ne
a
mapping
f
:
G
.
G
as
f(x)
7
.
gxg.
We
want
to
prove
that
the
given
function
is
an
isomorphism.
We
do
so
by
showing
that
it's
a
bijective
homomorphism.
23

f
is
a
homomorphism
..1
..1
f(xy)=
g(xy)g
=
g(xg
gy)g
..1
=(gxg
..1)(gyg
..1)=
f(x)f(y)
f
is
surjective
Choose
h
.
G.
Then
g..1hg
.
G
and
f(g..1hg)=
h.
Thus
f
is
surjective
f
is
injective:
..1
f(x)
=
1
iff
gxg..1
=
1
iff
x

=
gg
=
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
re
exive:
1
.
H
and
a
=1a
so
a
~
a.
transitive:
If
a
~
b
then
a
=
hb,
which
implies

b
=
h..1a,
which
implies
b
~
a.
symmetric:
If
a
~
b
and
b
~
c,
then
a
=
bh1
and
b
=
ch2,
which
implies
a
=
ch2h1,
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
Ox
be
the
orbit
of
x
under
the
action
of
H.
De

ne
the
map
f
:
H
!Ox
as
h
7
.
hx.
We
prove
that
f
is
a
bijection.
f
is
surjective:
Ox
is
de

ned
as
the
set
fhx
:
h
.
G}
which
is
exactly
the
image
of
f(H).
f
is
injective:
h(a)=
h(b)
iff
ha
=
hb
iff
a
=
b
(by
left-cancellation).
We've
shown
that
jH|
=
jOxj.
But
H
was
an
arbitrary
subgroup
and
x
was
an
arbitrary
element
of
G:
thus
jOx|
=

jOy|
for
all
x,
y
.
G.
Applying
the
preceeding
exercise,
we
conclude
that
G
can
be
partitioned
into
disjoint
sets
of
size
jOx|
=
jHj;
this
can
only
occur
if
jG|
is
an
integer
multiple
of
jHj.
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
de

ne
the
function
.
:
G
.
S4
so
that
'(g)
is
the
permutation
corresponding
to
the
rigid
motion
g.
It's
clear
that
.
is
a
homomorphism:
'(g1g2)=
'(g1)'(g2),
since
each
side
of
the
equation
represents
the
rigid
motion
g1
followed
by
the
rigid
motion
g2.
.
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
a1;a2;a3;a4.
On
the
opposite
face,
label
the
diagonally
opposite
points
to
be
b1;b2;b3;b4.
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
ai;bi
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
ai
with
its
corresponding
bi)
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)
(a1;b2;b3;a4),
then
none
of
the
15
other
ordered
4-tuples
we
could
form
from
replacing
ai
with
bi
(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
de

ne
the
function
.
:
G
.
S4
so
that
'(g)
is
the
permutation
corresponding
to
the
rigid
motion
g.
It's
clear
that
.
is
a
homomorphism:
'(g1g2)=
'(g1)'(g2),
since
each
side
of
the
equation
represents
the
rigid
motion
g1
followed
by
the
rigid
motion
g2.
.
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
S4
itself
(the
errata
for
this
textbook
eliminates
the
'subgroup
quali

cation,
although
S4
is
technically
a
subgroup
of
S4).
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:

r1
=(ABDE);r2
=(ACDF
);r3
=(CBF
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
r1
=(fA,
DgfB,
Eg);r2
=(fA,
DgfC,
F
g);r3
=(fC,
F
gfB,
Eg)
22
2
so
that
r1;r2,
and

rare
all
equal
to
the
identity
motion.
The
kernel
of
the
action
becomes
3
22
K
=
fi,
r1
2
;r
2;r
3}
(Any
product
of
elements
of
K
can
be
simpli

ed
to
one
of
these
elements).
Note
that
4j24,
in
concordance
with
exercise
19.
25

Potrebbero piacerti anche