You are on page 1of 85

Solutions to (Selected) HW Problems

Math 455 – Fall 2006

January 12, 2007

Please read: I will try to post here a few solutions. The new solutions will be added to
this same file. They might come with no explanation, just the “answer”. If yours do not
match mine, you can try to figure out again. (Also, read the disclaimer below!) You can
come to office hours if you want explanations for the unexplained answers. Be careful that
just because our “answers” were the same, it doesn’t mean that you solved the problem
correctly (it might have been a “fortunate” coincidence), and in the exams what matters is
the solution itself. I will do my best to post somewhat detailed solutions, though.

Disclaimer: I will have to put these solutions together rather quickly, so they are subject
to typos and conceptual mistakes. (I expect you to be a lot more careful when doing your
HW than I when preparing these.) You can contact me if you think that there is something
wrong and I will fix the file if you are correct.

Homework 1

Section 1.1
1. a2 1 = 2, a2 3 = 8.
 
" # −20 −28 −28
1 0  
2. (a) AB = , BA =  14 33 .
32 
0 1 
−9 −12 −11
" # " #
−2 4 2 16
(b) AB = , BA = .
0 0 1 −8

1
 
1 h 2i 1
 
(c) AB = 
 −1 −2 −1 , BA = −1 .

0 0 0
4. We have:
   
" # " #! 1 " # 1 " #
1 2 0 1 2   2 3 8   38
· ·
 4 =
 ·
 4 =
 ,
0 1 1 1 3 1 1 3 14
3 3
and   
" # " # 1 " # " # " #
1 2  0 1 2   1 2 10 38
· ·
 4  =
 · = ,
0 1  1 1 3 0 1 14 14
3

11. (a)  
d1 a1 1 d1 a1 2 . . . d1 a1 n
 
 d2 a2 1 d2 a2 2 . . . d2 a2 n 
DA =  ,
 
.. .. .. ..

 . . . . 

dn an 1 dn an 2 . . . dn an n
and  
d1 a1 1 d2 a1 2 . . . d n a1 n
 
 d1 a2 1 d2 a2 2 . . . d n a2 n 
AD =  .
 
.. .. .. ..

 . . . . 

d1 an 1 d2 an 2 . . . dn an n
(b)
     
a1 b1 a1 b 1
     
 a2   b2   a2 b 2 
· = .
     

 ...   ...   ... 
     
an bn an bn

(c) By the previous item one can see that a diagonal matrix is invertible if, and only if,
all the elements of the diagonal are non-zero. (Or, equivalently, if the determinant
is non-zero.) Then:
 −1  
a1 a−1
1

 ... 


= ... 

   
an a−1
n

2
Section 1.2
2. We can solve it in general, and then use the result for the different cases. We will
consider then the system:
   
1 2 1 1 a
   
 3 0
 0 4 ·X = b 
 
.
1 −4 −2 −2 c

By row reduction (Gaussian elimination):


     
1 2 1 1 a 1 2 1 1 a 1 2 1
1 a

     
 ∼  0 6 3 −1
 3 0 0 4 b   3a − b  ∼  0 6 3 −1 3a − b .
   

1 −4 −2 −2 c 0 6 3 3 a−c 0 0 0 4 −2a + b − c

Thus, if  
x1
 
 x 
 2 
X= ,
 x3 
 
x4
we obtain:
1 1 1
x4 = − a + b − c,
2 4 4
5 1 1 x3
x2 = a− b− c− ,
12 8 24 2
2 1
x1 = a + c.
3 3
[We use the last row of the matrix to find x4 , then replace the value found for x4 in
the second row of X and solve for x2 (leaving x3 as a parameter), etc.]

(a) We have a = b = c = 0, and thus we obtain:

x1 = 0, yx2 = −λ/2, x3 = λ, x4 = 0,

where λ is a free parameter. [Note that in this case we are computing the kernel of
the linear map given by the matrix. The kernel is then the vector space generated
by (0, −1/2, 1, 0).]

3
(b) We have a = b = 1 and c = 0. Thus, we obtain
2 7 1
x1 = , x2 = − λ/2, x3 = λ, x4 = − ,
3 24 4
where λ is a free parameter.
(c) We have a = 0 and b = c = 2. Thus, we obtain
2 1
x1 = , x2 = − − λ/2, x3 = λ, x4 = 0,
3 6
where λ is a free parameter.

12. Proof. One could use determinants here: a matrix is invertible if, and only if, the
determinant is non-zero. Moreover, the determinant is multiplicative, i.e., det(AB) =
det(A) det(B).
But, for Artin, determinants haven’t been covered yet. So here goes a proof in the
def
context of the book (no determinants): let (AB)−1 be the inverse of AB, and M =
B(AB)−1 . Thus AM = A(B(AB)−1 ) = (AB)(AB)−1 = Id. (Here Id denotes the
identity matrix.) Thus, M is a left inverse of A. By Proposition 2.23 (pg. 17), M is
the inverse of A.
def
In the same manner, let N = (AB)−1 A. Then N B = Id and, again by Proposition
2.23, N is the inverse of B.

14. (a) Proof. We just need to be careful with the indices. If A = (ai j ) and B = (bj k ),
where i ∈ {1, . . . , l}, j ∈ {1, . . . , m} and k ∈ {1, . . . , n}, then:
m
X
AB = (ci k ), where ci k = ai j bj k .
j=1

Also,
def
At = (ãj i ), where ãj i = ai j ,
def
B t = (b̃k j ), where b̃k j = bj k ,
def
(AB)t = (c̃k i ), where c̃k i = ci k ,

with i, j and k in the same ranges as before.


From the formulas above formulas we obtain
m
X m
X
B t At = (dk i ), where dk i = b̃k j ãj i = ai j bj k = ci k = c̃k i .
j=1 j=1

4
Hence dk i = ck i for all i and j, and thus B t At = (dk i ) = (c̃k i ) = (AB)t .
Also, observe that

˜i j ),
(At )t = (ã ˜i j def
where ã = ãj i = ai j .

Therefore, (At )t = A.

(b) Note that (AA−1 )t = Idt = Id. On the other hand, from item (a), we have
(AA−1 )t = (A−1 )t At . Putting the two formulas together, we have (A−1 )t At = Id.
Repeating the same argument with (A−1 A)t instead of (AA−1 )t , we obtain that
At (A−1 )t = Id.
Thus, (A−1 )t is the inverse of At , i.e., (At )−1 = (A−1 )t .

Section 1.3
1. (b) 1 · (−1) − 1 · 1 = −2.

(c) 2 · 1 · 2 + 0 · 0 · 1 + 1 · 0 · 0 − 1 · 1 · 1 − 0 · 0 · 2 − 2 · 0 · 0 = 3.

(c) Since the matrix is lower triangular, the determinant is the product of the elements
in the diagonal: 1 · 2 · 3 · 4 = 24.

2.

1 2 5 6 2 1 5 6



3 1 7 7 1 3 7 7
= − [Switch first and second columns.]



0 0 2 3
0 0 2 3
4 2 1 5 2 4 1 5



2 1 5 6−5



1 3 7 7−7
= − [Subtract the third column from the fourth.]



0 0 2 3−2

2 4 1 5−1



2 1 5 1



1 3 7 0
= − .


0 0 2 1

2 4 1 4

5
Homework 2

Section 1.4
1. (a)  
0 1 0 0
 
 0 0 0 1 
 
 
 1 0 0 0 
 
0 0 1 0

(b) We can do it by observing the changes of the rows:


   
1 0 0 0 0 0 1 0
   
 0 1 0 0  switch 1 and 3  0 1
  0 0 
 switch 2 and 1
 −−−−−−−−−−→   −−−−−−−−−−→


 0 0 1 0   1 0 0 0 
   
0 0 0 1 0 0 0 1
   
0 1 0 0 0 1 0 0
   
 0 0 1 0   switch 2 and 4  0 0 0 1 

 −−−−−−−−−−→ 
 
 
 1 0 0 0   1 0 0 0 
   
0 0 0 1 0 0 1 0

Let
def def def
f1 = 1 ↔ 3, f2 = 1 ↔ 2, f3 = 2 ↔ 4.

Then, Mf = Mf3 · Mf2 · Mf1 · Id = Mf3 · Mf2 · Mf1 = Mf3 ◦f2 ◦f1 . So, since f and
f3 ◦f2 ◦f1 have the same permutation matrix associated with them, f = f3 ◦f2 ◦f1 .
[Let’s double check:

f3 ◦ f2 ◦ f1 (1) = f3 ◦ f2 (3) = f3 (3) = 3;


f3 ◦ f2 ◦ f1 (2) = f3 ◦ f2 (2) = f3 (1) = 1;
f3 ◦ f2 ◦ f1 (3) = f3 ◦ f2 (1) = f3 (2) = 4;
f3 ◦ f2 ◦ f1 (4) = f3 ◦ f2 (4) = f3 (4) = 2.

Hence, indeed f = f3 ◦ f2 ◦ f1 .]

(c) Since f is product of three transpositions, it is odd, and hence sign(f ) = −1.

6
4. We have sign(p) = det Mp , which is either 1 or −1. But then, sign(p−1 ) = det Mp−1 =
det(Mp−1 ) = (det Mp )−1 = (sign(p))−1 . Since 1−1 = 1 and −1−1 = −1, we have
sign(p−1 ) = sign(p).

Section 2.1
1.(b) Let:

def def def


f1 = id, f2 = 1 ↔ 2, f3 = 1 ↔ 3,

def
f4 = 2 ↔ 3, f5 : 1 7→ 2 f6 : 1 7→ 3
2 7→ 3 2 7→ 1
3 7→ 1 3 7→ 2

Then the multiplication table is:

◦ f1 f2 f3 f4 f5 f6
f1 f1 f2 f3 f4 f5 f6
f2 f2 f1 f6 f5 f4 f3
f3 f3 f5 f1 f6 f2 f4
f4 f4 f6 f5 f1 f3 f2
f5 f5 f3 f4 f2 f6 f1
f6 f6 f4 f2 f3 f1 f5

def
3. Proof. Let G = {x ∈ S : x−1 ∈ S}.

(0) [Law of composition] We first prove that the law of composition (or operation) on
S defines a law of composition on G: let x, y ∈ G. So, we have x−1 , y −1 ∈ S. Since
“·” is a law of composition on S, y −1 · x−1 ∈ S. But then, (x · y)−1 = y −1 · x−1 ∈ S.
Therefore, x · y ∈ G.

(1) [Identity] By assumption, S has an identity, which we shall denote by 1. But,


since 1 · 1 = 1, we have 1 = 1−1 [which is always true: the identity is always its
own inverse! ], 1 ∈ S.

(2) [Associativity] Since G ⊆ S and “·” is associative in S [by assumption], it is also


associative in G.

7
(3) [Inverse Element] Note that since x·x−1 = x−1 ·x = 1, we have [by definition] that
(x−1 )−1 = x. [This is also always true: the inverse of the inverse of an element is
the element itself ! ] Hence, if x is invertible in S, so is x−1 . Thus, if x ∈ G, then
x−1 ∈ G.

4. We have:

x y z −1 w = 1 ⇒
y z −1 w = x−1 1 = x−1 ⇒
y z −1 = x−1 w−1 ⇒
y = x−1 w−1 z

7. Proof. Let a, b, c ∈ S. Then (a · b) = a [by definition] and (b · c) = b [also by defn.].


Hence
(a · b) · c = a · c = a

and
a · (b · c) = a · b = a.

Hence, the law of composition is associative.

11. Proof. (0) [Law of Composition] Let a, b ∈ G0 = G. Since “·” is a composition law
in G, we have that a ◦ b = b · a ∈ G = G0 . Thus “◦” is a composition law in G0 .

(1) [Identity] Let 1 be the identity of G. Then, for any a ∈ G0 = G, 1 ◦ a = a · 1 = a


and a ◦ 1 = 1 · a = a. So 1 is also the identity of G.

(2) [Associativity] Let a, b, c ∈ G0 . Then:

(a ◦ b) ◦ c = (b · a) ◦ c = c · (b · a) = (c · b) · a = (b ◦ c) · a = a ◦ (b ◦ c).

(c) [Inverse Element] Let a ∈ G0 = G and a−1 be the inverse of a in G [with respect
to “·”]. Then a ◦ a−1 = a−1 · a = 1G = 1G0 and a−1 ◦ a = a · a−1 = 1G = 1G0

8
Homework 3

Section 2.2
2. Proof. We have

a3 b = ba3 ⇒ a6 b = a3 ba3 [multiply by a3 on the left]


⇒ ab = a3 ba3 [since a5 = 1, a6 = a]
⇒ ab = ba6 [since a3 b = ba3 ]
⇒ ab = ba [a6 = a again].

3. (a) GLn (R) is a subgroup of GLn (C) since it is itself a a group [proved in class] also
with multiplication and is contained in GLn (C).
(b) {1, −1} is a subgroup of R:
(1) [Closed under Multiplication] Since 1·1, 1·(−1), (−1)·1, (−1)·(−1) ∈ {1, −1},
we have that {1, −1} is closed under multiplication.
(2) [Inverse] 1−1 = 1 and (−1)−1 = −1. So every element has an inverse.
(c) The set of positive integers {1, 2, 3, . . . } [with the addition as operation] is not
a subgroup of (Z, +), since not all elements have inverses: 2 has no [additive]
inverse in {1, 2, 3, . . . } [since −2 6∈ {1, 2, 3, . . . }].
def
(d) The set H = (0, ∞) [positive real numbers] is a subgroup of (R, ·):
(1) [Closed under Multiplication] The product of two real numbers is a real num-
ber and if both are positive, then the result is also positive. Hence, H is
closed under multiplication.
(2) [Inverse] If a ∈ H, then 1/a ∈ H.
(e) The set given is not a subset of GL2 (R), since their elements have determinant
zero. Hence it cannot be a subgroup of GL2 (R).

4. We again have to prove the same properties, but we can only assume that if x, y ∈ H,
then x y −1 ∈ H. Here, though, it is easier to do in the opposite order.

(2) [Inverse] Since H is non-empty [by hypothesis], there is some y ∈ H. By assump-


tion we know that y y −1 = 1 ∈ H. Thus, we now know that 1, y ∈ H. But then,
1 y −1 = y −1 ∈ H.

9
(1) [Closed under Multiplication] Let x, y ∈ H. [We need to show that x y ∈ H.]
Since y is H and we have already proved (2), we know that x, y −1 ∈ H. Thus,
x (y −1 )−1 = x y ∈ H.

11. We have:

(ab)m = 1 ⇔ ab(ab)m−2 ab = a(ba)m−1 b = 1


⇔ (ba)m−1 b = a−1 [multiply by a−1 on the left]

⇔ (ba)m−1 ba = a−1 a [multiply by a on the right]

⇔ (ba)m = 1.

Thus, the computation above proves that (ab)m = 1 if, and only if, (ba)m = 1. [Note
that I’ve used “⇔” instead of “⇒”, since every step is reversible, meaning, if you start
at the bottom, you can get to the top. That is why it is “if, and only if ”.]

[Case 1:] If ab has infinite order, then (ab)m 6= 1 for all positive integers m. So,
(ba)m 6= 1 for all positive integers m too [since if (ba)m = 1 for some integer, then
(ab)m = 1 for this same integer], and hence the order of ba is also infinite.

[Case 2:] If ab has finite order, say m, then m is the smallest positive integer such that
(ab)m = 1. By the computation above, we also know that (ba)m = 1 [for this same
positive integer m]. Hence, the order of ba is less than or equal to m. [We need to
show that is in fact “equal to”.] If there is an n < m, also a positive integer, such that
(ba)n = 1 [i.e., the order is in fact less than m], then the computation above tells us
that (ab)n = 1. But this contradicts the fact that m was the least positive integer such
that (ab)m = 1 [since by assumption n < m]. Hence, there can be no such n, and m is
the least positive integer such that (ba)m = 1. Therefore, m is [by definition] the order
of ba.

13. Let G = hxi and H be a subgroup of G. [Case 1:] If H = {1}, then H = h1i, and
hence H is cyclic.

[Case 2:] Suppose that H 6= {1}. Then there is an m 6= 0 such that xm ∈ H. If m < 0,
we know that −m > 0 and x−m = (xm )−1 ∈ H [since a subgroup contains the inverses
of all its elements]. Therefore, the set

S = {n ∈ Z : n > 0 and xn ∈ H}

10
is not empty since ±m ∈ S. Since S is non-empty and contained in the set of positive
integers [and hence bounded below by 0], by the Well Ordering Principle, there is
minimum element l ∈ S.


We now claim that H = xl . Clearly, since xl ∈ H [by the definition of S and since



l ∈ S], we have xl ⊆ H. If xl 6= H, there is an element xt such that xt ∈ H, but


xt 6∈ xl . [The latter means that t is not a multiple of l.] If we make the long division
of t by l: t = q · l + r, with r ∈ {0, 1, . . . , (l − 1)}, then r > 0, since t is not a multiple
of l.
Now, since xl ∈ H, then (xl )−q = x−q·l ∈ H. Also, since xt ∈ H [by our previous
assumption], we have xt · x−q·l = xt−q·l = xr ∈ H. [The last equality is a consequence
of the long division above.] But, since r > 0 and xr ∈ H, we have that r ∈ S [by
the definition of S]. But r < l, and by assumption l was the minimum element of S.
Hence we have a contradiction, which means that there is no such t [i.e., there is no



element xt ∈ H other than the elements of xl .] Thus H ⊆ xl . Since we already


knew that xl ⊆ H, we have that they are actually equal and H is cyclic.

16.(a) Since G = hxi and |G| = 6, we have then that:

G = {1, x, x2 , x3 , x4 , x5 }.

Clearly h1i = {1} =


6 G and hxi = G.
Since (x2 )3 = x6 = 1 [i.e., the order of x2 is less than or equal to 3], we have hx2 i =
{1, x2 , x4 } =
6 G.
In the same way, (x3 )2 = 1 and hx3 i = {1, x3 } =
6 G.
We also have that (x4 )2 = x8 = x6 · x2 = 1 · x2 = x2 , and (x4 )3 = x12 = (x6 )2 = 12 = 1.
Thus, hx4 i = {1, x2 , x4 } = hx2 i =
6 G.
Finally, we have: (x5 )2 = x10 = x6 · x4 = x4 ; (x5 )3 = x15 = (x6 )2 · x3 = x3 ; (x5 )4 =
x20 = (x6 )3 · x3 = x3 ; (x5 )6 = x30 = (x6 )5 = 1. Therefore, hx5 i = G.
Thus, only 2 elements generate G: x and x5 .

17. We need to prove that for all a, b ∈ G, where G is the group in question, we have
ab = ba.
Since every element of G has order 1 [the identity] or order 2 [the rest], we have that
for all x ∈ G, x2 = 1. Hence, for any a, b ∈ G, since ab ∈ G [law of composition], we

11
have (ab)2 = 1. Moreover, also a2 = 1 and b2 = 1. [In G, every element is its own
inverse.] Thus:

(ab)2 = 1 ⇒ abab = 1 [just expand power]


⇒ aba = b [multiply both sides by b on the right]
⇒ ab = ba [multiply both sides by a on the right],

which is what we needed to prove.

Homework 4

Section 2.3
def
1. Let φ(x) = ex . We will prove that φ is an isomorphism.
[(1) Bijection:] From calculus (or even precalculus), we know that ex has domain R,
range (0, ∞) [which is the set of positive real numbers] and is strictly increasing [for
instance, because φ0 (x) = ex > 0 for all x ∈ R]. Hence φ(x) is one-to-one [since it is
strictly increasing], and since we are considering the co-domain to be (0, ∞), it is also
onto. [Given any y ∈ (0, ∞), we have that eln(y) = y.] Therefore, φ is a bijection from
R onto (0, ∞).
[(2) Compatible with operations:] Also, observe that φ(x + y) = ex+y = ex · ey =
φ(x) · φ(y). Thus, φ is an isomorphism. [Note that the operation of R is “+”, while
the operation of (0, ∞) is “·”.]
def
[Note: φ(x) = ax for any a ∈ (0, ∞), except a = 1, also works.]

4. (a) We can do it for any n ∈ Z, even negatives! Let us do it by induction on n.


[First step:] For n = 1, the statement is clearly true. For n = −1, observe that

(aba−1 )(ab−1 a−1 ) = abb−1 a−1 = aa−1 = 1.

Hence, ab−1 a−1 is the inverse of aba−1 , i.e., (aba−1 )−1 = ab−1 a−1 . [We need the
case “−1” to make an induction also for negative integers.]
[Second step:] Assume that, for any positive n, we have that (aba−1 )n = abn a−1
and (aba−1 )−n = ab−n a−1 . [We need to prove that (aba−1 )n+1 = abn+1 a−1 , and
that (aba−1 )−(n+1) = ab−(n+1) a−1 .] We have:

(aba−1 )n+1 = (aba−1 )n (aba−1 ) = abn a−1 aba−1 = abn ba−1 = abn+1 a−1 ,

12
and

(aba−1 )−(n+1) = (aba−1 )−n (aba−1 )−1 = ab−n a−1 ab−1 a−1 = ab−n b−1 a−1 = ab−(n+1) a−1 .

(b) By (a), we have that (aba−1 )4 = ab4 a−1 . But, since we have that aba−1 = b2 , we
obtain b8 = (b2 )4 = ab4 a−1 . Then:

b8 = ab4 a−1 = a(b2 )2 a−1 = a(aba−1 )2 a−1 = a(ab2 a−1 )a−1


= a2 b2 a−2 = a2 (aba−1 )a−2 = a3 ba−3 .

5. [(1) Bijection:] Any bijection [not only for groups!] has an inverse and its inverse is
also a bijection. [This is easy to prove, and I assume you’ve done in Math 300 or
even in precalculus. So, we can just say it. Actually, I have used this fact already
when dealing with permutations.] So, if φ : G → G0 is an isomorphism [and hence a
bijection], then its inverse function φ−1 : G0 → G is also a bijection.

[(2) Compatible with operations:] Let a0 , b0 ∈ G0 . [We need to show that φ−1 (a0 · b0 ) =
φ−1 (a0 )·φ−1 (b0 ).] Since φ is a bijection [and hence onto], we know that there are a, b ∈ G
such that φ(a) = a0 and φ(b) = b0 . [Note that this also means that φ−1 (a0 ) = a and
φ−1 (b0 ) = b, since φ−1 is the inverse function of φ.] Therefore:

φ−1 (a0 · b0 ) = φ−1 (φ(a) · φ(b)) [by the previous remark]

= φ−1 (φ(a · b)) [ since φ is an isomorphism]

=a·b [since φ−1 is the inverse of φ]

= φ−1 (a0 ) · φ−1 (b0 ) [by def. of a and b].

9. Remember that as sets, G = G0 . The real difference is the operation. If the operation
def
of G is “·”, the operation of G0 is denoted by “◦” and defined as a ◦ b = b · a.

Let φ : G → G0 = G be defined as φ(x) = x−1 . [Since G0 = G as a set, we have


x−1 ∈ G = G0 , and the function is well defined.] We will show that φ is an isomorphism
between G and G0 .

[(1) Bijection:] To show that φ is a bijection we don’t use the group properties: being
bijective is a set theoretic property. So, as a function, we can think of φ and a function
from G to G [instead of from G to G0 ]. Then we can refer to problem 12(a) below,
where we carefully prove that φ(x) = x−1 is a bijection.

13
[(2) Compatible with operations:] To prove that φ is compatible with the operations
of G and G0 [so here the operations are important] we need to prove that φ(a · b) =
φ(a) ◦ φ(b). But:

φ(a · b) = (a · b)−1 [by def. of φ]

= b−1 · a−1 [property of inverses]

= φ(b) · φ(a) [def. of φ]

= φ(a) ◦ φ(b) [def. of “◦”].

12. (a) Let us first show that φ is one to one:

φ(x) = φ(y) ⇒ x−1 = y −1


⇒ xx−1 = xy −1
⇒ 1 = xy −1
⇒ y = xy −1 y
⇒ y = x.

Hence, φ is one-to-one.
Let us now show that φ is onto. Let y ∈ G. [We need to find x ∈ G such that
φ(x) = y.] Take x = y −1 . Then φ(x) = x−1 = (y −1 )−1 = y. Thus, φ is onto.
(b) Let x, y ∈ G. Since we know that φ is always a bijection [part (a)], we have that
φ is an isomorphism if, and only if, for all x, y ∈ G, we have φ(x · y) = φ(x) · φ(y).
Moreover, by definition, we know that G is Abelian if, and only if, for all x, y ∈ G,
we have x · y = y · x.
Let x, y ∈ G. Then:

φ is an isom. ⇔ φ(x · y) = φ(x) · φ(y)


⇔ (x · y)−1 = x−1 · y −1
⇔ y −1 · x−1 = x−1 · y −1
⇔ x−1 = y · x−1 · y −1
⇔ x−1 · y = y · x−1
⇔ y = x · y · x−1
⇔ y·x=x·y
⇔ G is Abelian.

14
Thus, φ is an isomorphism if, and only if, G is Abelian.

Homework 5

Section 2.4
def
4. Let φ : Z → Z be [an arbitrary] homomorphism. Also, let k = φ(1). Then, if n > 0,

φ(n) = φ(1| + 1 +{z· · · + 1}) = φ(1) + φ(1) + · · · + φ(1) = n · φ(1).


| {z }
n-times n-times

If n < 0, then n = −|n| and,

φ(n) = φ(−|n|) = −φ(|n|) = −φ(1| + 1 +{z· · · + 1})


|n|-times

= −(φ(1) + φ(1) + · · · + φ(1)) = −|n| · φ(1) = n · φ(1).


| {z }
|n|-times

For n = 0, we have φ(0) = 0 = 0 · φ(1) [where the first equality holds since φ is a
homomorphism]. Hence, for all n ∈ Z, φ(n) = n · φ(1).

[The above statement can be generalized in the following way:

Proposition. If φ : G → G0 is a homomorphism, then for all


n ∈ Z and a ∈ G, we have φ(an ) = φ(a)n .

With the additive notation, the formula becomes φ(n · a) = n · φ(a). The proof of the
proposition goes just like the above! You can use this proposition (without proving it)
from now on! ]

Hence, every homomorphism from Z to Z is of the form φ(n) = k · n for some k ∈ Z.


[Namely, k is just φ(1).] Moreover, for any k ∈ Z, the function φ(n) = k · n is a
homomorphism: φ(n + m) = k(n + m) = k · n + k · m = φ(n) + φ(m).

Injective: Let m, n ∈ Z. If φ(n) = φ(m), then k · m = k · n. If k 6= 0, then m = n.


Thus, every homomorphism from Z to Z, except φ(n) = 0 [for all n], is injective.

Surjective: If φ is surjective, then there is some n ∈ Z, such that φ(n) = 1 [since 1 ∈ Z].
But φ(n) = k · n = 1, implies that n = 1/k. But, 1/k ∈ Z if, and only if, k = ±1.

15
Thus the only surjective homomorphisms from Z to Z are φ(n) = n [the identity] and
φ(n) = −n.

Isomorphism: If φ is an isomorphism, then it is both injective and surjective. From


the above, we can see that the two surjective homomorphisms are also injective. Thus
the only automorphisms of Z are φ(n) = n [the identity] and φ(n) = −n.

[What is really behind all this is the fact that Z = h1i. By the proposition above, if we
know the value of a homomorphism at the generator of a cyclic group, then we know
its value at any other element.]

8.(b) The only subgroups of Q8 = {±1, ±i, ±j, ±k} are:

H0 = h1i = {1}, H1 = h−1i = {±1},


H2 = hii = h−ii = {±1 ± i}, H3 = hji = h−ji = {±1 ± j},
H4 = hki = h−ki = {±1 ± k}, H6 = Q 8 .

All of them are normal in Q8 . [There are some computations to be done to prove all
that. I won’t post them here, but feel free to come see me if you want to see them.]

Note that every proper subgroup of Q8 is cyclic, but Q8 itself is not cyclic [nor even
Abelian!].

9. (a) Let ψ : G → G0 and φ : G0 → G00 be homomorphisms, and a, b ∈ G. Then

φ ◦ ψ(ab) = φ(ψ(ab)) [by definition of composition]

= φ(ψ(a)ψ(b)) [since ψ is a homomorphism]

= φ(ψ(a))φ(ψ(b)) [since φ is a homomorphism]

= (φ ◦ ψ(a))(φ ◦ ψ(b)) [by definition of composition].

Hence, φ ◦ ψ is a homomorphism.

(b) If a ∈ G is in ker φ◦ψ, then φ◦ψ(a) = φ(ψ(a)) = 1G0 , i.e., ψ(a) ∈ ker φ. Therefore,
a ∈ ψ −1 (ker φ). [Remember from Math 300: if f : A → B is any function between
two sets, and S ⊆ B, then

def
f −1 (S) = {x ∈ A : f (x) ∈ S}.

f −1 (S) is called the pre-image of S by f .] Thus, ker φ ◦ ψ ⊆ ψ −1 (ker φ).

16
Now, if a ∈ ψ −1 (ker φ), then, by definition, ψ(a) ∈ ker φ. This means that
φ(ψ(a)) = 1G00 , and thus, a ∈ ker φ ◦ ψ. So, ψ −1 (ker φ) ⊆ ker φ ◦ ψ.

Hence, ψ −1 (ker φ) = ker φ ◦ ψ.


13.(a) We can do it with the one-step method. [Clearly, gHg −1 6= ∅.] Let a, b ∈ gHg −1 .
[Note that g, although arbitrary, is fixed after you choose it! It’s always the same
g.] We need to prove that ab−1 ∈ gHg −1 .
Since a, b ∈ gHg −1 , [by the definition of gHg −1 ] there are x, y ∈ H such that
a = gxg −1 and b = gyg −1 . Then

ab−1 = (gxg −1 )(gyg −1 )−1 = (gxg −1 )(gy −1 g −1 ) = g(xy −1 )g −1 .

Since H < G and x, y ∈ H, we know that xy −1 ∈ H. Then, ab−1 = g(xy −1 )g −1 ∈


gHg −1 .
14. Let n ∈ N and g ∈ G. [We need to show that g −1 ng ∈ N .] We have that g −1 ∈ G.
Since N / G, we have that g −1 n(g −1 )−1 = g −1 ng ∈ N . [Remember, N / G means
that for all x ∈ G and all n ∈ N , we have xnx−1 ∈ N . So we just took x = g −1 .]
17. [I had proved in class (by mistake) that Z(G) < G, so you only had to show
that it is normal. But note that, in general, to prove that something is a normal
subgroup, first you prove it is a subgroup, then you show that it is normal. For
sake of completeness, I will give the whole proof here.]
Let us first prove that Z(G) < G. First observe that Z(G) 6= ∅, since 1G ∈ Z(G).
We will do it with the two-step method :
(1) Closed under the operation: Let a, b ∈ Z(G). [We need to show that ab ∈
Z(G). By definition of Z(G), we need to show then that for all x ∈ G, we
have (ab)x = x(ab).] Let x ∈ G. Then

(ab)x = abx = axb [b ∈ Z(G)]


= xab = x(ab) [a ∈ Z(G)].

(2) Inverse: Let a ∈ Z(G). [We need to show that a−1 ∈ Z(G).] Let x ∈ G.
Since a ∈ Z(G), we have ax = xa. Then:

ax = xa ⇒ x = a−1 xa [multiply by a−1 on the left]

⇒ xa−1 = a−1 x [multiply by a−1 on the right]

17
Since x ∈ G was arbitrary, a−1 ∈ Z(G).

Therefore, Z(G) < G. It remains to show that it is normal. Then, let x ∈ G and
a ∈ Z(G). We have, xax−1 = axx−1 = a. [The first equality comes from the fact
that a ∈ Z(G).] Hence, xax−1 ∈ Z(G). Therefore, Z(G) / G.

22. (a) Let G = hxi. We will prove that G0 = hφ(x)i. Clearly, [since φ(x) ∈ G0 ] we
have that hφ(x)i ⊆ G0 . So, it suffices to show that G0 ⊆ hφ(x)i.
Let a0 ∈ G0 . [We need to show that a0 ∈ hφ(x)i, i.e., that there exists an n ∈ Z,
such that a0 = φ(x)n .] Since φ is onto, there a ∈ G such that φ(a) = a0 . Since,
G = hxi, there exists n ∈ Z such that a = xn . So, a0 = φ(a) = φ(xn ) = φ(x)n .
Thus a0 ∈ hφ(x)i.
(b) Let a0 , b0 ∈ G0 . [We need to show that a0 b0 = b0 a0 .] Since φ is onto, there are
a, b ∈ G such that a0 = φ(a) and b0 = φ(b). Hence:

a0 b0 = φ(a)φ(b) [a0 = φ(a) and b0 = φ(b)]

= φ(ab) [φ is a homomorphism]

= φ(ba) [G is Abelian]

= φ(b)φ(a) [φ is a homomorphism]

= b0 a0 [a0 = φ(a) and b0 = φ(b)].

Thus, G0 is Abelian.

Homework 6

Section 2.5
2. (i) Reflexive: Let G be a group. We need to show that G ∼ G, i.e., that G is
isomorphic to itself. [To prove that two groups are isomorphic, we need to exhibit
the isomorphism!] Let id : G → G be the identity function, i.e., for all g ∈ G, we
have that id(g) = g. Then, it is clearly bijective and a homomorphism. [It is easy
to prove, if you are not convinced.]

(ii) Symmetric: Suppose that G ∼ G0 . [We need to prove that G0 ∼ G.] We then
know [by definition] that there is an isomorphism φ : G → G0 . [Hence, φ is a
bijective homomorphism.] In problem 2.3.5 [from Homework 4], we saw that the

18
inverse function φ−1 : G0 → G [which exists since φ is a bijection] is also an
isomorphism. Hence, G0 ∼ G.

(iii) Transitive: Suppose that G ∼ G0 and G0 ∼ G00 . [We need to show that G ∼ G00 .]
By definition, we know that there are isomorphisms φ : G → G0 and φ0 : G0 → G00 .
def
Then, let φ00 = φ0 ◦ φ : G → G00 . On problem 2.4.9(a) [from Homework 5], you
showed that the composition of homomorphisms is a homomorphism. Hence φ00
is a homomorphism. Moreover, [as you probably have seen in Math 300, and
as I’ve mentioned before], the composition of bijective functions is also bijective.
Thus, φ00 is bijective, and hence an isomorphism. Therefore, we have proved that
G ∼ G00 .

5. I, again, accidentally solved this problem in class. I used the notation:

a≡b (mod H) iff b−1 a ∈ H.

But, here it is again anyway:

(i) Reflexive: Since H < G, we have that 1G = a−1 a ∈ H. Thus, by definition, a ∼ a.

(ii) Symmetric: Suppose that a ∼ b. Then b−1 a ∈ H. Since H < G, we have that
(b−1 a)−1 = a−1 b ∈ H. Thus, by definition, b ∼ a.

(iii) Transitive: Suppose that a ∼ b and b ∼ c. By definition, we know that b−1 a, c−1 b ∈
H. Since H < G, we have that (c−1 b)(b−1 a) = c−1 a ∈ H. Thus, by definition,
a ∼ c.

6. By definition, x is conjugate to y means that there is a g ∈ G such that x = gyg −1 . [It


is OK if you used y = gxg −1 instead, since you prove it is symmetric anyway...]

(a) (i) Reflexive: Since x = 1G x 1−1


G [and 1G ∈ G], we have that x ∼ x.

(ii) Symmetric: Suppose that x ∼ y. Then, there is g ∈ G such that x = gyg −1 .


[We need now to find g 0 such that y = g 0 x(g 0 )−1 .] But, multiplying the
equation by g −1 on the right, and by g on the left, we obtain y = g −1 xg.
Since g −1 ∈ G, we have that y ∼ x. [So, basically, we take g 0 = g −1 .]
(iii) Transitive: Suppose that x ∼ y and y ∼ z. By definition, we know that
x = gyg −1 and y = g 0 z(g 0 )−1 for some g, g 0 ∈ G. [We need now to find g 00

19
such that z = g 00 x(g 00 )−1 .] But replacing the second equation [y = g 0 z(g 0 )−1 ]
in the first [x = gyg −1 ], we obtain

x = g(g 0 z(g 0 )1 )g −1 = (gg 0 )z((g 0 )−1 g −1 ) = (gg 0 )z(gg 0 )−1 ,

and since gg 0 ∈ G, we have by definition that x ∼ z. [So, basically, g 00 = gg 0


in this case.]

(b) We have that the conjugacy class of a is given by:

def
Ca = {x ∈ G : x ∼ a}
= {x ∈ G : x = gag −1 , for some g ∈ G}
= {gag −1 : g ∈ G}.

So, Ca = {a}, if and only if, for all g ∈ G, we have gag −1 = a. Multiplying by
g on the right, we obtain ga = ag for all g ∈ G. So, Ca = {a} if, and only if,
a commutes with every element of G, i.e., if, and only if, a ∈ Z(G). [Remember
that Z(G) denotes the center of G.]

Homework 7

Section 2.6
1. The left cosets of nZ in Z are:

0 + nZ = nZ = {0, ±n, ±2n, ±3n, . . . }


1 + nZ = {1 ± n, 1 ± 2n, 1 ± 3n, . . . }
2 + nZ = {2 ± n, 2 ± 2n, 2 ± 3n, . . . }
..
.
(n − 1) + nZ = {(n − 1) ± n, (n − 1) ± 2n, (n − 1) ± 3n, . . . }.

To prove that these are all cosets of nZ in Z, it suffices to show that they make a
partition of Z [since clearly they are cosets].

[We first prove that those cosets do not intersect.] Given a coset i + nZ, we can always
find its smallest non-negative element [by the Well Ordering Principle]. Clearly, all
the cosets above have different smallest non-negative element [they are, respectively,

20
0, 1,. . ., (n − 1)], and hence no two are equal. Since they are cosets, no two of them
intersect.

[Now we prove that they cover all Z.] Let k ∈ Z. We need to show that k is in one of
the cosets listed above. By dividing k by n, we can write k = nq + r, where q ∈ Z and
the remainder r is in {0, 1, . . . , (n − 1)}. Thus, k ∈ r + nZ [since k is r plus a multiple
of n]. But the restriction on r tells us that r + nZ is one of the cosets listed above.

Hence, [Z : nZ] = n.

3. I will prove something more general:

If G is a group that has an element of order power of a prime p, then it has an element
of order p.

This is more general since if |G| = pn , then any element x ∈ G has order power of p,
since ord(x) = |hxi| | |G| = pn , and thus ord(x) = pr for some r ≤ n. So, if the above
statement is true, so is the statement of the problem!

Here is the proof: Let G be a group and x an element of order pr , where p is prime
and r is a positive integer. [Hence, x 6= 1G .]
def r−1
Let y = xp . We clearly have that y 6= 1G [since the order of x is pr and 0 < pr−1 < pr ]
r
and y p = xp = 1G . So, y has order greater than one and less than or equal to p. [We
need to show that it is in fact equal.]
r−1
Since y = xp , we clearly have that hyi < hxi. Hence, by Lagrange’s Theorem,
|hyi| | |hxi|. By definition of the order of an element, the order of y divides the order
of x. The latter is pr , and the only divisor of pr which is less than or equal to p and
greater that one is p.

5. If you look at the results from section 2.7, we see that if H, K < G and gcd(|H| , |K|) =
1 [where gcd denotes the greatest common divisor ], then H ∩ K = {1G }. Since
gcd(3, 5) = 1, one could just apply the result to this problem. But, since this was
done in a later section, you should not use that result in this problem.

We know that if a group has prime order, then it is cyclic, generated by any non-
identity element. Suppose then that x ∈ H ∩ K. Since the orders of H and K are
prime, if x 6= 1G , then x generates H and K [since x belongs to both]. But this says
that the order of x is 3 [since it generates H] and 5 [since it generates K], which clearly

21
cannot happen. Hence, H ∩K has no non-identity element. Since the identity is clearly
in H ∩ K [since H and K have the identity], we have H ∩ K = {1G }.

10.(a) Let a ∈ G. If a ∈ H, clearly aH = H. [Remember, if x ∈ yH, then xH = yH, since


then x ∈ xH ∩ yH, and the left cosets make a partition of G, and thus non-empty
intersection means equality.] But then, also Ha = H, [just as with left cosets, since
right cosets also make a partition].
If a 6∈ H, then aH =
6 H [since a ∈ aH, but a ∈ 6 H], and since [G : H] = 2, we have
that G = H ∪˙ aH. Hence, aH is the set of all elements of G that are not in H. But,
since a 6∈ H, and since we also only have two right cosets [by problem 2.6.9, which I
solved in class], we have G = H ∪˙ Ha, since again Ha 6= H [for a ∈ Ha, but a 6∈ H].
But then, all elements of G that are not in H are the elements of Ha. Thus aH = Ha,
and by Proposition 6.8, H / G.
[Alternatively, if you don’t want to use problem 2.6.9, you can prove, using the fact we
only have two left cosets, that aH = a−1 H, and then prove Ha−1 ⊆ aH = a−1 H.]

Section 2.7
1. We know that |G| = |ker φ| |im φ|. Hence |im φ| | |G|. Also, since im φ < G0 , we have
that |im φ| | |G0 |. Thus, |im φ| is a common [positive] divisor of |G| and |G0 |. Since
gcd(|G| , |G0 |) = 1, we have that |im φ| = 1, and thus im φ = {1G0 }.

3. (a) Suppose that xH ∩ yH 6= ∅. [If it were, we’d be done!] We need to find z ∈ G


such that xH ∩ yK = z(H ∩ K).
Let a ∈ xH ∩ yK. Then, we have a = xh = yk, for some h ∈ H and k ∈ K.
Since H ∩ K < G, the left cosets of H ∩ K make a partition of G. In particular,
they cover all of G. Since a ∈ G, there is one left coset, say z(H ∩ K), such that
a ∈ z(H ∩ K). Hence, there is an ` ∈ H ∩ K, such that a = z`.
We now show that xH ∩ yK ⊆ z(H ∩ K). [We need to show that if b ∈ xH ∩ yK,
then b ∈ z(H ∩ K).] Let b ∈ xH ∩ yK. Then, b = xh0 = yk 0 , for some h0 ∈ H
and k 0 ∈ K. By the previous paragraph, we have a = xh = z` [for the same
a used above!]. Thus, x = z`h−1 , and b = xh0 = z`h−1 h0 . In the same way,
since a = yk = z`, we obtain y = z`k −1 , and thus b = z`k −1 k 0 . We then
have b = z`h−1 h0 = z`k −1 k 0 , and canceling out “z`” in this equation we obtain
h−1 h0 = k −1 k 0 . Clearly the left hand side is in H while the right hand side is in

22
K. Since they are equal, we have that h−1 h0 = k −1 k 0 ∈ H ∩ K. But then, since
H ∩ K is a group and ` is also in H ∩ K, we have that `h−1 h0 ∈ H ∩ K and
b = z`h−1 h0 ∈ z(H ∩ K), as we wanted to show.
Now, we need to show z(H ∩ K) ⊆ xH ∩ yK. Then, let c ∈ z(H ∩ K). [We
need to show that c ∈ xH ∩ yK.] Hence c = z`0 for some `0 ∈ H ∩ K. Since
a = xh = z`, we have that z = xh`−1 , and then c = xh`−1 `0 . Since h, `, `0 ∈ H, we
have that c = xh`−1 `0 ∈ xH. In the same way, a = yk = z`, and then z = yk`−1 .
So, c = z`0 = yk`−1 `0 and since `, `0 , k ∈ K, we know that c = yk`−1 `0 ∈ yK.
Since we already proved that c ∈ xH, we have that c ∈ xH ∩ yK [as we needed
to prove].
Since xH ∩ yK ⊆ z(H ∩ K) and z(H ∩ K) ⊆ xH ∩ yK, we have that z(H ∩ K) =
xH ∩ yK

(b) [Note: We cannot assume that G is finite, since that would make the problem
trivial and the statement is true for infinite groups.]
Suppose that [G : H] = m and [G : K] = n. Then:

G = x1 H ∪˙ x2 H ∪˙ · · · ∪˙ xm H
= y1 K ∪˙ y2 K ∪˙ · · · ∪˙ yn K,

for some x1 , . . . , xm , y1 , . . . , yn ∈ G. We then have:

xi H = xi H ∩ G
= xi H ∩ (y1 K ∪˙ y2 K ∪˙ · · · ∪˙ yn K)
= (xi H ∩ y1 K) ∪˙ (xi H ∩ y2 K) ∪˙ · · · ∪˙ (xi H ∩ yn K),

for i = 1, . . . , m.
By item (a), each xi H ∩ yj K is either empty or a left coset of H ∩ K. Hence, xi H
is covered by at most n left cosets of H ∩ K. [Note that the xi H ∩ yj K’s, for i
fixed and j = 1, . . . , n, form a partition of xi H.] Doing this for i = 1, . . . , m, we
have that G is covered by at most mn left cosets of H ∩ K [at most n left cosets
for each one of the m possible xi H]. Hence [G : H ∩ K] ≤ nm < ∞.

23
Homework 8

Section 2.8
2. If S3 ∼
= H × K, then |H × K| = |H| · |K| = |S3 | = 6. [The first equality is just
the cardinality of the Cartesian product, which is the underlying set of H × K.] But
then, |H| = 2 and |K| = 3, or the other way around. [Since the groups H and K are
nontrivial, neither H nor K can have order 1 (or 6)!]

Since H × K ∼
= K × H [via the isomorphism (h, k) 7→ (k, h) – if you are not convinced
that this is an isomorphism, prove it] and isomorphisms of groups is an equivalence
relation [problem 2.5.2 from HW6], we can assume without loss of generality that
|H| = 2 and |K| = 3.

By Corollary 6.13, since 2 and 3 are prime, we have that H and K are cyclic. Hence,
by problem 2.8.3 below [or by the example done in class] H × K ∼
= C6 [where C6 is the
cyclic group of order 6]. Since S3 is not cyclic [nor Abelian], S3 is not isomorphic to
H × K [which is cyclic of order 6 – and hence also Abelian].

Therefore, S3 is not a product of two nontrivial groups. [Note that S3 ∼


= S3 × {1}, via
a 7→ (a, 1).]

3. Let:

Cr = hxi = {1, x, x2 , . . . , xr−1 },


Cs = hyi = {1, y, y 2 , . . . , y s−1 },
Cr = hzi = {1, z, z 2 , . . . , z rs−1 }.

Let’s first assume that gcd(r, s) = 1.


def
Define φ : Cr × Cs → Crs by φ(xm , y n ) = z rn+sm .

(1) φ is well defined: Suppose that (xm1 , y n1 ) = (xm2 , y n2 ). [We need to show that
φ(xm1 , y n1 ) = φ(xm2 , y n2 ).] Then xm1 = xm2 and y n1 = y n2 . Therefore, m2 =
m1 + q1 r [since the order of x is r] and n2 = n1 + q2 s [since the order of y is s] for

24
some q1 , q2 ∈ Z. Then:

φ(xm2 , y n2 ) = z rn2 +sm2 [by definition]

= z r(n1 +q2 s)+s(m1 +q1 r) [by the above remarks]

= z rn1 +sm1 +rs(q1 +q2 )


= z rn1 +sm1 [the order of z is rs]

= φ(xm1 , y n1 ) [by definition].

(2) φ is a homomorphism: We have:

φ((xm1 , y n1 )(xm2 , y n2 )) = φ(xm1 +m2 , y n1 +n2 ) [by definition of the product]

= z r(n1 +n2 )+s(m1 +m2 ) [by definition of φ]

= z rn1 +sm1 z rn2 +sm2


= φ(xm1 , y n1 )φ(xm2 , y n2 ) [by definition of φ]

(3) φ is onto: Let z k , where k ∈ Z, be an arbitrary element of Crs . [Since Crs =


hzi, that is an arbitrary element.] By Bezout’s Theorem [stated in class], since
gcd(r, s) = 1, there are a, b ∈ Z such that ar + bs = 1. Then

φ(xbk , y ak ) = z rak+sbk [by definition of φ]

= z (ar+bs)k = (z ar+bs )k
= zk [since ar + bs = 1].

(4) φ is one-to-one: Since Crs and Cr × Cs are finite [they both have rs elements],
and since φ is onto, by the Pigeonhole Principle, φ is also one-to-one.

Hence we can conclude that φ is an isomorphism, and hence Cr × Cs ∼


= Crs .

Now, we do the converse: assume that gcd(r, s) 6= 1. [We need to prove that Cr × Cs ∼
6=
def
Crs .] Let’s denote d = gcd(r, s). Then, since d > 1, we have that rs/d is an integer
[since d divides r (and s)] and less than rs. Also, observe that (r/d), (s/d) ∈ Z, since
d is a common divisor.

But then,

(xm , y n )rs/d = (xmrs/d , y nrs/d ) = ((xr )m(s/d) , (y s )n(r/d) ) = (1m(s/d) , 1n(r/d) ) = (1, 1).

25
[Here it is crucial that (rs/d), (r/d), (s/d) ∈ Z!!!] Therefore, every element of Cr × Cs
has order less than or equal to rs/d, which is strictly less than rs, and thus there is no
element of order rs in Cr × Cs . But z is an element of order rs in Crs . Thus, Cr × Cs
cannot be isomorphic to Crs .

4. (a) We can just apply Proposition 8.6(c). Since R× is Abelian, every subgroup is
normal. Hence H and K is normal. Clearly H ∩ K = {1}. Moreover, HK = R× :
given y ∈ R× , then y/ |y| ∈ H [y/ |y| is 1 if y > 0 and −1 if y < 0] and |y| ∈ K
[since y 6= 0, we have that |y| is a positive real number]. But y = (y/ |y|) · |y|, and
thus y ∈ HK and hence R× ⊆ HK. Also, since H, K < R× , we clearly have that
HK ⊆ R× , and thus R× = HK. Therefore, by Proposition 8.6(c), R× ∼
= H × K.
(b) We have that
" # " # " # " # " #
a 0 c 0 ac 0 c 0 a 0
· = = · .
0 b 0 d 0 bd 0 d 0 b

Also, " # " # " # " # " #


1 a 1 b 1 a+b 1 b 1 a
· = = · .
0 1 0 1 0 1 0 1 0 1
Hence, H and K are Abelian, and thus so is H × K. [That is easy to see. Try
to prove it if you don’t believe it. It should be very easy.] But, G is not Abelian,
since
" # " # " # " # " # " #
2 1 1 −1 2 −1 2 0 1 −1 2 1
· = 6= = · .
0 1 0 1 0 1 0 1 0 1 0 1

Hence, H × K cannot be isomorphic to G. [Isomorphic groups have the same


properties: if one is Abelian, the other also has to be Abelian.]

5. Let
C∞ = hxi = {1, x, x−1 , x2 , x−2 , . . . } = {xk : k ∈ Z}.

[Note that we have seen in class that C∞ ∼


= Z, so we could have used Z instead of C∞ .]
Since C∞ × C∞ has infinitely many elements, we need to show that it is not cyclic.

Let (xm , xn ) ∈ C∞ × C∞ . Then,

h(xm , xn )i = {(xm , xn )k = (xmk , xnk ) : k ∈ Z}.

26
If m = 0, then h(xm , xn )i = h(1, xn )i =
6 C∞ × C∞ , since [for instance] (x, x) 6∈ h(1, xn )i,
since all elements of h(1, xn )i are of the form (1, xkn ). In the same way, if n = 0, then
(x, x) 6∈ h(xm , 1)i.

So, assume that both m and n are non-zero. Then, (xm , x2n ) 6∈ h(xm , xn )i, since there
is no k ∈ Z such that (xmk , xnk ) = (xm , x2n ), for then we would need k = 1 for the first
coordinate [since m 6= 0] and k = 2 for the second [since also n 6= 0].

6. Let H and K be groups and Z(H) and Z(K) their respective centers.

Let (a, b) ∈ Z(H) × Z(K) [and thus, a ∈ Z(H) and b ∈ Z(K)] and (x, y) ∈ H × K
[and thus, x ∈ H and y ∈ K]. Then,

(a, b)(x, y) = (ax, by) [by the defn. of the prod.]


= (xa, yb) [since a ∈ Z(H) and b ∈ Z(K)]
= (x, y)(a, b) [by the defn. of the prod.].

Thus, (a, b) ∈ Z(H × K), and hence Z(H) × Z(K) ⊆ Z(H × K).

Let (a, b) ∈ Z(H × K), x ∈ H and y ∈ K. Then, (x, y) ∈ H × K, and (x, y)(a, b) =
(a, b)(x, y). This implies that (xa, yb) = (ax, by) [by definition of the product], and
hence, xa = ax and yb = by. Therefore, a ∈ Z(H) and b ∈ Z(K), and thus (a, b) ∈
Z(H) × Z(K). Hence Z(H × K) ⊆ Z(H) × Z(K).

Since Z(H)×Z(K) ⊆ Z(H ×K) and Z(H ×K) ⊆ Z(H)×Z(K), we have Z(H ×K) =
Z(H) × Z(K).

Homework 9

Section 2.9
1.

(7 + 14)(3 − 16) ≡ 21 · (−13) (mod 17)


≡ (4 + 17) · (4 − 17) (mod 17)
≡4·4 (mod 17)
≡ 16 (mod 17).

27
2. (a) Write a = 2q + r, where q ∈ Z and r ∈ {0, 1}. [I.e., divide a by 2. Note that
a ≡ r mod 2.] Then, a2 = 4q 2 + 4qr + r2 = 4(q 2 + qr) + r2 . Since (q 2 + qr) ∈ Z,
and r2 ∈ {0, 1}, we have that a2 ≡ 0, 1 (mod 4). [If r = 0, i.e., if a is even, then
a2 ≡ 0 (mod 4) and if r = 1, i.e., if a is odd, then a2 ≡ 1 (mod 4).]
(b) Let a = 4q + r, where q ∈ Z and r ∈ {0, 1, 2, 3}. Then, a2 = 16q 2 + 8qr + r2 =
8(2q 2 + qr) + r2 . Thus, a2 ≡ r2 (mod 8). So, if r = 0, then a2 ≡ 0 (mod 8), if
r = 1, then a2 ≡ 1 (mod 8), if r = 2, then a2 ≡ 4 (mod 8), and if r = 3, then
a2 ≡ 9 ≡ 1 (mod 8). Hence, a2 ≡ 0, 1, 4 (mod 8).

5. (a) We can do it by trial and error :

2 · 0 ≡ 0 6≡ 5 (mod 9)
2 · 1 ≡ 1 6≡ 5 (mod 9)
2 · 2 ≡ 4 6≡ 5 (mod 9)
2 · 3 ≡ 6 6≡ 5 (mod 9)
2 · 4 ≡ 8 6≡ 5 (mod 9)
2 · 5 ≡ 10 ≡ 1 6≡ 5 (mod 9)
2 · 6 ≡ 12 ≡ 3 6≡ 5 (mod 9)
2 · 7 ≡ 14 ≡ 5 (mod 9)
2 · 8 ≡ 16 ≡ 7 6≡ 5 (mod 9)

So, the only solution set is

7 + 9Z = {7, −2, 16, −9, 23, . . .} = {x ∈ Z : x ≡ 7 (mod 9)}.

(b) Again, we can do it by trial and error :

2 · 0 ≡ 0 6≡ 5 (mod 6)
2 · 1 ≡ 1 6≡ 5 (mod 6)
2 · 2 ≡ 4 6≡ 5 (mod 6)
2 · 3 ≡ 6 ≡ 0 6≡ 5 (mod 6)
2 · 4 ≡ 8 ≡ 2 6≡ 5 (mod 6)
2 · 5 ≡ 10 ≡ 4 6≡ 5 (mod 6)

So, there is no solution modulo 6.

28
Section 2.10
1. Let (" # )
def a b
G= : a, b, c ∈ R, with a, c 6= 0 .
0 c

(a) Let (" # )


def 1 x
H1 = : x, y ∈ R, with y 6= 0 .
0 y

Let: φ : G → R× defined by:


" #!
a b def
φ = a.
0 c

Since " # " # " #


a1 b1 a2 b2 a1 a2 a1 b2 + b1 c2
· = , (1)
0 c1 0 c2 0 c1 c2
we have that φ is a homomorphism.
It’s also onto, since given an x ∈ R× , we have that:
" #!
x 1
φ = x.
0 1

The kernel is clearly H1 , and hence, H1 / G [since the kernel is always a normal
subgroup!], and by the First Isomorphism Theorem, G/H1 ∼ = R× .
[Remark: You see that you don’t have to first prove that H1 is normal. But, on
the other hand, if it isn’t, there is no hope in finding the homomorphism φ. So,
if you don’t see the homomorphism, maybe you should find if the subgroup is
normal, at least in your scratch.]

(b) Let (" # )


def x 0
H2 = : x, y ∈ R× .
0 y

29
Then,
" # " #" #−1 " # " #" #
a b x 0 a b a b x 0 1/a −b/ac
· = ·
0 c 0 y 0 c 0 c 0 y 0 1/c
" #" #
ax by 1/a −b/ac
=
0 cy 0 1/c
" #
x b(−x + y)/c
= 6∈ H2 [if x 6= y and b 6= 0].
0 y

Hence H2 is not a normal subgroup of G.

(d) Let (" # )


def 1 x
H3 = : x∈R .
0 1

Let: ψ : G → R×× R× be defined as:


" #!
a b def
ψ = (a, c).
0 c

The computation (1) of part (a) shows that ψ is a homomorphism.


It’s also onto, since given an (x, y) ∈ R× ×R× , we have that:
" #!
x 1
ψ = (x, y).
0 y

The kernel is clearly H3 , and hence, H3 / G, and by the First Isomorphism The-
orem, G/H3 ∼ = R×× R× .

3. We will give two solutions. One uses the First Isomorphism Theorem, while the other
is more direct.

Solution 1.

Let N be the element of P [I will use P – “curly P” – instead of P for the partition]
such that 1 ∈ N . We first prove that P is a group with the operation “∗” defined as:

def
A ∗ B = C, where C is the element of P such that AB ⊆ C.

30
[We know that such C exists by hypothesis.] Hence, by definition, ∗ is an operation.

Before we proceed, we note that:

Claim: If a ∈ A and b ∈ B, then A ∗ B is the unique element of P that contains ab.


[Observe that a and b are arbitrary!]

Proof. Since P is partition, there is a unique C ∈ P such that ab ∈ C. Let A ∗ B = D.


[We need to show that C = D.] Since ab ∈ AB ⊆ D [by the definition of ∗], then
ab ∈ D, and thus ab ∈ C ∩ D. Since P is a partition and C, D ∈ P with C ∩ D 6= ∅,
we have C = D.

(1) Identity: Let A ∈ P [arbitrary] and a ∈ A [also arbitrary]. Then, since 1 ∈ N ,


we have that a ∈ AN, N A [since a = a · 1, 1 · a]. But, since a ∈ A, by the Claim
above, A ∗ N = N ∗ A = A. [So N is the identity.]

(2) Inverse: Let A ∈ P [arbitrary] and and a ∈ A [also arbitrary]. Let B ∈ P be the
[only] element of P such that a−1 ∈ B. [We will prove that the inverse of A is B.]
Hence, 1 = aa−1 = a−1 a ∈ AB, BA. Since 1 ∈ N , the Claim above tells us that
A ∗ B = B ∗ A = N.

(3) Associativity: Let A, B, C ∈ P and a ∈ A, b ∈ B and c ∈ C. Suppose that


A ∗ B = D1 and B ∗ C = D2 . Then, by definition of ∗, ab ∈ D1 and bc ∈ D2 .
Thus, by the Claim, (A ∗ B) ∗ C = D1 ∗ C is the only element of P that contains
(ab)c = abc. In the same way, A ∗ (B ∗ C) = A ∗ D2 is the only element of P that
contains a(bc) = abc. Therefore, (A ∗ B) ∗ C = A ∗ (B ∗ C).

Now, let φ : G → P where φ(g) is defined as the only element of P that contains g.
We now prove that φ is a homomorphism, surjective and ker φ = N .

(i) φ is a homomorphism: Let a, b ∈ G, φ(a) = A, and φ(b) = B. Then, by definition


of φ, φ(ab) is the only element of P that contains ab. But, by the Claim, this
element is A ∗ B. Thus, φ(ab) = A ∗ B = φ(a) ∗ φ(b).

(ii) φ is surjective: Let A ∈ P. Then, take a ∈ A. By definition of φ, we have


φ(a) = A.

(iii) ker φ = N : By the definition of φ, we have that φ(g) = N [and remember that N
is the identity of P] if, and only if, g ∈ N . Thus ker φ = N .

31
Since N = ker φ [and φ is a homomorphism], N / G.

Finally, we show that P = G/N [as sets, i.e., P is the set of cosets of N in G].

(i) [P ⊆ G/N :] Let A ∈ P and a ∈ A. [We need to show that A is a coset.] Then
[by the definition of φ], A = {g ∈ G : φ(g) = A = φ(a)}. But φ(g) = φ(a) iff
g ∈ a ker φ = aN [as we’ve seen before]. Therefore A = aN . [Note that since we
have “iff”, we get both inclusions, A ⊆ aN and aN ⊆ A, at once.]

(ii) [G/H ⊆ P:] Let aN ∈ G/N . Since P is a partition, there is A ∈ P such that
a ∈ A. Since P ⊆ G/N [by (i) above], there is b ∈ G such that A = bN . Hence
a ∈ aN ∩ bN , and since G/N is a partition, aN = bN = A. Hence aN = A ∈ P.

Solution 2.

Let N be the element of P [I will use P – “curly P” – instead of P for the partition]
such that 1 ∈ N .

Before we proceed, we prove two claims, which we shall use later on:

Claim 1: For all A ∈ P, we have AN, N A ⊆ A.

Proof. Indeed, since 1 ∈ N , we have that A ⊆ AN, N A. Let C1 , C2 ∈ P such that


AN ⊆ C1 and N A ⊆ C2 . [We know that these exist by hypothesis.] Then, A ⊆ C1 , C2 .
Since P is a partition [and A, C1 , C2 ∈ P], we have that A = C1 = C2 [since A ⊆
A ∩ C1 , A ∩ C2 ]. [Observe that in fact we have proved that AN, N A = A, but we don’t
need the other containment here.]

Claim 2: Let a ∈ G and A, B ∈ P, such that a ∈ A and a−1 ∈ B. Then, AB, BA ⊆ N .

Proof. Note that since P is a partition, any element of G is in some element of P.

Let C1 , C2 ∈ P such that AB ⊆ C1 and BA ⊆ C2 . [We know that these exist


by hypothesis.] Then, since a ∈ A and a−1 ∈ B, we have 1 = aa−1 ∈ AB ⊆ C1
and 1 = a−1 a ∈ BA ⊆ C2 . Thus, 1 ∈ N, C1 , C2 . Since P is a partition, we have
C1 = C2 = N .

32
We now show that N < G by the two-step method:

(i) Closed under the operation: By Claim 1 above, N N ⊆ N , i.e., for all a, b ∈ N ,
we have ab ∈ N .
(ii) Inverses: Let a ∈ N . [We need to show that a−1 ∈ N .] Since P is a partition of
G [and hence it covers all of G], there is some A ∈ P such that a−1 ∈ A. Then,
by the Claim 1 above, AN ⊆ A. By Claim 2, we have that AN ⊆ N . Hence
AN ⊆ N ∩ A, and thus [since P is a partition] A = N . Therefore and a−1 ∈ N .

[Notice that N is not empty since 1 ∈ N .]


Now we show that N / G: let g ∈ G and n ∈ N . [We need to show that gng −1 ∈ N .]
Let A ∈ P such that g ∈ A and B ∈ P such that g −1 ∈ B. Then, by the Claim
1, AN ⊆ A. Hence, using Claim 2, AN B ⊆ AB ⊆ N . Thus AN B ⊆ N , and so
gng −1 ∈ N [since g ∈ A, n ∈ N , and g −1 ∈ B].

Finally, we show that P = G/N [as sets, i.e., P is the set of cosets of N in G].

(i) [P ⊆ G/N :] Let A ∈ P and a ∈ A. [Note that a is an arbitrary element of A.]


We will now show that A = aN [and hence, A will be proved to be a coset.]
(a) [aN ⊆ A:] By Claim 1, we know that AN ⊆ A. Since a ∈ A, clearly
aN ⊆ AN ⊆ A.
(b) [A ⊆ aN :] Let b ∈ A. [We need to show that b ∈ aN .] Let B ∈ P such that
a−1 ∈ B. Then, by Claim 2, we have that BA ⊆ N . Thus a−1 b ∈ N [since
a−1 ∈ B and b ∈ A]. Therefore, b = a(a−1 b) ∈ aN [since a−1 b ∈ N ].
Hence, every element of P is a coset, i.e., P ⊆ G/N .
(ii) [G/H ⊆ P:] Let aN ∈ G/N . Since P is a partition, there is A ∈ P such that
a ∈ A. Since P ⊆ G/N [by (i) above], there is b ∈ G such that A = bN . Hence
a ∈ aN ∩ bN , and since G/N is a partition, aN = bN = A. Hence aN ∈ P.

9. Let ψ : G × G0 → G0 be the function defined by ψ(g, g 0 ) = g 0 . It is easy to prove that


this map is a homomorphism and ker ψ = G × {1}. [I actually sketched it in class, but
it is a very easy exercise.] Then, since the kernel is always a normal subgroup, we have
that G × {1} / G × G0 .
Also, ψ is onto, since given g 0 ∈ G0 , we have that ψ(1, g 0 ) = g 0 . By the First Isomor-
phism Theorem, (G × G0 )/(G × {1}) = (G × G0 )/ ker ψ ∼ = G0 .

33
One can easily verify that φ : G × {1} → G defined by φ(g, 1) = g is an isomorphism.
[I also sketched it in class, and again it is an easy exercise.]
def
11. Let φ : R → R/2πZ be the defined as φ(x) = 2πx + 2πZ.

(1) φ is a homomorphism: φ(x + y) = 2π(x + y) + 2πZ = (2πx + 2πy) + Z =


(2πx + 2πZ) + (2πy + 2πZ) = φ(x) + φ(y). [The third equality comes from the
operation of cosets.]

(2) ker φ = Z: φ(x) = 0 + 2πZ if, and only if, 2πx + 2πZ = 0 + 2πZ, and hence if,
and only if, 2πx ∈ 2πZ, i.e., if, and only if, x ∈ Z. Thus, ker φ = Z.

(3) φ is onto: given y + 2πZ ∈ R/2πZ, we have that y/(2π) ∈ R and φ(y/(2π) + Z) =
y + 2πZ. So, φ is onto.

By the First Isomorphism Theorem, R/Z = R/ ker φ ∼


= R/2πZ.

Homework 10

Section 5.5
1. (b) Let C6 = {1, x, x2 , x3 , x4 , x5 } and φ : C6 → C6 be an automorphism. Thus
φ(x) = xi for some i ∈ {0, . . . , 5}, and so φ(xk ) = xik [since φ is a homomorphism].
If i = 1, then φ = id [the identity function], and so clearly it is an automorphism.
If i = 5, then φ(xk ) = (x5 )k = x−k , and hence φ maps an element to its inverse.
As we have proved in problem 2.3.12 [and since C6 is Abelian], we have that φ is
an automorphism.
If i = 0, then φ(xk ) = 1, and hence φ is not injective [and hence not an automor-
phism].
If i = 2 or i = 4, then φ(1) = φ(x3 ) = 1, and hence φ is not injective [and hence
not an automorphism].
If i = 3, then φ(1) = φ(x2 ) = 1, and hence φ is not injective [and hence not an
automorphism].
Thus, Aut(C6 ) has only two elements [identity and a 7→ a−1 ], and hence Aut(C6 ) ∼
=
C2 .

34
(c) Let C2 = {1, x} and φ : C2 × C2 → C2 × C2 be an automorphism. Since φ(x, 1) ∈
C2 × C2 , we have four possibilities for it: φ(x, 1) = (1, 1), (x, 1), (1, x), (x, x).

(1) If φ(x, 1) = (1, 1), then φ(1, 1) = φ(x, 1) = (1, 1), and hence φ is not an
automorphism. Notice also that, in the same way, we must always have
φ(1, x) 6= (1, 1).
(2) If φ(x, 1) = (x, 1), then we must have φ(1, x) 6= (1, 1), (x, 1) [for φ to be
injective]. So either φ(1, x) = (1, x) or φ(1, x) = (x, x). If the former holds,
we have that φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (x, 1)(1, x) = (x, x),
and therefore φ = id, which is an automorphism. If the latter holds, we have
that φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (x, 1)(x, x) = (1, x). But this
gives an bijection. Also, one can easily verify that this is a homomorphism.
Hence this is a second automorphism.
(3) If φ(x, 1) = (1, x), then we must have φ(1, x) 6= (1, 1), (1, x) [for φ to be in-
jective]. So either φ(1, x) = (x, 1) or φ(1, x) = (x, x). If the former holds,
we have that φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (1, x)(x, 1) = (x, x).
But this gives an bijection and one can easily verify that it is a homomor-
phism. Hence this is a third automorphism. If the latter holds, we have that
φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (1, x)(x, x) = (x, 1). But this
gives an bijection and one can easily verify that this it is a homomorphism.
Hence this is a fourth automorphism.
(4) If φ(x, 1) = (x, x), then we must have φ(1, x) 6= (1, 1), (x, x) [for φ to be
injective]. So either φ(1, x) = (x, 1) or φ(1, x) = (1, x). If the former holds,
we have that φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (x, x)(x, 1) = (1, x).
But this gives an bijection and one can easily verify that it is a homomor-
phism. Hence this is a fifth automorphism. If the latter holds, we have that
φ(x, x) = φ((x, 1)(1, x)) = φ(x, 1)φ(1, x) = (x, x)(1, x) = (x, 1). But this
gives an bijection and one can easily verify that this it is a homomorphism.
Hence this is a sixth automorphism.

So Aut(C2 × C2 ) has 6 automorphisms [listed above]. Observe that each auto-


morphism is just a permutation of the set of three elements {(x, 1), (1, x), (x, x)},
and is naturally isomorphic to S3 .
[Alternatively, one can verify that Aut(C2 × C2 ) is not Abelian, since the second
and third automorphisms do not commute. Hence, since it has only 6 elements,

35
it has to be S3 .]

3. (i) Reflexive: Let s ∈ S. [We need to show that s ∼ s, i.e., for some g ∈ G, we have
s = fg (s).] We have that s = f1G (s) [by the definition of a group action], and
hence s ∼ s.
(ii) Symmetric: Suppose that s ∼ s0 . [We need to show that s0 ∼ s.] Hence, s0 =
fg (s), for some g ∈ G. Thus,

fg−1 (s0 ) = fg−1 (fg (s)) = fg−1 g (s) [property of group actions]

= f1G (s) = s [property of group actions]

Therefore, s = fg−1 (s0 ) and hence s0 ∼ s.


(iii) Transitive: Suppose that s ∼ s0 and s0 ∼ s00 . [We need to prove that s ∼ s00 .] By
definition, s0 = fg (s) and s00 = fg0 (s0 ), for some g, g 0 ∈ G. Then,

s00 = fg0 (s0 ) = fg0 (fg (s)) = fg0 g (s).

Since g 0 g ∈ G, we have that s ∼ s00 .

8. (a) If ~0 = (0, . . . , 0), then O~0 = {~0}, since for all M ∈ GLn (R), we have M · ~0 = ~0.
[Note that when we make this multiplication, we see the vector as a column
vector!]
~ ∈ Rn − {~0}. Since ~v 6= ~0, the set {~v } is linearly independent. By the
So, let ~v , w
first theorem of the hint [http://www.math.utk.edu/∼finotti/f06/hint 8a.
pdf], there is a basis of Rn of the form {~v , ~v2 , . . . , ~vn }. In the same way, since
~ 6= ~0, there is a basis of Rn of the form {w,
w ~ w ~ 2, . . . , w~ n }. Now, by the second
theorem of the hint, there is M ∈ GLn (R) such that M · ~v = w
~ [and M · ~vi = w
~ i,
but we don’t care about those]. Hence, given any two non-zero vectors, there is
an invertible matrix that takes one to the other.
Therefore
def
O~v = {M · ~v : M ∈ GLn (R)}
~ 6= ~0}.
~ ∈ Rn : w
= {w

So, the decomposition of Rn into orbits is given by the orbit of ~0 [which contains
only the vector ~0 itself] and the orbit of a non-zero vector ~v [which contains all
non-zero vectors].

36
(b) We have
def
(GLn (R))e1 = {M ∈ GLn (R) : M · e1 = e1 }.
But      
a11 a12 · · · a1n 1 a11
     
 a21 a22 · · · a2n   0   a21 
· =
     
.. .. . . .   .. ..
. .. 
 

 . .  
 .  
  . 

an1 an2 · · · ann 0 an1
Hence
(GLn (R))e1 = {(aij ) ∈ GLn (R) : (a11 , a21 , . . . , an1 ) = e1 }.

11.(a) We have
" # " # " #−1 " # " # " #
a b 1 0 a b a b 1 0 1 d −b
· · = · ·
c d 0 2 c d c d 0 2 ad − bc
−c a
" # " #
1 a 2b d −b
= ·
ad − bc c 2d −c a
" #
1 ad − 2bc ab
= .
ad − bc −cd 2ad − bc
Hence, ( " # )
1 ad − 2bc ab
OM = : ad − bc 6= 0
ad − bc −cd 2ad − bc
[A better theoretical description, for those who remember linear algebra, is that OM
is the set of all 2 by 2 matrices with eigen values 1 and 2.]
Now, a matrix is in the stabilizer if, and only if,
" # " #
1 ad − 2bc ab 1 0
=
ad − bc −cd 2ad − bc 0 2
Comparing the entries we have:
 
bc


 1 − ad−bc =1 

 bc = 0

 

 ab = 0  ab = 0



 −cd = 0 

 cd = 0
 
 ad + 1 = 2
 
 ad
=1
ad−bc ad−bc

The last equation tells us that a, d 6= 0, hence we must have b = c = 0. Therefore, the
stabilizer of the matrix is the set of invertible diagonal matrices.

37
Section 5.6
1. We have:

g ∈ GaH ⇔ gaH = aH
⇔ ga = ah for some h ∈ H
⇔ g = aha−1 for some h ∈ H
⇔ g ∈ aHa−1 .

Thus,
GaH = aHa−1 .

Homework 11

Section 5.6
2. Let g ∈ G. Since H = hxi, we need to show that gxn g −1 ∈ H, for all n ∈ Z.

By hypothesis, we have xg −1 H = g −1 H. But then, x2 g −1 H = x(xg −1 H) = x(g −1 H) =


g −1 H, and a simple induction shows that xn g −1 H = g −1 H for any positive integer n.
Having that, we can see that it also works for negative powers of x: if n > 0, then
x−n (g −1 H) = x−n (xn g −1 H) = g −1 H.

So, for any n ∈ Z, we have xn g −1 ∈ g −1 H, and thus gxn g −1 ∈ H.

6. (a) Let
def
X = {g ∈ G : fg (s) = s0 }.

[We need to show that X = aGs .]

“⊆”: Let x ∈ X. Then,

fa−1 x (s) = fa−1 (fx (s)) [group action]

= fa−1 (s0 ) [defn. of X]

=s [since fa (s) = s0 ].

Thus, a−1 x ∈ Gs and since x = a(a−1 x), we have x ∈ aGS .

38
“⊇”: Let g ∈ Gs . [Since g is arbitrary in Gs , we have that ag is arbitrary in aGs .]
Then,

fag (s) = fa (fg (s)) [group action]

= fa (s) [since g ∈ Gs ]

= s0 [since fa (s) = s0 ].

Thus, ag ∈ X.

(b)“⊆”: Let g 0 ∈ Gs0 . Then ,

fa−1 g0 a (s) = fa−1 (fg0 (fa (s))) [group action]

= fa−1 (fg0 (s0 )) [since fa (s) = s0 ]

= fa−1 (s0 ) [since g 0 ∈ Gs0 ]

=s [since fa (s) = s0 ].

Thus a−1 g 0 a ∈ Gs and hence g 0 ∈ aGs a−1 .


“⊇”: Let g ∈ Gs . [Since g is arbitrary in Gs , we have that aga−1 is arbitrary in
aGs a−1 .] Then,

faga−1 (s0 ) = fa (fg (fa−1 (s0 ))) [group action]

= fa (fg (s)) [since fa (s) = s0 ]

= fa (s) [since g ∈ Gs ]

= s0 [since fa (s) = s0 ].

Thus aga−1 ∈ G0s .

7. Let fg : S → S be the function associated with the g by the operation of G on S, and


fg0 : S 0 → S 0 be the function associated with the g by the operation of G on S 0 . [In
class I used “fg ” for both, but I think it might be better to emphasize that the actions
are different.] By definition, we have: for all s ∈ S,

φ(fg (s)) = fg0 (φ(s)).

(a) Let g ∈ Gs . Then, by definition, fg (s) = s. Then, fg0 (φ(s)) = φ(fg (s)) = φ(s).
Therefore g ∈ Gφ(s) .

39
(b) We need to show that φ(Os ) = Oφ(s) . But, we have φ(fg (s)) = fg0 (φ(s)), by the
definition of homomorphism of G-sets. Then,

x ∈ φ(Os ) ⇔ x = φ(fg (s)) for some g ∈ G


⇔ x = fg0 (φ(s)) for some g ∈ G
⇔ x ∈ Oφ(s) .

Section 5.7
5. [This one was not assigned, but you use it in number 6, so I decided to post the
solution.]

If G is finite, the problem is very simple: by Lagrange’s Theorem, [G : K] = |G| / |K|,


[G : H] = |G| / |H|, and [H : K] = |H| / |K|. Multiplying we obtain [G : K] = [G :
H][H : K].

Now, we prove it without assuming |G| < ∞. Let [G : H] = m,

G = a1 H ∪˙ . . . ∪˙ am H

[where ai ∈ G], [H : K] = n, and

H = b1 K ∪˙ . . . ∪˙ bn K.

[where bj ∈ H]. We shall prove that

G/K = {(ai bj )K : i ∈ {1, . . . , m} and j ∈ {1, . . . , n}},

with (ai bj )K = (ar bs )K iff i = r and j = s. [This proves that these ai bj K make the
partition of G by cosets of K, and hence [G : K] = mn = [G : H][G : K].]

Let g ∈ G. [We need to prove g ∈ ai bj K for some i and j, i.e., that the ai bj K cover G.]
Since G/H cover G, there is some i ∈ {1, . . . , m} such that g = ai h, for some h ∈ H.
Since H/K covers H, there is some j ∈ {1, . . . , n} such that h = bj k for some k ∈ K.
Thus, g = ai bj k, and so it is in ai bj K.

Now, suppose that (ai bj )K = (ar bs )K. Since bj K, br K ⊆ H, we have that ai H ∩ar H 6=
∅ [since ai (bj K) = ar (bs K) ⊆ ai H ∩ ar H]. Hence, by assumption, i = r. Hence, we
must have bj K = br K. But, by assumption, this implies that j = r.

40
6. (a) By Proposition 5.7.5 (pg. 181), [H : H ∩ K] ≤ [G : K] < ∞. Since also [G : H]
is finite, we have that [G : H][H : H ∩ K] is finite. By Problem 5.7.5, since we
clearly have H ∩ K < H < G, we have [G : H ∩ K] = [G : H][H : H ∩ K], and
hence [G : H ∩ K] is finite.
def def def
(b) Let G = S3 = {1, x, x2 , y, xy.x2 y}, H = hyi = {1, y} and K = hxyi = {1, xy}.
Hence, [G : K] = |G| / |K| = 3 and [H : H ∩ K] = |H| / |H ∩ K| = 2. Clearly,
2 - 3.

Homework 12

Section 6.1
1. Yes. We have:

f1G (x) = x1−1


G = x,

and

fg ◦ fh (x) = fg (xh−1 )
= xh−1 g −1
= x(gh)−1
= fgh (x).

2. The orbits of H are the left cosets:

OH = {fg (H) : g ∈ G}
= {gH : g ∈ G}.

4. We have that S can be written as a disjoint union of orbits, say:

S = Ox1 ∪˙ · · · ∪˙ Oxk .

Hence, [since the union is disjoint],


k
X
|S| = |Oxi | . (2)
i=1

41
We need to show that |Oxi | = 1 for some i ∈ {1, . . . , k}. [If that is the case, then
|Gxi | = |G| / |Oxi | = |G|, and hence G = Gxi .] Suppose then, to obtain a contradiction,
that |Oxi | > 1 for all i. Then, since |Oxi | | |G| = pr [for some r ∈ N], we have that
|Oxi | ≥ p. Hence p divides all terms in the left hand side of equation (2) above, and
so p divides the right hand side. Thus, p must also divide the left hand side. But, by
assumption, p does not divide |S|, yielding the desired contradiction.

6. The third equation cannot be the class equation, since 3, 4 - 10.

The first also cannot be, since it says that there are 3 elements with orbits of size one.
But, as seen in class, the orbit [when G acts on itself by conjugation] has only one
element if, and only if, this element in the center. So, the first equation says there
are 3 elements in the center. Since the center is a subgroup, by Lagrange’s Theorem,
its order has to divide the order of the group. Since 3 - 10, the center cannot have 3
elements.

The last equation also cannot happen. We know, as above, that the center has only
2 elements. [This seems at first fine, since 2 | 10.] But, we have elements with orbits
of size 2. By the counting formula |G| = |Os | |Gs |, we have that these elements have
stabilizers of order 5. But certainly the center is contained in the stabilizer of all
elements. [Remember that G is acting on itself by conjugation!] Thus, the center is a
subgroup of the stabilizer and hence, by Lagrange’s Theorem, its order should divide
the order of the stabilizer. But 2 - 5.

The second equation can happen. [After I define the dihedral group D2n , you will be
able to verify that D10 has a class equation precisely as the second equation.]

10. (a) We have:      


a b c 1 0 0 a 2b 3c
     
 d e f  ·  0 2 0  =  d 2e 3f 
     
g h i 0 0 3 g 2h 3i
and      
1 0 0 a b c a b c
     
 0 2 0  ·  d e f  =  2d 2e 2f 
     
0 0 3 g h i 3g 3h 3i

42
Thus, for this matrix to be in the centralizer, we must have:

a = a, 2b = b, 3c = c,
d = 2d, 2e = 2e, 3f = 2f,
g = 3g, 2h = 3h, 3i = 3i.

Hence, b = c = d = f = g = h = 0 and a, b, and c can be any elements of R×


[since we need abc 6= 0 for the matrix to be invertible]. Hence the centralizer of
the given matrix is the subgroup of invertible diagonal matrices.

(e) We have:      
a b c 1 0 0 a c b
     
 d e f · 0 0 1 = d f e 
     
g h i 0 1 0 g i h
and      
1 0 0 a b c a b c
     
 0 0 1 · d e f = g h i 
     
0 1 0 g h i d e f
Thus, for this matrix to be in the centralizer, we must have:

a = a, c = b, b = c,
d = g, f = h, e = i,
g = d, i = e, h = f.

Hence, a matrix in the centralizer must have the form:


 
a b b
 
 d e f ,
 
d f e

and for this matrix to be invertible, we need its determinant to be no zero, i.e.,

(e − f )(a(e + f ) − 2bd) 6= 0.

12. Assume that N is not contained in Z(G). [We will derive a contradiction.] Since the
order of N is prime, then it is cyclic generated by any of its non-identity elements.
So, Z(G) ∩ N = {1G }, since any non-identity element of the intersection is an element

43
of N that generates N , and since the intersection is a subgroup, it would contain all
powers of this element, and hence it would contain all of N .

Let x ∈ N −{1G }. Then, since N /G the orbit of x when G acts on itself by conjugation
is contained in N [since gN g −1 = N for all g ∈ G].

On the other hand, the orbit cannot contain only one element, since this would imply
that x ∈ Z(G), and we are assuming that N is not contained in Z(G) [and hence
Z(G) ∩ N = {1G }, as explained above].

It cannot have three elements either, since then the orbit of the remaining element of
N [i.e., not 1G and not the other three elements in the orbit of x] would then have
to have a single element in its orbit [due to normality and the fact the the orbits are
disjoint].

Therefore, the orbit of x has either 2 or 4 elements. But the order of the orbit divides
the order of the group [by the counting formula], but since |G| is odd, neither 2 nor 4
can divide it.

Thus, the assumption that N is not contained in G lead us to a contradiction, and


hence it must be false.

Section 6.3
4. Just observe that if we conjugate gH by g −1 , we obtain g −1 (gH)(g −1 )−1 = Hg.

5. No. If G is Abelian, then the stabilizer of any U ⊆ G is G itself [when G acts on itself
by conjugation].

[For example, let G = C3 = {1, x, x2 } and U = {1, x}. Then GU = G.]


def
7. The statement is false. Let G = S3 = {1, x, x2 , y, xy, x2 y} and H = hyi = {1, y}.
Then H < N (H), and hence 2 | |N (H)|. But |N (H)| | |S3 | = 6, and thus |N (H)| is
either 2 or 6.

Since x2 yx−2 = x2 yx = x2 x2 y = xy 6∈ H, we have that N (H) 6= S3 [since x2 6∈


N (H)]. Therefore |N (H)| = 2 [since it cannot be 6] and thus N (H) = H [since
H ⊆ N (H)]. But, as seen above, H, and hence N (H), is not a normal subgroup of S3
[since x2 Hx−2 6= H].

44
13. (a) Let H = {1G , x}. As always, the orbit of 1G , when G acts on itself by conjugation,
has only 1G itself. [1G is always in the center.] But for all g ∈ G, we have that
gHg −1 = H [since H / G], and thus gxg−1 is either 1G or x. It cannot be 1g ,
since then x would be in the orbit of 1G . Hence, gxg −1 = x for all g ∈ G. Thus,
x ∈ Z(G) and therefore H < Z(G).
(b) [Note that we did not use assume that G was finite in part (a), so part (b) does
not generalize part (a).]
As in part (a) [or 6.1.12 above], we have that an orbit of an element in a normal
group is contained in this group, and the orbit of 1G contains only 1G .
Let x ∈ H − {1G }. Then, Ox ⊆ H, but since 1G 6∈ Ox [since 1G is in its own orbit
by itself], |Ox | < |H| = p. Also, by the counting formula, |Ox | | |G|.
But, by hypothesis, p is the smallest prime divisor of |G|, and hence, it is the
smallest divisor of |G| that is greater than one. Since |Ox | is a divisor of |G| that
is less than p, it must be 1.
Therefore, x ∈ Z(G), and hence [since x was an arbitrary element of H − {1G }]
H ⊆ Z(G).

Homework 13

Section 6.4
1. Let sp [here and in the problems to come] denote the number of Sylow p-subgroups. We
have that 20 = 22 5. Then, by the Sylow Third Theorem, s5 divides 4 [so s5 ∈ {1, 2, 4}]
and s5 ≡ 1 (mod 5). Thus s5 = 1.
Now, any element of order 5 would generate a subgroup of order 5, which is, in this
case, a Sylow 5-subgroup. Since there is only one of those [by the above], only the 4
generators of this group have order 5.

2. Suppose that p = q, and hence |G| = p2 . Then, G ∼


= Cp × Cp or G ∼
= Cp2 [by Corollary
6.1.14]. In either case, it is Abelian, and both have non-trivial subgroups. [For the
former, we have Cp × {1}, while for the latter we have hxp i, where x is a generator of
Cp2 .]
Suppose then that p 6= q and [without loss of generality] that p < q. By the Sylow
Third Theorem, we have sq divides p [and so sq ∈ {1, p}], and sq ≡ 1 (mod q). Since

45
1 < p < q, we have that p 6≡ 1 (mod q), and hence sq = 1. Thus, G is not simple [since
the Sylow q-subgroup is normal].

3. [Case 1:] If p = q, then |G| = p3 . By Proposition 6.1.10, the center of G is order


greater than one. The center is always normal, so if G is not Abelian [i.e., Z(G) 6= G],
then Z(G) is a proper normal subgroup. If G is Abelian, every subgroup is normal,
and since |G| is not prime, G has a proper subgroup [e.g., by Corollary 6.4.3, there is
a cyclic subgroup of order p].

[Case 2:] If p > q, then sp cannot be equal to q, since q 6≡ 1 (mod p). But, since sp | q,
we must have that sp = 1, and hence G is not simple [since the Sylow p-subgroup is
normal].

[Case 3:] If q > p, then sq cannot be equal to p, since p 6≡ 1 (mod q). But, since
sq | p2 , we must have that sq = 1 or sq = p2 . In the former case, we see that the Sylow
q-subgroup is normal. Suppose then, that sq = p2 . This gives us p2 (q − 1) elements
of order q. [Note that the fact that the Sylow q-subgroups has prime order, not only
power of a prime, is essential to this counting argument!] This means that we have
only p2 elements of order different from q. Therefore, there can only be one subgroup
of order p2 [a Sylow p-subgroup], which is then normal.

13. If a = 1 and e > 1, we use an argument similar to the one in the first paragraph
of Problem 6.4.3 above: it has proper center [which is normal], or, since e > 1, it is
Abelian with a proper subgroup [or order p, for example].

Suppose then that a > 1. Since a < p, we have p - a. Then, by the Third Sylow
Theorem, that sp ≡ 1 (mod p) and sp | a. But since a < p, the second assertion tells
us that sp < p. But since sp ≡ 1 (mod p), we can only have sp = 1. Thus, the Sylow
p-subgroup is normal.

15. We have s3 ∈ {1, 11} and s11 ∈ {1, 3}. Since s3 ≡ 1 (mod 3), we must have s3 = 1,
and since s11 ≡ 1 (mod 11), we must have s11 = 1. So, we have normal subgroups, say
H and K of orders 3 and 11 respectively. Since H ∩ K = {1} [since they have relative
prime orders] and |H| · |K| = |G|, [and since H, K / G] we have, by Proposition 2.8.6,
that G ∼
=H ×K ∼ = C3 × C11 ∼= C33 .
Hence, the only group of order 33 [up to isomorphism] is C33 .

46
Homework 14

Section 5.3
1.

x2 yx−1 y −1 x3 y 3 = x2 yx−1 yx3 y [since y 2 = 1]


= x3 yyx3 y [since yx−1 = xy]
= x6 y [since y 2 = 1]

2. [We are using D2n for what the book uses Dn , i.e., for the dihedral group of order 2n.]
Remember:
D8 = {1, x, x2 , x3 , y, xy, x2 y, x3 y},

with x4 = 1, y 2 = 1, and yxn = x−n y. [The last is a consequence of yx = x−1 y.] We


have,

H0 = {1}, H1 = {1, x2 }, H2 = {1, y}, H3 = {1, xy},

H4 = {1, x2 y}, H5 = {1, x3 y}, H6 = {1, x, x2 , x3 },

H7 = {1, x2 , y, x2 y}, H8 = {1, x2 , xy, x3 y}, H9 = D8 .

Clearly, H0 , H9 / D8 . Also, since [D8 : H6 ] = [D8 : H7 ] = [D8 : H8 ] = 2, we have that


H6 , H7 , H8 / D8 .

We also have:
xi y · xj · xi y = xi yxi+j y = xi−(i+j) y 2 = x−j .

Hence, H1 / D8 [since x−2 = x2 ].

Moreover,
xi y · xj y · xi y = xi−j y 2 xi y = x2i−j y.

Hence, H3 and H5 are not normal in D8 [by taking i = 0 and j = 1, 3]. Neither are H2
and H4 [by taking i = 1 and j = 0, 2].

Thus, H0 , H1 , H6 , H7 , H8 and H9 are all the normal subgroups of D8 .

47
4. Remember:
D20 = {1, x, . . . , x9 , y, xy, . . . , x9 y},

with x10 = 1, y 2 = 1, and yxn = x−n y.

(a) The cosets are:

H = {1, x5 },
xH = {x, x6 },
x2 H = {x2 , x7 }
x3 H = {x3 , x8 }
x4 H = {x4 , x9 }
yH = {y, yx5 } = {y, x5 y},
xyH = {xy, xyx5 } = {y, x6 y},
x2 yH = {x2 y, x2 yx5 } = {x2 y, x7 y}.
x3 yH = {x3 y, x3 yx5 } = {x3 y, x8 y}.
x4 yH = {x4 y, x4 yx5 } = {x4 y, x9 y}.

(b) Since,

xi y · x5 · (xi y)−1 = xi y · x5 · xi y = xi x−5 yxi y = xi x5 x−i y 2 = x5 ,

and xi x5 x−i = x5 , we have that H /D20 . Hence we have the quotient group D20 /H
given by the cosets above. Clearly, xH and yH generate D20 /H, the order of xH
is 5, and the order of yH is two. Moreover,

yH · xH = (yx)H = (x9 y)H = x9 H · yH = x4 H · yH.

Hence, yH · xH = (xH)−1 · yH. These four properties define D10 , and hence
D20 /H ∼
= D10 .
(c) Yes. Let
D10 = {1, a, . . . , a4 , b, ab, . . . , a4 b},

where a5 = 1, b2 = 1, and ban = a−n b.


def def
Let X = (a, x5 ) and Y = (b, 1). [Hence, X, Y ∈ D10 × H. We need to show that
X and Y generate D10 × H and have the same properties as x and y in D20 .]

48
Then, X 10 = (a10 , x50 ) = (1, 1). So, the order of X divides 10. Since X 6= (1, 1),
X 2 = (a2 , 1) 6= (1, 1), and X 5 = (1, x5 ) 6= (1, 1), the order of X is 10.
Also, Y 6= (1, 1), but Y 2 = (1, 1), and so the order of Y is 2.
Moreover, Y X = (ba, x5 ) = (a−1 b, x5 ) = (a−1 , x5 )(b, 1) = (a, x5 )−1 (b, 1) = X −1 Y .
Finally, hX, Y i = D10 × H, since
 
X i , if i is even; X i , if i is odd;
i i 5
(a , 1) = (a , x ) =
X i+5 , if i is odd. X i+5 , if i is even.

 
X i Y, if i is even; X i Y, if i is odd;
(ai b, 1) = (ai b, x5 ) =
X i+5 Y, if i is odd. X i+5 Y, if i is even.

[Another way to see this last part, which is simpler, but less explicit, is to observe
that hXi ∩ hY i = 1, hXi / D20 [since its index is 2], and then use Proposition 2.8.6
to conclude that hXi hY i = D10 × H, and hence, X and Y generate G.]
[Also note that the actual isomorphism would be given by φ(X i Y j ) = xi y j .]

Section 6.7
1. False! The product of two cyclic groups is Abelian, while the free group on two gener-
ators is not.

Section 6.8
1. Let a, b ∈ G, H = ha, bi, and K = hbab2 , bab3 i. Since clearly bab2 , bab3 ∈ H [since they
are products of powers of a and b], we have that K ⊆ H. [Of course, since products of
powers of bab2 and bab3 will be products of power of a and b!]
So, we need to show that H ⊆ K. We have (bab2 )−1 (bab3 ) = (b−2 a−1 b−1 )(bab3 ) = b.
Hence b ∈ K. Now, since b, bab2 ∈ K, we have that b−1 (bab2 )b−2 = a is also in K.
Thus, a, b ∈ K and therefore H ⊆ K.
def
2. Let H be the smallest normal subgroup of G containing S [where S ⊆ G] and K =
h{gsg −1 : g ∈ G, s ∈ S}i.
Since H / G, for any g ∈ G and h ∈ H, ghg −1 ∈ H. Since S ⊆ H, for all s ∈ S,
gsg −1 ∈ H. Thus {gsg −1 : g ∈ G, s ∈ S} ⊆ H and therefore [since H is a group]
K ⊆ H.

49
Clearly S ⊆ K [since 1 ∈ G]. Let x ∈ G and k ∈ K. By definition, k is a [finite]
product of powers of elements of {gsg −1 : g ∈ G, s ∈ S}. Since (gsg −1 )n = gsn g −1 ,
we can write k as
k = g1 sn1 1 g1−1 · g2 sn2 2 g2−1 · · · · · gm snmm gm
−1
.

Therefore,

xkx−1 = x(g1 sn1 1 g1−1 · g2 sn2 2 g2−1 · · · · · gm snmm gm


−1 −1
)x
= x · g1 sn1 1 g1−1 · 1 · g2 sn2 2 g2−1 · 1 · · · · · 1 · gm snmm gm
−1
· x−1
= x · g1 sn1 1 g1−1 · x−1 x · g2 sn2 2 g2−1 · x−1 x · · · · · x−1 x · gm snmm gm
−1
· x−1
= (xg1 )sn1 1 (xg1 )−1 · (xg2 )sn2 2 (xg2−1 ) · · · · · (xgm )snmm (xgm )−1

Since xgi ∈ G for all i, we have that (xgi )si (xgi )−1 ∈ {gsg −1 : g ∈ G, s ∈ S}, and
thus xkx−1 ∈ K [since its a product of powers of elements in {gsg −1 : g ∈ G, s ∈ S}].
So, K / G. Since also S ⊆ K ⊆ H [by the previous paragraph], and since the H is the
smallest subgroup of G that contains S and is normal, we must have K = H [otherwise
K would be smaller than H, normal, and containing S].

6. [Note: There is a typo in the statement!! It should say “N and G/N are cyclic”, not
“G and G/N are cyclic”, since if G is cyclic, then G/N is automatically cyclic – proved
in class – and G is generated by a single element. So, the problem would be trivial...]
Let N = hai, and G/N = hbN i. Then, we claim that G = ha, bi. Indeed, let g ∈ G.
Then, since the left cosets of n make a partition of G, there is a coset containing g.
Since G/N = hbN i, there exists a power n ∈ Z such that g ∈ bn N . Then, since
N = hai, there exists m ∈ Z such that g = bn am . Thus g ∈ ha, bi. Since g was
arbitrary, G = ha, bi.

16. True. Suppose that G = {a1 , . . . , an }. Also, note that ai aj = ak(i,j) for some k(i, j) ∈
{1, . . . , n}, since ai aj ∈ G. [So, for each pair i, j ∈ {1, . . . , n}, we have k(i, j) ∈
{1, . . . , n}. Basically, this function k(i, j) gives us the multiplication table.] Then,


G = a1 , . . . , an : ai aj = ak(i,j) , for i, j ∈ {1, . . . , n} .

[So, basically we have that all elements of G generate G – of course – and the relations
are given to describe how every product should be computed, giving us the entire
multiplication table. We don’t need all those generators and relations, but they do
give us the group.]

50
Extra Credit
1. Prove that if |G| = 24, then G is not simple.

Proof. By the Sylow Theorems we have s2 ∈ {1, 3} and s3 ∈ {1, 4}. If either one is
equal to one, then G is not simple. So assume that s2 = 3 and s3 = 4. The latter
tells us that there are 4 · (3 − 1) = 8 elements of order 3 [since the Sylow 3-subgroups
cannot intercept since they have prime order].

Let H1 , H2 , and H3 be the three Sylow 2-subgroups [of order 8].

We have, by the Inclusion-Exclusion Principle,

|H1 ∪ H2 ∪ H3 | = |H1 |+|H2 |+|H3 |−(|H1 ∩ H2 |+|H1 ∩ H3 |+|H2 ∩ H3 |)+|H1 ∩ H2 ∩ H3 | .

Observe that since all those are 2-subgroups, we have the the orders have to be 1, 2, 4
or 8. Suppose that |Hi ∩ Hj | ≤ 2, for i 6= j. Then,

|H1 ∪ H2 ∪ H3 | ≥ 24 − (2 + 2 + 2) + 1 = 19,

But, since G has only 24 elements, and has 8 elements of order 3 [which are not in
H1 ∪ H2 ∪ H3 ], we would get a contradiction [at least 8 + 19 = 27 elements in G].

Thus, we can assume that |H1 ∩ H2 | = 4. [Note that this order cannot be 8 since
def
H1 6= H2 .] Therefore, K = H1 ∩H2 /H1 , H2 [since K has index 2]. So, H1 , H2 ⊆ N (K).
Hence
|N (K)| ≥ |H1 ∩ H2 | = |H1 | + |H2 | − |H1 ∩ H2 | = 12.

If |N (K)| = 12, then N (K) / G, since it has index 2. Then, N (K) is a proper normal
subgroup.

If |N (K)| > 12, we must have that N (K) = 24 [since |N (K)| divides |G| = 24].
Therefore, N (K) = G, which means that K / G. [And K is a proper normal subgroup.]

2. Prove that if |G| = 36, then G is not simple.

Proof. By the Sylow Theorems, we have s2 ∈ {1, 3, 9} and s3 ∈ {1, 4}. If s3 = 1, then
it has a normal Sylow 3-subgroup. So, assume s3 = 4.

51
Case 1: Assume that |H ∩ H 0 | = 1 for any two distinct Sylow 3-subgroups H, H 0 .
Then, the number of elements in all four Sylow 3-subgroups is 4 · (9 − 1) + 1 = 33.
[The “+1” takes into account the identity, which was excluded before.] Then, we have
only 3 elements left. Since we know we have a Sylow 2-subgroup with 4 elements, all
of which, expect the identity, haven’t been counted with the 33 elements coming from
the Sylow 3-subgroup, these three elements left, together with the identity, form the
unique [there are no more elements left to form others] Sylow 2-subgroup of G, which
then must be normal.

Case 2: Assume that |H ∩ H 0 | > 1 for some pair of distinct Sylow 3-subgroups H,
def
H 0 . Since |H| = |H 0 | = 32 and they are distinct, we must have that K = H ∩ H 0 has
order 3. Since H and H 0 have order 32 , by Proposition 6.1.13, they are Abelian. Thus
K / H, H 0 . So, H, H 0 ⊆ N (K) and |N (H)| ≥ |H ∪ H 0 | = 2 · 9 − 3 = 15. Since |N (H)|
divides |G| = 36, we have that N (H) has order 18 or 36.

If |N (K)| = 18, then N (K) / G [since it has index 2].

If |N (K)| = 36, then N (K) = G, and so K / G.

Homework 15

Section 10.1
6. It suffices to show that α, β ∈ Q[γ] and γ ∈ Q[α, β], since if α, β ∈ Q[γ] [and since
Q ⊆ Q[γ]], we would have that Q[α, β] ⊆ Q[γ], by the minimality of Q[α, β]. The
analogous argument for γ ∈ Q[α, β] would give us the other inclusion.

The latter is trivial, since α, β ∈ Q[α, β], and it is a ring [hence closed under addition],
we have that γ = α + β ∈ Q[α, β].

Now, we need to show that α, β ∈ Q[γ]. We have that γ 2 = (α + β)2 = 5 + 2 6 ∈ Q[γ]

[since it’s closed under multiplication]. So, 6 ∈ Q[γ] [since −5, 1/2 ∈ Q[γ]]. Hence,
√ √ √ √
6γ = 2 3 + 3 2 ∈ Q[γ]. Subtracting 2γ, we have that 2 = β ∈ Q[γ]. So,
γ − β = α ∈ Q[γ].

8. (a) (i) Identity: 1 = 1/1, and 1 is not divisible by 3. So, 1 ∈ S.

52
(ii) Add. subgroup: If a/b and c/d are such that neither b nor d is divisible by
3, then bd is also not divisible by 3 [since 3 is prime]. Hence, a/b − c/d =
(ad − cb)/(bd) ∈ S.
(iii) Closed under mult.: If a/b and c/d are such that neither b nor d is divisible
by 3, then bd is also not divisible by 3 [since 3 is prime]. Hence, (a/b)(c/d) =
(ac)/(bd) ∈ S.

(b) (i) Identity: The identity matrix is in S: just take a = 1 and b = 0.


(ii) Add. subgroup:
" # " # " #
a b c d (a − c) (b − d)
− =
−b a −d c −(b − d) (a − c).

(iii) Closed under mult.:


" # " # " #
a b c d (ac − bd) (ad + bc)
· =
−b a −d c −(ad + bc) (ac − bd).

9. (a) (i) Zero: The empty set ∅ is the zero: A ∪ ∅ = ∅ ∪ A = A.


(ii) Add. Assoc.: Unions are associative. [It is not difficult to prove, but I’ll
assume you know some set theory.]
(iii) Add. Inverse: FAILS! If A 6= ∅, there is no set B such that A ∪ B = ∅.
(iv) Add Comm.: We have A ∪ B = B ∪ A.
(v) Mult. Assoc.: Intersections are associative.
(vi) Distributive Law: We have A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C). [This is also
easy to prove.]
Note that ∩ is commutative and there is a multiplicative identity, namely U .

(b) (i) Zero: The constant function f ≡ 0 is the zero: [f + g](x) = f (x) + g(x) =
0 + g(x) = g(x).
(ii) Add. Assoc.: Since sums in R are associative, the sum of functions also are.
(iii) Add. Inverse: Given f (x), the function g(x) = −f (x) is its additive inverse.
(iv) Add Comm.: Since the sum in R is commutative, the sum of functions is
commutative.
(v) Mult. Assoc.: Composition of functions is associative.

53
(v) Distributive Law: FAILS! Let f (x) = x2 , g(x) = h(x) = x. Then

[f ◦ (g + h)](x) = f ([g + h](x)) = f (x + x) = (2x)2 = 4x2 ,

while,

[(f ◦ g) + (f ◦ h)](x) = f (g(x)) + f (h(x)) = f (x) + f (x) = 2x2 .

So, f ◦ (g + h) 6= f ◦ g + f ◦ h.
Note that ◦ is not commutative, but there is an multiplicative identity, namely
id(x) = x.

11. I’ve stated in class that in (Z/nZ)× = {ā : gcd(a, n) = 1}. But, since I did not prove
it, you can do it by “brute force”: given ā ∈ Z/nZ, compute all n multiples of it [i.e.,
m̄ · ā, for all m̄ ∈ Z/nZ] and see if you ever get 1̄. Below, I will just list the elements
that are in (Z/nZ)× .

¯
(a) (Z/12Z)× = {1̄, 5̄, 7̄, 11}. Note that each one of those is its own multiplicative
inverse.

(b) (Z/7Z)× = {1̄, 2̄, . . . , 6̄}:

1̄ · 1̄ = 1̄, 2̄ · 4̄ = 1̄,
3̄ · 5̄ = 1̄, 6̄ · 6̄ = 1̄.

13. In any ring with identity [even if it is non-commutative], we have that for all x ∈ R,

(1 + x)(1 − x + x2 − · · · + (−1)k xk ) = 1 + (−1)k xk+1 =


(1 − x + x2 − · · · + (−1)k xk )(1 + x).

[One can easily prove this by induction.] Then, if xn = 0, for some n ∈ N, then the
equation above with k = n shows that

(1 + x)−1 = (1 − x + x2 − · · · + (−1)n xn )

54
Solutions to (Selected) HW Problems

Math 456 – Spring 2007

April 26, 2007

Please read: I will try to post here a few solutions. The new solutions will be added to
this same file. They might come with no explanation, just the “answer”. If yours do not
match mine, you can try to figure out again. (Also, read the disclaimer below!) You can
come to office hours if you want explanations for the unexplained answers. Be careful that
just because our “answers” were the same, it doesn’t mean that you solved the problem
correctly (it might have been a “fortunate” coincidence), and in the exams what matters is
the solution itself. I will do my best to post somewhat detailed solutions, though.

Disclaimer: I will have to put these solutions together rather quickly, so they are subject
to typos and conceptual mistakes. (I expect you to be a lot more careful when doing your
HW than I when preparing these.) You can contact me if you think that there is something
wrong and I will fix the file if you are correct.

Homework 1

Section 10.3
1. First, observe that since φ is bijective, it has an inverse function which is also bijective.
[You should have seen this in Math 300.] Thus, we just need to show that φ−1 is a
homomorphism of rings.

(i) Since φ(1R ) = 1R0 [since φ is a homomorphism], we have that φ−1 (φ(1R )) =
φ−1 (1R ), and so φ−1 (1R0 ) = 1R .

(ii) Let a0 , b0 ∈ R0 . Since φ is onto, we have that there are a, b ∈ R such that a0 = φ(a)

1
and b0 = φ(b) [and so a = φ−1 (a0 ) and b = φ−1 (b0 )]. Therefore:

φ−1 (a0 + b0 ) = φ−1 (φ(a) + φ(b))


= φ−1 (φ(a + b)) [since φ is a homom.]

=a+b
= φ−1 (a) + φ−1 (b).

(iii) The product is just like the sum. [Just change “+” by “·” above.]

2. Let a ∈ R× [remember that R× denotes the set of units of R] such that a ∈ I. Then
there is a−1 ∈ R such that a · a−1 = 1R . Since a ∈ I, we have that 1R ∈ I [by the
previous equality]. Hence, for all r ∈ R, r = r · 1R ∈ I, which implies that I = R.

4. “⊇”: For all f ∈ Z[x], we have that (2x)f = 2(xf ) = x(2f ). Hence (2x)f is in both
(2) and (x). So, (2x) ⊆ (2) ∩ (x).

“⊆”: Let f ∈ (2) ∩ (x). Since f ∈ (2), we can write f = 2g1 , where g1 ∈ Z[x]. Since
f ∈ (x), f = xg2 , for some g2 ∈ Z[x]. Hence, 2g1 = xg2 . Then f (0) = 0,
which implies that g1 (0) = 0, and hence g1 has no constant term, i.e., g1 ∈ (x).
Therefore, f = 2g1 = 2xg3 , for some g3 ∈ Z[x], i.e., f ∈ (2x).
def def
6. No. Let f (x) = 2 [a constant polynomial] and g(x) = x. Then f is in the set, but
f (x)g(x) = 2x is not, since 22 - 2.

[Note that the set is closed under addition though.]

14. By the “Substitution Principle” (pg. 353) we have that if φ : R → R0 is an isomorphism


[and hence a homomorphism], then the induced map on polynomials [obtained by
applying φ to coefficients], say φ̃, is a homomorphism. Thus, it only remains to show
that the map is bijective.

Suppose that !
n
X
φ̃ ai xi = 0, ai ∈ R.
i=0
Pn
Then, by definition of φ̃, we have that i=0 φ(ai )xi = 0, i.e., φ(ai ) = 0 for i =
0, 1, . . . , n. Since φ is injective [since it is an isomorphism], we know that ker φ = {0},
and so ai = 0 for i = 0, 1, . . . , n. Therefore, ker φ̃ = {0}, and hence φ̃ is injective.

2
Now, let
n
X
bi xi ∈ R0 [x].
i=0

Since φ is onto, there are a0 , a1 , . . . , an ∈ R such that φ(ai ) = bi . Then,


n
! n n
X X X
φ̃ ai xi = φ(ai )xi = bi x i .
i=0 i=0 i=0

Thus, φ̃ is onto.

20. Let φ be an automorphism [i.e., an isomorphism from a ring onto itself] of Z[x]. Then,
since φ(1) = 1, for all n ∈ Z we have φ(n) = n. [Just as done in class.]

Let φ(x) = h(x). Then, for all f (x) ∈ Z[x], we have deg φ(f (x)) = deg f (h(x)) =
deg f · degh. But, since φ is onto, there must be some f (x) such that φ(f (x)) = x.
Thus, 1 = deg φ(f ) = deg f · deg, and so deg f = deg h = 1. So, f (x) = ax + b and
h(x) = cx + d, with a, b, c, d ∈ Z. Then,

φ(ax + b) = aφ(x) + b = a(cx + d) + b = (ac)x + (ad + b) = x.

So, ac = 1 and ad + b = 0. For the former, we need a = c = 1 or a = c = −1. Thus,


φ(x) = ±x + d.

We now prove that for all d ∈ Z, φ : Z[x] → Z[x] defined by φ(±x + d) is an automor-
phism.

By the Substitution Principle, φ is a homomorphism. So, we just need to show that φ


is a bijection. In order to do that it suffices to show that there is an inverse function,
def
i.e., ψ : Z[x] → Z[x] such that φ ◦ φ = ψ ◦ φ = id. So, just let ψ(f ) = f (±x − d).

Homework 2

Section 10.3
8.(a) We claim that ker φ = (x, y). Clearly, x, y ∈ ker φ, and so (x, y) ⊆ ker φ. [Remember
def
that (x, y) = {xf1 (x, y) + yf2 (x, y) : f1 , f2 ∈ R[x, y]}.]

Conversely, if f (x, y) ∈ ker φ, then f (x, y) has no constant term. Collecting all terms
that have an x in f (x, y), we can write f (x, y) = xf1 (x, y) + f2 (y). [So, all the terms
that have x are in xf1 (x, y) and the remaining, which have no terms in x at all,

3
form f2 (y).] Since f (x, y) has no constant term [and neither does xf1 (x, y)], f2 (y)
also has no constant term, and hence we can write it as f2 (y) = yf3 (y). Therefore,
f (x, y) = xf1 (x, y) + yf3 (y) ∈ (x, y).
def
9. We claim that ker φ = (x2 − 2x − 1). Again, clearly f (x) = x2 − 2x − 1 ∈ ker φ, and
so (f ) ⊆ ker φ.

Let g(x) ∈ ker φ. Then, since f is monic, we can divide g by f , obtaining g(x) =
f (x) · q(x) + r(x), where deg r(x) ≤ 1. Thus, r(x) = g(x) − f (x) · q(x), and so r ∈ ker φ.
Since deg r ≤ 1, we can write r(x) = ax + b, for some a, b ∈ Z, and since r ∈ ker φ,
√ √ √
r(1 + 2) = a(1 + 2) + b = 0. Thus, if a 6= 0, we have that 2 = −(a + b)/a ∈ Q,

which is not true. Therefore, a = 0 and since r(1 + 2) = b = 0, we have that r = 0,
i.e., g(x) = f (x) · q(x) ∈ (f (x)).
def
19. Let σ : R → R be this Frobenius map, i.e., σ(x) = xp .

Clearly, in any ring we have that σ(1R ) = 1pR = 1R , and in any commutative ring
[which we assume R is], we have that σ(ab) = (ab)p = ap bp = σ(a)σ(b).

So, we must show that σ(a + b) = σ(a) + σ(b), i.e., that (a + b)p = ap + bp , if the ring
R has characteristic p. Since R is commutative, we can use the Binomial Theorem to
expand (a + b)p :
p   p  
" #
p
X p p−i i p
X p p−i i
(a + b) = a b =a + a b + bp .
i=0
i i=0
i

Note that the binomial coefficients can be seen as elements of R by the standard
identification of elements in Z to elements in a ring [with unit] R.

So, to finish the proof it suffices to prove that, in R:


 
p p!
= = 0, for i = 1, . . . , (p − 1).
i i! · (p − i)!

Note that pi ∈ Z, so every term of the denominator is canceled by some term in the


numerator. But, since [for 0 < i < p] all terms in the numerator [i.e., i, (i − 1), (i −
2), . . . , 1, (p − i), (p − i − 1), . . . , 1] are less than p, and since p is prime [this is essential
here!!!], we have that p is not canceled by any term in the denominator. [You can think
about the decomposition into primes of the denominator: it is a product of numbers
less than p, so p cannot show up in its decomposition.] So, if 0 < i < p, then the

4
p

integer i
is a multiple of p. Since the characteristic of R is p, this integers becomes
0 in R.
Thus, σ(a + b) = (a + b)p = ap + bp = σ(a) + σ(b).

22. Let n be the characteristic of the field F . If n = n1 n2 , with n1 , n2 > 1, then n1 n2 = 0.


Since a field has no zero divisors, we have that either n1 or n2 is zero. Without loss of
generality, we may assume that n1 = 0. Since n2 > 1, we have that 0 < n1 < n. But
this contradicts the fact that n is the smallest positive integers that is equal to zero in
F [by the definition of characteristic].
def def def
29. Let R = Z, I = (2) and J = (3). Then,

I ∪ J = {0, ±2, ±3, ±4, ±6, ±8, ±9, . . . }.

So, 3, −2 ∈ (I ∪ J), but 3 + (−2) = 1 6∈ (I ∪ J). So, (I ∪ J) is not an ideal. [We


had seen last semester that union of subgroups are not subgroups. This is exactly the
same.]
To prove that (I + J) is an ideal, note that

(i) Since I, J 6= ∅, I + J 6= ∅.
(ii) If x1 , x2 ∈ I, y1 , y2 ∈ J, then (x1 + x2 ) ∈ I and (y1 + y2 ∈ J. Thus (x1 + y1 ) +
(x2 + y2 ) = ((x1 + x2 ) + (y1 + y2 )) ∈ I + J, and hence it is closed under addition.
(iii) Let r ∈ R, x ∈ I, and j ∈ J. Then rx ∈ I and ry ∈ J. Therefore, r(x + y) =
(rx + ry) ∈ I + J.

30. (a) Since intersection of subgroups are subgroups, I ∩ J is an additive subgroup of


R. Now, let r ∈ R and a ∈ I ∩ J. Since a ∈ I, we have that ra ∈ I. Since a ∈ J,
we have that ra ∈ J. So, ra ∈ I ∩ J.
def def def
(b) Let R = Z[x, y], I = (x, y), J = ((x − 1), (y − 1)). Let also,
def
S = {ab : a ∈ I and b ∈ J}.

If a ∈ I, then a = xf1 + yf2 , and if b ∈ J, then b = (x − 1)g1 + (y − 1)g2 , with


f1 , f2 , g1 , g2 ∈ Z[x, y]. So, an element of S has the form:

ab = (xf1 + yf2 )((x − 1)g1 + (y − 1)g2 )


= x(x − 1)f1 g1 + x(y − 1)f1 g2 + y(x − 1)f2 g1 + y(y − 1)f2 g2 .

5
If S is an ideal, then c = x(x − 1) + y(y − 1) ∈ S [since x(x − 1) and y(y − 1) are
in S]. But then, it must have the form above, say c = ab, with f1 g2 = f2 g1 = 0.
Thus, either f1 = 0, or g2 = 0. If f1 = 0, then ab(x, 0) = 0, but c(x, 0) = x(x−1) 6=
0. So, we must have g2 = 0. But then, ab(1, y) = 0, while c(1, y) = y(y − 1).
Hence, c 6= ab, i.e., c 6∈ S.

Now, let R be any [commutative] ring [with identity]. [In fact, this generalizes for
rings that might not have an identity and/or not be commutative, remembering
that an ideal in that case is such that for r ∈ R and a ∈ I, we have ar, ra ∈ I.]
Let ( )
def
X
IJ = aν bν : aν ∈ I and bν ∈ J .
ν

[Caveat emptor: The above notation is the standard notation (and hence we
cannot change it) for IJ, but be careful that it does not mean the same as the S
above. The elements of IJ are, by definition, finite sums of products of elements
in I with elements in J, and not just products of elements in I with elements in
J. This can be confusing, so be careful to remember the correct definition.]
Clearly, since I, J 6= ∅, IJ 6= ∅. Also, for all r ∈ R, and for aν ∈ I and bν ∈ J,
we have that raν ∈ I, and so
!
X X
r aν b ν = (raν )bν ∈ IJ.
ν ν

Finally the sum of two finite sums of products of elements of I with elements of
J is still a finite sum of products of elements of I with elements of J. So, IJ is
closed under multiplication. Therefore, IJ is an ideal.

(c) Let aν ∈ I and bν ∈ J. Since I and J are ideals, aν bν are in both I and J, and
hence in I ∩ J. Since I ∩ J is also an ideal [by (a)], it is closed under addition, so
P
ν aν bν ∈ I ∩ J.
def def def
(d) Let R = Z, I = (4), and J = (6). Then, (4)(6) = (24). [This is easy to check:
elements are sums of multiple of 24.] On the other hand, (4) ∩ (6) = (12). [Since
12 ∈ (4) ∩ (6), clearly (12) ⊆ (4) ∩ (6). Also, since 12 is the least common multiple
of 4 and 6, we have that (4) ∩ (6) ⊆ (12).]

6
Homework 3

Section 10.4
2. (a) We have Z[x]/(3) ∼
= F3 [x]. Now, in F3 [x], we have that x2 + 3 = x2 . So,

Z[x]/(3, x2 + 3) = F3 [x]/(x2 ) = {a + bx : a, b ∈ F3 , and x2 = 0}.

[Note that the resulting ring has zero divisors, e.g., x. Thus, it is not a field.]
(b) We have Z[x]/(5) ∼
= F5 [x]. Now, in F5 [x]/(x2 + 3), we have that x2 = −3 = 2.
[Note that the polynomial x2 + 3 has no root in F5 .] So,

Z[x]/(5, x2 + 3) = F5 [x]/(x2 + 3) = {a + bx : a, b ∈ F5 , and x2 = 2}.

Note that
1 a − bx a − bx a b
= 2 = = − x,
a + bx a − b2 x 2 a2 − 2b2 a2 − 2b2 a2 − 2b2
and that,
a2 − 2b2 = 0 iff a = b = 0.

To prove the last statement, observe that if b 6= 0, then a2 −2b2 = 0 iff (a/b)2 +3 =
0, but we have already established that x2 + 3 has no roots in F5 . If b = 0, then
a2 − 2b2 = 0 iff a2 = 0, and since there are no zero divisors in F5 , we need a = 0.
[Or, you could try all possibilities, since there are finitely many.]
Hence, the resulting ring is a field!

3. (a) We have that

Z[x]/(x2 − 3) = {a + bx : a, b ∈ Z, and x2 = 3}.



[So, basically Z[x]/(x2 − 3) ∼
= Z[ 3].]

Now in R = (Z[x]/(x2 − 3))/ 2x + 4 , x̄2 = 3 and 2x̄ = 4, but then, squaring this
last equation, we obtain 4x̄2 = 16, and so 12 = 16, or 4 = 0 [in R].
Then, 2x̄ + 4 = 2x̄ = 0. Thus, 2x̄2 = 0, or 6 = 0. Since also 4 = 0, we have that
2 = 0.
Hence,

Z[x]/(x2 − 3, 2x + 4) ∼
=R∼
= {a + bx : a, b ∈ F2 , and x2 = 1}.

7
(b) Let φ : Z → Z[i]/(2 + i). [Remember that there is only one way to define this
homomorphism, namely φ(n) = n + (2 + i) = n̄.]

If φ(n) = 0̄, then n ∈ (2 + i), i.e., n = (a + bi)(2 + i) = (2a − b) + (a + 2b)i. Thus,


a = −2b [since n has no imaginary part], and hence n = 2a − b = −5b. Therefore,
n ∈ 5Z. Conversely, if n = 5k, then n = k(2 − i)(2 + i), and so n ∈ (2 + i). So,
ker φ = 5Z.

Let a + bi ∈ Z[i]/(2+i). Then, 2 + i = 0̄, and so ī = −2. So, a + bi = a + b(−2) =


a − 2b. Since a − 2b ∈ Z, we have that φ(a − 2b) = a − 2b = a + bi, and φ is onto.

By the First Isomorphism Theorem, Z/5Z ∼


= Z[i]/(2 + i).

6. Let a ∈ I ∩ J. Then, a ∈ I and a ∈ J. Therefore a2 = a · a, and thus a2 ∈ IJ. So,


(a + IJ)2 = (a2 + IJ) = (0 + IJ), and (a + IJ) is nilpotent.

7. (a) Remember that IJ ⊂ I ∩ J, from problem 30(c) in section 10.3.


Let a ∈ I ∩ J. Since a ∈ R = I + J, there are x ∈ I and y ∈ J, such that
1 = x + y. Then, a = a · 1 = a(x + y) = ax + ay. But, since a ∈ J and x ∈ I, we
have ax ∈ IJ, and in the same way, since a ∈ I and y ∈ J, we have that ay ∈ IJ.
Thus, a = ax + ay ∈ IJ.

(b) [You were not asked to do this one, but since we need it for problem 8(a), I
thought you should see it.] Since I + J = R, there are r ∈ I and s ∈ J such that
1 = r + s.
Let x = br + as. Note that a = a · 1 = ar + as. Thus, x − a = br + as −
(ar + as) = (b − a)r ∈ I [since r ∈ I]. In the same way, b = br + bs, and
x − b = br + as − (br + bs) = (a − b)s ∈ J [since s ∈ J].

8. Let φ : R → (R/I) × (R/J), defined by φ(a) = (a + I, a + J).

Homomorphism: We have already seen that the reduction maps R → R/I and R →
R/J are homomorphisms. Moreover, it is easy to prove that if φ : R → R0 and
ψ : R → R00 are homomorphisms, then θ : R → R0 × R00 , defined by θ(x) = (φ(x), ψ(x))
is a homomorphism. [If you are not convinced, try to prove it.] Hence, the φ in the
definition above is a homomorphism.

Injective: If φ(a) = (0 + I, 0 + J), then a ∈ I ∩ J. But, since I + J = R, we have, by

8
7(a), that I ∩ J = IJ. Thus, since IJ = 0, we have a = 0. Therefore ker φ = {0}, and
φ is injective.

Surjective: Let (a + I, b + J) ∈ (R/I) × (R/J). By the Chinese Remainder Theorem


(problem 7(b), which you were not asked to do), since I + J = R, there is x ∈ R such
that x + I = a + I and x + J = b + J. Thus, φ(x) = (x + I, x + J) = (a + I, b + J),
and φ is surjective.

Hence, φ is an isomorphism.

Homework 4

Section 10.5
1. [Before we really write a proper solution, we figure out what is really happening:
α − 10 = 0 means that α = 10. On the other hand, 2α − 6 = 0 would imply that
2 · 10 − 6 = 14 = 0. So, we see that we need nothing but elements of Z, since the
element added was just “10”, but 14 = 0. We can then “guess” that Z[α] ∼
= Z/14Z,
and we will now prove it using the First Isomorphism Theorem.]

Let φ : Z → Z[x]/(x − 10, 2x − 6), defined in the only possible way, i.e., φ(n) = n̄ =
n+(x−10, 2x−6). [Note that Z[α] ∼ = Z[x]/(x−10, 2x−6).] By its standard definition,
we know that φ is a homomorphism.

We have that φ is onto, since given f (x) + (x − 10, 2x − 6), we have that f (x) =
q(x)(x − 10) + f (10) by the division algorithm, and hence, f (x) + (x − 10, 2x − 6) =
f (10) + (x − 10, 2x − 6). Since f (x) ∈ Z[x], we have that f (10) ∈ Z, and therefore
φ(f (10)) = f (x) + (x − 10, 2x − 6).

We now prove that ker φ = 14Z. First observe that 14 = −2(x − 10) − 1(2x − 6). So,
14 ∈ (x − 10, 2x − 6), and therefore 14Z ∈ ker φ. [Remember that if r ∈ ker φ, then so
is every multiple of r, for the kernel is an ideal.] On the other hand, if n ∈ ker φ, then
φ(n) = n + (x − 10, 2x − 6) = 0 + (x − 10, 2x − 6). So, n ∈ (x − 10, 2x − 6), which
means that there are f, g ∈ Z[x] such that n = f (x)(x − 10) + g(x)(2x − 6). But then,
substituting x by 10, we have n = f (10) · 0 + g(10) · 14. Since g(10) ∈ Z, we have
n ∈ 14Z. Thus, ker φ = 14Z.

By the First Isomorphism Theorem, Z/14Z ∼


= Z[x]/(x − 10, 2x − 6) ∼
= Z[α].

9
2. Let φ : R[x] → R × R defined by φ(f (x)) = (f (1), f (−1)). As observed in the solution
of 10.4.8, this map is a homomorphism [since it is a homomorphism – evaluation – in
each coordinate].

We now prove that φ is onto: given (a, b) ∈ R×R, we have φ(a(x+1)/2−b(x−1)/2) =


(a, b).

Now, φ(f (x)) = (0, 0) iff f (1) = f (−1) = 0 iff (x−1) and (x+1) divide f (x) iff (x2 −1)
divide f (x). Thus, ker φ = (x2 − 1).

By the First Isomorphism Theorem, R[α] ∼


= R[x]/(x2 − 1) ∼
= R × R, and φ(α) = φ(x) =
(1, −1). [So, (1, −1)2 = (1, 1) = 1R×R .]

3. Remember from class [or from the book] that if ab = 0 and we add an inverse of a,
then b becomes zero. Since (2, 0) · (0, b) = (0, 0), we have that by adjoining an inverse
of (2, 0) we make all elements of {0} × R zero. Then, intuitively, we only have to deal
with first coordinates. But, in the first coordinates, (2, 0)x − (1, 1) = (0, 0) becomes
2x − 1 = 0, or x = 1/2. So, adjoining this “new” element [the inverse of 2] really adds
nothing new [since 1/2 ∈ R]. Thus, it seems that what we obtain is R. We now make
this precise. [A well written argument similar to this might give you some good partial
credit, though.]

Let’s denote R = (R × R)[x]/((2, 0)x − (1, 1)), and use f (x) = f (x) + ((2, 0)x − (1, 1))
for all f (x) ∈ (R × R)[x].

First observe that (a, b) = (a, 0) in R, since:

(a, b) = (a, 0) + (0, b) = (a, 0) − (0, b) · [(2, 0)x − (1, 1)].

Moreover, x = (1/2, 0) in R, since:

x = (1, 1)x − (1/2, 0) + (1/2, 0)


= (1/2, 1)[(2, 1)x − (1, 0)] + (1/2, 0)
= (1/2, 0)[(2, 0)x − (1, 1)] + (1/2, 0) [remember (a, b) = (a, 0)]

= (1/2, 0) [since (2, 0)x − (1, 1) = 0]

So, we claim that φ : R → R = R × R[x]/((2, 0)x − (1, 1)), given by φ(a) = (a, 0) is an
isomorphism.

10
It is a homomorphism, since φ is the composition of a 7→ (a, 0) [a homomorphism],
with the inclusion of R × R in (R × R)[x] [another homomorphism], with the reduction
modulo the ideal ((2, 0)x − (1, 1)) [also a homomorphism].
It is onto since given f (x) ∈ R, we have that, using what was done above, f (x) =
f (1/2, 0) = (a, b) = (a, 0) for some a, b ∈ R. So, φ(a) = (a, 0) = f (x).
If φ(a) = (a, 0) = (0, 0), then (a, 0) ∈ ((2, 0)x − (1, 1)). So, there would be f (x) ∈
(R×R)[x] such that (a, 0) = f (x)((2, 0)x−(1, 1)). But then, substituting x by (1/2, 0),
we would obtain (a, 0) = f (1/2, 0)(0, 1). Since f (1/2, 0) ∈ R × R, we would have that
(a, 0) = f (1/2, 0)(0, 1) = (0, b) for some b ∈ R. But then a = 0, and ker φ = {0}.
Thus, φ is injective.

5. We have:

(α3 + α2 + α)(α + 1) = α4 + 2α3 + 2α2 + α


= −(α3 + α2 + α + 1) + 2α3 + 2α2 + α
= α3 + α2 − 1,

and

α5 = α4 · α
= −(α3 + α2 + α + 1)α
= −α4 − α3 − α2 − α
= (α3 + α2 + α + 1) − α3 − α2 − α
= 1,

or, alternatively,

α5 = (α5 − 1) + 1
= (α − 1)(α4 + α3 + α2 + α + 1) + 1
= (α − 1) · 0 + 1 = 1.

9. As before, every element that when multiplied by 2 gives 0, becomes equal to 0 in the
new ring. So, 6 becomes 0, and we can “guess” that the new ring is Z/6Z. But, after
further investigation, we realize that 6 being zero now means that 2 · 3 = 0, and since
2 is invertible, this means that 3 is also 0. We now prove that the new ring is indeed
Z/3Z.

11
Now we prove it.
def
Let Z → R = (Z/12Z)[x]/(2x − 1), defined in the only possible way, i.e., φ(n) = n =
n + (2x − 1). As always, this map gives us a homomorphism.
Note that in (Z/12/Z)[x], we have 6 = −6(2x − 1), and so 6 = 0 in R. But then,
3 · 2 = 0, and so, 0 = 3 · 2 · x = 3 · 1 = 3. So, 3 = 0. But then, in R, we have 2x − 1 = 0,
i.e., 2 · x = 1, or −1 · x = 1, and so x = −1.
So, now we prove that φ is onto. Let f (x) ∈ R. Then, f (x) = f (−1). Since f (−1) ∈
Z/12Z, there is some integer n such that φ(n) = n = f (−1) = f (x). Therefore, φ is
onto.
Now, by the work above, we see that φ(3) = 0, and so 3Z ⊆ ker φ. The only ideals [and
the kernel is an ideal] that contain 3Z are (1) = Z and 3Z itself. So, if ker 6= 3Z, then
ker φ = Z. So, suffices to show that 1 6∈ ker φ, i.e., that 1 6∈ (2x − 1) in (Z/12Z)[x].
Suppose it is. Then, 1 = f (x)(2x − 1) for some f (x) ∈ (Z/12Z)[x]. By substituting x
by 2, we have that 1 = 3f (2), now in Z/12Z, since we are only dealing with constants.
But, in Z/12Z, 3 is not a unit, and hence this cannot happen. Therefore, ker φ = 3Z.
By the First Isomorphism Theorem, (Z/12Z)[x]/(2x − 1) ∼
= Z/3Z.

Section 10.6
2. We just need to show that given a non-zero element a in a finite domain R, we have
an inverse of a in R, i.e., some element b such that ab = 1R .
Let then a ∈ R−{0} [this a is now fixed!] and define f : R → R by f (r) = ar. [We will
show that a has an inverse. Note also that f is just a function, not a homomorphism!]
The map f is injective: if f (r) = f (s), then ar = as, which implies that a(r − s) =
ar − as = 0. Since we are in a domain and a 6= 0, we have r − s = 0, or r = s, and f
is injective.
Since R is finite, by the Pigeonhole Principle [which only applies to finite sets], f is
also onto. Thus, there is an element b such that f (b) = 1R [since there is 1R ∈ R]. But
then f (b) = ba = ab = 1R , and a is invertible.

3. It will be useful to prove the following Lemma:

Lemma. If R is a domain and f, g ∈ R[x], then deg(f · g) = deg f + deg g. [Remember


that deg 0 = −∞.]

12
Proof. The lemma is trivial if either f or g is zero. So assume they are not. Let
deg f = n, deg g = m, the leading coefficient of f be an and the leading coefficient of
g be bm .

So, the term of highest power in the product f (x)g(x) is an · bm xn+m . Since we are in
a domain, and am , bm 6= 0, an · bm 6= 0, and so deg(f · g) = n + m.

Now, suppose that R is a domain, and f, g ∈ R[x] − 0. Then deg f, deg g ≥ 0. So,
deg(f · g) = deg f + deg g ≥ 0. Thus, f · g 6= 0 [or the degree would have been −∞.]

4. Let f ∈ R[x]. If f is invertible, then there exists g ∈ R[x] such that f · g = 1. But, by
the Lemma above, 0 = deg 1 = deg(f · g) = deg f + deg g ≥ deg f . So, deg f = 0, and
f ∈ R. [Note that deg f, deg g 6= −∞, since f 6= 0, by assumption, and g 6= 0, or f · g
would be 0, not 1.

Homework 5

Section 10.7
2. [See alternative solutions for parts (b), (c), and (d) below!]

(a) If (a, b) ∈ I, and a, b 6= 0, then ((a, b)) = R × R, since given (x, y) ∈ R × R,


(x/a, y/b)(a, b) = (x, y) ∈ ((a, b)). So, I = R × R. [Hence, I is not maximal.]
So, if M is a maximal ideal, there is no element in M with both entries non-
zero. Suppose that (a, 0) ∈ M , with a 6= 0. So, there is no (0, b) ∈ M , unless
b = 0, or M would be the whole ring, since (a, 0) + (0, b) would be in M . Hence,
M ⊆ R × {0}.
But, then, similar to the above, ((a, 0)) = R × {0}, and so R × {0} ⊆ M . So,
[since we already have the other inclusion] M = R × {0}.
In the same way, there is no (a, 0) ∈ M , except if a = 0, but (0, b) ∈ M , with
b 6= 0, then M = {0} × R.
Since, M = ((0, 0)) is not maximal [it is contained in both of the ideals above],
and, as observed above, there is no element with both entries non-zero in M , then
either M has an element of the form (a, 0), with a 6= 0, or of the form (0, b), with
b 6= 0. So, the only possibilities are the above.

13
Finally, observe that the possibilities are indeed maximal ideals, since in each case
R × R/M ∼ = R, a field. [Note that they are indeed ideals too.] Alternatively, we
proved that all ideals of R × R different from the zero ideal and the unit ideal are
the ones above, and hence, they are maximal.
def
(b) Note that in R = R[x]/(x2 ), every element can be represented by a + bx, where
a, b ∈ R and x2 = 0. Then, if a 6= 0, then (a + bx)(a−1 − ba−2 x) = 1 [since x2 = 0],
and so a + bx is a unit.
So, if we have an ideal I that is not the unit ideal [i.e., R] and not (0), its
elements are of the form bx. Since I 6= (0), there is an element with b 6= 0. So,
b−1 bx = x ∈ I. Thus, I = (x) = {bx : b ∈ R}. Since this is the only ideal
different from (0) and R, it is the only maximal ideal. [Also, R/I ∼
= R.]
(c) I’m pretty sure that I proved the following lemma in class [and that it is in the
book somewhere], but in any case, here it is:
Lemma. Let F be a field and f (x) ∈ F[x]. Then, ḡ(x) ∈ F [x]/(f ) is a unit iff
gcd(f, g) = 1.

Proof. We have: ḡ(x) ∈ (F [x]/(f ))× iff there exists r̄(x) ∈ F [x]/(f ) such that
ḡr̄ = 1̄ iff there exist r, s ∈ F [x] such that gr = 1 + f s iff there exist r, s ∈ F [x]
such that −f s+gr = 1 iff [by a Corollary of Bezout’s Theorem] gcd(f, g) = 1.
def
Now, back to the problem: Let f (x) = x2 − 3x + 2 = (x − 1)(x − 2), and
def
R = R[x]/(f ). By the lemma above, ax + b is a unit of R iff gcd(ax + b, f ) = 1.
By the factorization of f , this is the same as to say that (x − 1) or (x − 2) do not
divide ax + b, i.e., a 6= −b and a 6= −2b. Therefore, if M is a maximal ideal and
ax + b ∈ M , then either a = −b or a = −2b [otherwise, we would have M = R,
since ax + b would be a unit].
So, suppose that ax + b ∈ M , with b = −a and a 6= 0. Then, ax + b = a(x − 1)
and (x − 1) ⊆ M . But, since R/(x − 1) ∼
def
= R [by φ(ax + b) = a + b], (x − 1) = M
is maximal.
In the same way, we prove that if ax + b ∈ M , with b = −2a and a 6= 0, then
M = (x − 2).
Thus, these are all the maximal ideals.
def
(d) Since f (x) = x2 + x + 1 is irreducible [if it were not, it would have a factor of
degree one, and hence a real root, but one can check that it has no real roots], for

14
all ax + b ∈ R[x] − {0}, we have gcd(f, ax + b) = 1. Hence R[x]/(f ) is a field and
therefore contains only the zero and the unit ideals. So, the only maximal ideal
is (0).

Alternative solutions for (b), (c), and (d): Another way to approach quotient
rings is to use the Correspondence Theorem [Prop. 10.4.3 on pg. 360]. Problem 10.7.4,
which you did not have to do, is the main tool. [I give its solution below for your
reference, but as you will see, it is a very simple application of the Correspondence
Theorem.] It states that M is a maximal ideal of R containing the ideal I iff the
reduction modulo I of M is a maximal ideal of R/I.

(b) The maximal ideals of R[x]/(x2 ) are the reduction of maximal ideals of R[x] that
contain x2 . We know that the maximal ideals of R[x] are of the forms (x − a)
[with a ∈ R] or (x2 + ax + b) with a2 − 4b < 0 [and a, b ∈ R].
So, the only maximal ideal of R[x] that contains (x2 ) is (x). [Note that we know
that since every ideal of R[x] is principal, and if (x2 ) ⊆ (f ), then f (x) | x2 .] So,
(x) is the only maximal ideal of R[x]/(x2 ).
(c) In the same way, the only divisors of x2 − 3x + 2 in R[x] are (x − 1) and (x − 2).
Hence, the reductions of the ideals generated by each one give the maximal ideals
of R[x]/(x2 − 3x + 2).
(d) Finally, since (x2 + x + 1) is maximal [since 12 − 4 · 1 < 0], the only maximal ideal
of the quotient is the reduction of the ideal generated by the polynomial itself,
which clearly gives the zero ideal.

3. By the Hilbert’s Nullstellensatz, the ideal I is maximal iff I = (x − a, y − b) for some


a, b ∈ C. By inspection, one can easily guess that if that if the given ideal I is maximal,
then I = (x, y).
Indeed, we have
(y + x2 + 2xy 2 + y 4 ) − (x + y 2 )2 = y ∈ I.

So,
(x + y 2 ) − y 2 = x ∈ I.

Thus, (x, y) ⊆ I.
Now, one can easily see that I ⊆ (x, y), and hence I = (x, y). By the Nullstellensatz,
I is then maximal.

15
4. By the Correspondence Theorem, since I ⊆ M , there is an ideal J such that M & J & R
iff there is an ideal J¯ of R/I such that M̄ & J¯ & R/I. [The fact that it is a one-to-one
correspondence is crucial here!!]
def
7. Let f (x) = x3 + x + 1 ∈ F2 [x]. We then have f (0) = f (1) = 1, i.e., f has no root in
F2 . If f were reducible, it would have to factor as a polynomial of degree one times a
polynomial of degree two, or as a product of three polynomial of degree one. In either
case, f would have a root. Therefore, f is irreducible.

Here is another lemma that we can use whenever we need:

Lemma. Let F be a field and f ∈ F [x] irreducible. Then (f ) is maximal.

Proof. Suppose now that (f ) & I ⊆ R. Since every ideal of F [x] is principal [Proposi-
tion 10.3.21], there is g ∈ F [x] monic such that I = (g). But, if (f ) ⊆ (g), then f ∈ (g),
i.e., g divides f . On the other hand, f is irreducible, and so either g = 1 or g = f .
But we also have (f ) 6= (g), and thus (g) = (1) = R. Hence, (f ) is maximal.

So, we just apply the lemma for f above. Since (f ) is maximal, F2 [x]/(f ) is a field.

In F3 [x], we have that f (1) = 0. So, (x − 1) | f and (f ) & (x − 1) & F3 [x]. Therefore
(f ) is not maximal and hence F3 [x]/(f ) is not a field.

10. Suppose that M & I ⊆ R. Then, I has some element of R − M , and so, it has a unit.
Therefore I = R and M is then maximal.

Now suppose that M 0 is also maximal, with M 0 6= M . Since M 0 is maximal (and not
equal to M ), we cannot have M 0 ⊆ M . Thus, M 0 has some element of R − M , i.e., a
unit. This would mean that M 0 = R, and hence not a maximal ideal. [Contradiction.]

Section 11.1
1. Suppose that d | a, b, with d > 0. Then, d | (a + b) = p. Hence, d is either 1 or p. If
d = p, then a and b are positive multiples of p and hence cannot add to p. So, d = 1
is the only [positive] common divisor of a and b, i.e., gcd(a, b) = 1.

3. Suppose that e | a1 /d, . . . , an /d. Hence ai /d = ki e, i.e., ai = ki de, for i = 1, . . . , n.


Thus, de | a1 , . . . , an . If e were greater than 1, then d would not be the greatest common
divisor of the ai ’s, since de would be greater than d.

16
5.(a) If d | a, b, then d | (b − aq) = r [and d | a]. If d | a, r, then d | (aq + r) = b [and a]. So,
d is a common divisor of a and b iff d is a common divisor of a and r, i.e., the set of
common divisors of a and b is equal to the set of common divisors of a and r. So their
maxima, i.e., gcd(a, b) and gcd(a, r) respectively, are equal.

8. (a) As seen in problem 10.7.7, the polynomial is irreducible.

(a’) [I did p = 3 first by mistake. So, just so you have another example, I decided to
leave it.]
def
As seen in problem 10.7.7, 1 is a root of f (x) = x3 + x + 1 in F3 . Dividing by
(x−1) we obtain x3 +x+1 = (x−1)(x2 +x+2). Now, one can check that x2 +x+2
has no roots in F3 [by testing all three possible roots], and so it is irreducible.
[Note that if a polynomial of degree 2 is reducible, then it is a product of two
polynomials of degree one, and so it has a root.]

(b) Again, testing all possible roots, we see that x = 1, 2, so x3 −3x−3 = (x−1)(x−2)
in F5 [x].

Homework 6

Section 11.2
1. (a) FALSE. If R[x, y] were a Euclidean domain, it would be also a PID, but the ideal
(x, y) is not principal! Indeed, if (x, y) = (f (x, y)), then there are q1 , q2 ∈ R[x, y]
such that x = f q1 and y = f q2 . Looking at degrees in terms of x and y, we
have degx y = 0 = degx f + degx q2 , and degy = 0 = degy f + degy q1 . Hence
degx f = degy f = 0, i.e., f ∈ R× . But then (f ) = R[x, y], by (x, y) 6= R[x, y]
since if g(x, y) ∈ (x, y), then g(0, 0) = 0 [since then g(x, y) = xg1 (x, y)+yg2 (x, y)],
and so (x, y) has no constants other than 0.

(b) FALSE. The ideal (2, x) is not principal. Indeed, if (2, x) = (f (x)), then 2 =
f (x) q(x), for some q(x) ∈ Z[x]. But then, deg f = deg q = 0, i.e., f, q ∈ Z.
But, then, since 2 is prime, f = ±1 or f = ±2. If f = ±1, then (f ) = Z[x],
but (2, x) 6= Z[x] since 1 6∈ (2, x) [for if g ∈ (2, x), then g(x) = 2g1 (x) + xg2 (x),
and so g(0) = 2g1 (0) is even]. But also, f cannot be ±2, since we would need
x = ±2 · q(x) for some q ∈ Z[x], and looking a leading coefficients, we see that it
is impossible [the leading coefficient of ±2 · q(x) is even].

17
2.(b) We copy the idea from Gaussian integers. We will again use the size function σ(z) =
√ √
|z|2 . Let a = a1 + a2 −2 and b = b1 + b2 −2. Then,

a a1 + a2 −2
= √
b b1 + b2 −2
√ √
(a1 + a2 −2)(b1 − b2 −2)
=
b21 + 2b22
−a1 b2 + a2 b1 √
   
a1 b1 + 2a2 b2
= + −2.
b21 + 2b22 b21 + 2b22
Now, we take the long division (in Z) of the numerators by the denominators, with
the requirement that the remainder is in absolute value less than or equal to half of
the absolute value of the divisor [which we’ve seen in class is always possible], i.e., let
q1 , q2 , r1 , r2 ∈ Z such that:

a1 b1 + 2a2 b2 = (b21 + 2b22 )q1 + r1 , |r1 | ≤ (b21 + 2b22 )/2,

−a1 b2 + a2 b1 = (b21 + 2b22 )q2 + r2 , |r2 | ≤ (b21 + 2b22 )/2.

Then,

   
a a1 b1 + 2a2 b2 −a1 b2 + a2 b1
= + −2
b b21 + 2b22 b21 + 2b22
√ r2 √
 
r1
= (q1 + q2 −2) + + −2 ,
b21 + 2b22 b21 + 2b22
and so,
√ r2 √
 
r1
σ(a/b − (q1 + q2 −2)) = σ + −2
b21 + 2b22 b21 + 2b22
r12 r22
= + 2
(b21 + 2b22 )2 (b21 + 2b22 )2

≤ 1/4 + 1/4 · 2 = 3/4 < 1.



So, let q = q1 + q2 −2 and r = a − bq. Then, σ(r) = σ(a − bq) = |a − bq|2 =
|b|2 |a/b − q|2 = σ(b)σ(a/b − q) < σ(b). [This last inequality comes from the inequality
of the previous displayed equation.]

4. Let d and e be the greatest common divisors of m and n in Z and Z[i] respectively.
Since Z ⊆ Z[i], we have that d | e [property of the gcd], and so let e = de1 . We then

18
have that d = ar +bs, for some r, s ∈ Z [by Bezout’s Theorem], or, 1 = (a/d)r +(b/d)s.
[Note a/d, b/d ∈ Z.]

On the other hand, since e | a, b, we have that (a/e)r + (b/e)s ∈ Z[i], and so, (a/e)r +
(b/e)s = 1/e1 ((a/d)r + (b/d)s) = 1/e1 ∈ Z[i]. Thus, e1 is a unit, and d and e are
associates.

7. [Haven’t I done this in class??] Since R is a PID, given a, b ∈ R, there is d ∈ R such


that (a, b) = (d). Since d ∈ (a, b), there are r, s ∈ R such that d = ar + bs, which
proves (i).

Since a, b ∈ (a, b) = (d), we have that d | a, b, which proves (ii).

If e | a, b, then a, b ∈ (e), and hence (d) = (a, b) ⊆ (e). So d ∈ (e) and e | d, which
proves (iii).

Suppose now that d0 satisfy (i), (ii), (iii). Then, by (ii) for d0 , we have d0 | a, b and
hence d | d0 by (iii) for d. Reversing the argument, we get that d | d0 , and hence d
and d0 are associates. [Note that we did not need to use (i). So, the uniqueness, up
to unit, comes from (ii) and (iii). Hence, the gcd defined in the text for the case of
UFDs using the irreducible factors, coincides with the gcd defined here, which can only
be defined in PIDs, by Corollary 11.2.10.]

12. (a) We have that (a) ∩ (b) is an ideal of R. Hence, since R is a PID, there is m ∈ R
such that (a) ∩ (b) = (m). Since m ∈ (a), we have that a | m and in the same
way b | m. Now, a, b | r iff r ∈ (a) ∩ (b) = (m), i.e., iff m | r.
Suppose that m0 also satisfy the required properties. Then since a, b | m0 , we have
that m | m0 , and reversing the argument, m0 | m. Thus, m and m0 are associates.

[Alternatively, one could use the fact that PIDs are UFDs to define the lcm in
terms of irreducible factors, but this is a bit messy since we have to deal with
associates. But, basically you prove that you can write a = p1 · · · pr · q1 · · · qs and
b = p1 · · · pr · q10 · · · qt0 , where no qi is an associate of any qi0 . Then, a lcm of a and
b is an associate of p1 · · · pr · q1 · · · qs · q10 · · · qt0 .
def
Yet another possibility, would be to define lcm(a, b) = ab/ gcd(a, b).]
def def
(b) Let d = gcd(a, b) and m = lcm(a, b). [Note that those are not uniquely defined!
But it doesn’t matter, you can take any gcd and any lcm.] Let us write a = a1 d,
b = b1 d, m = m1 a, and m = m2 b.

19
Since ab = a1 db = ab1 d and R is a domain [so we can cancel non-zero factors], we
have that a1 b = ab1 , and this element is clearly a common multiple of a and b.
Thus m | a1 b = ab1 [property of lcm], i.e., there is m3 such that a1 b = ab1 = m3 m.
So, a1 b = m3 m2 b, and hence [since R is a domain] a1 = m3 m2 , i.e., m3 | a1 . In
the same way, ab1 = m3 m1 a, and we obtain b1 = m3 m1 , i.e., m3 | b1 . But, by
problem 11.1.3, we have that gcd(a1 , b1 ) = 1, and hence m3 has to be a unit.
Therefore, mm3 = a1 b implies [multiplying by d] that m3 md = ab, and since m3
is a unit, ab and md are associates.

13. Suppose b = aq, with a, b ∈ Z, with a 6= 0, and q = q1 + q2 i ∈ Z[i]. [I.e., a | b in Z[i].]


Then b + 0i = aq1 + aq2 i, i.e., b = aq1 and 0 = aq2 . Since a 6= 0 and Z is a domain, we
have that q2 = 0, and so q ∈ Z. So, a | b in Z.

Homework 7

Section 11.3
1. Suppose, that f (x) = f1 (x)f2 (x), with deg fi > 0. [In other words, f is reducible.]
Then, f (ax + b) = f1 (ax + b)f2 (ax + b). Since a 6= 0 and F is a field [all we really need
in fact it has no zero divisors], then deg fi (x) = deg fi (ax + b). [This is true because
in an integral domain, if deg f = n and deg g = m, then deg f (g(x)) = m · n. I won’t
prove it here, but it is easy to prove: just use the Newton’s Binomial Theorem: in any
commutative ring R, and for all a, b ∈ R and positive integer n,
n  
n
X n i n−i
(a + b) = a b .]
i=0
i
So, f (ax + b) also factors non-trivially [i.e., neither f1 (ax + b) nor f2 (ax + b) is a unit,
since they have positive degrees], and thus f (ax + b) is reducible.
def
For the converse, suppose that g(x) = f (ax+b) is reducible. We can apply the previous
part to show that g(a−1 x + (−a−1 b)) = f (x) is also reducible. [Note that since a 6= 0
and F is a field, then a−1 exists and is non-zero.]

2. We have that C(x) is the field of fractions of C[x], which is a UFD. Hence, we just
need to apply the [main application of] Gauss’s Lemma in the general case of UFDs,
namely Theorem 11.3.9. So, noticing that (C[x])[y] = C[x, y], part (c) of this theorem
[with R = C[x]] gives the desired result.

20
4. Let f, g ∈ Z[x]. Since Q[x] is a Euclidean domain [since Q is a field], we have that f
and g are relatively prime iff [by a Corollary of Bezout’s Theorem] there are r, s ∈ Q[x]
such that f r + gs = 1. So, it suffices to prove that there is an integer in (f, g) [here
considering the ideal generated by f and g in Z[x]] iff there are such r and s.
Let d1 , d2 be the denominators of the contents of r and s respectively. Hence d1 r, d2 s ∈
Z[x] [by the definition of content]. So, if f g + rs = 1, then f d2 (d1 r) + gd1 (d2 s) = d1 d2 ,
and the d1 d2 ∈ (f, g). [We basically “cleared denominators” here.]
Conversely, if d ∈ Z is in (f, g), then there are r, s ∈ Z[x] such that f r + gs = d. Thus,
in Q[x], we have f (r/d) + g(s/d) = 1.

8. Suppose that xw − yz = f (x, y, z, w)g(x, y, z, w). Looking at degw , i.e., the degrees
with respect to w only, we have that [since C[x, y, z] is a domain] either degw f = 1
and degw g = 0 or degw g = 1 and degw f = 0. Let’s assume [without loss of generality]
the former.
We now write, f (x, y, z, w) = f1 (x, y, z)w + f0 (x, y, z) and observe that g(x, y, z, w) =
g(x, y, z) [i.e., g does not depend on w]. We thus have that f1 g = x and f0 g = yz [since
f g = xw − yz]. Looking at degrees in y and z in f1 g, we conclude that g [and f1 ] do
not depend on those variables. So, g can depend only on x. But looking at degrees in
x of f0 g, we see that g does not depend on x either. Hence, g ∈ C.
So, we can only factor xw − yz if one of the factors is a constant, and therefore, this
polynomial must be irreducible.
[Note that C was not important here. It could have been any field, or even any domain.]
def
9. Let’s denote the kernel of this homomorphism by I. Clearly f (x) = (x − 1)2 − 2 =
x2 − 2x − 1 ∈ I, and so (f ) ⊆ I. We now prove that I = (f ).
It suffices to prove that I ⊆ (f ) [since the above shows (f ) ⊆ I]. Suppose that g ∈ I.
[We need to show that f | g.] Since f is monic [which is crucial here since Z is not
a field], we can divide g by f and obtain g = f q + r, where g, r ∈ Z[x] and deg r <
deg f = 2. Let’s then write r(x) = ax + b with a, b ∈ Z. [Note that they can be zero.]
√ √
Since f, g ∈ I, and r = g − f q, we have that r ∈ I, i.e., r(1 + 2) = a(1 + 2) + b = 0.
√ √
But then, if a 6= 0, we have 2 = −(b + a)/a, which is a contradiction since 2 is
irrational. [Do you know how to prove that fact??] Thus, we must have a = 0, and

then also b = 0 [since b = r(1 + 2) = 0], i.e., f | g. Therefore g ∈ (f ), and since g
was an arbitrary element of I, we have that I ⊆ (f ).

21
Homework 8

Section 11.4
1. (a) Use Eisenstein’s Criterion with p = 3.
(b) Use Eisenstein’s Criterion with p = 3.
(c) Since the polynomial has degree 3, it is irreducible iff it has no roots in Q. The
possible roots are ±1, ±1/2, ±1/4, ±1/8. Since none of those are actual roots [you
have to compute them all!!], it is irreducible.
(d) As in (c), it is irreducible iff it has a root. Here the possible roots are ±1, ±7.
Since all the coefficients are positive, the root would have to be negative. But −1
and −7 are not roots. [Again, you have to compute those.]
(e) Use Eisenstein’s Criterion with p = 3.

3. Let f (x) = x3 + x + 1.

p = 2: Since the polynomial has degree 3, it is irreducible iff it has no roots in F2 . We


have that f (0) = f (1) = 1. So f is irreducible in F2 [x].
p = 3: Observe that f (1) = 0. So, dividing f (x) by x − 1 [in F3 [x]], we obtain: f (x) =
(x2 + x + 2)(x − 1). Now, since x2 + x + 2 has no root in F3 [check the three
possible roots], that polynomial is irreducible, and so the expression above is the
complete factorization of f (x).
p = 5: f (x) has no roots in F5 [check the five possible roots], so it is irreducible.

4. We can find the roots of the polynomial, since it gives us a biquadratic equation. We
then have that √ √
−1 ± 12−4·1·1 1 3
x2 = =− ± i.
2 2 2
So, x is plus or minus the square root of those. One can do that [geometrically] by
noticing that: √ √
1 3 i2π/3 1 3
− + =e , − − = ei4π/3 ,
2 2 2 2
√ √
and hence the solutions are ±eiπ/3 = ±(1/2 + i 3/2) and ±ei2π/3 = ±(−1/2 + i 3/2).
Thus, x4 + x2 + 1 factors [in C[x]] as:
√ !! √ !! √ !! √ !!
1 3 1 3 1 3 1 3
x− + i x− − i x− − + i x− − − i
2 2 2 2 2 2 2 2

22
Multiplying out the first two terms and the second two terms, we obtain:

x4 + x2 + 1 = (x2 − x + 1)(x2 + x + 1).

Since we know that x2 + x + 1 and x2 − x + 1 do not have roots in Q [we know


their factorization in C[x]], these are irreducible, and we have given the complete
factorization of x4 + x2 + 1.

8. Use Eisenstein with the prime p.

10. [We follow the proof of the Eisenstein’s Criterion.] Suppose that f (x) = g(x)h(x), with
deg g, deg h > 0. Reducing modulo p we have that ā2n+1 x2n+1 = ḡ h̄. Hence, ḡ = b̄k xk
and h̄ = c̄l xl , with b̄n , c̄l 6= 0̄ and k + l = 2n + 1. Then, we have g(x) = bk xk + pg1 (x),
h(x) = cl xl + ph1 (x), where p does not divide bk or cl , g1 , h1 ∈ Z[x], deg g1 < k,
deg h1 < l, and k + l = 2n + 1. Let’s also write

g1 (x) = bk−1 xk−1 + · · · + b0 , h1 (x) = cl−1 xl−1 + · · · + c0 .

Since k + l = 2n + 1, either k or l must be greater than or equal to n + 1, while the


other is less than or equal to n. Without loss of generality, we may assume k ≥ (n + 1)
[and so, l ≤ n].

We then have:

f (x) = (bk xk + pg1 )(cl xl + ph1 ) = bk cl x2n+1 + p(bk xk h1 + cl xl g1 ) + p2 (g1 h1 ).

We now look at the term of degree l in the above expression. On the one hand, it is
al xl [by the representation of f (x)]. On the other hand [by looking at the right hand
side of the equation above], it is the sum of the terms in xl of p(bk xk h1 + cl xl g1 ) and
p2 (g1 h1 ). Since k ≥ n + 1 ≥ l, the term in xl of p(bk xk h1 + cl xl g1 ) must come from
pcl xl g1 [since the terms in pbk xk h1 have powers of x greater than or equal to k] and it
is in fact pcl b0 . Denoting the coefficient of xl in g1 h1 by d [which could be zero], we
have:
al = p(cl b0 ) + p2 d.

Since l ≤ n, we have that p2 | al [by hypothesis]. So, p | cl b0 . Since p is prime and


p - cl [see above], we have that p | b0 .

Now, the constant term of f (x) is in the one hand a0 , but on the other hand [again,
using the right hand side of the equation above] it is p2 b0 c0 , i.e., a0 = p2 b0 c0 . Since

23
p | b0 , we conclude that p3 | a0 , which is a contradiction with the hypothesis that
p3 - a0 . Therefore, f must be irreducible.

Homework 9

Section 13.1
1. We have that a−1 = a if, and only if, a2 = 1. So, a must be a solution if x2 − 1 =
(x − 1)(x + 1) = 0. Since a filed is a integral domain [i.e., no zero divisors], we must
have that a = 1 or a = −1, and hence these are all elements such that a−1 = a. [Note
that if the field has characteristic 2, there is only one element, since 1 = −1 in that
case.]

Section 13.2
3. (a) We have:
√ √ √
α= 3+ 5 ⇒ α2 = 8 + 2 15 ⇒ (α2 − 8)2 = 60,
def
so, α is a root of f (x) = x4 − 16x2 + 4. We now show that this polynomial
is irreducible. Since this polynomial is biquadratic, we can find all its complex
roots. In C[x] it factors as:
√ √ √ √ √ √ √ √
f (x) = (x − ( 3 + 5))(x − ( 3 − 5))(x − (− 3 + 5))(x − (− 3 − 5)).
√ √
[One could easily “guess” that, since we knew 3+ 5 was a root.] So, it clearly
does not have a root in Q, and hence, if it factors, it factors as a product of two
polynomials of degree 2. But, any product a two pairs of factors of f (x) will not
be in Q[x], and hence it must be irreducible and f (x) = minα,Q (x) [the minimal
polynomial of α over Q].

(b) Note that f (x) ∈ Q[ 5][x]. But, it is not irreducible! Note that:
def √ √ √ √ √
f1 (x) = (x − ( 3 + 5))(x − (− 3 + 5)) = x2 − 2 5x − 2.

Hence f1 (α) = 0 and f1 ∈ Q[ 5][x].
√ √ √
Now if f1 is not irreducible, then we must have that 5 ± 3 ∈ Q[ 5] [since
√ √ √ √
the f1 would have to factor in Q[ 5][x]]. But then, 3 ∈ Q[ 5], since 3 =

24
√ √ √ √
( 3 + 5) − 5 [and Q[ 5] is closed under addition]. But then
√ √
3 = a + b 5, with a, b ∈ Q.

Note that b 6= 0, for otherwise 3 ∈ Q. Also, the above formula implies that
√ √ √ √
( 3 − b 5)2 = a2 , i.e, 9 − 2b 15 + 5b2 = a2 . So, [since b 6= 0] we have 15 =

−(a2 − 5b2 − 9)/(2b) ∈ Q, which is also a contradiction. Thus, 3 cannot be in
√ √
Q[ 5], and hence f1 is irreducible in Q[ 5][x] and f1 (x) = minα,Q[√5] (x).
√ √
(c) Again, note that f (x) ∈ Q[ 10][x]. Note that if there is a root of f (x) in Q[ 10],

say β, then we have Q ⊆ Q[β] ⊆ Q[ 10]. But we know that the dimension
√ √
of Q[ 10] as a Q-vector space is 2 [the basis is {1, 10}], but since f (x) is
irreducible over Q, we have that Q[β] is a Q-vector space of dimension 4 [the
basis is {1, β, β 2 , β 3 }]. But a vector space of dimension 2 cannot have a subspace

of dimension 4. So, there are no roots of f (x) in Q[ 10]. Thus, if f (x) factors in

Q[ 10][x], then if factors as a product of two quadratic polynomials.
√ √ √
If f1 [as above] is in Q[ 10][x], then 5 ∈ Q[ 10] [since the coefficients must be
√ √ √
in Q[ 10]]. But, with an argument similar to the one above [for 3 6∈ Q[ 5]],

one can see that this is not true [since 50 6∈ Q].
So, one tries the other possible products of pairs of factors of f (x), and verify

that no such product will have coefficients in Q[ 10]. So, f (x) is irreducible also

in Q[ 10] and f (x) = minα,Q[√10] (x).

(d) Note that

def √ √ √ √ √
f2 (x) = (x − ( 3 − 5))(x − (− 3 + 5)) = x2 + 2 15x − 8,
√ √
and so f2 (α) = 0 and f2 ∈ Q[ 15][x]. If f2 is not irreducible, then α ∈ Q[ 15].
But, in that case:
√ √ √
α= 3+ 5 = a + b 15, with a, b ∈ Q.
√ √
Then, squaring both sides we have that 8 + 2 15 = (a2 + 15b2 ) + 2ab 15. Hence

ab = 1 and a2 + 15/a2 = 8, or a4 − 8a2 + 15 = 0. But this implies that a = ± 3

or a = ± 5, which contradicts the fact that a ∈ Q. Hence, f2 is irreducible and
f2 (x) = minα,Q[√15] (x).

25
Section 13.3
1. If α2 ∈ F , then α is a root of x2 −α2 ∈ F [x]. Hence, we would have [F [α] : F ] ≤ 2, which
contradicts the assumption that α generates an extension of degree 5. So, α2 6∈ F .

Since [F [α] : F ] = 5 is prime, α2 ∈ F [α], and α2 6∈ F , we have [by Corollary 13.3.7]


F [α2 ] = F [α].

2. We have that φ7 (ζ) = 0, where φ7 (x) = x6 + x5 + · · · + 1 is a cyclotomic polynomial


[and hence irreducible]. So, [Q[ζ] : Q] = 6.

In the same way, φ5 (η) = 0, where φ5 (x) = x4 + x3 + · · · + 1 is also a cyclotomic


polynomial [and hence also irreducible]. So, [Q[η] : Q] = 4. Since 4 - 6, by Corollary
13.3.6, η 6∈ Q[ζ].

3. (b) We have that ζ6 is a root of x6 − 1. This factors as x6 − 1 = (x3 − 1)(x3 + 1). Since

ζ63 = −1 [just compute ζ63 = (cos(π/6) + i sin(π/6))3 = ( 3/2 + i/2)3 ], we have
that ζ6 is a root of x3 +1. But, −1 is a root of x3 +1, so x3 +1 = (x+1)(x2 −x+1).
So, one sees that minζ3 ,Q = x2 − x + 1 [since it has ζ3 as a root and is irreducible,
since it has no roots in Q].

(f) We have that ζ6 is a root of x12 − 1. This factors as x12 − 1 = (x6 − 1)(x6 + 1).
6
We have that ζ12 6= 1, since 12 is the smallest positive power that makes ζ12 one.
[Just remember that geometric interpretation of ζ12 , and its powers.] Hence ζ12
is a root of x6 + 1 = (x2 )3 + 1.
Comparing with (b), we see that x6 + 1 = (x2 )3 + 1 = ((x2 ) + 1)((x2 )2 − x2 + 1) =
(x2 + 1)(x4 − x2 + 1). Since ζ6 = ζ12
2
, we see that ζ12 is a root of x4 − x2 + 1.
def
We now prove that f (x) = x4 − x2 + 1 is irreducible. One can quickly check that
it has no rational roots since the only possibilities are ±1. So, if it would factor,
it would factor as a product of two polynomials of degree 2 [and coefficients in
5
Q]. But since f is biquadratic, we can find all complex roots, which are ζ12 , ζ12 ,
7 5 11
ζ12 = ζ12 and ζ12 = ζ12 . But, the products of (x − α)(x − β), where α and β are
roots of f [listed above] never give you a polynomial with rational coefficients.
Hence, it f is irreducible [over Q].
Therefore, f = minζ12 ,Q .

5. If K has dimension one, then every non-zero “vector” [here just an element of K] forms
as basis, i.e., all elements of K are F -multiples of this non-zero element. In particular,

26
we can take this element to be 1 [since 1 6= 0 in a field], and all F -multiples of 1 is
obviously F . Hence K = F .
def def √
6. We have that f = x4 − a has α = 4
a as a root and coefficients in Q. So, we must
show that this polynomial is irreducible over Q.

This polynomial f factors in C as (x − α)(x + α)(x − iα)(x + iα). First observe that f

has no roots in Q, since ±iα 6∈ R and if ±α ∈ Q, then (±α)2 = a ∈ Q, which means
that a is a square in Q, contradicting the hypothesis. So, if f factors, it must factor
as a product of two polynomials of degree 2.
√ √
Since (x − α)(x + α) = x2 − a, and a 6∈ Q [as discussed above], this cannot be a
factor.

Also, (x − α)(x ± iα) = x2 − (α ± iα)x ± iα2 , has non-real coefficients, so it cannot be


a factor either.

Hence, f has no factor of degree 2 in Q[x], and therefore, f is irreducible.



7. (a) We have that [Q[ −2] : Q] = 2, since x2 + 2 is irreducible. Suppose that i ∈
def √ √ √ √
K = Q[ −2]. Then, 2 = (i 2)/i ∈ K, and thus, Q ⊆ Q[ 2] ⊆ K. But since
√ √ √
[K : Q] = 2, by Corollary 13.3.7, Q[ 2] = K [since 2 6∈ Q]. But Q[ 2] ⊆ R,
while K * R. So, this is a contradiction and i 6∈ K.
def
(c) First note that since f = x3 +x+1 has no roots in Q [the only possible roots would
def
be ±1], it must be irreducible [since f has degree 3]. Therefore, if K = Q[α], we
have [K : Q] = 3.
If i ∈ K, then, by Corollary 13.3.6, deg(mini,Q ) = deg(x2 + 1) = 2 must divide
[K : Q] = 3. Since this is false, i 6∈ K.

8. We have F ⊆ F [α] ⊆ F [α, β] = K, so [F [α] : F ] = m | [K : F ]. In the same way,


F ⊆ F [β] ⊆ F [α, β] = K, so [F [α] : F ] = n | [K : F ]. Since m and n are relatively
primes, we have that mn | [K : F ].
def
Also, since f = minα,F ∈ F [x] ⊆ F [β][x], and f (α) = 0, we have that [F [α, β] :
F [β]] ≤ deg(minα,F ) = m. So, [K : F ] = [K : F [β]] · n ≤ mn. But, since mn | [K : F ],
we must have [K : F ] = mn.

11. We have that f is irreducible in K[x] iff f = minα,K , since f (α) = 0 and f ∈ Q[x] ⊆
K[x]. In the same way, g is irreducible in F [x] iff g = minβ,F .

27
So,

f is irreducible in K[x] ⇔
f = minα,K ⇔
[Q[α, β] : K] = deg f ⇔
[Q[α, β] : Q] = [Q[α, β] : K][K : Q] = deg f · deg g [since g = minβ,Q ] ⇔
[Q[α, β] : Q] = [Q[α, β] : F ][F : Q] = deg f · deg g ⇔
[Q[α, β] : F ] = deg g [since f = minα,Q and so [F : Q] = deg f ] ⇔
g = minβ,F ⇔
g is irreducible in F [x].

Section 13.5
4. Let f = (x − α1 ) · · · (x − αn ) ∈ F [x]. [Note that the αi ’s might not be in F .] We shall
prove, by induction on n = deg f , that [F [α1 , . . . , αn ] : F ] ≤ n!

For n = 1, we clearly have that α1 ∈ F [since f = x − α1 ∈ F [x]], and the bound is


satisfied, since [F [α1 ] : F ] = 1 ≤ 1!.

Assume now that if g = (x − β1 ) · · · (x − βn−1 ) ∈ K[x] [where K is any field], then


[K[β1 , . . . , βn−1 ] : K] ≤ (n − 1)!. [Induction Hypothesis.]

Let f = (x − α1 ) · · · (x − αn ) ∈ F [x]. Then, we have that [F [αn ] : F ] ≤ deg f = n


[since f (αn ) = 0], and we have equality iff f is irreducible [since then f = minαn ,F ].
def
Now, in F [αn ][x], we can divide f by x − αn , and so g = (x − α1 ) · · · (x − αn−1 ) ∈
F [αn ][x]. By the induction hypothesis [since deg g = (n−1) and taking K = F [αn ]], we
have [F [αn ][α1 , . . . , αn−1 ] : F [αn ]] ≤ (n − 1)!. But F [αn ][α1 , . . . , αn−1 ] = F [α1 , . . . , αn ]
and hence, [F [α1 , . . . , αn ] : F ] = [F [α1 , . . . , αn ] : F [αn ]] · [F [αn ] : F ] ≤ (n − 1)! · n = n!,
which concludes the proof.

√ √ √
Note: For f = x3 − 2, we have that f = (x − 3 2)(x − ζ3 3 2)(x − ζ32 3 2), where
def √ √ √ √
ζ3 = e2πi/3 , and [Q[ 3 2, ζ3 3 2, ζ32 3 2 : Q] = [Q[ 3 2 : ζ3 ] : Q] = 6 = 3!. So, the bound
seems to be sharp [i.e., cannot be improved in this generality]. In fact, one can prove
that for all n, there is f ∈ Q[x] whose roots give an extension of degree equal to n!.
[I.e., the bound is indeed sharp.]

28
Homework 10

Section 15.1
√ √ √ √
2. We have Q ⊆ Q[ 2] ⊆ Q[ 2, i]. Since, Q[ 2] ⊆ R, but Q[ 2, i] * R, we have
√ √
that [Q[ 2, i] : Q[ 2]] > 1 [since they cannot be equal]. But, since i is a root of
√ √ √
x2 + 1 ∈ Q[ 2][x], we have that [Q[ 2, i] : Q[ 2]] ≤ 2. So, the degree is 2 and {1, i}
√ √
is a Q[ 2]-basis of Q[ 2, i].
√ √ √
Also, [Q[ 2] : Q] = 2 and {1, 2} is a Q-basis of Q[ 2].

As seen in the proof of Theorem 13.3.4, the Q-basis of Q[ 2, i] is given by the products
√ √ √ √
of element of these two basis, i.e., {1 · 1, 1 · 2, i · 1, i · 2} = {1, 2, i, i · 2}.

3. This is just an application of the example done in class for biquadratic extensions.
def √ √ def
Let’s denote K = Q[ 2, 3], and G = G(K/Q). Then, since K/Q is biquadratic,
G∼= K4 [the Klein 4-group], and G = {id, σ, τ, σ ◦ τ }, where
√ √ √ √
σ : 2 7→ − 2 τ : 2 7→ 2
√ √ √ √
3 7→ 3 3 7→ − 3
def def
[and both fix Q, of course]. The subgroups of G then are {id}, H1 = {id, σ}, H2 =
def
{id, τ }, H3 = {id, σ ◦ τ }, and G.
By the Main Theorem of Galois Theory, and since K/Q is Galois [as we’ve seen in
class, or it is shown in the book], we have three proper intermediate extensions of

degree 2, which are associated to H1 , H2 , and H3 . We have K H1 = Q[ 3], since this

is an extension of degree 2 fixed by H1 , and in the same way, we have K H2 = Q[ 2].
√ √ √ √
Finally, we have that 2 · 3 = 6 is fixed by σ ◦ τ , and so K H3 = Q[ 6]. [Also, we
have the trivial intermediate extensions K {id} = K and K G = Q.]
√ √ √
So, the intermediate extensions of K/Q are Q, Q[ 2], Q[ 3], Q[ 6], and K.
def
6. (a) We have f = x4 − 1 = (x2 − 1)(x2 + 1) = (x − 1)(x + 1)(x − i)(x + i). So, the
splitting of x4 − 1 is Q[i] and hence has degree 2 over Q.
(c) We have

x4 + 1 = (x2 − i)(x2 + i)
√ ! √ ! √ ! √ !
2 2 2 2
= x− (1 + i) x− (−1 − i) x− (−1 + i) x− (1 − i)
2 2 2 2

29
def √ √
Let α = 2/2·(1+i). Then i = α2 , and so i ∈ Q[α]. Then, (1−i)·α = 2 ∈ Q[α].
√ √ √
So, Q[ 2, i] ⊆ Q[α]. Clearly, we also have α ∈ Q[ 2, i], and so Q[α] = Q[ 2, i],
and we shall denote this field by K.
Thus, since the splitting field clearly contains K = Q[α] [it contains all roots of
√ √
f ], and K = Q[ 2, i] contains all roots of f [since we have both 2 and i], K is
the splitting field of f over Q.
√ √
Note that since Q & Q[ 2] & Q[ 2, i] = K, and each intermediate extension has
degree at most 2, we have [K : Q] = 4. [In particular, since K = Q[α], this shows
that f = minα,Q , and hence irreducible.]

9. Let β 2 = b, with β ∈ K and b ∈ F . Since K = F [α] is quadratic over F , there are


c, d ∈ F such that β = c + dα. Then, we have b = (c + dα)2 = ((c2 + d2 a) + 2cdα).
Thus, (c2 + d2 a) = b and 2cd = 0 [since {1, α} is an F -basis], and then either c = 0 or
d = 0 [since a field has no zero divisors].

If d = 0, then β ∈ F , which obviously has its square in F .

If c = 0, then β = dα, and all these elements indeed have their square in F .

So, the elements of K whose squares are in F are the elements of F and the elements
of the form dα, with d ∈ F .

13. Let {α1 , . . . , αn } be an F -basis of K. [Since it has finite degree, we have a finite basis.]
Thus, K = F [α1 , . . . , αn ].
def def def
For i ∈ {1, . . . , n}, let fi = minαi ,F and di = deg fi . Let σ ∈ G = G(K/F ). Clearly,
σ is completely determined by the values of σ(αi ) [since σ is an automorphism which
fixes F and the αi ’s generate K over F ].

On the other hand, for each i, σ(αi ) must be a root of fi [by Proposition 14.1.14(a)],
and fi has at most di distinct roots, giving us at most di possible values for σ(αi ).

So, we have at most d1 possibilities for σ(α1 ), d2 possibilities for σ(α2 ), etc. Since σ is
determined by those values, we have at most d1 · · · dn possible elements of G. Hence,
G is finite.

15. Let K/F such that G = G(K/F ) ∼


def
= K4 . So, we have G = {id, σ, τ, σ ◦ τ }, with
def def
σ 2 = τ 2 = id and σ ◦ τ = τ ◦ σ. The subgroups are {id}, H1 = {id, σ}, H2 = {id, τ },
def def
H3 = {id, σ ◦ τ }, and G. Let L1 = K H1 .

30
Since |H1 | = 2, by the Main Theorem of Galois Theory [and K/F is Galois by hypoth-
esis], we have [K : L1 ] = 2 and, and since [K : F ] = 4 [since |G| = |K4 | = 4 and K/F
is Galois], we have hence [L1 : F ] = 2. By Proposition 13.3.3 [done in class], [note
that the book assumes the fields have characteristic 0, but it suffices for this exercise
that the characteristic is not 2] we have that there is α ∈ L1 such that α2 ∈ F and
def
L1 = F [α]. In the same way, if L2 = K H2 , then [L2 : F ] = 2 and there is β ∈ L2 such
that β 2 ∈ F and L2 = F [β].

Note that σ(β) 6= β, or else, since β ∈ L2 = K H2 , and we would have β ∈ K G = F


[since β would be fixed by both σ and τ ], which cannot be since [F [β] : F ] > 1. Thus,
β 6∈ K H1 = F [α], i.e., [F [α, β] : F [α]] > 1. So F [α] & F [α, β] ⊆ K, and thus, since
[K : F [α]] = 2, we must have K = F [α, β].

Thus, K = F [α, β], with α2 , β 2 ∈ F and [K : F ] = 4. Thus, K/F is biquadratic.

31

You might also like