Ma103 Notes
Ma103 Notes
2018/19
MA103
Introduction to Abstract Mathematics
Second part, Analysis and Algebra
Amol Sasane
Revised by Jozef Skokan, Konrad Swanepoel, and Graham Brightwell
Copyright
c London School of Economics 2016
All rights reserved. No part of this work may be reproduced in any form, or by any
means, without permission in writing from the author. This material is not licensed for
resale.
Contents
1 Analysis 1
1.1 The real numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Intervals and absolute values . . . . . . . . . . . . . . . . . . . . . 6
1.2 Sequences and limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Limit of a convergent sequence . . . . . . . . . . . . . . . . . . . 10
1.2.3 The sequence (xn )n∈N . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2.4 Bounded and monotone sequences . . . . . . . . . . . . . . . . . . 16
1.2.5 Algebra of limits . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.2.6 The Sandwich Theorem . . . . . . . . . . . . . . . . . . . . . . . 25
1.2.7 Subsequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.3 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.3.1 Definition of continuity . . . . . . . . . . . . . . . . . . . . . . . . 30
1.3.2 Continuous functions preserve convergent sequences . . . . . . . . 34
1.3.3 Restrictions and compositions of functions . . . . . . . . . . . . . 36
1.3.4 Intermediate value theorem . . . . . . . . . . . . . . . . . . . . . 37
1.3.5 Extreme value theorem . . . . . . . . . . . . . . . . . . . . . . . . 40
1.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2 Algebra 55
2.1 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.1.1 Definition of a Group . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.1.2 Proving theorems, laws of exponents and solving equations in groups 63
2.1.3 Abelian groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.1.4 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.1.5 The order of an element of a group . . . . . . . . . . . . . . . . . 68
2.1.6 Homomorphisms and isomorphisms . . . . . . . . . . . . . . . . . 69
2.1.7 Cosets and Lagrange’s theorem . . . . . . . . . . . . . . . . . . . 72
2.1.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.2 Vector spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.2.1 Definition of a vector space . . . . . . . . . . . . . . . . . . . . . 83
2.2.2 Subspaces and linear combinations . . . . . . . . . . . . . . . . . 87
2.2.3 Basis of a vector space . . . . . . . . . . . . . . . . . . . . . . . . 90
iii
Contents
iv
Chapter 1
Analysis
Analysis is the theory behind real numbers, sequences, and functions. The word ‘theory’
is important. You might, for example, have a good idea of what we mean by a ‘limit’ of
a convergent sequence of numbers, or the notion of a ‘continuous’ function, but in this
part of the course we aim to formalize such notions.
does not have a largest element in Q. So we see that the rational number system has
“gaps”. The real number system R includes numbers to fill all of these gaps. Thus the set
n o
T = x ∈ R x2 ≤ 2
has a largest element. This is a consequence of a very important property of the real
numbers, called the least upper bound property. Before we state this property of R, we
need a few definitions.
1
1 Analysis
Examples.
3. The set S = {(−1)n | n ∈ N} is bounded. Note that S = {−1, 1}. It is bounded above
by 1 and bounded below by −1.
More generally, any finite set S is bounded.
n o
4. The set S = n1 n ∈ N is bounded. Any real number x satisfying 1 ≤ x is an upper
5. The sets Z and R are neither bounded above nor bounded below.
7. The set ∅ is bounded. For instance, 1 is an upper bound for ∅, since the condition
“for every element x of ∅, x ≤ 1” is satisfied: there is certainly no element x of ∅
that doesn’t satisfy x ≤ 1. Indeed, every real number is an upper bound for ∅, and
similarly every real number is a lower bound for ∅.
We now introduce the notions of a least upper bound (also called supremum) and a
greatest lower bound (also called infimum) of a subset S of R.
2
1.1 The real numbers
Theorem 1.1.1. If the least upper bound of a subset S of R exists, then it is unique.
Proof. The statement of the theorem means that the set S cannot have two different
least upper bounds. Suppose then that u∗ and u0∗ are two least upper bounds of S. Then
in particular u∗ and u0∗ are also upper bounds of S. Now since u∗ is a least upper bound
of S and u0∗ is an upper bound of S, it follows that
u∗ ≤ u0∗ . (1.1)
Definitions.
When the supremum and the infimum of a set belong to the set, we give them the
following familiar special names:
3
1 Analysis
Definitions.
Examples.
1. If S = {x ∈ R | 0 ≤ x < 1}, then sup S = 1 6∈ S and so max S does not exist. But
inf S = 0 ∈ S, and so min S = 0.
2. If S = N, then sup S does not exist, inf S = 1, max S does not exist, and min S = 1.
5. For the sets Z and R, sup, inf, max, min do not exist.
In the above examples, we note that if S is non-empty and bounded above, then its
supremum exists. In fact this is a fundamental property of the real numbers, called the
least upper bound property of the real numbers, which we now state:
1. S 6= ∅, and
Examples.
1. Use the least upper bound property to show that there exists a number s ∈ R such
that s > 0 and s2 = 2.
Solution. Let S = {x ∈ R | x2 < 2}. We’ll show that S has a least upper bound
s = sup S and that this s is the real number we want: s > 0 and s2 = 2.
In order to apply the least upper bound property, we first have to show that S 6= ∅
and that S has an upper bound. This is easy: since 12 ≤ 2, 1 ∈ S and therefore S 6= ∅.
Also, since for any x ∈ S, x2 ≤ 2 < 4, it follows that x < 2, which shows that 2 is an
upper bound of S.
4
1.1 The real numbers
By the least upper bound property, S has a least upper bound s = sup S. Since we
have already seen that 1 ∈ S, and we know that s is an upper bound of S, s ≥ 1 > 0. It
remains to prove that s2 = 2. We do this by showing that each of the two alternatives
s2 > 2 and s2 < 2 leads to a contradiction.
Suppose first that s2 > 2. In this case, our plan is to show that, for a suitably small
ε > 0, the real number s − ε < s is also an upper bound for S. In order to find a
suitable ε, calculate
(s − ε)2 = s2 − 2sε + ε2 > s2 − 2sε.
Since s2 > 2, the right-hand side will be greater than 2 if ε is chosen sufficiently small
2 2
that s2 − 2sε > 2, which can be rearranged as ε < s 2s−2 . Thus if we choose ε = s 4s−2 ,
2
for instance, then we have ε < s 2s−2 which gives (s − ε)2 > s2 − 2sε > 2 > x2 for all
x ∈ S. Thus s − ε > x for all x ∈ S, which means that s − ε is an upper bound of S
smaller than s. This is a contradiction to the choice of s as the least upper bound.
Suppose next that s2 < 2. In this case we show that by adding a suitably small ε > 0,
we obtain that s + ε ∈ S, which contradicts that s is an upper bound. To find a
suitable ε, calculate as follows:
(Note that in this case we couldn’t throw away the positive number ε2 , but instead
we bounded it using ε2 < ε, which holds as long as ε < 1.) The right-hand side
2−s2 2−s2
s2 + (2s + 1)ε is less than 2 if ε < 2s+1 . Choosing ε = min{ 2(2s+1) , 21 }, we have ε < 1
2−s 2
and ε < 2s+1 , which implies (s + ε)2 < s2 + (2s + 1)ε < 2. Thus s + ε ∈ S. Since
s + ε > s, s is not an upper bound of S, which is a contradiction.
The only remaining possibility is that s2 = 2. We have shown the existence of s ∈ R
such that s > 0 and s2 = 2.
Formally, we treat R as a number system, satisfying all the usual rules of arithmetic,
together with a total order <. We also treat the least upper bound property as a
fundamental property of R. Given these principles, we now deduce some other familiar
properties of R.
The following theorem is called the Archimedean property of the real numbers.
Proof. Suppose that the conclusion is false, that is, that there does not exist an n ∈ N
such that y < nx. This means that for all n ∈ N, y ≥ nx. In other words, y is an upper
bound of the non-empty set S = {nx | n ∈ N}. By the least upper bound property of the
5
1 Analysis
reals, S has a least upper bound u∗ . Note that u∗ − x is not an upper bound of S, since
u∗ − x is smaller than u∗ (x is positive) and u∗ is the least upper bound. Hence there
exists an element mx ∈ S (with m ∈ N) such that u∗ − x < mx, that is, u∗ < (m + 1)x.
Then (m + 1)x is also an element of S with (m + 1)x > u∗ : this contradicts the fact that
u∗ is an upper bound of S.
As a consequence of the Archimedean property we are now able to prove that the set
N of natural numbers is not bounded above.
Examples.
1. Show that the set N is not bounded above.
6
1.1 The real numbers
0 ≤ x < 1. (Note that the use of the symbol ∞ in the notation for intervals is simply a
matter of convenience and is not be taken as suggesting that there is a real number ∞.)
We do not give any special name to intervals of the form [a, b) or (a, b].
In analysis, in order to talk about notions such as convergence and continuity, we will
need a notion of ‘closeness’ between real numbers. This is provided by the absolute value
| · |, and the distance between real numbers x and y is |x − y|. We give the definitions
below.
Definitions.
1. For a real number x, the absolute value |x| of x is defined as follows:
x if x ≥ 0,
|x| =
−x if x < 0.
2. The distance between two real numbers x and y is the absolute value |x − y| of their
difference.
Note that |x| ≥ 0 for all real numbers x, and that |x| = 0 iff1 x = 0. Thus |1| = 1,
|0| = 0, | − 1| = 1, and the distance between the real numbers −1 and 1 is equal to
| − 1 − 1| = | − 2| = 2. The distance gives a notion of closeness of two points, which is
crucial in the formalization of the notions of analysis.
We can now specify regions comprising points close to a certain point x0 ∈ R in terms
of inequalities in absolute values, that is, by demanding that the distance of the points
of the region, to the point x0 , is less than a certain positive number δ, say δ = 0.01 or
δ = 0.0000001, and so on. See Exercise 12 and Figure 1.1.
1
The word “iff” is in common use in mathematics as an abbreviation for “if and only if”.
7
1 Analysis
I
z }| {
x0 − δ x0 x x0 + δ
x > 0 and y > 0. Then |x| = x and |y| = y, and so |x| |y| = xy. On the other hand,
as x > 0 and y > 0, it follows that xy > 0 and so |xy| = xy.
x > 0 and y < 0. Then |x| = x and |y| = −y, and so |x| |y| = x(−y) = −xy. On the
other hand, as x > 0 and y < 0, it follows that xy < 0 and so |xy| = −xy.
x < 0 and y > 0. This follows from 1.1.1 above by interchanging x and y.
x < 0 and y < 0. Then |x| = −x and |y| = −y, and so |x| |y| = (−x)(−y) = xy. On
the other hand, as x < 0 and y < 0, it follows that xy > 0 and so |xy| = xy.
x + y < 0. Then |x + y| = −(x + y). Since |x| ≥ −x and |y| ≥ −y, it follows that
|x| + |y| ≥ −x + (−y) = −(x + y) = |x + y|.
8
1.2 Sequences and limits
a1 , a 2 , a 3 , . . .
of real numbers, where there is the first number (namely a1 ), the second number (namely
a2 ), and so on. For example,
1 1
1, , , . . .
2 3
is a sequence of real numbers. The first number is 1, the second number is 12 and so on.
(There may not be a connection between the numbers appearing in a sequence.) If we
think of a1 as f (1), a2 as f (2), and so on, then it becomes clear that a sequence of real
numbers is a special type of function, namely one with domain N and co-domain R. This
leads to the following formal definition.
Only the notation is somewhat unusual. Instead of writing f (n) for the value of f at a
natural number n, we write an . The entire sequence is then written in any one of the
following ways:
(an )n∈N , (an )∞
n=1 , (an )n≥1 , (an ).
In (an )∞
n=1 , the ∞ symbol indicates that the assignment process 1 7→ a1 , 2 7→ a2 , . . .
continues indefinitely. In these notes, we shall normally use the notation (an )n∈N . In
general, the terms of a sequence need not be real numbers, but in this guide we shall
only be dealing with sequences whose entries are real numbers, so we shall simply refer
to them as sequences from now on.
The nth term an of a sequence may be defined explicitly by a formula involving n, as
in the example given above:
1
an = , n ∈ N.
n
It might also sometimes be defined recursively. For example,
n
a1 = 1, an+1 = an for n ∈ N.
n+1
(Write down the first few terms of this sequence.)
9
1 Analysis
Examples.
1
1. n n∈N
is a sequence with the nth term given by n1 , for n ∈ N. This is the sequence
1 1
1, , , ....
2 3
2. 1 + n1 is a sequence with the nth term given by 1 + n1 , for n ∈ N. This is the
n∈N
sequence
3 4 5 6 7
2, , , , , , . . . .
2 3 4 5 6
3. (−1)n 1 + n1 is a sequence with the nth term given by (−1)n 1 + 1
n
, for n ∈ N.
n∈N
This is the sequence
3 4 5 6 7
−2, , − , , − , , . . . .
2 3 4 5 6
4. ((−1)n )n∈N is a sequence with the nth term given by (−1)n , for n ∈ N. This sequence
is simply
−1, 1, −1, 1, −1, 1, . . .
with the nth term equal to −1 if n is odd, and 1 if n is even.
5. (1)n∈N is a sequence with the nth term given by 1, for n ∈ N. This is the constant
sequence
1, 1, 1, . . . .
6. (n)n∈N is a sequence with the nth term given by n, for n ∈ N. This is the strictly
increasing sequence
1, 2, 3, . . . .
1
7. 11
+ 212 + 313 + · · · + n1n is a sequence with the nth term given by 1
11
+ 1
22
+ 1
33
+
n∈N
· · · + n1n , for n ∈ N. This is the sequence of ‘partial sums’
1 1 1 1 1 1
1
, 1 + 2, 1 + 2 + 3, ....
1 1 2 1 2 3
lim an = L,
n→∞
10
1.2 Sequences and limits
1
0.5
0
0 1 2 3 4 5 6 7
1
Figure 1.2: First 7 points of the graph of the sequence n n∈N
.
where L denotes the limit of the sequence. If there is no such finite number L to which
the terms of the sequence get arbitrarily close, then the sequence is said to diverge.
The problem with this characterization is its imprecision. Exactly what does it mean for
the terms of a sequence to get “closer and closer”, or “as close as we like”, or “arbitrarily
close” to some number L? Even if we accept this apparent ambiguity, how would one use
the definition given in the preceding paragraph to prove theorems that involve sequences?
Since sequences are used throughout analysis, the concepts of their convergence and
divergence must be carefully defined.
2
1.5
1
0.5
0
0 1 2 3 4 5 6 7
1
Figure 1.3: First 7 points of the graph of the sequence 1 + n n∈N
.
For example, the terms of 1 + n1 get “closer and closer” to 0 (indeed the distance
n∈N
to 0 keeps decreasing),
but
its
limit is 1. See Figure 1.3.
n 1
The terms of (−1) 1 + n get “as close as we like” or “arbitrarily close” to 1,
n∈N
but the sequence has no limit. See Figure 1.4.
Definition. The sequence (an )n∈N is said to converge to L if for every real number
ε > 0, there exists an N ∈ N (possibly depending on ε) such that for all n > N ,
|an − L| < ε.
Then we say that (an )n∈N is convergent with limit L and write
lim an = L.
n→∞
11
1 Analysis
0
1 2 3 4 5 6 7 8
−1
−2
1
Figure 1.4: First eight points of the graph of the sequence (−1)n 1 + n
.
n∈N
L+ε
L−ε
... ...
N N +1 N +2 N +3
Examples.
1
1. Use the definition of the limit of a sequence to show that n n∈N
is a convergent
sequence with limit 0.
Solution. Given ε > 0, we need to find N such that, for all n > N ,
1 1
|an − L| = − 0 = < ε.
n n
12
1.2 Sequences and limits
that is,
13
1 Analysis
Fix ε = 1. (It is not obvious that ε = 1 will work, but it has been found by trial and
error.) Now we will show that
|an − L1 | < ε.
Since L2 is a limit of (an )n∈N , ∃N2 ∈ N such that for all n > N2 ,
|an − L2 | < ε.
2ε = |L1 − L2 | = |L1 − an + an − L2 |
≤ |L1 − an | + |an − L2 | = |an − L1 | + |an − L2 |
< ε + ε = 2ε,
a contradiction. Therefore the sequence (an )n∈N does not have two different limits, as
required.
14
1.2 Sequences and limits
15
1 Analysis
Definition. A sequence (an )n∈N is said to be bounded if there exists a real number
M > 0 such that
for all n ∈ N, |an | ≤ M. (1.5)
Note that a sequence is bounded iff the set S = {an | n ∈ N} is bounded. (See
Exercise 15.)
Examples.
|an | = an
1 1 1 1
= 1 + 2 + 3 + ··· + n
1 2 3 n
1 1 1 1
< 1 + 2 + 3 + ··· + n
1 2 2 2
1 1 1 1 1 1 1
= 1+ 1− + 2 1− + · · · + n−1 1 −
1 2 2 2 2 2 2
1 1 1 1 1 1
= 1 + − 2 + 2 − 3 + − · · · + n−1 − n
2 2 2 2 2 2
1 1
=1+ − n
2 2
3
< .
2
(Write down a detailed proof using induction on n.) Thus all the terms are bounded
by 32 , and so the sequence is bounded.
16
1.2 Sequences and limits
f) Show that the sequence (an )n∈N given by an = n for n ∈ N, is not bounded.
Solution. Given any M > 0, there exists an N ∈ N such that M < N (Archimedean
property with y = M and x = 1). Thus
convergent bounded
sequences sequences
|an | ≤ M
17
1 Analysis
a1 ≤ a2 ≤ a3 ≤ . . . ,
strictly increasing if
a1 < a2 < a3 < . . . ,
monotonically decreasing if
a1 ≥ a2 ≥ a3 ≥ . . . ,
and strictly decreasing if
a1 > a2 > a3 > . . . .
Examples.
monotonically strictly monotonically strictly
Sequence monotone?
increasing? increasing? decreasing? decreasing?
1
n n∈N No No Yes Yes Yes
1 + n1 No No Yes Yes Yes
n∈N
(−1) 1 + n1
n No No No No No
n∈N
(1)n∈N Yes No Yes No Yes
(n)n∈N Yes Yes No No Yes
1 1 1 1
11
+ 22
+ 33
+ ··· + nn n∈N Yes Yes No No Yes
The following theorem can be useful in showing that sequences converge when one
does not know the limit beforehand.
Proof. Let (an )n∈N be a monotonically increasing sequence. Since (an )n∈N is bounded,
it follows that the set
S = {an | n ∈ N}
has an upper bound and so sup S exists. We show that in fact (an )n∈N converges to
sup S.
18
1.2 Sequences and limits
Let ε > 0 be given. Since sup S − ε < sup S, it follows that sup S − ε is not an upper
bound for S and so ∃aN ∈ S such that sup S − ε < aN , that is
sup S − aN < ε.
Since (an )n∈N is a monotonically increasing sequence, for n > N , we have aN ≤ an . Since
sup S is an upper bound for S, an ≤ sup S and so |an − sup S| = sup S − an . Thus for
n > N we obtain
(see Exercise 6). So given ε > 0, ∃N ∈ N such that for all n > N , | − an − (− inf S)| < ε,
that is, |an − inf S| < ε. Thus (an )n∈N is convergent with limit inf S.
Examples.
for all n ∈ N) and bounded (see Example e on page 16). Thus it follows from Theorem
1.2.5 that this sequence2 is convergent.
2
Although it is known that this sequence is convergent to some limit L ∈ R, it is so far not even
known if the limit L is rational or irrational, and this is still an open problem in mathematics! Also
∞ Z 1
X 1 1
associated with this sequence is the interesting identity n
= x
dx, the proof of which is
n=1
n 0 x
beyond the scope of this course.
19
1 Analysis
2. The following table gives a summary of the valid implications, and gives counterexam-
ples to implications which are not true.
Question Answer Reason/Counterexample
4n2 + 9
an =
3n2 + 7n + 11
converges to 43 . However, it is simpler to observe that
9
n2 4 + n2 4 + n92
an = = ,
3 + n7 + n112
7 11
n2 3 + n
+ n2
where the terms n92 , n7 , n112 all have limit 0, and by a repeated application of Theorem
1.2.6 given below, we obtain that
9 9
lim 4 + 2 lim 4 + lim 2 4+0 4
lim an =
n→∞ n n→∞ n→∞ n
7 11 = 7 11 = 3 + 0 + 0 = 3 .
n→∞
lim 3 + + 2 lim 3 + n→∞
lim + n→∞
lim 2
n→∞ n n n→∞ n n
20
1.2 Sequences and limits
Theorem 1.2.6. If (an )n∈N and (bn )n∈N are convergent sequences, then the following
hold:
1. For all α ∈ R, (αan )n∈N is a convergent sequence and n→∞
lim αan = α n→∞
lim an .
2. (|an |)n∈N is a convergent sequence and lim |an | = lim an .
n→∞ n→∞
k
5. For all k ∈ N, (akn )n∈N is a convergent sequence and lim ak = lim an .
n→∞ n n→∞
1
6. If for all n ∈ N, bn 6= 0 and lim bn 6= 0, then is convergent and moreover,
n→∞ bn n∈N
1 1
lim = .
n→∞ b
n lim
n→∞
b n
Proof. Let (an )n∈N and (bn )n∈N converge to La and Lb , respectively.
1. Let ε > 0 be given. By the definition of limn→∞ an = La , there exists an N ∈ N such
that for all n > N ,
ε
|an − La | < .
|α| + 1
(We added 1 to |α| in the denominator to deal with the possibility that α = 0.) Then
ε
|αan − αLa | = |α| |an − La | ≤ |α| < ε,
|α| + 1
and (again by the definition of a convergent sequence) (αan )n∈N is convergent with
limit αLa , that is,
lim αan = αLa = α lim an .
n→∞ n→∞
|an − La | < ε.
21
1 Analysis
4. Note that
|an bn − La Lb | = |an bn − La bn + La bn − La Lb |
≤ |an bn − La bn | + |La bn − La Lb |
= |an − La | |bn | + |La | |bn − Lb |. (1.6)
Given ε > 0, we need to find N such that for all n > N ,
|an bn − La Lb | < ε.
This can be achieved by finding an N such that each of the summands in (1.6) is less
than 2ε for n > N . This can be done as follows.
Step 1. Since (bn )n∈N is convergent, by Theorem 1.2.4 it follows that it is bounded:
∃M > 0 such that for all n ∈ N, |bn | ≤ M . Let N1 ∈ N be such that, for all n > N1 ,
ε
|an − La | < .
2M
Step 2. Let N2 ∈ N be such that for all n > N2 ,
ε
|bn − Lb | < .
2|La | + 1
(Note that it is possible that La = 0.) Thus, for n > N := max{N1 , N2 }, we have
|an bn − La Lb | ≤ |an − La | |bn | + |La | |bn − Lb |
ε ε
< M + |La |
2M 2|La | + 1
ε ε
< +
2 2
= ε.
It follows that (an bn )n∈N is a convergent sequence with limit La Lb , that is,
lim an bn = La Lb = lim an lim bn .
n→∞ n→∞ n→∞
22
1.2 Sequences and limits
5. This can be shown by using induction on k and from part 4 above. It is trivially true
with k = 1. Suppose that it holds for some k, then (akn )n∈N is convergent and
k
lim ak = lim an .
n→∞ n n→∞
Hence by part 4 above applied to the sequences (an )n∈N and (akn )n∈N , we obtain that
the sequence (an · akn )n∈N is convergent and
k k+1
lim an akn = lim an lim akn = lim an lim an = lim an .
n→∞ n→∞ n→∞ n→∞ n→∞ n→∞
Thus (ak+1
n )n∈N is convergent and
k+1
lim ak+1 = lim an .
n→∞ n n→∞
7. Exercise 38.
8. Since an ≥ 0 for each n, we have L ≥ 0 (see Exercise 24). The case L = 0 is left as an
exercise.
In the case L > 0, we use a technique called “rationalising the numerator”: note that
√ √
√ √ √ √ an + L a −L
an − L = ( an − L) √ √ =√ n √ .
an + L an + L
√
Now, given ε > 0, choose N ∈ N so that, for n > N , |an − L| < ε L. Then we have,
for n > N , √
√ √ |an − L| ε L
| an − L| = √ √ ≤ √ = ε.
an + L L
√ √
Thus ( an )n∈N is convergent, with limit L,
23
1 Analysis
Examples. 1. Determine whether the following sequence is convergent and find its limit.
!
n2 − 24n3 + 3n4 − 12
1 + 7n + 21n4 n∈N
1
Solution. By Example 1 on page 12 we know that lim = 0. We now use Theo-
n→∞ n
rem 1.2.6, we obtain
1 24
n2 − 24n3 + 3n4 − 12 n4 n2
− n
+ 3 − n124
lim = lim · 1
n→∞ 1 + 7n + 21n4 n→∞ n4
n4
+ n73 + 21
2 4
lim 1
n→∞ n
lim n1 + 3 − 12 n→∞
− 24 n→∞ lim n1
= 4 3
1 1
lim + 7 lim n + 21
n→∞ n n→∞
02 − 24 · 0 + 3 − 12 · 04
=
04 + 7 · 03 + 21
3 1
= = .
21 7
2. Determine whether the following sequence is convergent and find its limit.
n
2 + 3n + 1
3n+1 + 3 n∈N
Solution. Divide top and bottom by the fastest growing term appearing, which is
3n , and use that limn→∞ xn = 0 for |x| < 1:
2n + 3n + 1 (2/3)n + 1 + (1/3)n
lim =
n→∞ 3n+1 + 3 3 + 3(1/3)n
limn→∞ (2/3)n + limn→∞ 1 + limn→∞ (1/3)n
=
limn→∞ 3 + 3 limn→∞ (1/3)n
0+1+0 1
= = .
3+0 3
Remark. It follows from Theorem 1.2.6.3 that if we have three convergent sequences
(an )n∈N , (bn )n∈N , (cn )n∈N , then their sum (an + bn + cn )n∈N is also convergent with limit
24
1.2 Sequences and limits
This is also true for the sum of four convergent sequences, the sum of five convergent
sequences, and by an easy induction proof, the sum of any number of convergent sequences.
However, the following reasoning is clearly incorrect (because it would show that 1 = 0):
1 1 1
1= + + ··· +
|n n {z n}
n terms
and therefore,
1 1 1
1 = n→∞
lim 1 = n→∞
lim + + ··· +
n n n
! 1 1 1
= lim + lim + · · · + lim
n→∞ n n→∞ n n→∞ n
| {z }
n limits
= |0 + 0 +{z· · · + 0}
n terms
= 0.
The incorrect step is marked with a ‘!’. The problem here is that the number of terms is
not a fixed value, but depends on n, and therefore Theorem 1.2.6.3 cannot be applied.
for all n ∈ N, an ≤ cn ≤ bn ,
then (cn )n∈N is also convergent with the same limit, that is,
lim an = n→∞
n→∞
lim cn = n→∞
lim bn .
Proof. Let L denote the common limit of (an )n∈N and (bn )n∈N :
lim an = L = lim bn .
n→∞ n→∞
Given ε > 0, let N1 ∈ N be such that for all n > N1 , |an − L| < ε. Hence for n > N1 ,
L − an ≤ |L − an | = |an − L| < ε,
25
1 Analysis
Examples.
n
1. Use the Sandwich theorem to show that lim = 0.
n→∞ 10n
Solution. It can be shown by induction that for all n ∈ N, n2 < 10n .
Consequently, we have
n n 1
0≤ n
≤ 2 = .
10 n n
1
Since n→∞
lim 0 = 0 = n→∞
lim , from the Sandwich theorem it follows that the sequence
n
n
10n is convergent and
n∈N
n
lim n = 0.
n→∞ 10
1 2 3 4
Thus the sequence 10 , 100 , 1000 , 10000 , . . . is convergent with limit 0.
1
2. Use the Sandwich theorem to show that for any a, b ∈ R, lim (|a|n + |b|n ) n =
n→∞
max{|a|, |b|}.
Solution. Clearly,
(max{|a|, |b|})n ≤ |a|n + |b|n ≤ (max{|a|, |b|})n + (max{|a|, |b|})n
and so
1 1
max{|a|, |b|} ≤ (|a|n + |b|n ) n ≤ 2 n max{|a|, |b|}.
1
lim (|a|n + |b|n ) n = max{|a|, |b|}.
So using the Sandwich theorem, it follows that n→∞
1
In particular, with a = 24 and b = 2005, we have that lim (24n + 2005n ) n = 2005,
n→∞
that is, the sequence
2029, 2005.1436, 2005.001146260873, . . .
is convergent with limit 2005.
26
1.2 Sequences and limits
n n n
3. Show that n→∞
lim 2
+ 2 + ··· + 2 = 1.
n +1 n +2 n +n
n n
Solution. There are n terms in the sum: the smallest is n2 +n
and the largest is n2 +1
.
Thus, for all n ∈ N, we have
n2 n n n n2
≤ + + · · · + ≤ ,
n2 + n n2 + 1 n2 + 2 n2 + n n2 + 1
and since
n2 n2
lim = 1 = lim 2 .
n→∞ n2 + n n→∞ n + 1
1.2.7 Subsequences
In this section we prove an important result in analysis, known as the Bolzano-Weierstrass
theorem, which says that every bounded sequence has a convergent ‘subsequence’. We
begin this section by defining what we mean by a subsequence of a sequence.
Definition. Let (an )n∈N be a sequence and let (nk )k∈N be a strictly increasing sequence
of natural numbers. Then (ank )k∈N is called a subsequence of (an )n∈N .
Examples.
1 1 1 1 1
1. , , and are all subsequences of .
2n n∈N n2 n∈N n! n∈N nn n∈N n n∈N
1, 1, 1, . . .
and the sequence ((−1)2n−1 )n∈N , that is, the constant sequence
27
1 Analysis
Theorem 1.2.8. If (an )n∈N is a convergent sequence with limit L, then any subsequence
of (an )n∈N is also convergent with the limit L.
Proof. Let (ank )k∈N be a subsequence of (an )n∈N . Given ε > 0, let N ∈ N be such that
for all n > N , |an − L| < ε. Since the sequence n1 < n2 < n3 < . . . is not bounded
above, it follows that there exists a K ∈ N such that nK > N . Then for all k > K,
nk > nK > N . Hence for k > K, |ank − L| < ε, and so (ank )k∈N is convergent with limit
L.
Examples.
1 1 1 1
1. 2n n∈N
, n2 n∈N
, n! n∈N
and nn n∈N
are convergent sequences with limit 0.
2. The sequence ((−1)n )n∈N is divergent since the subsequence 1, 1, 1, . . . has limit 1,
while the subsequence −1, −1, −1, . . . has limit −1.
The following theorem is a very important step in the proof of the Bolzano-Weierstrass
theorem. It gives us a fact about arbitrary sequences of real numbers.
Since we do not assume anything about the sequence, other than that it is a sequence
of real numbers, there is virtually no obvious way to start the proof. Nevertheless, even
though it is very difficult to discover such a proof, it is not so difficult to understand it,
once found.
Before doing the formal proof, we first illustrate the idea behind it. Let (an )n∈N be the
given sequence. Imagine that an is the height of a hotel with number n, which is followed
by hotel n + 1, and so on, along an infinite line, where the sea is in the far distance. A
hotel is said to have the seaview property if it is higher than all hotels following it.
See Figure 1.7.
...
...
1 2 3 4 5 6
First possibility. There are infinitely many hotels with the seaview property. Then their
heights form a strictly decreasing subsequence.
28
1.2 Sequences and limits
Second possibility. There is only a finite number of hotels with the seaview property.
Then after the last hotel with the seaview property, one can start with any hotel
and then always find one that is at least as high, which is taken as the next hotel,
and then finding yet another that is at least as high as that one, and so on. The
heights of these hotels form a monotonically increasing subsequence.
that is, S is the set of all natural numbers m such that an < am for all n > m. Then we
have the following two cases.
S is infinite. Arrange the elements of S in strictly increasing order: n1 < n2 < n3 < . . . .
Then (ank )k∈N is a strictly decreasing subsequence of (an )n∈N .
Proof. Let (an )n∈N be a bounded sequence. Then there exists M > 0 such that for all
n ∈ N, |an | ≤ M . From Theorem 1.2.9 above, it follows that the sequence (an )n∈N has
a monotone subsequence (ank )k∈N . Then, clearly, for all k ∈ N, |ank | ≤ M and so the
sequence (ank )k∈N is also bounded. Since (ank )k∈N is monotone and bounded, it follows
from Theorem 1.2.5 that it is convergent.
√
Example. Consider the sequence (an )n∈N of fractional parts of integral multiples of 2,
defined by
√ √
an = n 2 − bn 2c, for n ∈ N,
3
Bernhard Bolzano (1781–1848)
4
Karl Weierstrass (1815–1897)
29
1 Analysis
The sequence (an )n∈N is bounded: indeed, 0 ≤ an < 1 for every n ∈ N. So by the
Bolzano-Weierstrass theorem this sequence has a convergent subsequence5 .
1.3 Continuity
A function f : R → R is a rule of correspondence that assigns to each real number a
unique real number. The functions we are most familiar with are actually unusually
well-behaved, and most functions are impossible to describe fully, let alone to work with.
Many bizarre functions make their appearance in analysis, and in order to avoid falling
into pitfalls with simplistic thinking based on our experience with “nice” functions, we
need our definitions and the hypotheses (assumptions) of theorems to be stated carefully
and clearly.
Within the huge collection of functions, there is an important subset: the continuous
functions. Continuous functions play a prominent role in analysis since they include all
the most familiar and useful functions, and because the all continuos functions share
many useful properties.
In this section we give the formal definition of a continuous function and prove two of
the most important properties of continuous functions: the Extreme value theorem and
the Intermediate value theorem.
30
1.3 Continuity
If a function has a break at a point, say c, then even if points x are close to c, the
points f (x) do not get close to f (c). See Figure 1.8.
This motivates the following definition of continuity, which guarantees that if a function
is continuous at a point c, then we can make f (x) as close as we like to f (c), by choosing
x sufficiently close to c. See Figure 1.9.
Definitions.
1. Let I be an interval in R and let c ∈ I. A function f : I → R is continuous at c
if for every ε > 0, there exists a δ > 0 such that for all x ∈ I satisfying |x − c| < δ,
|f (x) − f (c)| < ε.
f (c)
Figure 1.8: A function with a break at c. If x lies to the left of c, then f (x) is not close
to f (c), no matter how close x comes to c.
31
1 Analysis
f (c) + ε
f (x)
f (c)
f (c) − ε
c−δ c x c+δ x
Figure 1.9: The definition of the continuity of a function at point c. If the function is
continuous at c, then given any ε > 0 (which determines a strip around the
line y = f (c) of width 2ε), there exists a δ > 0 (which determines an interval
of width 2δ around the point c) such that whenever x lies in this interval (so
that x satisfies c − δ < x < c + δ, that is, |x − c| < δ), then f (x) satisfies
f (c) − ε < f (x) < f (c) + ε, that is, |f (x) − f (c)| < ε.
Solution. Fix c ∈ R. Given ε > 0, we again need to fnd a suitable δ > 0. Here we
set δ = ε/2. Then, if x ∈ R and |x − c| < δ, we have:
32
1.3 Continuity
0 x
Solution. Suppose that f is continuous at 0. Then for any given ε > 0 there exists
a δ > 0 such that whenever |x − 0| < δ, |f (x) − f (0)| < ε. In particular, setting
ε = 12 , we see that there is a δ > 0 such that for all x ∈ R, if |x| = |x − 0|
< δ, then
|f (x) − f (0)| = |f (x) − 0| = |f (x)| < ε = 2 . Take x = 2 ∈ R: then |x| = 2 = 2δ < δ,
1 δ δ
δ 1
but |f (x)| = f = |1| = 1 > = ε, which is a contradiction. So f is not
2 2
continuous at 0.
Next we show that for all c ∈ R \ {0}, f is continuous at c. Let ε > 0 be given. Take
δ = |c|
2
> 0. Then if x ∈ R and |x − c| < δ, we have
|c|
|c| − |x| ≤ ||c| − |x|| ≤ |c − x| = |x − c| < δ =
2
and so
|c|
|x| > > 0.
2
Thus x 6= 0 and so f (x) = 1. Hence if x ∈ R and |x − c| < δ, we obtain
Consequently f is continuous at c.
1
5. Show that the function f : (0, ∞) → R given by f (x) = x
for all x ∈ R is continuous.
n o
c εc2
Solution. Let c ∈ (0, ∞). Given ε > 0, let δ = min ,
2 2
(> 0). Then if x ∈ (0, 1)
and |x − c| < δ, we have
c c
c − x ≤ |c − x| < δ ≤ , and so x > > 0.
2 2
33
1 Analysis
1 1 |c − x| 1 1 2 1 2δ
− = = |x − c| · · < δ · · = 2 ≤ ε.
x c xc x c c c c
So f is continuous at c. Since the choice of c ∈ (0, ∞) was arbitrary, it follows that f
is continuous on (0, ∞).
Note: In practice, the choice of δ is the last part of the proof to be filled in. As
we go through the rest of the proof, we see the need to have (a) δ ≤ |c|/2, and then
(b) 2δ
c2
≤ ε. Only after we have collected all the conditions do we make the choice of
δ.
Proof. “Only if” (=⇒) direction: Assume that f is continuous at c ∈ I and let
(xn )n∈N be a convergent sequence contained in I with limit c. We have to show that
(f (xn ))n∈N converges to f (c).
Since f is continuous at c ∈ I, given any ε > 0, ∃δ > 0 such that for all x ∈ I satisfying
|x − c| < δ, |f (x) − f (c)| < ε. Since (xn )n∈N is convergent with limit c, ∃N ∈ N such
that for all n > N , |xn − c| < δ.
Consequently for all n > N , |f (xn ) − f (c)| < ε. Thus (f (xn ))n∈N is convergent with
limit f (c).
“If” (⇐=) direction: Suppose that (1.7) holds. We have to show that f is continuous
at c. We prove this by contradiction. Assume that f is not continuous at c, that is,
¬ [∀ε > 0 ∃δ > 0 such that ∀x ∈ I such that |x − c| < δ, |f (x) − f (c)| < ε,]
or equivalently,
∃ε > 0 such that ∀δ > 0 ∃x ∈ I such that |x − c| < δ but |f (x) − f (c)| ≥ ε.
34
1.3 Continuity
The above theorem allows us to easily prove the following useful Theorem 1.3.2. Before
stating it, we introduce some convenient notation.
Definitions. Let I be an interval in R. Given functions f : I → R and g : I → R, we
define the following:
1. If α ∈ R, then we define the function αf : I → R by (αf )(x) = α · f (x), x ∈ I.
2. We define the absolute value of f to be the function |f | : I → R given by |f |(x) =
|f (x)|, x ∈ I.
3. The sum of f and g is the function f +g : I → R defined by (f +g)(x) = f (x)+g(x),
x ∈ I.
4. The product of f and g is the function f g : I → R defined by (f g)(x) = f (x)g(x),
x ∈ I.
5. If k ∈ N, then we define the kth power of f , to be the function f k : I → R given by
f k (x) = (f (x))k , x ∈ I.
1 1 1
6. If for all x ∈ I, g(x) 6= 0, then we define the function g
: I → R by g
(x) = g(x)
,
x ∈ I.
Theorem 1.3.2. Let I be an interval in R and let c ∈ I. Suppose that f : I → R and
g : I → R are continuous at c. Then:
1. For all α ∈ R, αf is continuous at c.
2. |f | is continuous at c.
3. f + g is continuous at c.
4. f g is continuous at c.
5. For all k ∈ N, f k is continuous at c.
1
6. If for all x ∈ I, g(x) 6= 0, then g
is continuous at c.
Proof. Suppose that (xn )n∈N is a convergent sequence contained in I, with limit c. Since
f and g are continuous at c, from Theorem 1.3.1, it follows that (f (xn ))n∈N and (g(xn ))n∈N
are convergent with limits f (c) and g(c), respectively. Each one of the statements now
follows easily from Theorem 1.2.6 and a second application of Theorem 1.3.1. For
example, consider statement number 4 (the other cases may be proved as exercises). By
Theorem 1.2.6, (f (xn )g(xn ))n∈N is convergent with limit f (c)g(c), that is, ((f g)(xn ))n∈N
is convergent with limit (f g)(c). Thus by Theorem 1.3.1, f g is continuous at c.
35
1 Analysis
36
1.3 Continuity
√
Solution. Let g : (0, ∞) → R be defined by g(x) = x. By Exercise 68, g is continuous.
Let h : R → (0, ∞) be defined by
x2 − 1
h(x) = , x ∈ R.
x2 + 1
We have already seen that all polynomial functions are continuous. Therefore, p1 , p2 : R →
R defined by p1 (x) = x2 − 1 and p2 (x) = x2 + 1 are continuous. Since p2 (x) 6= 0 for all
1
x ∈ R, it follows by Theorem 1.3.2(6) that is continuous. Then by Theorem 1.3.2(4),
p2
1
p1 · = h is continuous. Note that the restriction of h to (1, ∞), h|(1,∞) assumes only
p2
positive values: if x ∈ (1, ∞) then x > 1, therefore x2 − 1 > 0, and since x2 + 1 > 0
anyway, h(x) > 0. This shows that the composition g ◦ h|(1,∞) = f is properly defined.
By Theorem 1.3.3, h|(1,∞) is continuous. Finally, by Theorem 1.3.4, g ◦ h|(1,∞) = f is
continuous at each c > 1, and therefore, continuous.
f (a)
a c x
b
The Intermediate value theorem was first proved by Bernhard Bolzano in 1817.
Theorem 1.3.5. (Intermediate value theorem). If f : [a, b] → R is continuous and y
is such that f (a) ≤ y ≤ f (b) or f (b) ≤ y ≤ f (a), then there exists c ∈ [a, b] such that
f (c) = y.
Proof. Suppose first that f (a) ≤ y ≤ f (b). The case f (b) ≤ y ≤ f (a) is similar, and
will be discussed at the end of the proof below.
Let S = {x ∈ [a, b] | f (x) ≤ y}. We want to prove that
1. sup S exists, and
37
1 Analysis
We have shown that f (c) ≤ y and f (c) ≥ y. In conclusion, we have found c ∈ [a, b]
such that f (c) = y.
Suppose now that y ∈ R is such that f (b) ≤ y ≤ f (a). To prove the theorem in this
case, it is easy to modify the above proof. Alternatively, we could use what we have just
proved, as follows. Note that we have
so we may apply the above proof to the function −f (which is continuous by The-
orem 1.3.2(1) with α = −1). Then we obtain the existence of c ∈ [a, b] such that
(−f )(c) = −y, that is, f (c) = y.
Examples.
1. Show that every polynomial of odd degree with real coefficients has at least one real
root.
where a2m+1 6= 0. If we divide p by its leading coefficient a2m+1 then the resulting
polynomial will still have the same roots (if any). We may therefore assume without
loss of generality that a2m+1 = 1.
In order to show that p has a real root, we will choose a large enough n ∈ N such
that p(n) > 0 and p(−n) < 0 and restrict our attention to an interval [−n, n]. Then
appealing to the Intermediate Value Theorem, we can conclude that p must vanish at
some point in this interval, that is, for some real c ∈ [−n, n], p(c) = 0.
38
1.3 Continuity
It remains to find an n ∈ N such that p(n) > 0 > p(−n). We expect such an n to be
large in general. Intuitively, the term in the polynomial that dominates when x is
large is the term with highest exponent x2m+1 . We now compare p(x) with x2m+1 by
considering the quotient:
p(x) a0 a1 a2m
2m+1
= 2m+1 + 2m + · · · + +1 (1.8)
x x x x
(recall we assumed a2m+1 = 1). If we make the substitution x = n and let n → ∞, we
obtain
p(n) a0 a1 a2m
lim = lim + 2m + · · · + + 1 = 1.
n→∞ n2m+1 n→∞ n2m+1 n n
By choosing ε = 1/2 in the definition of a limit, we obtain that there exists N1 ∈ N
such that for all n > N1 ,
p(n) 1
2m+1 − 1 < ε = ,
n 2
and since |A| ≥ −A for any A ∈ R, it follows that
p(n) 1
1− 2m+1
< ,
n 2
n2m+1
or p(n) > 2
> 0 as long as n > N1 .
To obtain the inequality p(−n) < 0 for suitably large n, we next make the substitution
x = −n in (1.8) and take the limit as n → ∞ to obtain
!
p(−n) a0 a1 a2m
lim = lim + + ··· + + 1 = 1.
n→∞ (−n)2m+1 n→∞ (−n)2m+1 (−n) 2m −n
Again applying the definition of a limit with ε = 1/2, we obtain as before that there
exists an N2 ∈ N such that for all n > N2 ,
p(−n) 1
1− < .
(−n)2m+1 2
n2m+1
n2m+1 + p(−n) < ,
2
2m+1
or p(−n) < − n 2
< 0 as long as n > N2 .
Thus if we choose n = 1 + max{N1 , N2 }, we obtain both inequalities p(n) > 0 and
p(−n) < 0, which finishes the proof.
1
For example, the polynomial p(x) = x3 − x2014 + 399 has a real root in [−1, 1]: indeed
1 1
p(1) = 399 > 0 and p(−1) = −2 + 399 < 0 and so ∃c ∈ [−1, 1] such that p(c) = 0.
39
1 Analysis
2. Show that at any given time, there exists a pair of diametrically opposite points on
the equator which have the same temperature.
Solution. Let T (Θ) denote the surface temperature at the point at longitude Θ. See
Figure 1.11. (Note that Θ(0) = Θ(2π).) Assuming that Θ is a continuous function of
d
a c x
b
Theorem 1.3.6 (Extreme value theorem). Let [a, b] be any closed and bounded interval
and let f : [a, b] → R be a continuous function. Then there exists c ∈ [a, b] such that
40
1.3 Continuity
Since c, d ∈ [a, b] in the above theorem, the supremum and infimum in (1.9) and (1.10)
are in fact the maximum and minimum, respectively. This proof is a beautiful application
of the Bolzano-Weierstrass theorem.
Proof. We prove the first half of the theorem, leaving the second half as an exercise.
Let S = {f (x) | x ∈ [a, b]}. We have to prove that there exists c ∈ [a, b] such that
f (c) = sup S. The plan of the proof is as follows.
Then, by the definition of S, we may conclude that sup S = f (c) for some c ∈ [a, b].
To show that sup S exists, we use the l.u.b. property of R. First, S 6= ∅, since for
example f (a) ∈ S. Secondly, S is bounded above, as shown by the following proof by
contradiction.
Suppose S is not bounded above. For each n ∈ N, choose an f (xn ) for some xn ∈ [a, b]
such that f (xn ) > n. Then (xn ) is a bounded sequence. By the Bolzano-Weierstrass
Theorem, (xn ) has a convergent subsequence (xnk )k∈N . Let its limit be c = lim xnk . Since
k→∞
a ≤ xn ≤ b and limits preserve inequalities, a ≤ c ≤ b. By Theorem 1.3.1, (f (xnk )k∈N ) is
convergent (we assumed that f is continuous). By Theorem 1.2.4, (f (xnk )k∈N ) is bounded.
On the other hand, for each k ∈ N, f (xnk ) > nk ≥ k (since 1 ≤ n1 < n2 < n3 < . . . ).
Therefore, (f (xnk )k∈N ) is unbounded, a contradiction.
Next we show that sup S ∈ S. Recall that S = {f (x) | x ∈ [a, b]}. Write M = sup S.
We have to prove that M ∈ S. Since M is the least upper bound of S, it follows that
for all n ∈ N, M − n1 is not an upper bound of S. Therefore, there exists f (xn ) ∈ S
(with xn ∈ [a, b]) such that M − n1 < f (xn ). The sequence (xn ) is bounded. As before,
by the Bolzano-Weierstrass Theorem it has a convergent subsequence (xnk )k∈N with limit
c ∈ [a, b], say. We next show that this c satisfies f (c) = M = sup S.
By Theorem 1.3.1, (f (xnk )k∈N is convergent with limit f (c). Since c ∈ [a, b], f (c) ∈ S.
Finally, because
1 1
M− < f (xnk ) ≤ M and lim M − = M,
nk k→∞ nk
the Sandwich Theorem shows that f (c) = lim f (xnk ) = M .
k→∞
We have found c ∈ [a, b] such that f (c) = M = sup S. This finishes the proof of the
first half of the theorem.
The proof that there exists d ∈ [a, b] such that f (d) = inf S is left as an exercise
(Exercise 80).
41
1 Analysis
Example. Show that the statement of Theorem 1.3.6 does not hold if [a, b] is replaced
by (a, b).
Solution. The function f : (0, 1) → R given by f (x) = x is continuous on (0, 1). If
1.4 Exercises
1. Fill in the following table for each of the sets S = A, B, . . . , K below.
a) A = {x ∈ R | 0 ≤ x < 1} g) G = {x ∈ R | x2 + 2x ≤ 1}
b) B = {x ∈ R | 0 ≤ x ≤ 1} h) H = {0, 2, 5, 2016}
c) C = {x ∈ R | x ∈ N and xprime} n o
1
i) I = (−1)n 1 + n∈N
n o
1 n
d) D = n ∈ Z \ {0}
n
j) J = {x2 | x ∈ R}
n o
e) E = − n1 n ∈ N
n o
x2
f) F = {x ∈ R | x > x2 } k) K = x∈R
1+x2
2. Determine whether the following statements are TRUE or FALSE. Give reasons for
your answers in each case: either a (brief) proof or a counterexample. (For instance, if
you think statement (d) is false, you should give an example of a bounded non-empty
subset of R that does not have a maximum, and explain (briefly) why your set has
the required properties.)
a) Let S be a subset of R such that, for each x ∈ S, there exists u ∈ R with x ≤ u.
Then S is bounded above.
b) If u is an upper bound of a subset S of R, and u0 < u, then u0 is not an upper
bound of S.
c) If u∗ is the supremum of a subset S of R, and ε is any positive real number, then
u∗ − ε is not an upper bound of S.
d) Every bounded non-empty subset of R has a maximum.
e) Every bounded subset of R has a supremum.
42
1.4 Exercises
6
f) A non-empty subset of R is bounded above iff it has a supremum.
g) For every set that has a supremum, the supremum belongs to the set.
h) For every non-empty bounded set S, if inf S < x < sup S, then x ∈ S.
i) For S a non-empty subset of R, {x2 | x ∈ S} is bounded iff S is bounded.
3. For any non-empty bounded set S, prove that inf S ≤ sup S, and that the equality
holds iff S is a singleton set (that is a set with exactly one element).
4. Let A and B be non-empty subsets of R that are bounded above and such that A ⊆ B.
Prove that sup A ≤ sup B.
5. Let A and B be non-empty subsets of R that are bounded above and define
A + B = {x + y | x ∈ A and y ∈ B} .
6. Let S be a non-empty subset of real numbers which is bounded below. Let −S denote
the set of all real numbers −x, where x belongs to S.
a) Show that −S is bounded above, and therefore has a supremum.
b) Prove that inf S exists, and that inf S = − sup(−S).
Note: in this exercise, we deduce the “greatest lower bound principle” from the least
upper bound principle. Accordingly, you may not assume the greatest lower bound
principle in this exercise. (However, once we have established this result, we can use
the greatest lower bound principle henceforth.)
7. In this exercise it is shown that for any positive real number r there exists a positive
real number s such that s2 = r. In other words, any positive real number has a
positive square root.
Given r ∈ R with r > 0, let S = {x ∈ R | x2 < r}.
a) Prove that S 6= ∅.
b) Prove that S is bounded above.
Hint: Distinguish between the cases r > 1 and r ≤ 1.
c) Conclude that S has a least upper bound s. Show that s > 0.
6
If you think the statement is TRUE, you have two things to prove: (i) if a non-empty set S is bounded
above, then it has a supremum; (ii) if a non-empty set S has a supremum, then it is bounded above.
If you think the statement is FALSE, you should provide a counterexample to one of (i) and (ii).
43
1 Analysis
d) Show that if s2 > r then there exists ε > 0 such that s − ε is an upper bound of S.
e) Show that if s2 < r then there exists ε > 0 such that s + ε ∈ S.
f) Conclude that s2 = r by explaining why each of the conclusions in 7d) and 7e)
gives a contradiction.
n o
8. Let S be a non-empty set of positive real numbers, and define S −1 = 1
x∈S .
x
9. In this exercise we define the floor and ceiling functions. For any x ∈ R define the
set Sx = {n ∈ Z | n ≤ x}.
a) Show that, for any x ∈ R, the set Sx is non-empty and bounded above.
Hint: To show that Sx 6= ∅, you will need the Archimedean property.
b) Show that sup Sx exists and sup Sx ∈ Sx for any x ∈ R. Explain why we obtain as
a consequence that the following gives a proper definition of the floor function
b·c : R → Z
bxc = max {n ∈ Z | n ≤ x} , x ∈ R.
dxe = min {n ∈ Z | n ≥ x} , x ∈ R.
√
e) Show that b k 2 + kc = k for all k ∈ N.
10. Let S be a subset of R that does not have a maximum. Show that, for every x ∈ S,
there is y ∈ S with y > x.
Hint: Deal separately with the two cases: (i) S has a supremum; (ii) S has no
supremum.
11. Determine whether the following statements are TRUE or FALSE. Give reasons (brief
proofs or counterexamples) for your answers in each case.
a) If a ≤ b then (−∞, a) ⊆ (−∞, b).
b) For real numbers a and b, a ≤ b iff (b, ∞) ⊆ (a, ∞).
c) (a, b) ⊆ (c, d) iff c ≤ a and b ≤ d.
d) (a, b) ⊆ [c, d] iff c < a and b < d.
e) For any x ∈ R, −|x| ≤ x ≤ |x|.
f) For any x ∈ R, −x ≤ |x| ≤ x.
g) For any x, y ∈ R, if x < y then |x| ≤ |y|.
h) For any x, y ∈ R, if x2 ≤ y 2 then |x| ≤ |y|.
44
1.4 Exercises
45
1 Analysis
24. Let (an )n∈N be a sequence such that for all n ∈ N, an ≥ 0. Prove that if (an )n∈N is
convergent with limit L, then L ≥ 0.
25. A sequence (an )n∈N is said to be a Cauchy sequence if for every ε > 0, there exists
an N ∈ N such that for all n, m > N , |an − am | < ε.
Show that every convergent sequence is Cauchy.
Hint: |an − am | = |an − L + L − am | ≤ |an − L| + |am − L|.
26. In the proof of Theorem 1.2.2, where did we use the assumption that x ≥ −1? Is the
inequality true for x = −2? Is it true for x = −3?
√
27. For a real number x such that |x| < 1, show that the sequence (xn n)n∈N is convergent,
with limit 0.
1
b) Prove that (n n )n∈N is convergent with limit 1.
n+a
29. Let a and b be positive real numbers. Show that the sequence is monotone.
n+b n∈N
31. Use Bernoulli’s Inequality (Theorem 1.2.2) to show that, for x a real number with
x > 1, the sequence (xn )n∈N is divergent.
32. Suppose that the sequence (an )n∈N converges to 0 and the sequence (bn )n∈N is bounded.
Use the definition to prove that lim an bn = 0.
n→∞
33. For each n ∈ N, let Hn = 1 + 12 + 13 + · · · + n1 . The sequence (Hn )n∈N is the sequence
of harmonic numbers.
a) Show that for any m, n ∈ N with m > n,
m−n
Hm − Hn > .
m
1 1 1
(Hint: note that Hm − Hn = n+1
+ n+2
+ ··· + m
, a sum of m − n terms.)
1
b) Deduce that H2k+1 − H2k > 2
for any k ∈ N.
3 4 k+2
c) Show that H2 = 2
, H4 > 2
, H8 > 52 , H16 > 26 , H32 > 27 , and in general, H2k > 2
for all k ≥ 2.
46
1.4 Exercises
d) Let M > 0 be given. Show that there exists n ∈ N such that Hn > M .
e) Show that (Hn )n∈N is divergent.
34. a) (∗) Let (an )n∈N be a convergent sequence with limit L. Prove that the sequence
(sn )n∈N , where
a1 + · · · + an
sn = , n ∈ N,
n
is also convergent with limit L.
Hint: use the fact that (an )n∈N is bounded.
b) Give an example of a sequence (an )n∈N such that (sn )n∈N is convergent but (an )n∈N
is divergent.
`k = inf {an | n ≥ k} , k ∈ N.
Show that the sequence (`n )n∈N is bounded above and increasing, and conclude
that it is convergent. (Its limit is denoted by lim inf an .)
n→∞
36. Recall the definition of a Cauchy sequence from Exercise 25. Prove that every Cauchy
sequence is bounded. (Hint: follow the proof of Theorem 1.2.2.)
37. Determine whether the following sequences are convergent and, if so, find their limits:
√ !
n+ n
1/2n
√ 3 n
√ ; 1 + n + n( ) .
3n2 − 1 n∈N 4 n∈N
38. Let (an )n∈N be a convergent sequence with limit L, and let k ∈ N. Show that (an+k )n∈N
is also convergent with limit L.
39. Recall the convergent sequence (an )n∈N from Exercise 30 defined by
2n + 1
a1 = 1 and an = an−1 for n ≥ 2.
3n
Determine its limit.
Hint: Use the previous exercise.
40. a) Use Theorem 1.2.6 and induction to show that for any fixed k ∈ N,
1 1 1 1 1
lim + + + ··· + + = 0.
n→∞ n+1 n+2 n+3 n+k−1 n+k
Indicate step-by-step which parts of Theorem 1.2.6 you use.
47
1 Analysis
41. Suppose that the sequence (an )n∈N is bounded. Prove that the sequence (cn )n∈N
defined by
a3 + 5n
cn = n2
an + n
is convergent, and find its limit.
42. Let (an )n∈N be a convergent sequence with limit 0 and suppose that an ≥ 0 for all
√
n ∈ N. Prove that the sequence ( an )n∈N is also convergent, with limit 0.
√
43. Show that n2 + n − n is a convergent sequence and find its limit.
n∈N
Hint: ‘Rationalise the numerator’ (as in the proof of Theorem 1.2.6.(8)).
44. Prove that if (an )n∈N and (bn )n∈N are convergent sequences such that for all n ∈ N,
an ≤ bn , then
lim an ≤ lim bn .
n→∞ n→∞
45. Show that, for each k ∈ N and each real number x with |x| < 1, limn→∞ xn nk = 0.
√
Hint: Write the elements of the sequence as (y n n)2k , for some suitably chosen y,
and use Exercise 27 as well as the Algebra of Limits.
46. Prove that the sequence (3 + (−1)n )1/n is convergent, with limit 1.
n∈N
47. Recall the floor function defined in Exercise 9: bxc = max{n ∈ Z | n ≤ x}.
bnxc
Let x be any real number. Set an = , for each n ∈ N. Show that the sequence
n
(an )n∈N is convergent, with limit x.
7 1 1 1 1 1
In fact, lim + + + ··· + + = ln 2, the proof of which is outside the
n→∞ n + 1 n+2 n+3 2n − 1 2n
scope of this course.
48
1.4 Exercises
nk+2 n∈N
is convergent and
1k + 2k + 3k + · · · + nk
lim = 0.
n→∞ nk+2
1
51. Suppose that (an )n∈N is a sequence with n
≤ an ≤ n for each n ∈ N. Show that
limn→∞ a1/n
n = 1.
Hint: See Exercise 28.
52. a) Show that a monotonically increasing sequence always has a constant subsequence
or a strictly increasing subsequence.
b) Show that any sequence of real numbers always has either a constant subsequence
or a strictly increasing subsequence or a strictly decreasing subsequence.
53. a) For x ∈ R, show that x(1 − x) ≤ 14 .
b) Define the sequence (an )n∈N recursively by a1 = 1/3 and an+1 = 4an (1 − an ) for
each n ∈ N. Write down a2 , a3 and a4 . Show that (an )n∈N has a convergent
subsequence.
54. a) Let L ∈ R and let (an )n∈N be a sequence of real numbers that does not converge to
L (that is, it is either divergent or its limit is not equal to L). Use the definition of
convergence to L to show that for some ε > 0, (an )n∈N has a subsequence (ank )k∈N
such that ank ∈ / (L − ε, L + ε) for all k ∈ N.
b) Use 54a) and the Bolzano-Weierstrass theorem to show that given any L ∈ R, any
bounded, divergent sequence has a convergent subsequence with limit not equal
to L.
c) Use 54b) to show that any bounded, divergent sequence has two subsequences
which converge to different limits.
49
1 Analysis
50
1.4 Exercises
63. Let f : R → R be a continuous function. Prove that if for some c ∈ R, f (c) > 0, then
there exists a δ > 0 such that for all x ∈ (c − δ, c + δ), f (x) > 0.
64. Recall Exercise 59. The function f : R → R is given by f (x) = x2 . Using the
characterization of continuity provided in Theorem 1.3.1, prove that f is continuous
on R.
66. Let f : R → R be a function that preserves divergent sequences, that is, for every
divergent sequence (xn )n∈N , (f (xn ))n∈N is divergent as well. Prove that f is one-to-one.
Hint: Let x1 , x2 be distinct real numbers, and consider the sequence x1 , x2 , x1 , x2 , . . . .
51
1 Analysis
q
72. a) Suppose g : I → R is continuous, and g(x) ≥ 0 for x ∈ I. Let f (x) = g(x). Show
that f is continuous on I.
b) Prove that the function f : R → R defined by
v s
u
u 1 + 2x + x2
f (x) = 1+ , x ∈ R,
t
2 − 2x + x2
is continuous.
73. Suppose that f : [0, 1] → R is a continuous function such that for all x ∈ [0, 1],
0 ≤ f (x) ≤ 1. Prove that there exists at least one c ∈ [0, 1] such that f (c) = c.
Hint: Consider the continuous function g(x) = f (x) − x, and use the intermediate
value theorem.
74. At 8:00 a.m. on Saturday, a hiker begins walking up the side of a mountain to his
weekend campsite. On Sunday morning at 8:00 a.m., he walks back down the mountain
along the same trail. It takes him one hour to walk up, but only half an hour to walk
down. At some point on his way down, he realizes that he was at the same spot at
exactly the same time on Saturday. Prove that he is right.
Hint: Let u(t) and d(t) be the position functions for the walks up and down, and
apply the intermediate value theorem to f (t) = u(t) − d(t).
75. Show that the polynomial function p(x) = 2x3 − 5x2 − 10x + 5 has a real root in the
interval [−1, 2].
76. Set p(x) = x8 + bx3 − 5, where b is a real number. Show that p(x) has a root in [0, 1]
iff b ≥ 4.
77. Let f : R → R be continuous. If S := {f (x) | x ∈ R} is neither bounded above nor
bounded below, prove that S = R.
Hint: If y ∈ R, then since S is neither bounded above nor bounded below, there exist
x0 , x1 ∈ R such that f (x0 ) < y < f (x1 ).
78. Let f : [0, 1] → R be an arbitrary continuous function. Show that there exists a
c ∈ [0, 1] such that
f (c) − f (1) = (f (0) − f (1))c.
Hint: Find an appropriate function g on which the Intermediate Value theorem can
be applied.
79. a) Show that, given any continuous function f : R → R, there exists x0 ∈ [0, 1] and
m ∈ Z such that f (x0 ) = mx0 . In other words, the graph of f intersects some line
y = mx with integer slope, at some point x0 in [0, 1].
Hint: If f (0) = 0, take x0 = 0 and any m ∈ Z. If f (0) > 0, then choose N ∈ N
satisfying N > f (1), and apply the intermediate value theorem to the continuous
function g(x) = f (x) − N x on the interval [0, 1]. If f (0) < 0, then proceed in a
similar manner.
52
1.4 Exercises
b) Prove that there does not exist a continuous function f : R → R such that f
assumes rational values at irrational numbers, and irrational values at rational
numbers, that is,
f (Q) ⊆ R \ Q and f (R \ Q) ⊆ Q.
Hint: Show that, if f is such a function, then there do not exist m ∈ Z and
x0 ∈ R such that f (x0 ) = mx0 .
80. Complete the proof of Theorem 1.3.6 by proving that if f : [a, b] → R is continuous,
then there exists d ∈ [a, b] such that f (d) = inf {f (x) | x ∈ [a, b]}.
81. Show that the statement of Theorem 1.3.6 does not hold if [a, b] is replaced by [a, b).
82. Give examples of functions f1 , f2 , f3 : [0, 1] → R, continuous on (0, 1], with the following
properties.
a) f1 is not continuous at 0, and {f (x) | x ∈ [0, 1]} is not bounded above;
b) f2 is not continuous at 0, and {f (x) | x ∈ [0, 1]} is bounded above, but does not
have a maximum;
c) f3 is not continuous at 0, and {f (x) | x ∈ [0, 1]} has a maximum.
85. A function f : R → R is a periodic function if there exists T > 0 such that for all
x ∈ R, f (x + T ) = f (x). If f : R → R is continuous and periodic, then prove that f
is bounded, that is, the set S = {f (x) | x ∈ R} is bounded.
53
1 Analysis
54
Chapter 2
Algebra
We have already seen many different “algebraic structures”, that is, different sets with
algebraic operations defined on them. For example, we know of the natural numbers
N on which addition (+) and multiplication (·) are defined, the integers Z on which
subtraction (−) is additionally defined, and the rational numbers Q on which division (|)
is defined in addition to the previous operations. What distinguishes the real numbers R
from Q is not so much an algebraic property as an order property (the l.u.b. property).
On the other hand, the difference between R and the complex numbers C is that certain
equations with no real solutions, such as x2 + 1 = 0, have solutions in C. Thus C has an
algebraic property not shared by R.
When we consider the operations on the above sets, we see for example that addition
and multiplication are commutative:
x + y = y + x and xy = yx
and associative
(x + y) + z = x + (y + z) and (xy)z = x(yz).
Subtraction and division do not share these properties. On the other hand, they are not
really operations in their own right, but are derived from addition and multiplication
along with the notion of inverses. Also, addition and multiplication have identity elements
(0 and 1, respectively).
We have also seen other algebraic systems, such as matrices, on which addition and
multiplication have been defined. We know that matrix addition is commutative and
associative (A + B = B + A and AB = BA), and matrix multiplication is associative:
(AB)C = A(BC),
AB 6= BA in general.
Again, there are identity elements (the zero matrix O for addition and the identity matrix
I for multiplication).
55
2 Algebra
There are many more different algebraic structures in mathematics. Abstract algebra
is the study of general properties of these algebraic structures. In this part of the course,
we study two important abstract algebraic structures: groups and vector spaces. We
begin with a discussion about groups, one of the simplest algebraic structures.
2.1 Groups
In this section, we study one of the most basic algebraic objects, namely a group. A
group is a set on which an operation or law of composition is defined, such that certain
properties hold. The precise definition is given in the next subsection.
with a fixed choice being made for the particular law in question. In this notation we
then have that a ∗ b is a law of composition on S iff a ∗ b ∈ S for any a, b ∈ S. For later
reference, we state the property of ∗ being a law of composition (which any particular ∗
may either have or not have), as an axiom.
G0 (Law of Composition) For all a, b ∈ G, a ∗ b ∈ G.
Examples.
1. The addition of integers is a law of composition on Z. Indeed, the sum of two integers
is yet another integer, and addition is the function from the set Z × Z to the set Z
that assigns a + b to the pair (a, b), denoted by (a, b) 7→ a + b.
2. The multiplication of real numbers is a law of composition on R.
3. If a, b are rational numbers, then let a ∗ b = a + b − ab. The function from Q × Q to
Q given by (a, b) 7→ a ∗ b is a law of composition on Q.
√
4. If a, b are real numbers, then define a ∗ b = a2 √+ b2 . Then (a, b) 7→ a ∗ b is not a law
of composition on Q, since 1 ∈ Q, but 1 ∗ 1 = 2 6∈ Q. However, (a, b) 7→ a ∗ b is a
law of composition on R.
5. Let n ∈ N, and let S denote the set of all matrices of size n × n with real entries.
Then matrix multiplication is a law of composition on S.
56
2.1 Groups
6. Let n ∈ N, and let GL(n, R) denote the set of all invertible matrices of size n × n
with real entries. Then matrix multiplication is a law of composition on GL(n, R).
Indeed, if A, B ∈ GL(n, R), then the matrix AB is again a matrix of size n × n with
real entries, and moreover, since A and B are invertible, it follows that AB is also
invertible.
7. Let a, b be real numbers such that a < b. Let C[a, b] denote the set of all continuous
functions on the interval [a, b]. Let addition of functions be defined as follows: if f, g
belong to C[a, b], then
Then addition of functions is a law of composition on C[a, b], since the sum of
continuous functions is again continuous; see Theorem 1.3.2.
8. If a, b ∈ N, then let
a
a∗b= .
b
∗ is not a law of composition on N, since 1 ∗ 2 = 12 6∈ N. The function (a, b) 7→ ab is
also not a law of composition on Q, since (1, 0) is not mapped to any rational number
by the function.
On a set there may be many different laws of compositions that can be defined. Some
laws of compositions are nicer than others, that is, they possess some desirable properties.
A group is a set G together with a law of composition on G that has three such desirable
properties, and we give the definition below.
Definition. A group is a set G together with a law of composition (a, b) 7→ a∗b : G×G →
G, which has the following properties:
G3 (Inverse axiom) For every a ∈ G, there exists an element a−1 ∈ G such that
a ∗ a−1 = e = a−1 ∗ a. Such an element a−1 is called an inverse of the element a
in the group G.
G1, G2, G3 are called the group axioms. Sometimes we use the notation (G, ∗) for the
group.
Remark. Note that hidden in the definition of a group is the axiom G0 stating that ∗ is
actually a law of composition on G. Hence when checking that a certain set G is a group
with respect to a certain operation ∗ that combines pairs of elements from G, we have
to check first of all that the axiom G0 is satisfied as well: for every a, b ∈ G, a ∗ b
belongs to G.
57
2 Algebra
Examples.
1. Z with addition is a group, as may be seen by checking each of the axioms G0, G1,
G2 and G3:
G0. For all a, b ∈ Z, a + b ∈ Z.
G1. For all a, b, c ∈ Z, (a + b) + c = a + (b + c).
G2. 0 serves as an identity element: for all a ∈ Z, a + 0 = a = 0 + a.
G3. If a ∈ Z, then −a ∈ Z and a + (−a) = 0 = −a + a.
3. For n ∈ N, let GL(n, R) denote the set of all invertible matrices of size n × n with
real entries. Then GL(n, R) is group with matrix multiplication:
G0. For all A, B ∈ GL(n, R), since A and B are invertible n × n matrices, det A 6= 0
and det B = 6 0. Therefore, det(AB) = (det A)(det B) 6= 0, which gives that AB
is also an invertible n × n matrix. That is, AB ∈ GL(n, R).
G1. For all A, B, C ∈ GL(n, R), (AB)C = A(BC), since matrix multiplication is
associative.
58
2.1 Groups
4. Let a, b be real numbers such that a < b. Then C[a, b] is a group with addition of
functions:
G0. For all f, g ∈ C[a, b], f + g ∈ C[a, b] since the sum of continuous functions is
continuous.
G1. For all f, g, h ∈ C[a, b], and any x ∈ [a, b] we have
Hence f + (g + h) = (f + g) + h.
G2. The constant function 0 defined by 0(x) = 0 for all x ∈ [a, b], serves as an identity
element. 0 is a continuous function on [a, b] and so 0 ∈ C[a, b]. Moreover, for all
f ∈ C[a, b], we have for all x ∈ [a, b]:
Hence f + 0 = f = 0 + f .
59
2 Algebra
G3. If f ∈ C[a, b], then define −f by (−f )(x) = −f (x), for x ∈ [a, b]. Given
f ∈ C[a, b], we have for all x ∈ [a, b]:
Hence f + (−f ) = 0 = −f + f .
5. For any set S, its power set P(S) is defined to be the set consisting of all subsets of
S:
P(S) = {A | A ⊆ S} .
Define the following law of composition on P(A), called the symmetric difference
operation:
A 4 B = (A \ B) ∪ (B \ A), A, B ∈ P(S).
A B
A B
G2. Since A 4 ∅ = A = ∅ 4 A for all A ∈ P(S), the empty set ∅ serves as an identity
element.
G3. Since A 4 A = ∅, each element is its own inverse: A−1 = A for all A ∈ P(S).
60
2.1 Groups
6. Given n ∈ N, let Sn be the set of all bijections from the set {1, 2, . . . , n} to itself,
and let ◦ denote the composition of functions, so α ◦ β means the function with
(α ◦ β)(x) = α(β(x)). Then (Sn , ◦) is a group, called the symmetric group on n
symbols.
G0. You have seen that the composition of two bijections is a bijection, so indeed if
α and β are in Sn then α ◦ β is also in Sn .
G1. If α, β, γ are three elements of Sn , and x ∈ {1, . . . , n}, then
(α ◦ β) ◦ γ (x) = (α ◦ β)(γ(x)) = α(β(γ(x))) = α((β ◦ γ)(x)) = α ◦ (β ◦ γ) (x),
Definitions. A group (G, ∗) is said to be a finite group if the set G has finite cardinality.
The order of a finite group (G, ∗) is the cardinality of G. A group is said to be an
infinite group if it is not finite.
Examples.
3. The set P(S) with the symmetric difference operation 4 is a finite group if S is a
finite set, and an infinite group if S is an infinite set.
Definition. The group table of a finite group is a table that displays the law of
composition as follows: the elements of the group are listed in the first row and the
first column. Conventionally, the two lists have the group elements in the same order,
with the identity element first. Given a, b ∈ G, the element a ∗ b is entered in the row
corresponding to a and the column corresponding to b, as shown below.
∗ ··· b ···
..
.
a a∗b
..
.
We clarify this further by considering a few examples.
61
2 Algebra
Examples.
1. The finite group {−1, 1} with multiplication can be described by the group table given
below.
· 1 −1
1 1 −1
−1 −1 1
We can now form the group table for (S3 , ◦). For example, since
α ◦ β(1) = α(β(1)) = α(3) = 2,
α ◦ β(2) = α(β(2)) = α(2) = 3,
α ◦ β(3) = α(β(3)) = α(1) = 1,
we obtain that α ◦ β = δ. In a similar way, the whole group table may be calculated:
◦ ι α β γ δ ε
ι ι α β γ δ ε
α α ι δ ε β γ
β β ε ι δ γ α
γ γ δ ε ι α β
δ δ γ α β ε ι
ε ε β γ α ι δ
3. The finite group P({1, 2}) with the symmetric difference operation 4 has four elements:
P({1, 2}) = {∅, {1}, {2}, {1, 2}}
and has the following group table:
4 ∅ {1} {2} {1, 2}
∅ ∅ {1} {2} {1, 2}
{1} {1} ∅ {1, 2} {2}
{2} {2} {1, 2} ∅ {1}
{1, 2} {1, 2} {2} {1} ∅
Again the table completely describes the law of composition. For example,
{1, 2} 4 {2} = ({1, 2} \ {2}) ∪ ({2} \ {1, 2}) = {1} ∪ ∅ = {1}.
62
2.1 Groups
Proof. Let e and e0 be identity elements in (G, ∗). Since e ∈ G and e0 is an identity, we
obtain
e = e ∗ e0 .
Moreover, since e0 ∈ G and e is an identity, we also have
e ∗ e0 = e0 .
Consequently, e = e0 .
Theorem 2.1.2. Let (G, ∗) be a group and let a ∈ G. Then a has a unique inverse.
Proof. Let the group have the identity e. If a1 and a2 are inverses of a, then we have
Examples.
1. Let a and b be elements of a group (G, ∗). Find all x ∈ G such that:
a ∗ x = b.
Solution. Left-multiply both sides of the equation by a−1 (which exists by G3 and
satisfies a ∗ a−1 = e = a−1 ∗ a where e is the identity element of G):
a−1 ∗ (a ∗ x) = a−1 ∗ b.
Then by G1, (a−1 ∗ a) ∗ x = a−1 ∗ b,
by G3, e ∗ x = a−1 ∗ b,
and finally by G2, x = a−1 ∗ b.
We must now check that x = a−1 ∗ b is a solution to the equation a ∗ x = b, i.e., that
a ∗ (a−1 ∗ b) = b. This follows via a very similar calculation.
63
2 Algebra
Solution. Note in each step how the associativity axiom (G1) is applied:
(a ◦ b) ◦ (c ◦ d) = a ◦ (b ◦ (c ◦ d))
= a ◦ ((b ◦ c) ◦ d)
= (a ◦ (b ◦ c)) ◦ d.
In the previous example we saw that by the associativity axiom, two ways of using
parentheses on a ◦ b ◦ c ◦ d resulted in equal expressions. It can be shown that in
general, for any elements a1 , a2 , . . . , an of a group (G, ∗), any two ways of parenthesising
a1 ∗ a2 ∗ · · · ∗ an will result in equal elements in the group. We therefore in general do
not show the parentheses. Instead we write a ◦ b ◦ c ◦ d or a1 ∗ a2 ∗ · · · ∗ an , or even just
abcd or a1 a2 . . . an , when it is clear which law of composition is meant.
Moreover, if n ∈ N, define
a−n = (an )−1 .
It can be shown that the usual laws of exponents hold: for all m, n ∈ Z,
Example. Let b be an element of a group G with identity element e. Find all solutions
x to the following simultaneous equations:
x2 = b and x5 = e.
64
2.1 Groups
2. The group (R∗ , ×) of non-zero real numbers with multiplication is an abelian group.
3. Let m, n ∈ N. The set of matrices Rm×n of size m × n with entries in R with matrix
addition is an abelian group.
4. Let n ∈ N. The set GL(n, R) with matrix multiplication is a group, but it is not an
abelian group if n ≥ 2.
6. The symmetric group S3 is not an abelian group. For instance, referring to the group
table given earlier, we see that α ◦ β = δ, but β ◦ α = ε 6= δ.
We normally use additive notation when referring with abelian groups. We then
use a symbol such as ‘+’, ‘⊕’, ‘’, etc., to denote the law of composition,
65
2 Algebra
2.1.4 Subgroups
Definition. A subset H of a group G, with operation ∗, is called a subgroup of G if it
has the following properties:
H1 (Closure) If a, b ∈ H, then a ∗ b ∈ H.
H2 (Identity) e ∈ H.
The three properties of a subgroup H ensure that (H, ∗) is itself a group, as we now
explain.
H1. This condition tells us that the law of composition ∗ on the group G can be used
to define a law of composition on H, namely, the function from H × H to H
given by (a, b) 7→ a ∗ b, which is called the induced law of composition. Since ∗ is
associative, it follows that the induced law of composition is associative as well: for
all a, b, c ∈ H, a ∗ (b ∗ c) = (a ∗ b) ∗ c.
H2. This shows that H, with the induced law of composition, has an identity element.
Indeed the identity element from G (which also belongs to H) serves as the identity
element also in H: for every a ∈ H, a ∗ e = a = e ∗ a.
H3. Finally this shows that every element in H possesses an inverse element in H. Of
course, since G is a group, we already knew that a possesses an inverse element
a−1 ∈ G. But now H3 says that this inverse element is in H. Thus for all a ∈ H,
∃a−1 ∈ H such that a ∗ a−1 = e = a−1 ∗ a.
Thus the conditions H1, H2, H3 imply that the subset H, with the induced law of
composition, is a group. Thus, a subgroup is itself a group which sits in a larger group.
Examples.
2. The group of integers with addition (Z, +) is a subgroup of the group of rational
numbers with addition (Q, +), which in turn is a subgroup of the group of real numbers
with addition (R, +).
66
2.1 Groups
is a subgroup of the set of all 2 × 2 matrices with real entries together with matrix
addition. Indeed, given any two symmetric matrices
" # " #
a b a0 b 0
and ,
b d b0 d0
their sum " # " # " #
a b a0 b 0 a + a0 b + b 0
+ =
b d b0 d 0 b + b0 d + d0
is also symmetric, and so H1 holds. Clearly the identity element
" #
0 0
0 0
is symmetric, and so H2 also holds. Finally, the inverse (with respect to matrix
addition) of any symmetric matrix
" #
a b
,
b d
is the element " #
−a −b
−b −d
which is also symmetric, and so H3 holds.
5. The subset of upper triangular invertible matrices,
(" # )
a b
U T (2, R) =
a, b, d ∈ R and ad 6= 0
0 d
67
2 Algebra
ord(a) = min {m ∈ N | am = e} .
3. If a does not have finite order, then a is said to have infinite order.
Examples.
1. The element −1 has order 2 in the group of non-zero real numbers with multiplication.
2. The element 2 has infinite order in the group of integers with addition.
0 0 1
3. The element 1 0 0 is an element of order 3 in the group GL(3, R).
0 1 0
We now prove that given any element from a group, the set of its powers forms a
subgroup of the group.
Theorem 2.1.3. Suppose that (G, ∗) is a group and a ∈ G. Let hai = {an | n ∈ Z}.
Then:
Proof.
68
2.1 Groups
Example. The element [1] in the group Z5 with addition modulo 5 has finite order, since
and the subgroup h[1]i = {[0], [1], [2], [3], [4]} is in fact the whole group.
The above example motivates the following definition.
Definition. A group G is said to be a cyclic group if there exists an element a ∈ G
such that G = hai. Such an element a is then called a generator of the group G.
2. Let G be the group GL(2, R) with matrix multiplication, and G0 be the group of non-
zero real numbers with multiplication. Then the determinant function det : G → G0 is a
homomorphism, since for all 2×2 real matrices A, B we have det(AB) = det(A) det(B).
69
2 Algebra
Theorem 2.1.4. Let G be a group with identity e and G0 be a group with identity e0 . If
ϕ : G → G0 is a homomorphism, then:
1. ϕ(e) = e0 .
Proof. We have
e0 ∗0 ϕ(e) = ϕ(e) = ϕ(e ∗ e) = ϕ(e) ∗0 ϕ(e),
and so cancelling ϕ(e) on both sides, we obtain e0 = ϕ(e). Next,
and similarly,
e0 = ϕ(e) = ϕ(a ∗ a−1 ) = ϕ(a) ∗0 ϕ(a−1 ).
Thus ϕ(a−1 ) ∗0 ϕ(a) = e0 = ϕ(a) ∗0 ϕ(a−1 ), and by the uniqueness of the inverse of ϕ(a)
in G0 , we obtain (ϕ(a))−1 = ϕ(a−1 ).
Every group homomorphism ϕ determines two important subgroups: its image and its
kernel.
1. ker(ϕ) is a subgroup of G.
2. im(ϕ) is a subgroup of G0 .
70
2.1 Groups
Examples.
1. Let G be R with addition, and G0 be the set of positive reals with multiplication. The
exponential function from G to G0 , given by x 7→ 2x , is a homomorphism with kernel
the trivial subgroup comprising the element 0, and the image is the whole group of
positive reals with multiplication, since it can be shown that given any y > 0, there
exists a unique real number (called the logarithm of y to the base 2, denoted by log2 y)
such that y = 2log2 y .
2. Let G be the group GL(2, R) with matrix multiplication, and G0 be the group of non-
zero real numbers with multiplication. Then the determinant function det : G → G0
is a homomorphism. Its kernel is the set of all invertible matrices with determinant
equal to 1, and we denote this subgroup by SL(2, R), and it is called the special
linear group:
SL(2, R) = {A ∈ GL(2, R) | det(A) = 1} .
The image of this homomorphism is the whole group of non-zero reals: indeed, given
any real number a not equal to zero, we have that
" #
1 0
A := ∈ GL(2, R),
0 a
and det(A) = 1 · a − 0 · 0 = a.
3. Let G be the group C[0, 1] of continuous functions on the interval
[0, 1] with addition,
0
and G be the group R with addition. Then the function f 7→ f 12 is a homomorphism,
and its kernel is the set of all continuous functions on the interval [0, 1] that have
a root at 12 (for instance the straight line f (x) = x − 12 belongs to the kernel). The
image is the set of all real numbers, since given any a ∈ R, the constant function
f (x) = a for all x ∈ [0, 1] is continuous, and f 12 = a.
In Example 1 above, the homomorphism between the two groups was also bijective.
We give a special name to such homomorphisms.
Definition. Let G, G0 be groups. A homomorphism ϕ : G → G0 is said to be an iso-
morphism if it is bijective.
Examples.
1. Let G be R with addition, and G0 be the set of positive reals with multiplication. The
exponential function from G to G0 , given by x 7→ 2x , is an isomorphism.
2. If G is a group, then the identity function ι : G → G defined by ι(a) = a for all a ∈ G
is an isomorphism.
" #
1 x
3. Let G be the subgroup of GL(2, R) comprising all matrices of the form ,
0 1
where x ∈" R. Let# G0 be the group R with addition. Then the function from G to G0 ,
1 x
given by 7→ x, is an isomorphism.
0 1
71
2 Algebra
Isomorphisms are important because their existence between two groups means that
the two groups are essentially the “same”, in the sense that as far as algebraic properties
go, there is no real difference between them.
Definition. Two groups G and G0 are called isomorphic if there exists an isomorphism
ϕ : G → G0 .
{x ∈ G | a R x} = {x ∈ G | ∃h ∈ H such that x = a ∗ h} = {a ∗ h | h ∈ H} ,
We know that the equivalence classes of an equivalence relation partition the set. Recall
that by a partition of a set S, we mean a subdivision of the set S into nonoverlapping
subsets:
72
2.1 Groups
G b∗H
a ∗ H = a0 ∗ H
(a R a0 )
e∗H =H
Corollary 2.1.6. Let H be a subgroup of a group G. Then the left cosets of H partition
the group G.
Remarks.
a ∗ H = b ∗ H iff a R b.
Corollary 2.1.6 says that if a ∗ H and b ∗ H have an element in common then they are
equal.
a R b; b R a; a ∗ H = b ∗ H; a ∈ b ∗ H; b ∈ a ∗ H.
Examples.
1. Consider the group (Z6 ⊕) of integers modulo 6, with addition modulo 6, and let H be
the subgroup h[2]i = {[0], [2], [4]}. The left cosets, which we now denote by a ⊕ H are
Note that the cosets {[0], [2], [4]} and {[1], [3], [5]} form a partition of G:
G = {[0], [1], [2], [3], [4], [5]} = {[0], [2], [4]} ∪ {[1], [3], [5]},
and {[0], [2], [4]} ∩ {[1], [3], [5]} = ∅.
73
2 Algebra
2. Consider the group (Z, +), and let H be the subgroup of even numbers {2m | m ∈ Z}.
The left cosets, which we now denote by a + H are
· · · = −4 + H = −2 + H = 0 + H
= {2m | m ∈ Z} = {. . . , −4, −2, 0, 2, 4, . . . }
= 2 + H = 4 + H = 6 + H = ···,
and
· · · = −3 + H = −1 + H = 1 + H
= {2m + 1 | m ∈ Z} = {. . . , −3, −1, 1, 3, . . . }
= 3 + H = 5 + H = 7 + H = ···.
Note that the cosets {. . . , −4, −2, 0, 2, 4, . . . } and {. . . , −3, −1, 1, 3, . . . } do indeed
partition the set of all integers.
Note that in Example 1 above, there are only two distinct cosets, and
In particular the order of H (namely 3) divides the order of G (namely 6). This is not a
coincidence. We now prove an important result concerning the order of a group G and
the number of cosets of a subgroup H, due to Lagrange2 .
Theorem 2.1.7 (Lagrange’s Theorem). Let H be a subgroup of a finite group G. Then
the order of H divides the order of G.
Proof. We claim that, for any a ∈ G, the function φ from the subgroup H to the coset
a ∗ H, given by φ(h) = a ∗ h, for h ∈ H, is a bijection. Indeed, it is a surjection by
definition of the coset a ∗ H; to see that it is an injection, we note that φ(h) = φ(h0 ) ⇒
a ∗ h = a ∗ h0 ⇒ h = h0 .
Consequently, each coset a ∗ H has the same number of elements as H does.
Since G is the union of the cosets of H, and since these cosets do not overlap, we
obtain the counting formula
74
2.1 Groups
The following theorem characterises all groups whose order is a prime number.
Corollary 2.1.9. If G is a group with prime order p, then G is cyclic, and G = hai for
every a ∈ G \ {e}.
Proof. If a =6 e, then a has order > 1, say m. Since m divides p, and p is prime, it
follows that m = p. As G itself has order p, it now follows that G = hai, and so G is
cyclic.
2.1.8 Exercises
1. For each of the following definitions for a ∗ b and a given set, determine which of the
axioms G0, G1, G2, G3 are satisfied by a ∗ b. In which cases do we obtain a group?
a+b i) a + b − ab on the set R.
a) on the set Z.
ab
q j) max{a, b} on the set N.
b) |ab| on the set Q.
k) min{a, b} on the set Z.
c) a − b on the set Z.
l) a + b on R.
d) |a − b| on the set N ∪ {0}.
e) −ab on the set Q. m) a + b on [0, ∞).
3. a) Show that the set Z6 of integers modulo 6, with addition modulo 6 is a group.
Write down its group table.
b) Is Z6 with multiplication modulo 6 also a group? If, instead, we consider the set
Z∗6 of non-zero integers modulo 6, then is Z∗6 a group with multiplication modulo
6?
c) Let m be an integer such that m ≥ 2, and let Z∗m denote the set of non-zero
integers modulo m. Prove that Z∗m is a group with multiplication modulo m iff m
is a prime number.
Hint: If m is a prime number, then any r ∈ {1, . . . , m − 1} is coprime to m, and
so there exist integers s and t such that sm + tr = 1.
75
2 Algebra
6. Let G be a group with identity element e, and let a and b be elements of G. For each
of the following pairs of equations, find all solutions x, and determine conditions on a
and b for there to be any solutions to the equations.
a) x2 a = bx and ax = xa.
b) ax2 = b and x3 = e.
c) (xax)3 = bx and x2 a = (xa)−1 .
Note that, in this and some later exercises, the symbol for the group operation is
omitted – so ab is used instead of a ∗ b.
7. Let a and b be elements of a group (G, ∗). Show that (a ∗ b)2 = a2 ∗ b2 iff a ∗ b = b ∗ a.
Give an example of a group (G, ∗) and two elements a, b ∈ G such that (a∗b)2 6= a2 ∗b2 .
8. Determine if the following statements are TRUE or FALSE: Give reasons for your
answers in each case.
In any group G with identity element e,
a) for any x ∈ G, if x2 = e then x = e.
b) for any x ∈ G, if x2 = x then x = e.
c) for any x ∈ G there exists y ∈ G such that x = y 2 .
d) for any x, y ∈ G there exists z ∈ G such that y = xz.
9. Prove the following statements for any group G with identity element e:
a) For all a, b, c ∈ G, if abc = e then cab = e and bca = e.
b) Suppose that a = a−1 , b = b−1 and c = c−1 (that is, each of a, b and c is its own
inverse). For all a, b, c ∈ G, if ab = c then bc = a and ca = b.
c) For all a ∈ G, a = a−1 iff a2 = e.
76
2.1 Groups
10. a) Show that in any group, the inverse of the inverse of any element is itself.
b) (∗) Let (G, ∗) be a finite group with identity element e. Show that if the order
of G is even, then there exists an element x ∈ G such that x 6= e and x−1 = x.
(That is, in a group of even order there always exists an element other than the
identity which is its own inverse.)
Hint: Use 10a) to show that there is an even number of elements that are not
their own inverses.
11. Show that the general linear group GL(3, R) with matrix multiplication is not an
abelian group.
12. Prove that a group G is abelian iff (ab)−1 = a−1 b−1 for all a, b ∈ G.
14. Determine if the following statements are TRUE or FALSE. Give reasons for your
answers in each case.
a) The nonnegative integers form a subgroup of Z with addition.
b) The odd integers form a subgroup of Z with addition.
c) The set {4k | k ∈ Z} is a subgroup of Z with addition.
d) If G is abelian and H is a subgroup of G, then H is abelian.
e) Every group has an abelian subgroup.
f) If H and K are subgroups of G and H ⊆ K, then H is a subgroup of K.
17. For n ∈ N, show that the set H = {α ∈ Sn | α(1) = 1} is a subgroup of (Sn , ◦).
18. Show that the set H = {x ∈ R | sin x ∈ Q and cos x ∈ Q} is a subgroup of the group
(R, +) of real numbers with addition.
You may use any properties of the sine and cosine functions that you know (or can
look up).
19. Let S and T be sets such that S ⊆ T . Note that then P(S) ⊆ P(T ), that is, every
subset of S is also a subset of T . Show that P(S) is a subgroup of the group (P(T ), 4).
20. a) Let (G, +) be an abelian group. Show that the set H = {x ∈ G | x = −x} (that
is, the set of all elements of G that are inverses of themselves) is a subgroup of G.
b) Show that the set H = {θ ∈ S3 | θ = θ−1 } is not a subgroup of (S3 , ◦).
77
2 Algebra
21. a) Show that, if H and K are subgroups of a group G, then H ∩ K is also a subgroup
of G.
b) Consider the group of integers Z with addition. Suppose that H is the subgroup of
Z comprising multiples of 4, and let K be the subgroup of Z comprising multiples
of 6. What is H ∩ K?
c) If H and K are subgroups of a group G, is H ∪ K necessarily a subgroup of G? If
true, give a proof. If false, provide a counterexample.
22. Let C[0, 1] denote the group comprising the set of continuous functions on the interval
[0, 1] with addition of functions defined pointwise: if f, g belong to C[0, 1], then for
all x ∈ [0, 1], (f + g)(x) = f (x) + g(x). Prove that each of the following subsets of
C[0, 1] are subgroups of C[0, 1].
n o
1
a) H1 = f ∈ C[0, 1] f =0 .
2
b) The set H2 of all polynomial functions, that is: the set of all functions p : [0, 1] → R
such that there exists an n ∈ N ∪ {0} and real numbers a0 , a1 , a2 , . . . , an such that
p(x) = a0 + a1 x + a2 x2 + · · · + an xn , for all x ∈ [0, 1].
c) H3 = {f ∈ C[0, 1] | ∀x ∈ [0, 1] f (1 − x) = −f (x)}.
23. Let G be a group. The centre of G is the set Z(G) = {z ∈ G | ∀a ∈ G, z ∗ a = a ∗ z}.
a) Show that Z(G) is not empty.
b) If G is abelian, then determine Z(G).
c) Show that Z(G) is a subgroup of G.
d) (∗) If G is the group GL(2, R) with matrix multiplication, then determine Z(G).
" # " #
1 1 1 0
Hint: Consider A = ,B = . Which elements Z of GL(2, R)
0 1 1 1
satisfy both ZA = AZ and ZB = BZ?
24. If G is a finite group, then show that every element in the group has finite order,
which is at most equal to |G|.
Hint: If a ∈ G, then consider the set S = {e, a, a2 , a3 , . . . , a|G| }, and use the pigeon-
hole principle.
" #
1 1
25. Determine the order of in the group GL(2, R).
−1 0
26. Determine the orders of all the elements of the following groups:
a) the group Z6 with addition modulo 6.
b) the symmetric group S3 .
c) the group (P(S), 4).
27. Let H be a subgroup of a group (G, ∗), and let h be an element of H. Is it necessarily
true that the order of h in (H, ∗) is the same as the order of h in (G, ∗)? Justify your
answer, by means of a proof or a counterexample.
78
2.1 Groups
28. Prove that in any group (G, ∗), and for any a, b in G, the orders of a ∗ b and b ∗ a are
the same. (Note that it is not given that (G, ∗) is abelian.)
29. Is the group of integers with addition cyclic? What is a generator of this group? Is it
unique?
30. Consider the group (Z, +), and let a be a positive integer. What are the elements of
the subgroup hai?
Show that if a and b are two positive integers, then the smallest subgroup of (Z, +)
containing both a and b is hdi, where d = gcd(a, b).
31. Show that any cyclic group is abelian. Give an example of an abelian group that is
not cyclic.
32. Define φ : (R, +) → (C∗ , ×) by φ(x) = eix = cos x + i sin x. Show that φ is a
homomorphism, and find its kernel and image. You may use any properties you know
about the function x 7→ eix .
33. a) Let G, G0 , G00 be groups and let ϕ : G → G0 and ψ : G0 → G00 be group homomor-
phisms. Prove that the composition ψ ◦ϕ : G → G00 defined by (ψ ◦ϕ)(a) = ψ(ϕ(a)),
a ∈ G is a group homomorphism.
b) Describe the kernel of ψ ◦ ϕ.
35. Let G be a group and let a be an element of G. Prove that the function from Z to
hai, given by m 7→ am , is a homomorphism from the group (Z, +) to G. What is the
image of this homomorphism? What is the kernel of the homomorphism: (a) if a has
finite order m, (b) if a has infinite order.
36. Show that the function ϕ(n) = 2n is an isomorphism from (Z, +) to its subgroup of
even integers {2k | k ∈ Z}.
37. Let G be the group defined on R with law of composition a ∗ b = a + b + 1. Show that
G and (R, +) are isomorphic.
Hint: Try ϕ(x) = x + 1.
38. Let G be the group defined on R \ {1} with law of composition a ∗ b = a + b − ab.
Show that G is isomorphic to the group R \ {0} with multiplication.
Hint: Try ϕ(x) = 1 − x.
40. Show that the function from G to G given by a 7→ a−1 is an isomorphism iff G is
abelian.
79
2 Algebra
41. Let G, G0 be groups and let ϕ : G → G0 be an isomorphism. Prove that the inverse
function ϕ−1 : G0 → G is also an isomorphism.
44. Determine if the following statements are TRUE or FALSE. Give reasons for your
answers in each case.
a) If H is a subgroup of G and a, b ∈ G are such that a 6= b, then (a ∗ H) ∩ (b ∗ H) = ∅.
b) If H is a subgroup of G and a ∈ G is such that a ∗ H has 4 elements, then H has
4 elements.
c) If H is a subgroup of a finite group G, then for any a ∈ G, the left coset a ∗ H has
the same number of elements as the right coset H ∗ a.
d) All groups of order 7 are isomorphic to (Z7 , +).
e) All groups of order 6 are isomorphic to (Z6 , +).
f) The group S4 does not have an element of order 13.
g) The group S4 does not have an element of order 24.
h) (∗) The group S4 does not have an element of order 12.
is a subgroup of G.
80
2.1 Groups
" #
x y
c) An element of G can be represented by the point (x, y) in the right
0 1
half-plane {(x, y) | x, y ∈ R and x > 0}. Draw the partitions of the right half-plane
into left and into right cosets of H.
(a) Show that G is closed under matrix multiplication, with the addition and multi-
plication carried out in Z5 .
" #
a b
(b) Find the inverse of the matrix in (G, ⊗), where ⊗ denotes matrix
0 a−1
multiplication as in (a).
You may assume without further proof that (G, ⊗) is a group.
(c) What is the order of G?
(d) Let H be the subset of G consisting of the diagonal matrices. Show that H is a
subgroup of G.
" #
2 1
(e) Let m = . Find all the elements of the left coset m ⊕ H.
0 3
47. Let H and K be subgroups of a group G of orders 3 and 5 respectively. Prove that
H ∩ K = {e}.
48. a) Let p be a prime, and let (Z∗p , ⊗) be the group of non-zero integers modulo p, with
multiplication modulo p. Show that, if a is an integer such that a is not divisible
by p, then [a]p−1 = [1].
b) Prove Fermat’s little theorem: for any integer a, ap ≡ a (mod p).
Hint: If a ∈ Z is not divisible by p, then [a] 6= [0], and so by part (48a) above,
[a]p−1 = [1]. Hence p | (ap−1 − 1), and so p | (ap − a).
c) Show that 7 divides 22225555 + 55552222 .
Hint: Note that 2222 = 7 · 317 + 3, so that in Z7 , [22225555 ] = [3]5555 . Now use
the fact that [3]6 = [1] to conclude that [22225555 ] = [35 ]. Now simplify [55552222 ]
in a similar fashion.
81
2 Algebra
82
2.2 Vector spaces
V3 and V4 are called the distributive laws. The elements of a vector space are called
vectors.
We observe that since (V, +) is an abelian group, a vector space also has the following
properties built into its definition: (note the use of additive notation)
G2. There exists an element o ∈ V (called the4 zero vector) such that for all v ∈ V ,
v + o = v = o + v.
G3. For every v ∈ V , there exists a unique5 element in V , denoted by −v, such that
v + (−v) = o = −v + v.
A. For all v1 , v2 ∈ V , v1 + v2 = v2 + v1 .
Examples.
1. Let m, n ∈ N, and consider the set Rm×n of m × n matrices having real entries. It
is easy to check that Rm×n , with matrix addition, is an abelian group: we omit the
verification of G1, G2, G3 and A. Define scalar multiplication as follows: if α ∈ R and
a11 . . . a1n
. .. m×n
A = ..
∈R
. ,
am1 . . . amn
4
Since there is a unique identity element in a group, the zero vector is unique!
5
In a group, every element has a unique inverse!
83
2 Algebra
then
αa11 . . . αa1n
α · A = ... ..
.
. (2.2)
αam1 . . . αamn
84
2.2 Vector spaces
a11 . . . a1n a11 . . . a1n
. . . ..
. . + β · ..
.
=α· .
.
am1 . . . amn am1 . . . amn
= α · A + β · A.
Hence Rm×n is a vector space with matrix addition and with scalar multiplication
defined by (2.2). If n = 1, then we denote the vector space of column vectors Rm×1
by Rm .
2. Let a, b be real numbers with a < b. The set C[a, b] of continuous functions on the
interval [a, b] with addition of functions is an abelian group. Let scalar multiplication
be defined as follows:
if α ∈ R and f ∈ C[a, b], then (α · f )(x) = αf (x), x ∈ [a, b]. (2.3)
(We say that scalar multiplication, and addition, are defined “pointwise”.) Then
α · f ∈ C[a, b], and moreover V1, V2, V3, V4 are satisfied:
V1. Let f ∈ C[a, b]. For all x ∈ [a, b], we have
(1 · f )(x) = 1f (x) = f (x),
and so 1 · f = f .
85
2 Algebra
V2. Let α, β ∈ R and f ∈ C[a, b]. For all x ∈ [a, b], we have
and so (α · (β · f ) = (αβ) · f .
V3. Let α, β ∈ R and f ∈ C[a, b]. For all x ∈ [a, b], we have
and so (α + β) · f = α · f + β · f .
V4. Let α ∈ R and f, g ∈ C[a, b]. For all x ∈ [a, b], we have
and so α · (f + g) = α · f + α · g.
Hence C[a, b] with addition and scalar multiplication is a vector space, called the
vector space of continuous functions on [a, b].
1. For all v ∈ V , 0 · v = o.
2. For all α ∈ R, α · o = o.
0 · v + 0 · v = (0 + 0) · v = 0 · v.
86
2.2 Vector spaces
3. Finally, we have
S1. o ∈ U .
S2. If v1 , v2 ∈ U , then v1 + v2 ∈ U .
Subspaces of a vector space are analogous to subgroups of a group in the sense that
a subspace of a vector space is itself a vector space with the same addition and scalar
multiplication as with V (this is easy to check). So a subspace is really a smaller vector
space sitting inside a larger vector space.
Examples.
1. If V is a vector space, then the subset U comprising only the zero vector, namely
U = {o}, is a subspace of V .
Also, the entire vector space, that is U = V , is a subspace of V .
If a subspace U of V is neither {o} nor V , then it is called a proper subspace of V .
2. Consider the vector space R2×2 with matrix addition and scalar multiplication defined
by (2.2). Then the set of upper triangular matrices
(" # )
a b
U1 = a, b, d ∈ R
0 d
is a subspace of R2×2 .
Also, the set of symmetric matrices
(" # )
a b
U2 = a, b, d ∈ R
b d
is a subspace of R2×2 .
87
2 Algebra
∃n ∈ N ∪ {0} and
a0 , a1 , a2 , . . . , an ∈ R such that
P [a, b] = p : [a, b] → R
.
p(x) = a0 + a1 x + a2 x2 + · · · + an xn
for all x ∈ [a, b]
Then U = P [a, b] is a subspace of the vector space C[a, b] with addition of functions
and scalar multiplication defined by (2.3).
Definitions. Let V be a vector space.
1. If v1 , . . . , vn are vectors in V and α1 , . . . , αn belong to R, then the vector α1 · v1 +
· · · + αn · vn is called a linear combination of the vectors v1 , . . . , vn .
lin(S) = {α1 · v1 + · · · + αn · vn | n ∈ N, v1 , . . . , vn ∈ S, α1 , . . . , αn ∈ R} .
Hence lin({e1 , e2 . . . , em }) = Rm .
2. Let a, b ∈ R with a < b. Any polynomial p on the interval [a, b] is a linear combination
of the functions from the set S = {1, x, x2 , . . . }. Hence lin(S) = P [a, b] in the vector
space C[a, b].
The span of a set of vectors turns out to be a special subspace of the vector space.
Theorem 2.2.2. Let V be a vector space and S be a subset of V . Then lin(S) is a
subspace of V , and moreover lin(S) is the smallest subspace of V that contains S.
Proof. The result is true in the case where S = ∅, as lin(∅) = {o} is the smallest
subspace of V .
Suppose that S is non-empty, and take any v ∈ S. Then o = 0 · v ∈ lin(S). If
u, v ∈ lin(S), then we know that
u = α1 · u1 + · · · + αn · un and v = β1 · v1 + · · · + βm · vm
6
Note that although S might be infinite, a linear combination, by definition, is always a linear
combination of a finite set of vectors from S.
88
2.2 Vector spaces
u + v = α1 · u1 + · · · + αn · un + β1 · v1 + · · · + βm · vm ∈ lin(S).
β · v = β · (α1 · v1 + · · · + αn · vn )
= β · (α1 · v1 ) + · · · + β · (αn · vn )
= (βα1 ) · v1 + · · · + (βαn ) · vn ∈ lin(S).
0 · v1 + · · · + 0 · vk−1 + 1 · vk + 0 · vk+1 + · · · + 0 · vn = o.
2. The vectors
1 0 0
0
1
0
e1 = .. , e2 = .. , . . . , em = ..
. . .
0 0 1
7
Note that (α1 , α2 , . . . , αn ) 6= (0, 0, . . . , 0) iff not all of α1 , . . . , αn are equal to 0, that is, iff at least
one of them is non-zero.
89
2 Algebra
and so α1 = · · · = αm = 0.
3. Let a, b ∈ R with a < b. The functions 1, x on the interval [a, b] are linearly independent.
Indeed, if for all x ∈ [a, b], α · 1 + β · x = 0(x), then in particular, we have
α + βa = 0 and α + βb = 0,
4. It follows from the definition that the empty set is not linearly dependent, since it
does not contain a non-empty subset. Thus the empty set is a linearly independent
subset of any vector space.
Examples.
1. The vectors
1 0 0
0
1
0
e1 = .. , e2 = .. , . . . , em = ..
. . .
0 0 1
form a basis of Rm .
2. The empty set ∅ is a basis of the subspace {o} of any vector space.
Theorem 2.2.3. Let V be a vector space and B be a basis of V such that B has n
elements. Then for any linearly independent set S of m ≤ n vectors there exists a subset
B 0 ⊂ B of n − m vectors such that S ∪ B 0 is also a basis of V .
90
2.2 Vector spaces
α1 , α2 , . . . , αm−1 , β1 , β2 , . . . , βn−m+1 ∈ R
such that
α1 v1 + α2 v2 + · · · + αm−1 vm−1 − vm = o,
which gives
α1 αm−1 1 β1
uj = − v1 − · · · − vm−1 + vm − u1 − . . .
βj βj βj βj
(2.5)
βj−1 βj+1 βn−m+1
− uj−1 − uj+1 − · · · − un−m+1 .
βj βj βj
Now remove uj from B10 to obtain B 0 = B10 \ {uj }. Then B 0 contains n − m vectors. It
remains to prove that S ∪ B 0 is a basis of V . For this we have to show that
91
2 Algebra
1. lin(S ∪ B 0 ) = V , and
2. S ∪ B 0 is linearly independent.
First we show that lin(S ∪ B 0 ) = V . Let v ∈ V . Since S1 ∪ B10 is a basis, there exist
λ1 , . . . , λm−1 , µ1 , . . . , µn−m+1 ∈ R such that
Use (2.5) to eliminate uj in the right-hand side. It is not necessary to calculate this
explicitly; it is sufficient to know that this will give v as a linear combination of
v1 , . . . , vm , u1 , . . . , un−m+1 except for uj , and then we may conclude that v ∈ lin(S ∪ B 0 ).
Thus we have shown that V ⊂ lin(S ∪ B 0 ). The opposite inclusion lin(S ∪ B 0 ) ⊂ V is
obvious. Thus lin(S ∪ B 0 ) = V .
Next we show that S ∪ B 0 is a linearly independent set. To do this, write o as a linear
combination of the vectors in S ∪ B 0 :
o = γ1 v1 + · · · + γm vm + δ1 u1 + · · · + δj−1 uj−1
(2.6)
+ δj+1 uj+1 + · · · + δn−m+1 un−m+1 .
Since {v1 , . . . , vm−1 , u1 , . . . , un−m+1 } = S1 ∪ B10 is a basis, all the coefficients in the latter
linear combination equal 0:
Thus all the coefficients in (2.6) are equal to 0, which shows that S ∪ B 0 is linearly
independent.
It follows that S ∪ B 0 is a basis, and the induction step is complete.
Corollary 2.2.4. Let V be a vector space and B be a basis of V such that B has n
elements. Then any linearly independent set S of n vectors is also a basis of V .
92
2.2 Vector spaces
Corollary 2.2.5. If a vector space V has a basis with n elements, then every basis of V
has the same number of elements.
Proof. Suppose that B is a basis of V with n elements and suppose that B 0 is another
basis of V .
Let B 0 have more than n elements. Then take any n distinct elements v1 , . . . , vn from B 0 .
From the previous corollary, it follows that these span V , and so if v ∈ B 0 \{v1 , . . . , vn }, it
can be written as a linear combination of {v1 , . . . , vn }, which contradicts the independence
of B 0 .
If B 0 has fewer than n elements, then by interchanging the roles of B and B 0 and
proceeding as above, we once again arrive at a contradiction.
The above result motivates the following natural definitions.
1. If there exists a basis B of V such that B has n elements, then n is called the
dimension of V , and it is denoted by dim(V ).
2. If a vector space has a basis with a finite number of elements, then it is called a finite
dimensional vector space.
Examples.
Just as group homomorphisms are functions that respect group operations, linear
transformations are functions that respect vector space operations.
93
2 Algebra
Examples.
1. Let m, n ∈ N. Let
a11 . . . a1n
. .. m×n
..
A= ∈R
. .
am1 . . . amn
Then the function TA : Rn → Rm defined by
n
X
a1k xk
x1 a11 x1 + · · · + a1n xn
k=1
. ..
..
.
TA = =
(2.7)
. .
.
n
xn am1 x1 + · · · + amn xn X
amk xk
k=1
x1
.. n n
for all ∈ R , is a linear transformation from the vector space R to the vector
.
xn
space Rm . Indeed, we have
x1 y1 x1 + y1
. .. ..
.. +
TA = TA
.
.
xn yn xn + yn
n
X
a1k (xk + yk )
k=1
..
=
.
Xn
amk (xk + yk )
k=1
n
X
n
X n
X
(a1k xk + a1k yk ) a1k xk + a1k yk
k=1 k=1 k=1
.. ..
=
.
=
.
n
X n
X n
X
(amk xk + amk yk )
amk xk + amk yk
k=1 k=1 k=1
Xn n
X
a1k xk a1k yk
k=1 k=1
.. ..
= +
. .
Xn Xn
amk xk
amk yk
k=1 k=1
x1 y1
. ..
.
= TA
. + TA
,
.
xn yn
94
2.2 Vector spaces
for all
x1 y1
. .
. , . ∈ Rn ,
. .
xn yn
and so L1 holds. Moreover,
n
X
a1k αxk
x1 αx1
k=1
. .
..
. = TA . =
TA α ·
. .
.
n
xn αxn X
amk αxk
k=1
n
X
n
X
α a1k xk a1k xk
x1
k=1
k=1
.. .. .
= = α· = α · TA ..
,
.
.
n n
X X xn
α amk xk
amk xk
k=1 k=1
Just as in the case of homomorphisms between groups, there are two important subsets
associated with a linear transformation between vector spaces, whose definitions are
given below.
95
2 Algebra
2. The image of T is defined to be the set im(T ) = {v ∈ V | ∃u ∈ U such that T (u) = v}.
Examples.
a11 x1 + · · · + a1n xn = 0
..
.
am1 x1 + · · · + amn xn = 0
is simultaneously satisfied.
The range of TA is the set of all vectors
y1
. m
y = ..
∈R
ym
such that
a11 x1 + · · · + a1n xn = y1
..
.
am1 x1 + · · · + amn xn = ym .
2. The function T : C[0, 1] → R given by T f = f 12 for all f ∈ C[0, 1], is a linear
transformation from the vector space C[0, 1] to the vector space R, with kernel equal
to the set of continuous functions on the interval [0, 1] that vanish at the point 12 . The
range of T is the whole vector space R.
96
2.2 Vector spaces
Analogous to Theorem 2.1.5 for homomorphisms between groups, we now prove the
following result.
1. ker(T ) is a subspace of U .
2. im(T ) is a subspace of V .
Our final result shows a connection between the nullity and the rank of a linear transfor-
mation.
Theorem 2.2.7. Let U be a finite dimensional vector space, V be a vector space, and
let T : U → V be a linear transformation. Then:
Proof. Let U be a vector space with dimension n and basis B, V be a vector space,
and T : U → V be a linear transformation. By the previous theorem, ker(T ) is a
subspace of U , thus it is finite dimensional (see Exercise 19), with dimension k ≤ n. Let
S = {u1 , u2 , . . . , uk } be a basis of ker(T ). By Theorem 2.2.3, there exists a subset B 0 of
B with n − k elements such that S ∪ B is a basis of U . Denote the elements of B 0 by
97
2 Algebra
α1 · v1 + · · · + αn−k · vn−k = β1 · u1 + · · · + βk · uk .
u = β1 · u1 + · · · + βk · uk + α1 · v1 + · · · + αn−k · vn−k .
v = T (u)
= T (β1 · u1 + · · · + βk · uk + α1 · v1 + · · · + αn−k · vn−k )
= β1 · T (u1 ) + · · · + βk · T (uk ) + α1 · T (v1 ) + · · · + αn−k · T (vn−k ).
2.2.5 Exercises
1. a) Is the set of invertible 2 × 2 matrices having real entries with matrix addition and
with scalar multiplication defined by (2.2) a vector space?
b) Is the set of invertible 2 × 2 matrices having real entries with matrix multiplication
and with scalar multiplication defined by (2.2) a vector space?
98
2.2 Vector spaces
2. Let V be a vector space. Prove, directly from the axioms, that if α ∈ R and v ∈ V
are such that α · v = o, then either α = 0 or v = o.
Hint: If α 6= 0, how can you manipulate the equation α · v = o?
U × V = {(u, v) | u ∈ U, v ∈ V } .
if (an )n∈N , (bn )n∈N ∈ R∞ , then (an )n∈N + (bn )n∈N = (an + bn )n∈N , (2.8)
Prove that R∞ is a vector space with the above addition and scalar multiplication.
6. Determine if the following statements are TRUE or FALSE. Give reasons for your
answers in each case.
a) The union of two subspaces of a vector space V is a subspace of V .
b) The intersection of two subspaces of a vector space V is a subspace of V .
1 1
0
c) 2 ∈ lin 3 , 1 in the vector space R3 .
3
4 1
1 1 0
d) The vectors 0 , 2 , 3 are linearly independent in the vector space R3 .
1 1 0
e) If v1 , v2 , v3 , v4 are linearly independent vectors, then v1 , v2 , v3 are linearly inde-
pendent.
f) If v1 , v2 , v3 , v4 are linearly dependent vectors, then v1 , v2 , v3 are linearly dependent.
g) If v1 , v2 are linearly independent, and v1 , v2 , v3 are linearly dependent, then v3 is
a linear combination of v1 and v2 .
h) If v1 , v2 are linearly dependent, then there is a real number α such that v1 = αv2 .
X + Y = {x + y | x ∈ X, y ∈ Y } .
99
2 Algebra
8. Consider the space C[0, π] of continuous functions on [0, π], and the four functions
f1 , f2 , f3 , f4 , where f (x) = 1 (i.e., f1 is the constant function 1), f2 (x) = cos x,
f3 (x) = cos 2x and f4 (x) = cos2 x.
a) Show that {f1 , f2 , f3 , f4 } is linearly dependent. You may use any properties of the
trigonometric functions that you know.
b) Show that {f1 , f2 , f3 } is linearly independent.
Hint: If α1 · f1 + α2 · f2 + α3 · f3 = o, then α1 f1 (x) + α2 f2 (x) + α3 f3 (x) = 0 for
all x ∈ [0, 1], and so in particular for x = 0, x = π/2 and x = π.
11. Let R∞ be the vector space of all sequences, with addition and scalar multiplication
defined by (2.8) and (2.9), respectively. We define the following subsets of R∞ :
a) `∞ is the set of all bounded sequences.
b) c is the set of all convergent sequences.
c) c0 is the set of all convergent sequences with limit 0.
d) c00 = {(an )n∈N ∈ R∞ | ∃N ∈ N such that ∀n > N, an = 0}, is the set of all se-
quences that are eventually zero.
Prove that c00 ⊂ c0 ⊆ c ⊆ `∞ ⊆ R∞ , and that each is a subspace of the next one.
12. Consider the vector space C[0, 1] with addition of functions and scalar multiplication
defined by (2.3). Let S(y1 , y2 ) = {f ∈ C[0, 1] | f (0) = y1 and f (1) = y2 }. Show that
S(y1 , y2 ) is a subspace of C[0, 1] iff y1 = 0 = y2 .
100
2.2 Vector spaces
13. Consider the vector space C[−1, 1] with addition of functions and scalar multiplication
defined by (2.3).
a) Let Se be the set of all even functions in C[−1, 1]:
14. Consider the vector space C[0, 1], with the usual addition and scalar multiplication.
Show that the set S = {1, x, x2 , x3 , . . . } of functions in C[0, 1] is linearly independent.
What is lin(S)?
Hint: In order to show that a function f in C[0, 1] is not equal to the zero function,
one approach is to try and show that there is some δ > 0 such that f (x) > 0 for all x
with 0 < x < δ.
16. Prove that if B is a basis of a finite-dimensional vector space V , then every element
v ∈ V can be written as a unique linear combination of the vectors from B.
18. Is the following statement true or false? Justify your answer, by means of a proof or a
counterexample.
Let V be a vector space of dimension 3. Let u, v, w be vectors in V such that neither
v nor w is in lin({u}). Then {u, v, w} is a basis of V .
19. a) Suppose that S is a linearly independent set of vectors in a vector space V such
that lin(S) 6= V . Show that there is a vector v in V such that S ∪ {v} is linearly
independent.
b) Suppose U is an infinite-dimensional vector space. Show that, for each n ∈ N, U
has an independent set with n elements.
c) Suppose that V is a finite-dimensional vector space and that U is a subspace of V .
Show that U does not contain an independent set with more than dim(V ) elements.
Deduce that U is finite-dimensional, and dim(U ) ≤ dim(V ).
Show also that, if dim(U ) = dim(V ), then U = V .
101
2 Algebra
21. a) For k ∈ N, let ek denote the sequence with the k th term equal to 1, and all other
terms equal to zero:
1 if n = k,
ek = (an )n∈N , where an =
0 6 k,
if n =
and let B = {ek | k ∈ N}. Prove that B is a basis for the vector space c00 ,
comprising all sequences that are eventually zero.
b) Is B = {ek | k ∈ N} also a basis for the vector space R∞ ?
Hint: Consider the constant sequence (1)n∈N .
22. Find the kernel and image of the linear transformations TA : R2 → R2 , where A ∈ R2×2
is given by:
" #
1 1
a) A = .
0 1
" #
1 1
b) A = .
0 0
" #
0 0
c) A = .
0 0
" #
x
Each vector in R2 represents a point in the plane. In each of the cases, draw a
y
picture of the subspaces ker(TA ) and im(TA ) in the plane.
23. Consider the vector space R2 with matrix addition and the usual scalar multiplication
defined by (2.2). Define the function T : R2 → R2 as follows:
" #
0
if x2 6= 0,
" # " #
x2
x1 x1
if ∈ R2 , then T = "
x2 x2 #
x1
if x2 = 0.
x2
Show that T satisfies L2, but not L1, and hence it is not a linear transformation.
24. Let P [0, 1] denote the space of polynomial functions on [0, 1], with pointwise addition
and scalar multiplication. For a function p ∈ P [0, 1], with p(x) = a0 + a1 x + a2 x2 +
· · · + an xn , define p∗ (x) = a1 + a2 x + · · · + an xn−1 .
Show that the function T : P [0, 1] → P [0, 1] given by T (p) = p∗ is a linear transfor-
mation.
102
2.2 Vector spaces
25. Let c denote the vector space of all convergent sequences. Consider the function
T : c → R given by
T ((an )n∈N ) = lim an .
n→∞
a) Prove that T is a linear transformation from the vector space c to the vector
space R.
b) What is the kernel of T ?
c) Show that im(T ) = R.
26. Recall the definitions of the subspaces of even functions Se and odd functions So of
C[−1, 1] in Exercise 13.
a) Define, for each function f ∈ C[−1, 1], the function fe : [−1, 1] → R by
f (x) + f (−x)
fe (x) = , x ∈ [−1, 1].
2
Show that fe ∈ Se for each f ∈ C[−1, 1].
b) Define the function T : C[−1, 1] → C[−1, 1] by T (f ) = fe . Show that T is a
linear transformation.
c) Find ker(T ) and im(T ).
27. In this exercise, the associativity of matrix multiplication A(BC) = (AB)C is proved
using linear transformations.
a) Show that for any linear transformation T : Rn → Rm there is a unique m × n
matrix A such that T (x) = Ax for all x ∈ Rn .
Hint: If {e1 , e2 , . . . , en } is the standard basis of Rn , then explain why the only
possibility for A is
.. .. ..
. . ··· .
A=
T (e1 ) T (e2 ) · · · T (en ) .
.. .. ..
. . ··· .
103
2 Algebra
Explain why there is a linear transformation that satisfies (∗), and show that
there is only one such linear transformation.
Find ker(T ) and im(T ).
104
Written by:
Professor Amol Sasane
Revised by:
Professor Jozef Skokan, Professor Konrad
Swanepoel and Professor Graham
Brightwell