You are on page 1of 9

Homework 1 Solutions

Math 150A
Prepared by Jeff Ferreira

1.1.13 In each case, find all real 2 × 2 matrices that commute with the given matrix.
 
1 0
(a)
0 0
 
a b
Sol: To do this problem, write out the commutation relation between an arbitrary matrix and the
c d
given matrix:
         
a b 1 0 1 0 a b a 0 a b
= ⇔ = .
c d 0 0 0 0 c d c 0 0 0
From this we get the equations:
a=a 0=b
c=0 0=0

 
a 0
Which tells us that all matrices which commute with the given matrix are of the form for a, d ∈ R.
0 d
 
1 3
(d)
0 1
Sol: Do the same as above:
         
a b 1 3 1 3 a b a b + 3a a + 3c b + 3d
= ⇔ = .
c d 0 1 0 1 c d c d + 3c c d
Which gives a + 3c = a, b + 3d = b + 3a, c = c, and d = d + 3c as our
 equations.
 From this we conclude that
a b
all matrices which commute with the given matrix are of the form with a, b ∈ R.
0 a

1.1.16 Prove that if A is nilpotent, then I + A is invertible.


Proof 1 (Using the hint on the discussion board): To show I+A is invertible, consider the following (remember
A is nilpotent):
(I +A)(I −A+A2 −A3 +A4 −· · ·±Ak ) = I −A+A2 −A3 +· · ·±Ak +A−A2 +A3 −· · ·∓Ak ±Ak+1 = I ±Ak+1 = I
where I±Ak+1 = I because A is nilpotent. You can do the same for (I−A+A2 −A3 +A4 −· · ·±Ak )(I+A) = I,
so this shows the inverse of I + A is (I − A + A2 − A3 + A4 − · · · ± Ak ), and so I + A is invertible. Note: we
didn’t include terms that had A to a higher power than k because if Ak = 0, then Am = 0 whenever m ≥ k.
Proof 2: Let v be a vector in the kernel of I + A. That means (I + A)(v) = 0 which gives Av = −v. Since
A is nilpotent, Ak = 0, we get Ak v = (−1)k v which tells us v = 0. We conclude that ker(I + A)=0 and thus
I + A is invertible.
   
1 2 5 6 2 1 5 1
3 1 7 7 1 3 7 0
1.3.2 Prove det0 0 2 3 = -det 0 0 2 1.
  

4 2 1 5 2 4 1 4
Sol: Notice that to get from the first matrix to the second matrix we swap columns 1 and 2, and add a
multiple of column 3 to column 4. From Proposition (3.18) and property (3.9’) of Artin, we can deduce the
equality needed.

1.3.4b Compute the determinant of the following n × n matrix by induction on n.


1
 
2 −1 0 0 ··· 0
−1 2 −1 0 ··· 0 
 
 0 −1 2 −1 ··· 0 
An = 
· · ·

 ··· ··· ··· ··· ···

 0 ··· 0 −1 2 −1
0 ··· 0 0 −1 2
Claim: det(An )=n + 1.
Proof: To prove by induction, we first need to establish a base case. Clearly,  det(A  1 )=2. We will also need
2 −1
the base case when n = 2 for our induction, which gives det(A2 )=det = 3. Now assume the
−1 2
statement is true for k ≤ n − 1. We wish to compute det(An ) expanding by minors along the first column.
 
2 −1 0 0 ··· 0    
−1 2 −1 0 · · · 2 −1 0 · · · 0 −1 −1 0 ··· 0
0 
−1 2 −1 · · ·

 0 −1 2 −1 · · ·
 0   0 2 −1 ··· 0 
0    
 = 2det · · · · · · · · · · · · · · · −(−1)det  0 −1 2

det(An ) = det 
· · · · · · · · · · · · · · · · · ·     −1 ···

  · · · 0 −1 2 −1 · · · · · · · · · ··· ···
 0 ··· 0 −1 2 −1
··· 0 0 −1 2 0 ··· 0 −1 2
0 ··· 0 0 −1 2
The first matrix on the right hand side is An−1 and computing the determinant in the second term we get,
for a similar reason, −1detAn−2 . In the end we get det(An )=2det(An−1 ) - (-1)(-1)det(An−2 )=2(n) − (n − 1),
this last equality coming from the inductive hypothesis. So det(An )=n + 1.

1.4.1 Consider the permutation p defined by (1 3 4 2) in cycle notation.


(a) The permutation matrix of p is  
0 1 0 0
0 0 0 1
P = 1 0 0 0

0 0 1 0
(b) One solution is p = (12)(14)(13). For the corresponding matrix product, let [ij] be the matrix obtained
by swapping columns i and j of the identity matrix. Then P = [12][14][13].
(c) Denote by sgn(p) the sign of p. We can use part (b) and the definition of sign to say sgn(p)=det(P ) =
det([12][14][13])=det[12]det[14]det[13] = (−1)3 = −1.

1.4.4 Let p be a permutation. Prove that sign(p)=sign(p−1 ).


Proof 1: Recall that sign(p)=det(P ) where P in the permutation matrix associated to p. Also, recall that
1
det(P )=±1. That means det(P )det(P )=1, so det(P )= det(P ) . From properties of determinants we conclude
−1 −1
det(P )=det(P ) so we get sign(p)=sign(p ).
Proof 2: From problem 1.4.5 we know that for a permutation matrix P , P −1 =P t . So we write
sign(p)=det(P )=det(P t )=det(P −1 )=sign(p−1 )

1.4.5 Prove that the transpose of a permutation matrix P is its inverse.


Pn
Proof: We want to show P t P = I. Write P as j=1 ep(j)j , where eij is the matrix with a 1 in the (i, j)
position and zeros elsewhere (see Artin, pg. 26). Now, since etij = eji ,
n
X n
X n X
X n
P tP = ejp(j) ep(k)k = ejp(j) ep(k)k .
j=1 k=1 j=1 k=1

Since eij ei0 j 0 is eij 0 when j = i0 and is zero otherwise, and p(j) = p(k) if and only if j = k (p is bijective),
n
X
t
P P = ejj = I.
j=1

So P t = P −1 .
2
Homework 2 Solutions
Math 150A
Prepared by Jeff Ferreira

2.1.1
(a) Verify (1.17) and (1.18) by explicit computation.
Proof. Given    
0 1 0 0 1 0
x = 0 0 1 and y = 1 0 0
1 0 0 0 0 1
we need to show that {1, x, x2 , y, xy, x2 y} are the six permutations of S3 and that x3 = 1, y 2 = 1,
and yx = x2 y. First, we can check we do get the six permutations of S3 . To do this, we just compute
using matrix multiplication
x ↔ (132)
y ↔ (12)
 2
0 1 0
y 2 = 1 0 0 =I ↔ id
0 0 1
 2  
0 1 0 0 0 1
x2 = 0 0 1 = 1 0 0 ↔ (123)
1 0 0 0 1 0
    
0 1 0 0 1 0 1 0 0
xy = 0 0 1 1 0 0 = 0 0 1 ↔ (23)
1 0 0 0 0 1 0 1 0
    
0 0 1 0 1 0 0 0 1
x2 y = 1 0 0 1 0 0 = 0 1 0 ↔ (13)
0 1 0 0 0 1 1 0 0
The only thing we have left to check is x3 = 1 and yx = x2 y, which can be quickly checked:
  
0 0 1 0 1 0
x3 = 1 0 0 0 0 1 = I
0 1 0 1 0 0
    
0 1 0 0 1 0 0 0 1
yx = 1 0 0 0 0 1 = 0 1 0 = x2 y
0 0 1 1 0 0 1 0 0

(b) Make a multiplication table for S3 .
The way to read it is this: the entry in the row labelled by a and the column labelled by b is the product ab.
· 1 x x2 y xy x2 y
1 1 x x2 y xy x2 y
x x x2 1 xy x2 y y
x2 x2 1 x x2 y y xy
2
y y x y xy 1 x2 x
xy xy y x2 y x 1 x2
x2 y x2 y xy y x2 x 1

2.1.4 Solve for y, given that xyz −1 w = 1 in a group.
1
Proof. Working with the given equation, we can write the following:
xyz −1 w = 1
−1
xyz = w−1
−1
yz = x−1 w−1
y = x−1 w−1 z
where the steps were as follows: from (1) to (2), multiply on the right by w−1 , from (2) to (3) multiply on
the left by x−1 , and from (3) to (4) multiply on the right by z. 

2.1.5 Assume that the equation xyz = 1 holds in a group G. Does it follow that yzx = 1? That yxz = 1?

Proof.
xyz = 1
yz = x−1
z = y −1 x−1
zx = y −1
yzx = 1
So the answer to the first question is YES. To answer the second question, assume both xyz = 1 and yxz = 1
are true in our group G. That means xyz = yxz or that xy = yx, which is not necessarily true in an arbitrary
group. So the answer to the second question is NO. Another good way to see the answer to the second question
is no is to exhibit a group where the property fails. For example, let x = (12), y = (123), z = (23) in S3 . 

2.1.9 Show that if ab = a in a group, then b = 1. And if ab = 1 then b = a−1 .

Proof. Recall that every group has an identity element, call it 1. So in the first question, we assume ab = a,
and we also know a ∗ 1 = a, meaning ab = a ∗ 1 or that b = 1 by multiplying both sides on the left by a−1 .
Next, recall that in a group, every element has an inverse. So in the second question we assume ab = 1, and
we also know aa−1 = 1, meaning ab = aa−1 , or that b = a−1 , again by multiplying on the left by a−1 . 

2.1.10 Let a, b be elements of a group G. Show that the equation ax = b has a unique solution in G.

Proof. We are asked to prove two things. First, that the equation has a solution. Second, that the solution
is unique. To accomplish the first question we can solve for x and get x = a−1 b as a solution to the equation.
To answer the question of uniqueness, assume we have a second solution x0 to the equation. That means
ax0 = b. But we can then write ax0 = ax which gives x0 = x, and that means our solution x = a−1 b is the
only one. 
 
1 1
2.2.1 Determine the elements of the cyclic group generated by the matrix explicitly.
−1 0

Proof. Let A denote the matrix. Start by calculating a few powers:


   
1 1 2 0 1
A= ,A = , A3 = −I.
−1 0 −1 −1

Since A3 = −I, we can calculate that A4 = −A and A5 = −A2 and A6 = (A3 )2 = I. Since A6 is the first
power of A which equals the identity, we can say that the cyclic group has order 6 and the elements are
I, A, A2 , A3 = −I, A4 = −A, A5 = −A2


2.2.2 Let a, b be elements of a group G. Assume that a has order 5 and that a3 b = ba3 . Prove that ab = ba.
2
Proof. Take a and b as above. a having order 5 means a5 = 1. Using this fact and that a3 b = ba3 , we can
write

(a3 )(a3 b) = (a3 )(ba3 )


ab = (a3 b)(a3 )
ab = (ba3 )(a3 )
ab = ba

2.2.3 Which of the following are subgroups?

Proof. For each, we need to check the three conditions to be a subgroup.


(a) GLn (R) ⊂ GLn (C).
Subgroup. Clearly it’s a subset containing the identity, and clearly it is closed under multiplication (multi-
plying real numbers gets you real numbers). It is also closed under taking inverses because when we go from
[A|I] to [I|A−1 ] via row reduction, all the numbers involved in our row operations are real as long as A is a
real matrix.
(b) {1, −1} ⊂ R× .
Subgroup. This one is easy to check. Just remember R× is the group of nonzero real numbers under
multiplication.
(c) The set of positive integers in Z+ .
Not a subgroup. The element 1 ∈ Z+ has no additive inverse, as such a number would be negative.
(d) The set of positive reals in R× .
Subgroup. Contains 1, closed
 under multiplication and 1/x is positive if x is.
a 0
(e) The set of matrices , for a 6= 0, in GL2 (R).
0 0
Not a subgroup. Not even a subset (check the determinant). 

2.2.6 (a) Find generators and relations analogous to (2.13) for the Klein four group.
   
−1 0 1 0
Proof. Take as generators the matrices A = ,B= . Now A, B satisfy three rules:
0 1 0 −1
(1) A2 = 1,
(2) B 2 = 1,
(3) AB = BA.
Any string of As and Bs or their inverses can now be reduced to one of 1, A, B or AB using only the
three rules I wrote down. To see why notice that As and Bs can move past each other freely according to
AB = BA. Also note that A = A−1 and ditto for B so we can eradicate any inverses from the string. So in
the end can be turned into An B m for some integers n and m and now applying the first two rules, we can
force n and m to be either 0 or 1.
So the group determined by these relations has at most four elements, and the four group has two generators
satisfying those relations, so they must be the same group. 

(b) Find all subgroups of the Klein four group.

Proof. Let K denote the Klein four group. There are always two obvious subgroups namely K itself and h1i,
the trivial group generated by the identity. Notice now that any subgroup not in the ones we listed must
consist of either 2 or 3 elements, and so must be cyclic. So we need only look at powers of single elements
of K to find them. It’s easy to see that there are only three of these, generated by A, B and AB, and each
of them has order two. We can diagram this situation as follows:
3
Order Subgroup

4 K KK
ss sss KKK
KKK
sss KKK
sss
2 hAi J hABi hBi
JJ tt
JJ tt
JJ tt
JJ tt
J tt
1 h1i

2.2.7 Let a and b be integers.
(a) Prove that the subset aZ + bZ is a subgroup of Z+ .
Proof. First note that aZ + bZ = {ar + bs|r, s ∈ Z}. We need to check the three conditions to be a subgroup.
Clearly, 0 ∈ aZ + bZ by taking r = s = 0. aZ + bZ is closed, for if ar + bs and ar0 + bs0 are two elements in
aZ + bZ, then (ar + bs) + (ar0 + bs0 ) = a(r + r0 )b(s + s0 ) is in aZ + bZ. Finally, a(−r) + b(−s) is the inverse
of ar + bs. 

(b) Prove that a and b + 7a generate the subgroup aZ + bZ.


Proof. We need to show that any element ar+bs in aZ+bZ can be written as the sum of the given generators,
and that the generators a and b + 7a are in aZ + bZ. This first task can be done by taking an element ar + bs
and writing it as a(r − 7s) + (b + 7a)s, and the second task by noting the generators are already in the form
ar + bs. 
2.2.12 Describe all groups G which contain no proper subgroups.
Proof. Note that every group G contains the proper subgroup {1}, so lets describe all groups G which contain
no proper nontrivial subgroups. Suppose we are given a group G containing no proper nontrivial subgroups.
Take any element a 6= 1 in G. Then the cyclic subgroup generated by a, hai, is a nontrivial subgroup, and
so must be all of G. Hence, G is cyclic with generator a. If |a| = ∞, ha2 i is a proper subgroup, and so G
must have finite order. So say |a| = |G| = n. Then given any d that divides n, han/d i is a proper subgroup
of order d. So we must have n prime, and hence groups containing no proper nontrivial subgroups must be
cyclic groups of prime order p. Conversely, given a cyclic group of prime order p, if k does not divide p,
|ak | = p and we have hak i = G, and since any nontrivial subgroup H of G contains some ak , H = G. 
2.2.13 Prove that every subgroup of a cyclic group is cyclic.
Proof. Let G be a cyclic group generated by a, and let H be a subgroup of G. If H = {1}, the H is cyclic
with generator 1 (where 1 here means identity). If H 6= {1}, then every element of H is of the form am for
some integer m. Let k be the smallest positive integer such that ak ∈ H. I claim H is generated by ak .
To prove the claim, we need to show every element of H can be written as (ak )l for some integer l. So let
b ∈ H. We can write b = am for some integer m, and since we picked k smallest, we can write m = kq + r
where q, r are integers, and 0 ≤ r < k. So we have
b = am = akq+r = akq ar .
Multiplying by the inverse of akq on the left we get
a−kq am = ar .
Now since a−kq and am are both in H, so is their product (H is a subgroup), so that means ar ∈ H. But
0 ≤ r < k, so we must have r = 0. Thus we have shown any element b ∈ H is of the form akq = (ak )q , and
so H is cyclic. 

4
Homework 3 Solutions
Math 150A
Prepared by Jeff Ferreira
2.2.14 Let G be a cyclic group of order n, and let r be an integer dividing n. Prove that G contains exactly
one subgroup of order r.
n
First we need the following Claim: ord(as ) = gcd(n,s) .

Proof. We know that (as )m = 1 if and only if n divides ms. We want the smallest m such that n divides
ms. Let d = gcd(n, s). By Artin, there exists integers u, v such that d = un + vs, and dividing by d we see
1 = u( nd ) + v( ds ). This last equation tells us that nd and ds are relatively prime. We want the smallest m such
that n divides ms, and this is equivalent to wanting the smallest m such that
ms m(s/d)
=
n n/d
is an integer. But since n/d and s/d are relatively prime, we get that n/d divides m, and so the smallest m
n
is equal to n/d. Therefore ord(as ) = gcd(n,s) . 

Proof of 2.2.14. Let H = han/r i. By the lemma above, ord(an/r )=r, so H is a subgroup of G of order
r. Suppose K is another subgroup of G of order r. Since G is cyclic, so must be K, so let K = haj i. So
n
r = gcd(j,n) , so in fact gcd(j, n)=n/r. Again by Artin, there exists integers x, y such that n/r = xj + yn,
and from this it follows an/r = (aj )x , which shows an/r ∈ K. But we already showed ord(an/r )=r, so
K = han/r i = H. 

Remark: It is a common mistake to say that since ajr = 1 and an = 1, then jr = n. This is note
true. Be careful!

2.2.18b According to Chapter 1 (2.18), the elementary matrices generate GLn (R). The special
linear group SLn (R) is the set of real n × n matrices whose determinant is 1. Show that SLn (R)
is a subgroup of GLn (R).
Proof. Clearly, the identity matrix In has determinant 1, so In ∈ SLn (R). Let A, B ∈ SLn (R),
then we compute det(AB) = det(A)det(B)=1, so SLn (R) is closed under matrix multiplication.
Let A ∈ SLn (R), we compute that det(A−1 )= det(A)
1
= 1, so SLn (R) is closed under taking inverses.
Therefore SLn (R) is a subgroup of GLn (R). 
2.2.19 Determine the number of elements or order 2 in the symmetric group S4 .
Proof. There are 9 elements of order 2 in S4 , and they are (written in cycle notation): (12), (13),
(14), (23), (24), (34), (12)(34), (13)(24), (14)(23). 
2.2.20a Let a, b be elements of an abelian group of orders m, n respectively. What can you say
about the order of their product ab?
Proof. What follows below shows (ab)lcm(m,n) = 1. Let s = lcm(m, n).
(ab)s = (ab)(ab)(ab) · · · (ab) s factors
= aaaa · · · abbbb · · · b s factors of each. Here we used the group is abelian.
= 1.
But this doesn’t tell us that the order of ab is equal to lcm(m, n). For example, let b = a−1 , then
the order of ab is 1. The most we can say is that the order of ab is a divisor of lcm(m, n). 
2.4.3 Prove that the kernel and image of a homomorphism are subgroups.
Proof. Let φ : G −→ G0 be a homomorphism. From Artin,
1
ker(φ)={a ∈ G|φ(a) = 1G0 } and im(φ)={x ∈ G0 |x = φ(b) for some b ∈ G}.
It is important to notice that we must show ker(φ) is a subgroup of G while im(φ) is a subgroup
of G0 .
(1) For ker(φ): Since we showed in class that for any homomorphism φ, φ(1G ) = 1G0 , 1G ∈
ker(φ). Now let a, b ∈ ker(φ) and consider φ(ab) = φ(a)φ(b)=1G0 1G0 = 1G0 , so ker(φ) is
closed. Now suppose a ∈ ker(φ) and consider φ(a−1 ) = (φ(a))−1 = (1G0 )−1 = 1G0 , so ker(φ)
is closed under taking inverses. So ker(φ) is a subgroup of G.
(2) For im(φ): 1G0 ∈ im(φ) since φ(1G ) = φ(1G0 ). Let x, y ∈ im(φ), then x = φ(a) and y = φ(b)
for some a, b ∈ G. Then xy = φ(a)φ(b) = φ(ab) so xy ∈ im(φ). Again, let x = φ(a) be in
im(φ). Then x−1 = (φ(a))−1 = φ(a−1 ) and we conclude x−1 is in im(φ). Thus, im(φ) is a
subgroup of G0 .

2.4.5 Let G be an abelian group. Prove that the nth power map φ : G −→ G defined by φ(x) = xn
is a homomorphism from G to itself.
Proof. We check the condition for φ to be a homomorphism: Let x, y ∈ G, then
φ(xy) = (xy)n
= (xy)(xy) · · · (xy) with n factors
= (xx · · · x)(yy · · · y) with n factors of both x and y. Here we used G is abelian
= (xn )(y n )
= φ(x)φ(y)
So φ is a homomorphism. 
2.4.8 (a) Find all subgroups of S3 , and determine which are normal.
Proof. S3 has a normal subgroup of order 3, which is {(id, (123), (132)}. There are also three
subgroups of order 2 which are not normal: {id, (12)}, {id, (13)}, {id, (23)}. Also there are the
trivial subgroups {id} and S3 , both of which are normal. 
(b) Find all subgroups of the quaternion group H, and determine which are normal.
Proof. The quaternion group has one subgroup of order 2: {±1}. And three subgroups of order 4:
{±1, ±i}, {±1, ±j}, {±1, ±k}. There are also the trivial subgroups {1} and H. And all subgroups
of the quaternion group are normal. 
2.4.17 Prove that the center of a group G is a normal subgroup of G.
Proof. Recall the center of a group G is defined by Z(G) = {x ∈ G|gx = xg, ∀g ∈ G}. First we check
that Z(G) is indeed a subgroup: Clearly 1 ∈ Z(G) since g1 = 1g for all g ∈ G. Let x, y ∈ Z(G).
We can then say g(xy) = gxy = xgy = xyg = (xy)g for all g ∈ G. So xy ∈ Z(G). Let x ∈ Z(G).
Then g(x−1 ) = (xg −1 )−1 = (g −1 x)−1 = (x−1 )g. So x−1 ∈ Z(G). So Z(G) is a subgroup. To see
that it is normal, we must check that for any x ∈ Z(G) and for all g ∈ G we have gxg −1 ∈ Z(G).
So let h be any element of G. We see h(gxg −1 ) = h(xgg −1 ) = hx = xh = xgg −1 h = (gxg −1 )h. So
indeed, gxg −1 ∈ Z(G), and therefore Z(G) is a normal subgroup. 
2.3.13 (a) Let G be a group of order 4. Prove that every element of G has order 1, 2, or 4.
Proof. Note, if a ∈ G, then ord(a) ≤ 4 since |hai| = ord(a). We need to show there can be no
element of order 3. Suppose x ∈ G has order 3. This means G consists of the four elements
1, x, x2 , y, where y is some other element of G. But now the multiplication table for G is almost
determined. What can xy be? By cancellation, it can’t be y, so it must be a power of x, i.e.
2
xy = xp . Now cancel x and we find that y is also a power of x. But this contradicts our assumption
that x has order 3, since there are four distinct powers of x in G. 
(b) Classify groups of order 4 by considering the following two cases: (i) G contains the element of
order 4, and (ii) Every element of G has order < 4.
Proof. If G contains an element of order 4 then G is cyclic, and all cyclic groups of the same order
are isomorphic. Thus G is isomorphic to C4 .
If G does not contain an element of order 4 (i.e. we are in case (ii)), then, since only the identity
has order 1, we must have at least two distinct elements x, y of order two (x2 = 1, y 2 = 1). So
we have accounted for three out of the four elements: 1, x, y. Now consider xy and yx, these must
be different from 1, x, y since otherwise we arrive at a contradiction [by cancellation]. So xy is the
fourth element of the group and so is yx. This means: x, y generate G and xy = yx, x2 = 1, y 2 = 1.
This is the Klein four group. One can write an explicit isomorphism taking the generators to the
corresponding generators. 

You might also like