2017 Bookmatter FromNaturalNumbersToQuaternion
2017 Bookmatter FromNaturalNumbersToQuaternion
2017 Bookmatter FromNaturalNumbersToQuaternion
In this chapter we present solutions to some exercises along with hints for
helping to solve others.
Exercise 1.8. We prove first the validity of the associative law of addition,
that is, the equality
n + (m + p) = (n + m) + p (1)
N
for all n, m, p ∈ . We do this by induction on p. For p = 0, the assertion is
clear. Suppose, then, that assertion (1) holds for an arbitrary but fixed p ∈ N
N
and for all n, m ∈ . Then from the definition of addition and the induction
hypothesis, we have the equality
(n + m) · p = n · p + m · p (2)
N
for all n, m, p ∈ , can be proved using the associativity and commutativity
of addition along with induction on p. For p = 0, the assertion is clear. Sup-
N
pose now that (2) holds for an arbitrary but fixed p ∈ and for all n, m ∈ . N
Then we have
(n + m) · p∗ = (n + m) · p + (n + m) = (n · p + m · p) + (n + m)
= (n · p + n) + (m · p + m) = n · p∗ + m · p∗ ,
as desired. This proves the asserted distributivity. Using the first distributive
law, one can then prove the commutativity of multiplication by induction.
This then also yields the validity of the second distributive law. Finally, one
can prove the associativity of multiplication by induction.
N
Exercise 1.10. Let m, n ∈ . If m = 0 or n = 0, then one immediately has
the equality m · n = 0 by Definition 1.5 (2) and the commutativity of mul-
m · n = m · b ∗ = ( m · b ) + m = ( m · b ) + ( a + 1) = ( m · b + a ) + 1 = ( m · b + a ) ∗ ,
that is, the natural number m · n is the successor of the natural number
m · b + a. By the third Peano axiom, we must then have m · n 6= 0.
Exercise 1.14. The power law from Lemma 1.13 can by proved by induction.
Exercise 1.17. The proof of properties (i), (ii), and (iii) of Remark 1.16 are left
to the reader.
Exercise 1.20. Properties (i) and (ii) of Remark 1.19 can be proved by induc-
tion.
m = n + x = x + n < y + n = n + y = m,
that is, we conclude that m < m, a contradiction. If x > y, then there fol-
lows analogously the contradiction m > m. Therefore, we must have x = y,
proving the asserted uniqueness.
Exercise 2.5. (a) By assumption we have 3 | ( a1 · · · ak + 1), that is, there exists
n∈ N with a1 · · · ak + 1 = 3 · n. If now 3 | a j for some j ∈ {1, . . . , k }, then
by Lemma 2.4 (ix), we have 3 | ( a1 · · · ak ), that is, there exists m ∈ N with
a1 · · · ak = 3 · m. We therefore have the equality
1 = 3 · n − a1 · · · a k = 3 · n − 3 · m = 3 · ( n − m ),
k
a1 · · · a k − 1 = ∏ (3 · n j + 1) − 1
j =1
N
We claim that we have the equality an = 5 · bn + rn with bn ∈ and rn ∈ {2, 3}
N
(n ∈ , n ≥ 1). This can be proved by induction on n. If n = 1, the assertion is
clear. We now assume that the assertion holds for arbitrary but fixed n ∈ , N
n ≥ 1. Then we have
a n + 1 = ( a n − 1 ) · a n + 1 = ( 5 · bn + r n − 1 ) · ( 5 · bn + r n ) + 1
= 5 · (5bn2 + 2bn rn − bn ) + rn2 − rn + 1.
Exercise 2.13. Let us assume, contrary to the assertion, that there are only
N
finitely many prime numbers p1 , . . . , pn in 2 + 3 · . We then consider the
natural number
a : = 3 · p1 · · · p n − 1 .
We have that a > 1, and by Lemma 2.9, a has a prime divisor p. Since 3 - a,
N
it follows that p 6= 3. We now show that p ∈ 2 + 3 · ; that is, we must show
that 3 | ( p + 1). If p = a, we are done. If p < a, then there exists q ∈ , q > 1, N
with a = p · q. Since 3 | ( p · q + 1), it follows by Exercise 2.5 (b) that 3 | ( p + 1)
or 3 | (q + 1). In the first case, we are done. In the second case, we repeat the
process for a prime divisor of q. Finally, after finitely many steps, we obtain
N
a prime divisor p of a with p ∈ 2 + 3 · . We now proceed as in Euclid’s
proof, for on the assumption that there are only finitely many prime num-
N
bers in the set 2 + 3 · , we must have p ∈ { p1 , . . . , pn }. In particular, we have
p | ( p1 · · · pn ). However, since we have the divisibility relation p | a, we must
have p | 1 from the divisibility rules, which is a contradiction.
(i) Suppose that n is not a prime. Then there exist natural numbers a, b ∈ N
with n = a · b and 1 < a, b < n. We thus obtain
Exercise 2.18. Let { a1 , . . . , an } and {b1 , . . . , bm } denote the sets of all divisors
of a and all divisors of b. Since a and b are relatively prime, the set of all
divisors of a · b is equal to { a j · bk | j = 1, . . . , n; k = 1, . . . , m}. It follows that
n m
S( a) · S(b) = ( a1 + · · · + an ) · (b1 + · · · + bm ) = ∑ ∑ a j · bk = S ( a · b ) .
j =1 k =1
Therefore, we have S(220) = 220 + 284 = S(284), which proves that the
numbers 220 and 284 are amicable.
(b) We must show that S( a) = a + b = S(b). Since x, y, z are distinct odd
primes, it follows that
= 3 · 5 · 17 · 257 · 65 537 .
Exercise 3.7. Let a = 232 − 1 and b = 255 with prime factorizations (see Ex-
ercise 3.2)
Exercise 4.6. With the help of Theorem 4.3, we obtain (3 600, 3 240) = 360,
(360360 , 540180 ) = ((23 · 32 · 5)360 , (22 · 33 · 5)180 ) = 2360 · 3540 · 5180 , (232 −
1, 38 − 28 ) = 5, where for the last equality, we used the prime factorization
from Exercise 3.2 and the prime factorization 38 − 28 = (32 − 22 ) · (32 + 22 ) ·
(34 + 24 ) = 5 · 13 · 97.
Exercise 4.13. We have (2 880, 3 000, 3 240) = (120, 3 240) = 120 and [36, 42, 49]
= [252, 49] = 1 764.
Exercise 4.17. Let a1 , . . . , an ∈ N. We leave to the reader the proof of the fol-
lowing equivalence:
Exercise 5.4. This process can be carried out for arbitrary natural numbers
g > 1. One obtains the unique representation
n = q` · g` + q`−1 · g`−1 + · · · + q1 · g1 + q0 · g0
( a ⊕ b) ⊕ c = Rn ( a + b) ⊕ c = Rn (Rn ( a + b) + c),
a ⊕ (b ⊕ c) = a ⊕ Rn (b + c) = Rn ( a + Rn (b + c)).
a + b = q1 · n + R n ( a + b ) and b + c = q2 · n + R n ( b + c ),
whence follows
Rn (Rn ( a + b) + c) = Rn ( a + b + c − q1 · n) = Rn ( a + b + c)
= Rn ( a + b + c − q2 · n) = Rn ( a + Rn (b + c)).
N N
Exercise 1.4. (a) The set 2 · = {2 · n | n ∈ } of even natural numbers is a
N
nonempty subset of . If 2 · m, 2 · n ∈ 2 · , then N
2 · m + 2 · n = 2 · (m + n) ∈ 2 · N, (2 · m) · (2 · n) = 2 · (m · 2 · n) ∈ 2 · N.
N
Thus both + and · are operations on 2 · . Since the operations + on and N
N
· on are associative, it follows that in particular, the operations + on 2 · N
N N
and · on 2 · are associative. Therefore, both (2 · , +) and (2 · , ·) are N
semigroups.
N N
The set 2 · + 1 = {2 · n + 1 | n ∈ } of odd natural numbers is a
N
nonempty subset of . If 2 · m + 1 and 2 · n + 1 are in 2 · + 1, then N
(2 · m + 1) + (2 · n + 1) = 2 · (m + n + 1) ∈ 2 · N,
(2 · m + 1) · (2 · n + 1) = 2 · (m · 2 · n + m + n) + 1 ∈ 2 · N + 1.
Solutions to Exercises in Chapter II 253
N
Therefore, while · is an operation on 2 · + 1, we see that + is not an opera-
N N
tion on 2 · + 1. Therefore, (2 · + 1, +) is not a semigroup. The operation ·
N
on is associative, and so in particular, the operation · on 2 · + 1 is asso- N
N
ciative. Therefore, (2 · + 1, ·) is a semigroup.
N N N
(b) Let k ∈ , k > 1. The set k · = {k · n | n ∈ } is a nonempty subset of
N N N
. One shows, as in (a), that both (k · , +) and (k · , ·) are semigroups.
map( A1 ) = {id},
map( A2 ) = {id, f , g, . . . },
( f ◦ g)( a1 ) = f ( g( a1 )) = f ( a1 ) = a2 6= a1 = g( a2 ) = g( f ( a1 )) = ( g ◦ f )( a1 ),
Exercise 1.12. Let e` be a left identity element and er a right identity element
of H. Then
e ` = e ` ◦ er = er ,
where the first equality follows from the fact that er is a right identity el-
ement of H, and the second from the fact that e` is a left identity element
of H.
N N
Exercise 1.14. (a) By Exercise 1.4, (2 · , +) and (2 · , ·) are semigroups.
It remains to show that there exists an additive identity element in 2 · . N
By the definition of addition, 0 is this element. Since 1 6∈ 2 · , there is noN
N
multiplicative identity element, so that (2 · , ·) is only a semigroup.
(b) We leave it to the reader to find other examples of semigroups that are
not monoids.
254 Solutions to Exercises
Exercise 2.3. (a) Suppose that g0 and g00 are two inverse elements to an ele-
ment g ∈ G. Then
where the second equality follows from the fact that g00 is in particular a
right inverse to g, and the fourth equality follows from the fact that g0 is in
particular a left inverse to g.
(b) Let g`0 be a left inverse and gr0 a right inverse to an element g ∈ G. Then it
follows that
g ◦ g−1 = e = g−1 ◦ g.
Thus h−1 ◦ g−1 is the inverse element to g ◦ h, that is, ( g ◦ h)−1 = h−1 ◦ g−1 .
The calculational rules (c) and (d) follow directly from the definition.
Exercise 2.9. We compare only the groups that have the same numbers of el-
ements. The Cayley tables of the groups (R4 , ⊕), (R5 \ {0}, ), and ( D4 , ◦)
have, reading from left to right, the following form:
⊕ 0 1 2 3 1 2 3 4 ◦ d0 d1 s0 s1
0 0 1 2 3 1 1 2 3 4 d0 d0 d1 s0 s1
1 1 2 3 0 2 2 4 1 3 d1 d1 d0 s1 s0
2 2 3 0 1 3 3 1 4 2 s0 s0 s1 d0 d1
3 3 0 1 2 4 4 3 2 1 s1 s1 s0 d1 d0
One may conclude from the Cayley tables that all three groups under con-
sideration are abelian. We now determine the smallest nonzero natural num-
ber n such that gn = e for g 6= e. In the group (R4 , ⊕), we have e = 0 and
12 = 2, 13 = 3, 14 = 0; 22 = 0; 32 = 2, 33 = 1, 34 = 0.
22 = 4, 23 = 3, 24 = 1; 32 = 4, 33 = 2, 34 = 1; 42 = 1.
Thus in each group there are two elements with n = 4 and one element with
n = 2. In the group ( D4 , ◦), however, d21 = s20 = s21 = e with e = d0 , that is,
there is no element with n = 4.
The Cayley tables for (R6 , ⊕) and ( D6 , ◦), reading from left to right, have
the following form:
⊕ 0 1 2 3 4 5 ◦ d0 d1 d2 s0 s1 s2
0 0 1 2 3 4 5 d0 d0 d1 d2 s0 s1 s2
1 1 2 3 4 5 0 d1 d1 d2 d0 s2 s0 s1
2 2 3 4 5 0 1 d2 d2 d0 d1 s1 s2 s0
3 3 4 5 0 1 2 s0 s0 s1 s2 d0 d1 d2
4 4 5 0 1 2 3 s1 s1 s2 s0 d2 d0 d1
5 5 0 1 2 3 4 s2 s2 s0 s1 d1 d2 d0
From these tables, one can see that the group (R6 , ⊕) is abelian. The group
( D6 , ◦) is nonabelian, since s0 ◦ s1 = d2 6= d1 = s1 ◦ s0 . We again determine
the smallest nonzero natural number n such that gn = e for g 6= e. In (R6 , ⊕),
we have e = 0 and
12 = 2, 13 = 3, 14 = 4, 15 = 5, 16 = 0; 22 = 4, 23 = 0; 32 = 0;
42 = 2, 43 = 0; 52 = 4, 53 = 3, 54 = 2, 55 = 1, 56 = 0.
12
1 12
2 21
for (R3 \ {0}, ) and the Cayley table from Exercise 2.9 for (R5 \ {0}, )
that (R3 \ {0}, ) and (R5 \ {0}, ) are groups.
(b) We leave to the reader the task of verifying the assertions of Exam-
ple 2.8 (iii) regarding the dihedral group ( D2n , ◦).
(c) Let n ≥ 3. We consider the elements
1 2 3 ··· n 1 2 3 ··· n
π1 = , π2 =
2 1 3 ··· n 3 1 2 ··· n
256 Solutions to Exercises
1 2 3 ··· n 1 2 3 ··· n
π1 ◦ π2 = 6= = π2 ◦ π1 ,
3 2 1 ··· n 1 3 2 ··· n
where for n > 3, each of the elements 4, . . . , n is mapped to itself. This proves
that (Sn , ◦) for n ≥ 3 is nonabelian.
123 123 12 3
π1 = , π2 = , π3 = ,
123 231 31 2
123 123 12 3
π4 = , π5 = , π6 = .
132 321 21 3
h π1 i = { π1 }, h π2 i = { π1 , π2 , π3 } = h π3 i ,
h π4 i = { π1 , π4 }, h π5 i = { π1 , π5 }, h π6 i = { π1 , π6 } ,
and the subgroup S3 itself, which is not cyclic. One can see that S3 has no
other subgroups.
f ( d0 ) = π1 , f ( d1 ) = π3 , f ( d2 ) = π2 ,
f ( d0 ◦ s0 ) = π6 , f ( d1 ◦ s0 ) = π5 , f ( d2 ◦ s0 ) = π4 ,
It follows that
k k
sgn d j1 ◦ s01 ◦ d j2 ◦ s0k2 = k1 ⊕ k2 = sgn d j1 ◦ s01 ⊕ sgn d j2 ◦ s0k2 .
The group S4 has, therefore, only elements of orders 1, 2, 3, and 4. There can
therefore be no group isomorphism between D24 and S4 .
whence follows ker( f 1 ) = R4 , ker( f 2 ) = {0}, ker( f 3 ) = {0, 2}, ker( f 4 ) = {0},
im( f 1 ) = {0}, im( f 2 ) = R4 , im( f 3 ) = {0, 2}, im( f 4 ) = R4 . This shows in par-
ticular that f 2 and f 4 are bijective.
(b) Since R p = h1i, every group homomorphism g : (R p , ⊕) −→ (Rn , ⊕)
is uniquely determined by specifying the image g(1). One now shows that
only g(1) = 0 is possible, since n and p are relatively prime. There is, there-
fore, only one group homomorphism g, which is given by the assignment
g(m) = 0 (m ∈ R p ). We have ker( g) = R p and im( g) = {0}.
Exercise 4.3. (a) The verification of the statement of Example 4.2 is left to the
reader.
(b) The order relation ≤ is not an equivalence relation on , since ≤ is not N
symmetric.
N
(c) The relation ∼ is not an equivalence relation on , since ∼ is not transi-
tive.
Solutions to Exercises in Chapter II 259
π1 ◦ U = U = π4 ◦ U,
π2 ◦ U = {π2 , π6 } = π6 ◦ U,
π3 ◦ U = {π3 , π5 } = π5 ◦ U.
◦ e a b c ◦ e a b c
e e a b c e e a b c
a a b c e a a e c b
b b c e a b b c e a
c c e a b c c b a e
Both groups are abelian.
One shows further that up to group isomorphism, the only groups of order
6 are (R6 , ⊕) and ( D6 , ◦). Thus every abelian group of order 6 is isomorphic
to (R6 , ⊕), and every nonabelian group of order 6 is isomorphic to ( D6 , ◦).
Exercise 4.19. Exercises 4.11 and 4.12 can be solved analogously for right
cosets.
ker( f ) = { g ◦ K | f ( g ◦ K ) = H } = { g ◦ K | g ◦ H = H }
= { g ◦ K | g ∈ H } = H/K.
Solutions to Exercises in Chapter III 261
as claimed.
Exercise 6.6. (a) The solution to this exercise is left to the reader.
N N N
(b) On (2 · + 1) × (2 · + 1) = {( a, b) | a, b ∈ 2 · + 1} we define the
relation
( a, b) ∼ (c, d) ⇐⇒ a · d = b · c ( a, b, c, d ∈ 2 · N + 1).
N
If we write ba for the equivalence class [ a, b] of ( a, b) ∈ (2 · + 1) × (2 · + 1), N
N N
then the group G := ((2 · + 1) × (2 · + 1))/ ∼ can be identified with the
N
set of all fractions of the form ba , where a, b ∈ 2 · + 1 and a, b are relatively
prime. We leave the detailed construction from Theorem 6.5 to the reader.
Z
Exercise 7.6. The generalization to of the addition and multiplication rules
in Remark 1.19 of Chapter I is left to the reader.
Exercise 7.9. The verification of the assertions of this example are left to the
reader.
Exercise 1.2. The proof of the calculational laws from Lemma 1.1 are left to
the reader.
262 Solutions to Exercises
a = e · p1 · · · pr = e · q1 · · · q s
p2 · · · pr = q2 · · · q s . (3)
Since p2 divides the left-hand side of (3), p2 must also divide the right-hand
side. As in the first step, we conclude that p2 = q2 . Proceeding in this way,
we obtain the equalities r = s and p j = q j for j = 1, . . . , r, which proves the
asserted uniqueness.
Exercise 2.5. We leave it to the reader to prove that the polynomial ring
( R[ X ], +, ·) is a commutative ring if and only if ( R, +, ·) is commutative.
(0 + f )( a) = 0( a ) + R f ( a ) = 0R + R f ( a) = f ( a)
def. of + def. of 0 0R id. el.
w.r.t. + R
and analogously the equality ( f + 0)( a) = f ( a) for all a ∈ A. The other ring
properties of map( A, R) are proved similarly using the ring properties of R.
Exercise 2.11. If n > 1 is not prime, then there exist natural numbers a, b ∈
Rn , a > 1, b > 1, with a · b = n. But then we have a b = 0. Therefore, a and
b are zero divisors of Rn .
Exercise 2.12. We show here only that the lack of zero divisors in ( R, +, ·)
implies the lack of zero divisors in ( R[ X ], +, ·). We assume that the ring R[ X ]
has zero divisors and show that then the ring R must have zero divisors. Let,
then, f ( X ) = an · X n + · · · + a1 · X + a0 (an 6= 0) and g( X ) = bm · X m + · · · +
b1 · X + b0 (bm 6= 0) be zero divisors in R[ X ], so that
f ( X ) · g( X ) = ( an · bm ) · X n+m + · · · + ( a1 · b0 + a0 · b1 ) · X + a0 · b0 = 0,
where 0 denotes the zero element of R[ X ], i.e., the zero polynomial. In par-
ticular, we must have an · bm = 0, which proves that R has zero divisors.
Exercise 2.13. The ring (map( A, R), +, ·) from Exercise 2.6 has as its zero
element 0 : A −→ R, a 7→ 0R . If now, for example, we have R = (R6 , ⊕, ),
then for f : A −→ R, a 7→ 2 (a ∈ A) and g : A −→ R, a 7→ 3 (a ∈ A), we have
the equality
( f · g)( a) = f ( a) g( a) = 2 3 = 0 = 0( a ),
def of · def of f , g
that is, f and g are zero divisors of (map( A, R), +, ·). Therefore, the ring
(map( A, R), +, ·) also has zero divisors if R has zero divisors. We note that
the ring (map( A, R), +, ·) can possess zero divisors even if R has no zero
divisors.
Z
Exercise 2.19. In the polynomial ring ( [ X ], +, ·), the unit element is given
by the unit polynomial 1. Let f ( X ) = an · X n + · · · + a1 · X + a0 (an 6= 0) be a
Z
unit in [ X ]. Then there exists a polynomial g( X ) = bm · X m + · · · + b1 · X +
Z
b0 (bm 6= 0) in [ X ] such that
f ( X ) · g( X ) = ( an · bm ) · X n+m + · · · + ( a1 · b0 + a0 · b1 ) · X + a0 · b0 = 1.
Exercise 2.20. That the units of a ring ( R, +, ·) with unit element 1 form a
group under multiplication with multiplicative identity element 1 follows
directly from how units are defined.
Z
Exercise 3.14. No, since for every subring U of ( , +, ·), we have in particu-
Z
lar that (U, +) is a subgroup of ( , +). Therefore, we must have U = n for Z
N
some n ∈ , which proves that (U, +) is an ideal of . Z
Exercise 3.15. For example, Z is a subring of the polynomial ring (Z[X], +, ·)
Z
that is not an ideal of [ X ].
Solutions to Exercises in Chapter III 265
Z Z
Exercise 3.16. The principal ideals of [ X ] are of the form { h · f | f ∈ [ X ]}
Z
for some h ∈ [ X ]. We leave it to the reader to show that the ideal
Z
a := {2 · f + X · g | f , g ∈ [ X ]}
Z
is not a principal ideal of ( [ X ], +, ·).
Z Z Z
Exercise 3.27. Let a ∈ . Then the mapping f : [ X ] −→ given by the as-
signment f ( X ) 7→ f ( a) is a surjective ring homomorphism with ker( f ) =
( X − a) by Exercises 3.2, 3.6, and 3.18. Invoking Corollary 3.25, we see that
Z [ X ]/( X − a) ∼ Z
= is a ring isomorphism, as desired.
Exercise 4.4. This is impossible, since one can show that every skew field
with finitely many elements is in fact a field.
Z N
Exercise 6.3. Let r = s/t for s ∈ and t ∈ \ {0}. Then r corresponds to the
Z Z
element [s, t] ∈ × ( \ {0}). If now d = (s, t) > 0 is the greatest common
divisor of s and t, then we may write s = d · a and t = d · b with a ∈ , Z
N
b ∈ \ {0}, and a, b relatively prime. From the equality s · b = (d · a) · b =
(d · b) · a = t · a we infer the equality [s, t] = [ a, b]. This proves the existence
266 Solutions to Exercises
Exercise 6.4. Since one usually learns this proof in a first course in real anal-
ysis, we leave this exercise to the reader.
Z
Exercise 7.4. We obtain in [ X ] the decompositions 20X = 22 · 5 · X and
10X 2 + 4X − 6 = 2 · ( X + 1) · (5X − 3). Therefore, 2 is the greatest common
divisor, and 22 · 5 · X · ( X + 1) · (5X − 3) = 100X 3 + 40X 2 − 60X the least
common multiple, of the polynomials 20X and 10X 2 + 4X − 6 in [ X ]. Z
Exercise 7.9. We leave it to the reader to come up with relevant examples.
Exercise 7.15. Part (ii) of the proof of Lemma 7.14 is left to the reader.
a := { X · f + Y · g | f , g ∈ K [ X, Y ]}
is not principal.
Exercise 7.38. (a) We obtain (123 456 789, 555 555 555) = 9.
(b) We calculate
X 4 + 2X 3 + 2X 2 + 2X + 1 = ( X + 1) · ( X 3 + X 2 − X − 1) + (2X 2 + 4X + 2) ,
1 1
X3 + X2 − X − 1 = X− · (2X 2 + 4X + 2) + 0 .
2 2
We thereby obtain ( X 4 + 2X 3 + 2X 2 + 2X + 1, X 3 + X 2 − X − 1) = 2X 2 +
4X + 2.
Solutions to Exercises in Chapter IV 267
123 123 41
= = .
103 − 1 999 333
Exercise 2.2. (a) Without loss of generality, we may assume that e ∈ Q, 0 <
N
e < 1. For n ∈ , we then have
1 1−e
n + 1 < e ⇐⇒ e < n.
If we set N (e) := [(1 − e)/e], where [ x ] is the greatest integer less than or
N
equal to x, then for all n ∈ with n > N (e), we have the inequality
1
n + 1 < e.
1
This proves that the sequence n+ 1 n≥0 is a rational null sequence. Using
the inequality
n 1
<
2n n+1
N
for n ∈ , n ≥ 5, one shows analogously that 2nn n≥0 is also a rational null
sequence.
268 Solutions to Exercises
1
(b) Other examples of rational null sequences are the sequences
N
with k ∈ , k ≥ 2.
( n +1) k n ≥0
N
for all n ∈ with n > N (q0 ) := max( N (q), N
e (e)). This proves the claimed
independence of representative.
So if we set
2 − a20 2 + a20
a1 : = a0 + = ,
2a0 2a0
√
then because a0 > 0, we have the inequality a21 > 2, that is, a1 > 2. This
implies 2 − a21 /(2a1 ) < 0, and we therefore have for
2 − a21 2 + a21
a2 : = a1 + =
2a1 2a1
Solutions to Exercises in Chapter IV 269
√
both a1 > a2 and a22 > 2, that is, a2 > 2. We consider now the rational se-
Q
quence ( an )n≥0 with a0 ∈ , a0 > 0, arbitrary and
2 + a2n
a n +1 : = (n ∈ N, n ≥ 1). (5)
2an
One first shows by induction
√ that both an > an+1 for all n ∈ , n ≥ 1, and N
2 N
an > 2, that is, an > 2, for all n ∈ , n ≥ 1. Using these inequalities, one
then shows in a second step that ( an )n≥0 is a rational Cauchy sequence. This
R
rational Cauchy sequence has limit α ∈ , α > 0. Because of the recurrence
formula (5), we see that α satisfies the equation
2 + α2 √
α= ⇐⇒ α = 2,
2α
as desired.
Exercise 3.12. (a) The decimal representation 0.101 001 000 100 001 . . . is nei-
ther terminating nor periodic. Therefore, this number cannot be rational.
One can find analogous examples, such as 0.121 331 222 133 331 . . . .
(b) Using the rational
√ Cauchy √series ( an )n≥0 constructed in Exercise 2.30
for calculating 2, one obtains 2 ≈ 1.4142135623, accurate to ten decimal
places, by choosing, for example, a0 = 1 and iterating four times.
n2 + 2
( a n ) n ≥0 : = .
2n n ≥0
n2 + 2 ( n + 1)2 + 2
2(n2 + 2) > (n + 1)2 + 2 ⇐⇒ n
> ⇐⇒ an > an+1
2 2n +1
N
for n ∈ , n ≥ 3, it follows that the sequence ( an )n≥0 is monotonically de-
creasing, but not strictly. One shows analogously that the sequences
n3 + 3
1
12 n+1 ,
n ≥0 3n n ≥0
n3 − 2
n2 − 2 n ≥0
270 Solutions to Exercises
Q
Exercise 4.6. The subset M ⊆ consisting of all sequence terms an (n > 0)
in the rational Cauchy sequence ( an )n≥0 constructed
√ in Exercise 2.30, which
is bounded below, has greatest lower bound 2 ∈ / . Q
√
Exercise 4.7. The greatest lower bound of the set { x x | x ∈ , x √
≥ 0} is at- Q
tained when x = 0 and is equal to zero. The least upper bound is e e.
Exercise 6.7. We leave the completion of the details in the sketch of the proof
of Theorem 6.5 to the reader.
Exercise 1.8. The completion of the proof of Theorem 1.7 is left to the reader.
C √
Exercise 1.10. Let α = α1 + α2 i ∈ , α 6= 0. If α2 = 0 and α1 > 0, then ± α1 are
the solutions to the equation x2 = α. If α2 = 0 and α1 < 0, then ± |α1 |i are
p
√
R
Since now β 1 ∈ , we need consider only the nonnegative solution y1 . With
√
β 1 = ± y1 and β 2 = ±α2 /(2 y1 ), we obtain the solution formula
α2 i
p
α1 + | α |
β=± √ ±q .
2 2 α1 + | α |
1+i
x1,2 = ± √ ,
2
Exercise 1.11. Since by Exercise 1.10 we have the equality ((1 + i )/2)2 = i/2,
we obtain, on completing the square,
1 + i 2 i
x2 + (1 + i ) · x + i = 0 ⇐⇒ x + + = 0.
2 2
If we now substitute y := x + (1 + i )/2, we obtain the quadratic equation
y2 = −i/2. With the solution formula from Exercise 1.10, we obtain the so-
lution
1+i 1 i 1+i
x1,2 = y1,2 − =± ∓ − ,
2 2 2 2
that is, x1 = −i and x2 = −1 are the solutions of x2 + (1 + i ) · x + i = 0, which
can be easily checked by substitution. One can prove analogously that the
equation x2 + (2 − i ) · x − 2i = 0 has the solutions x1 = i and x2 = −2.
272 Solutions to Exercises
Exercise 1.14. (a) First, one verifies the equality α · β = α · β for all α, β ∈ C.
One then has
| α · β |2 = ( α · β ) · α · β = α · α · β · β = | α |2 · | β |2 ,
Exercise 2.3. The statement of Remark 2.2 results immediately from use of
the product rule from Exercise 1.14.
R
Exercise 2.5. Let A ∈ M2 ( ) be an invertible matrix. One may convince one-
C R
self that the mapping f A : ( , +, ·) −→ (M2 ( ), +, ·) given by
α1 α2
α = α1 + α2 i 7 → A · · A −1
− α2 α1
where for the second equality we have invoked the addition theorems for
sine and cosine. From this follows for m ∈ the equalityN
αm = |α|m · cos(mϕ) + i sin(mϕ) .
α · β = ( α1 β 1 − α2 β 2 − α3 β 3 − α4 β 4 ) + ( α1 β 2 + α2 β 1 + α3 β 4 − α4 β 3 ) i
+ ( α1 β 3 − α2 β 4 + α3 β 1 + α4 β 2 ) j + ( α1 β 4 + α2 β 3 − α3 β 2 + α4 β 1 ) k
and
β · α = ( β 1 α1 − β 2 α2 − β 3 α3 − β 4 α4 ) + ( β 1 α2 + β 2 α1 + β 3 α4 − β 4 α3 ) i
+ ( β 1 α3 − β 2 α4 + β 3 α1 + β 4 α2 ) j + ( β 1 α4 + β 2 α3 − β 3 α2 + β 4 α1 )k,
however, we have
If α2 6= 0, then from the third equality, it follows that β 3 = (α3 α2−1 ) · β 2 for
H
every β ∈ . This contradiction implies that we must have α2 = 0. Similarly,
one shows that we must also have α3 = 0 and α4 = 0. Altogether, therefore,
R
we have that α = α1 ∈ . This proves the inclusion Z ( ) ⊆ . H R
Exercise 1.14. (a) Let α = α1 + α2 i + α3 j + α4 k. We then calculate
274 Solutions to Exercises
α · β = ( α2 i + α3 j + α4 k ) · ( β 2 i + β 3 j + β 4 k )
= (−α2 β 2 − α3 β 3 − α4 β 4 ) + (α3 β 4 − α4 β 3 )i
+ (−α2 β 4 + α4 β 2 ) j + (α2 β 3 − α3 β 2 )k
= − Im(α)t , Im( β)t + Im(α)t × Im( β)t · i,
2 · hα, βi · β = α · β · β + β · α · β,
| α · β |2 = ( α · β ) · α · β = α · ( β · β ) · α = | β |2 · ( α · α ) = | β |2 · | α |2 = | α |2 · | β |2 ,
Exercise 1.21. The completion of the proof of Theorem 1.20 is left to the
reader.
Solutions to Exercises in Chapter VI 275
Exercise 2.3. The statement of Remark 2.2 is an immediate result of the prod-
uct rule from Exercise 1.19.
C
Exercise 2.5. Let A ∈ M2 ( ) be an arbitrary invertible matrix. One may con-
H C
vince oneself that the mapping f : ( , +, ·) −→ (M2 ( ), +, ·) given by
α1 + α2 i α3 + α4 i
α = α1 + α2 i + α3 j + α4 k 7 → A · · A −1
− α3 + α4 i α1 − α2 i
that is, f is R-linear. The verification of the remaining assertions are left to
the reader.
R
Exercise 3.6. We first prove that every -linear mapping v 7→ A · v (v ∈ 3 ) R
R R
with A ∈ SO3 ( ) is an orientation-preserving rotation of 3 about an axis
passing through the origin. To this end, we begin by noting that for A ∈
R
SO3 ( ), on account of
10 0
S −1 · A · S = 0 α β
0γδ
R R
for certain α, β, γ, δ ∈ . But since because of A, S ∈ SO3 ( ), we have also
R
the inclusion S−1 · A · S ∈ SO3 ( ) and thereby
R
αβ
∈ SO2 ( ),
γδ
1 0 0
R
The mapping v 7→ D ϕ · v (v ∈ 3 ) is an orientation-preserving rotation in the
x2 , x3 -plane through the angle ϕ about the x1 -axis (counterclockwise if a1
points toward the observer). Altogether, we have shown that the mapping
R
v 7→ A · v (v ∈ 3 ) is an orientation-preserving rotation in the a2 , a3 -plane
through the angle ϕ about the a1 -axis (counterclockwise if a1 points toward
the observer).
We now describe, conversely, an arbitrary orientation-preserving rotation
R
of 3 through the angle ϕ ∈ [0, 2π ) about an axis passing through the origin
that is determined by the vector (normed to have length 1) a1 := (ν1 , ν2 , ν3 )t ∈
R 3 . To this end, we first consider the matrices
√ ν21 2 0 √ ν23 2
q
ν + ν ν + ν ν12 + ν32 ν2 0
R
1 3 1 3
D1 := 0 1 0 , D2 := ∈ SO3 ( ).
q
− ν ν 2 + ν2 0
2 1 3
− √ ν23 2 0 √ ν21 2
ν1 +ν3 ν1 +ν3 0 0 1
Solutions to Exercises in Chapter VI 277
where we have set µ := 1 − cos( ϕ); this can now be easily decomposed as
with
0 −ν3 ν2
N := ν3 0 −ν1 ,
−ν2 ν1 0
as asserted.
Selected Literature
The following list of books on (elementary) number theory and algebra can
serve to fill in some of the gaps in this book’s presentation. Some of these
books will take the reader much deeper into various topics. The literature
on the concept of number and the representation of numbers is of cultural-
historical significance, while the works of a historical nature provide insight
into the historical development of algebra and number theory. Finally, we
offer the interested reader two books on approaches to the teaching of alge-
bra and number theory.
Selected literature for the appendices is listed at the end of the respective
appendix.
[37] A. Arcavi, P. Drijvers, K. Stacey: The learning and teaching of algebra. IM-
PACT Series, Routledge, 2016.
[38] J. D. Sally, P. J. Sally: Integers, fractions, and arithmetic: a guide for teachers.
American Mathematical Society, Providence, RI, 2012.
Index