You are on page 1of 79

1X

Algebra
2011-2012
Contents
1 Numbers 1
1.1 Number Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Geometric Interpretation of . . . . . . . . . . . . . . . . . . . . 10
1.2.2 Regions in the Complex Plane . . . . . . . . . . . . . . . . . . . . 13
1.2.3 Multiplication of Complex Numbers . . . . . . . . . . . . . . . . . 14
2 Methods of Proof 17
2.1 Direct Proof . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Examples and Counterexamples . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Proof By Contradiction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Converse and Contrapositive . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Proof By Induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Variations of the Principle of Induction . . . . . . . . . . . . . . . . . . . 25
3 Sequences and Series 28
3.1 Standard Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Arithmetic Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3 Geometric Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.4 Two Special Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4 The Binomial Theorem and Applications 37
4.1 Permutations and Combinations . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 The Binomial Theorem and Applications . . . . . . . . . . . . . . . . . . 40
4.3 Application to Trigonometric Calculations . . . . . . . . . . . . . . . . . 46
5 Polynomials and roots in the polar form 48
5.1 Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.2 Roots in the Polar Form . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6 Matrices 55
6.1 Scalars, Matrices and Arithmetic . . . . . . . . . . . . . . . . . . . . . . 55
6.2 Systems of Linear Equations . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.3 Inverse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.4 Transpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.5 Transformations of the Plane . . . . . . . . . . . . . . . . . . . . . . . . 74
6.6 Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
i
Chapter 1
Numbers
1.1 Number Systems
Number systems are sets of numbers closed under addition and multiplication.
For example,
= set of natural numbers (positive integers) = {1. 2. 3. 4. . . .}.
This set has the following properties:
1. is closed under addition, i.e. if :. : then : + : .
2. is closed under multiplication, i.e. if :. : then : : .
Notice that
1. is not closed under substraction, e.g. 1 2 = 1 .
2. is not closed under division, e.g.
1
2
.
To get more closure properties working we need to enlarge the number system. So, we
look at
= set of integers = {. . . . 2. 1. 0. 1. 2. 3. 4. . . .}.
is closed under addition, subtraction and multiplication but not division since, for ex-
ample,
1
2
.
Let :, : be integers. We write : : if : divides :, i.e. if : = /: for some
integer /. We write : : if : does not divide :. For example,
4 20 (because 20 = 5 4).
(4) 20.
3 3.
3 8.
Example 1.1. Show that if 3 : and 3 : then 3 (: + :).
1
Solution. Suppose 3 : and 3 :. Then : = 3/ and : = 3| for some integers /, |. It
follows that
: + : = 3/ + 3| = 3(/ + |).
where / + | is an integer. Therefore 3 (: + :) as required.
Note. It is crucial that in our arguments /, | and / + | are integers. To illustrate by
example: 3 8 because there is no integer / such that 8 = 3/. There is however a number
/ such that 8 = 3/, namely / =
8
3
, but / is not an integer.
Special Types of Integers
A number : is called even if 2 :, i.e. if : = 2/ for some / .
A number : is called odd if 2 :. This is the same as having the presentation
: = 2/ +1 for some / . Equivalently, every odd number is 1 more than an even
number.
A number : is called prime if : is greater than 1 and : has no positive integer
divisors other than 1 and :. The rst few primes are 2, 3, 5, 7, 11, 13, 17, 19,. . .
Question. Are there innitely many primes? (Euclid c300B.C.)
Greatest Common Divisor
We say c is a common divisor of c. / if c c and c /. The greatest common
divisor (g.c.d.) of c and / is denoted gcd(c. /), e.g. gcd(8. 12) = 4. We say two numbers,
c. / , are coprime if gcd(c. /) = 1.
Division Algorithm
For given c. / there exists unique . : with 0 and / : 0 such that
c = / + :
where : is called the remainder and is called the quotient.
Indeed, this can be seen as follows. Let be such that
c
/
< + 1 ( is
uniquely determined by these inequalities). Then / c < / + /. Dene : = c /.
Then we have 0 : < / as required.
2
Euclidean Algorithm (by example)
This algorithm allows us to nd the g.c.d. of two integers and also to express the g.c.d.
as a combination of these two integers. Thus it consists of two steps.
1. Find the g.c.d. of 96 and 36.
We have
96 = 36 2 + 24 division of 96 by 36 with the remainder (1.1)
36 = 24 1 + 12 division of 36 by the remainder from (1.1) (1.2)
24 = 12 2 + 0. continue and stop when the remainder is 0 (1.3)
The g.c.d. is the number we divide by in the last equation, i.e. 12.
2. Represent the greatest common divisor of 96 and 36 as gcd(96. 36) = 96: + 36:.
We start with the second to the last line (equation (1.2)) and express the remainder
in the circle as
12 = 36 24 1. (1.4)
Then we use the previous equation (1.1) to express the remainder in the box as
24 = 96 36 2.
and we plug these into equation (1.4) to get
12 = 36 (96 36 2) 1.
Upon collecting the terms with 36 and 96 we get
12 = 36 3 96
which is the required form with : = 3 and : = 1.
The Euclidean algorithm allows us to establish an important property of prime num-
bers (Proposition 1.1 below). Let us rst state the following Lemma.
Lemma 1.1. Let :. :. j be such that j: and j:. Then j(c: + /:) for any
c. / .
Proof. We have : = j/, : = j| for some /. | . Then
c: + /: = cj/ + /j| = j(c/ + /|).
Since c/ + /| we get j(c: + /:) as required.
Proposition 1.1. Let :. : , and let j be prime. Suppose that j::. Then at
least one of : and : is divisible by j.
3
Proof. Suppose that j :. Then gcd(j. :) = 1. By the Euclidean algorithm there exists
:. : such that 1 = :j +::. Hence : = :j:+:::. We have j:j: and j:::, hence
by Lemma 1.1 j:.
A corollary of this Proposition is the Fundamental Theorem of Arithmetic.
Theorem 1.2 (Fundamental Theorem of Arithmetic). Let : 1 be an integer. Then
: = j
I
1
1
j
I
2
2
. . . j
I

n
(1.5)
where j
i
. /
i
, j
i
are prime such that j
1
< j
2
< . . . < j
n
, and the presentation of (1.5)
is unique.
Exercise 1. Prove the Fundamental Theorem of Arithmetic.
Rational Numbers
In order to have the division we enlarge our number system from to
=set of rational numbers=
_
:
:

:. : and : = 0
_
.
( stands for quotients).
Recall that for /. d = 0 we have
c
/
=
c
d
if and only if cd = /c.
_
e.g.
15
12
=
5
4
because 15 4 = 12 5
_
.
Also,
:
1
= : for any : . Therefore and clearly .
The set of rational number has the following properties:
1. is closed under addition.
Indeed,
c
/
+
c
d
=
cd + /c
/d
.
and cd +/c is an integer if c. /. c. d are integers, also /d is a nonzero integer if /. d
are.
For example,
3
5
+
1
2
=
6 + 5
10
=
11
10
.
also
3
5
+
2
4
=
12 + 10
20
=
22
20
=
11
10
(as expected!).
4
2. is closed under subtraction.
Indeed,
c
/

c
d
=
cd /c
/d
and cd /c is an integer if c. /. c. d are, also /d is a nonzero integer if /. d are.
3. is closed under multiplication.
Indeed,
c
/

c
d
=
cc
/d
and cc is an integer if c. c are integers; /d is a nonzero integer if /. d are.
4. is closed under division by non-zero numbers.
Indeed,
c
/
c
d
=
c
/

d
c
=
cd
/c
and cd is an integer, /c is a non-zero integer where c. /. c. d are integers and /. d
and c are non-zero.
But does not have all the solution of equations that we need. For example, although
the equation r
2
= 4 has the solutions r = 2 in , the equation r
2
= 2 has no solutions
in . This is because

2 is irrational (i.e. it cannot be expressed as


:
:
with :. : .
We will prove this later). So we add irrational numbers and enlarge our number system
from to
= set of real numbers.
stands for reals; it is usually pictured as a line.
Figure 1.1
Illustration of as a line. Here we marked

2, , as some examples of well known irrational numbers.


Calculations With Real Numbers
We would normally like to present numbers in the simplest way. For the fractions
involving roots this normally means rationalizing the denominators, i.e. getting rid of
irrational numbers in the denominators.
Example 1.2. Simplify
2 +

3
4

3
.
5
Solution. First we get rid of the

3 in the denominator. We have


2 +

3
4

3
=
2 +

3
4

4 +

3
4 +

3
(Multiplying by
c
c
= 1 leaves the value unchanged but allows us to use the dierence of
two squares formula, ( 1)( + 1) =
2
1
2
, to simplify the denominator). So we
continue,
(2 +

3)(4 +

3)
(4

3)(4 +

3)
=
8 + 3 + 6

3
16 3
=
11 + 6

3
13
.
Example 1.3. Simplify

72

32
.
Solution.

72

32
=

8
=

4
=
3
2
.
Example 1.4. Show that

9 4

5 = 2.
Solution. Recall that

c/ =

/ if c. / are positive.
Consider

5 2. We have
(

5 2)
2
= 5 + 4 4

5 = 9 4

5. (1.6)
Notice that

5 2 0 (as 5 4) and 9 4

5 0 (as 9 4

5 = (

5 2)
2
). Therefore

5 2 =

9 4

5, on taking the positive square root of both sides of (1.6). Hence


result.
6
Recall the formula and derivation for the solution of quadratic equation. Let c. /. c
, c = 0. Then
cr
2
+ /r + c = 0 r
2
+
/
c
r +
c
c
= 0

_
r +
/
2c
_
2

/
2
4c
2
+
c
c
= 0 (completing the square)

_
r +
/
2c
_
2
=
/
2
4c
2

c
c

_
r +
/
2c
_
2
=
/
2
4cc
4c
2
r +
/
2c
=

/
2
4cc
2c
r =
/

/
2
4cc
2c
Example 1.5. Solve the equation 2r
2
+ 7r 4 = 0.
Solution. We have r =
7

81
4
=
7 9
4
, so the solutions are r = 4 and r =
1
2
.
It is easy to construct a quadratic equation with the required roots.
Example 1.6. Find a quadratic equation whose roots are r = 3 and 7.
Solution. An equation is (r 3)(r 7) = 0, i.e. r
2
10r + 21 = 0.
Example 1.7. Find a quadratic equation whose roots are r = 3

7.
Solution. An equation is
(r 3

7)(r 3 +

7) = 0.
i.e. (r 3)
2
7 = 0
(here we used the dierence of two squares formula (1)(+1) =
2
1
2
). Equiv-
alently, the equation is r
2
6r + 2 = 0.
Notice that not all quadratic equations have real solutions. Indeed if /
2
4cc < 0
then there are no real solutions. What do we do then?
Trick: Introduce i =

1. This helps solving quadratic equations. For example,


r
2
+ 3 = 0 can be solved by
r =

3 =

(1)3 =

3 = i

3.
7
More generally the solution to cr
2
+ /r + c = 0 where /
2
4cc < 0 is given by
r =
/

/
2
4cc
2c
=
/

(1)(4cc /
2
)
2c
=
/ i

4cc /
2
2c
and now 4cc /
2
0 so the square root

4cc /
2
is dened.
Example 1.8. Solve the equation r
2
10r + 40 = 0.
Solution.
r =
10

100 160
2
=
10

60
2
=
10

4 (1) 15
2
=
10

15
2
=
=
10 2

15
2
= 5

15 = 5 i

15 where i =

1.
Is this trick justiable? Yes: this is in our next section.
1.2 Complex Numbers
The set of complex numbers, is a number system that contains and a special number
i such that i
2
= 1. consists of all numbers of the form r + i where r. .
The number system has the following properties.
1. is closed under addition.
Indeed,
(r + i) + (n + i) = (r + n) + i( + )
and r+n and + are real if r. . n. are real. (For example, (3 +4i) +(7 2i) =
10 + 2i). Note that is closed under subtraction by similar reasons.
2. is closed under multiplication. Indeed,
(r + i)(n + i) = rn + i(r + n).
Here we used i
2
= 1, and rn . r + n are both real if r. . n. are real.
For example,
(2 + i)(5 + 3i) = 10 3 + (6 + 5)i = 7 + 11i.
(2 + 3i)(5 4i) = 10 + 12 + (8 + 15)i = 22 + 7i.
i(7 i) = 7i + 1 = 1 + 7i.
8
3. is closed under division by non-zero numbers.
Indeed,
r + i
n + i
=
(r + i)(n i)
(n + i)(n i)
=
rn + + i(n r)
n
2
+
2
=
rn +
n
2
+
2
+ i
n r
n
2
+
2
and
a&
&
2
+
2
.
&a
&
2
+
2
, provided r. . n. and n
2
+
2
= 0, i.e. provided n. are
both nonzero, i.e. n + i = 0, as required.
By simplifying a complex number we normally mean rearranging it to the form r+i
where r. and r. are presented in the simplied form.
Example 1.9. Simplify
2 + i
3 4i
.
Solution.
2 + i
3 4i
=
2 + i
3 4i

3 + 4i
3 + 4i
=
6 4 + 11i
9 + 16
=
2 + 11i
25
=
2
25
+
11
25
i.
If . = r + i where r. , then:
r is called the real part of ., denoted Re(.),
is called the imaginary part of ., denoted Im(.),
r i is called the complex conjugate of ., denoted by ..
Example 1.10. Find the real and imaginary parts of
3 + i
(7 3i)(1 + 2i)
.
Solution. First we simplify the denominator, then we arrive to the number in the form
r + i . We have,
3 + i
(7 3i)(1 + 2i)
=
3 + i
7 + 6 + 11i
=
3 + i
13 + 11i
=
3 + i
13 + 11i

13 11i
13 11i
=
39 + 11 20i
169 + 121
=
=
50 20i
290
=
5 2i
29
=
5
29
+
_

2
29
_
i.
Therefore the real part is
5
29
and the imaginary part is
2
29
.
Now we consider how a simple equation for a complex number can be solved.
Example 1.11. Solve for . the equation
.
3 .
= 1 + i.
Solution. We consider two dierent methods of the solution.
9
Method 1. The equation is
. = (1 + i)(3 .) . = 3 . + 3i i. 2. + i. = 3 + 3i
.(2 + i) = 3 + 3i . =
3 + 3i
3 + i
.
so that,
. =
3(1 + i)
2 + i

2 i
2 i
=
3(2 + i + 1)
4 + 1
=
3(3 + i)
5
=
9
5
+
3
5
i.
Method 2. The equation is . = (1 +i)(3 .). Let . = r +i, where r. . We have
r + i = (1 + i)(3 r i)
r + i = 3 r i + 3i ir +
r + i = (3 r + ) + i(3 r ).
Now, the LHS equates the RHS which means that the real part of LHS equals the
real part of the RHS and the imaginary part of the LHS equals the imaginary part
of the RHS. Thus, we obtain
r = 3 r + and = 3 r .
i.e.
_
2r = 3.
r + 2 = 3.
By multiplying the rst equation by 2 and adding the second equation we get
5r = 9, i.e. r =
9
5
. Substituting this in to the rst equation we get =
3
5
.
Therefore . =
9
5
+
3
5
i.
1.2.1 Geometric Interpretation of
Complex numbers can be represented as the points in a plane by identifying the number
r + i with the point with coordinates (r. ).
10
Figure 1.2
An example of various complex numbers represented as points on the plane.
The horizontal axis, OX, is called the real axis; it contains the real numbers. The
vertical axis, OY, is called the imaginary axis; it contains the numbers of the form i,
where ; these numbers are called the imaginary numbers. The non-zero imaginary
numbers are called purely imaginary. When a plane is used this way it is called the
complex plane or Argand diagram.
Note. A formal proof of the existence of might start by taking a plane and then
dening addition and multiplication of points in the plane in such a way that the plane
becomes .
Let . = r + i where r. . Then

r
2
+
2
is called the modulus of ., denoted
.. It is the distance of the point . from the origin in the complex plane.
For example,
3 + 2i =

3
2
+ 2
2
=

13.
i = 1.
1 + 3i =

10 (see Figure 1.3).


Note also that the conjugate, . = r i, is the mirror image of . in the real axis (see
Figure 1.4). In particular . = . if and only if . is on the real axis, i.e. . = . if and only
if . is real. We also have .. = .
2
. Indeed,
.. = (r + i)(r i) = r
2
+
2
= .
2
.
Any point in the plane has unique cartesian coordinates (r. ) (named after Descartes
1596 1650). It can also be represented by its polar coordinates, (:. o) (see Figure 1.5)
11
that are dened by:
: =

r
2
+
2
. (1.7)
tan o =

r
. (1.8)
r = : cos o. = : sin o. (1.9)
So the complex number . = r + i can also be written as . = :(cos o + i sin o) where
: = . is the modulus of ., and o is usually called the argument of . and is denoted
arg(.). This is called the polar form of the complex number .. Note that : 0 and o is
dened up to 2:, : , and it can be chosen such that 0 o < 2.
Figure 1.3
Complex numbers and their moduli.
Figure 1.4
Complex conjugation.
Figure 1.5
Polar coordinates.
Example 1.12. Find the modulus and argument of the complex numbers 1 + i, 1 i,
i, 1,

3 i.
Solution.
(a)
. = 1 + i has the modulus

2 and argu-
ment

4
.
12
(b)
. = i 1 has the modulus

2 and argu-
ment

4
(or
7
4
).
(c) . = i has the modulus 1 and argument

2
.
(d)
. = 1 has the modulus 1 and argument
.
(e)
. =

3i has modulus . =

3 + 1 =
2 and tan o =
1

3
=
1

3
. Note also that
o is in the third quadrant as r = cos o < 0
and = sin o < 0. Hence o = +

6
=
7
6
.
Thus . =

3 i has modulus 2 and


argument
7
6
.
1.2.2 Regions in the Complex Plane
Sometimes it is convenient to dene various regions in the plane using complex numbers.
Recall that . =

r
2
+
2
is distance of . from the origin. More generally, if .
1
= r
1
+i
1
and .
2
= r
2
+ i
2
(r
1
.
1
. r
2
.
2
) then
.
2
.
1
=

(r
2
r
1
)
2
+ (
2

1
)
2
= distance from .
1
to .
2
.
Example 1.13. Sketch the regions in the complex plane given by
13
(a) {. : . 3 < 1} . (b) {. : . 3 . i} .
(c)
_
. : 0 < arg . <

4
_
.
Solution.
(a)
(b)
(c)
1.2.3 Multiplication of Complex Numbers
Polar coordinates are particularly useful when multiplying complex numbers. Let . =
:(cos o + i sin o) and u = :(cos o + i sin o) be complex numbers in the polar form. Then
.u = ::(cos o cos o sin o sin o + i(sin o cos o + cos o sin o)) = ::(cos (o + o) + i sin (o + o)).
So, when complex numbers are multiplied,
moduli are multiplied, arguments are added.
By repeated application of this principle we also get De Moivres Theorem.
Theorem 1.3 (De Moivres Theorem). Let . = :(co:o +i sin o) be a complex number in
the polar form. Then
.
a
= :
a
(cos :o + i sin :o)
for every positive integer :.
14
Now we consider a few examples on multiplication and taking powers of complex
numbers.
Example 1.14. By calculating (1 + i)(1 + i

3) in two dierent ways show that


cos
7
12
=

6
4
and sin
7
12
=

2 +

6
4
.
Solution. We have
(1 + i)(1 + i

3) = 1

3 + i(1 +

3). (1.10)
But also,
1 + i =

2(cos

4
+ i sin

4
) (from Example 1.12)
and
1 + i

3 = 2(cos

3
+ i sin

3
) (similar to Example 1.12).
So
(1 + i)(1 + i

3) = 2

2
_
cos
_

4
+

3
_
+ i sin
_

4
+

3
__
= 2

2
_
cos
7
12
+ i sin
7
12
_
. (1.11)
From (1.10) and (1.11) we obtain
(1

3) + i(1 +

3) = 2

2
_
cos
7
12
+ i sin
7
12
_
.
Equating real and imaginary parts we obtain
1

3 = 2

2 cos
7
12
and 1 +

3 = 2

2 sin
7
12
.
Thus
cos
7
12
=
1

3
2

2
=

6
4
. and sin
7
12
=
1 +

3
2

2
=

2 +

6
4
as required.
Example 1.15. Find the real and imaginary parts of (

3 i)
8
.
Solution. From Example 1.12

3 i = 2
_
cos
7
6
+ i sin
7
6
_
.
15
Therefore, by De Moivres theorem
(

3 i)
8
= 2
8
_
cos
8 7
6
+ i sin
8 7
6
_
= 2
8
_
cos
4
3
+ i sin
4
3
_
as
8 7
6
=
28
3
= 8 +
4
3
. Hence
(

3 i)
8
= 2
8
_

1
2
i

3
2
_
= 2
7
i2
7

3 = 128 128

3i.
so that the real and imaginary parts are 126 and 126

3, respectively.
Example 1.16. For which positive integers : is (1 + i

3)
a
real?
Solution. We have 1 + i

3 = 2
_
cos
2
3
+ i sin
2
3
_
(Check this). Therefore by de
Moivres theorem
(1 + i

3)
a
= 2
a
_
cos
2:
3
+ i sin
2:
3
_
.
This is real if and only if sin
2:
3
= 0, i.e. if and only if
2:
3
is an integer, i.e. if and only
if : is divisible by 3.
Now we mention a few more properties in connection with polar coordinates and de
Moivres theorem. Let o . Note rstly that (cos o +i sin o)(cos o i sin o) = 1, hence,
1
cos o + i sin o
= cos o i sin o.
Equivalently
1
.
= . when . = 1.
Then we note the following relation
(cos o i sin o)
a
= cos :o i sin :o (1.12)
for any : . Indeed
(cos o i sin o)
a
= (cos(o) + i sin (o))
a
and by de Moivres theorem
(cos(o) + i sin(o))
a
= cos (:o) + i sin (:o) = cos :o i sin :o
as required. Finally, note that
(cos o + i sin o)
a
=
_
1
cos o + i sin o
_
a
= (cos o i sin o)
a
= cos :o i sin :o
for any : .
16
Chapter 2
Methods of Proof
2.1 Direct Proof
A direct proof is a style of proof laid out as in the following:
hypothesis logical argument conclusion.
Example 2.1. Prove that if :. : are even integers then so is 3: + 5:.
Solution. We have that :. : are even integers. Then : = 2/ and : = 2| for some
integers /. |. It follows that 3:+5: = 3 2/ +5 2| = 2(3/ +5|) and 3/ +5| is an integer.
Therefore 3: + 5: is even.
The the following propositions are proven by direct proofs.
Proposition 2.1. Let : . Prove that
(a) if : is even then :
2
is even,
(b) if : is odd then :
2
is odd.
Proof. (a) Let : be even. Then : = 2/ for some / . It follows that :
2
= 4/
2
=
2(2/
2
) and 2/
2
. Therefore :
2
is even.
(b) Let : be odd. Then : = 2/ + 1 for some / . It follows that
:
2
= (2/ + 1)
2
= 4/
2
+ 4/ + 1 = 2(2/
2
+ 2/) + 1.
and 2/
2
+ 2/ . Therefore :
2
is odd.
Proposition 2.2. Let :. :. . Prove that if :: and : then :.
Proof. We have :: and hence : = ::
1
for some :
1
. We also have : so = ::
2
for some :
2
. Therefore = ::
2
= : (:
1
:
2
) with :
1
:
2
. Thus : as required.
17
Proposition 2.3. Show that if .. u then .u = . u.
Solution. Let .. u . Then . = r + i and u = n + i for some r. . n. . It
follows that
.u = (r + i)(n + i) = (rn ) + i(r + n).
and .u = (rn ) i(r + n).
On the other hand
. u = (r i)(n i) = (rn ) i(r + n)
so .u = . u as required.
2.2 Examples and Counterexamples
We often use examples to illustrate results. For instance, 3 is odd and 3
2
= 9 is odd is
an example that illustrates the general result that if : is odd then :
2
is odd but does
not prove that result. A proof must deal with all :, not just : = 3 (see the proof of
Proposition 2.1(b) above).
However, if we want to prove that a general statement is false, one counterexample is
enough.
Example 2.2. Show that the statement if : is odd then :
2
is even is false.
Solution. : = 3 is a counterexample. Indeed 3 is odd and 3
2
= 9 is not even.
Example 2.3. Show that the following statement is false: the sum of two primes is
never prime.
Solution. 2 + 3 = 5 is a counterexample as 2. 3. 5 are all primes.
2.3 Proof By Contradiction
Proof by contradiction is a style of proof showing that a proposition is true. The format
is as follows:
Assume that a statement 1 is false logical argument a contradiction.
Therefore we can conclude 1 is true.
18
We will demonstrate proof by contradiction in a number of proofs below. Note rst
that any rational number can always be expressed in its lowest terms, i.e. in the form
:
:
where :. : are integers having no common factor grater than 1. This can always
be achieved by canceling all common factors of the numerator and denominator, e.g.
15
20
=
3
4
, in lowest terms.
Proposition 2.4. Let : be such that :
2
is odd. Then : is odd.
Proof. Suppose the statement is false, that is we have an integer : such that :
2
is odd
but : is even. Then : = 2/ for some / . Then :
2
= (2/)
2
= 2 (2/
2
), and we have
2/
2
so :
2
is even. This is a contradiction hence the statement is true.
Proposition 2.5. Let : be such that :
2
is even. Then : is even.
Proof. We have : , :
2
is even. Suppose that : is odd. Then : = 2/ + 1 for some
/ . Hence :
2
= (2/ + 1)
2
= 4/
2
+ 4/ + 1 = 2(2/
2
+ 2/) + 1. We have 2/
2
+ 2/
thus :
2
is odd. This is a contradiction, hence : is even.
Recall that a number : is called irrational if : , i.e. : cannot be represented
as : =
:
:
where :. : .
Example 2.4. Prove that

2 is irrational.
Solution. Suppose

2 is rational. Then we can express it in its lowest terms, i.e.

2 =
:
:
for some integers :. : having no common factors greater than 1. Then : =

2:
so that :
2
= 2:
2
. So :
2
is even. Therefore : is even (see Proposition 2.5). So : = 2/
for some / . Then (2/)
2
= 2:
2
, i.e. 4/
2
= 2:
2
, i.e. :
2
= 2/
2
. So :
2
is even.
Therefore : is even. But this means that : and : have 2 as a common factor, which is
a contradiction. Therefore

2 is irrational.
Example 2.5. Prove that if r is irrational then r + 2 is irrational.
Solution. Let r be irrational. Let us assume that r + 2 is rational. Then the equality
r = (r + 2) 2 expresses r as the dierence of two rational numbers which is rational
since is closed under subtraction. Hence r is rational, which is a contradiction. Thus
r + 2 must be irrational.
Now we consider two statements on the prime numbers.
Proposition 2.6. Every positive integer greater than 1 has a prime factor.
Proof. Let : be a positive integer greater than 1. Let : be the smallest integer greater
than 1 that divides : (there will always be such an integer: if all else fails, : divides :).
Suppose : is not prime. Then : = /|, where : /, | 2, /. | . We have :: and
/: hence /: (see Proposition 2.2), but this is not possible by the choice of :, so we
have a contradiction. Hence : must be prime.
19
Theorem 2.1 (Euclid). There are innitely many prime numbers.
Proof. Suppose there is only a nite number of primes, say j
1
. j
2
. . . . . j
n
. Let ` =
j
1
j
2
. . . j
n
+ 1. ` must have a prime factor (from Proposition 2.6), i.e. j
i
` for some
i. But j
i
j
1
j
2
. . . j
n
, i.e. j
i
(` 1). By Lemma 1.1 j
i
(` (` 1)) i.e. j
i
1, which is
impossible. Therefore there must be an innite number of primes.
2.4 Converse and Contrapositive
Consider the statement
If 1 then Q (also written 1 Q, spoken: 1 implies Q)
where 1 and Q are some statements.
Its converse is the statement
If Q then 1 (also written Q 1).
Examples below show that whether the converse is true or not is completely independent
of whether the original statement is true or not. We assume that :. : in the examples
below.
Example A
[Statement] If 4: then 4:
2
. (True: prove it)
Notice that in this case 1 is the statement 4: and Q is the statement 4:
2
.
[Converse] If 4:
2
then 4:. (False: : = 2 is a counterexample)
Example B
[Statement] If : is even then :
2
is even. (True: see Proposition 2.1)
[Converse] If :
2
is even then : is even. (Also true: see Proposition 2.5)
Example C
[Statement] If :
2
9 then : 3. (False: : = 4 is a counterexample)
[Converse] If : 3 then :
2
9. (True)
The contrapositive to the statement if 1 then Q is the statement
If (not Q) then (not 1). (also written (not Q)(not 1))
Contrapositives of the statements given in examples A, B and C are
Example A
[Contrapositive] If 4 :
2
then 4 : (True)
Example B
[Contrapositive] If :
2
is not even then : is not even.
Equivalent formulation is If :
2
is odd then : is odd. (True)
20
Example C
[Contrapositive] If : 3 then :
2
9.
Equivalent formulation is If : 3 then :
2
9 (False: : = 4 is a counterexam-
ple)
Exercise 2. Explain why the contrapositive is true if and only if the original statement
is true.
2.5 Proof By Induction
Firstly we introduce some notations on sequences and series.
Denition 2.1. A sequence is a list of numbers c
1
. c
2
. c
3
. . . . (sometimes denoted as
{c
a
}
a
or just {c
a
}). c

is called the :-th term of the sequence. A series is a sum of


terms of a sequence.
notation is used for summing up successive terms of a sequence, e.g.
25

=4
c

= c
4
+ c
5
+ . . . + c
25
.
where {c
a
} is some sequence. Another example is

6
=3
: = 3 + 4 + 5 + 6 = 18.
Now proof by induction is a proof using the following principle (or its variations).
Principle of Mathematical Induction
Let 1(:) denote some statement. That is the statement involves a parameter : .
Suppose 1(1) is true. Suppose also that 1(:) is true for : = / + 1 whenever it is true
for : = / (here / ). Then 1(:) is true for any : .
Why does this principle work? We have that 1(1) is true. Then it follows from our
assumptions that 1(2) is true. Hence also 1(3) is true. By repeating the process we get
1(:) is true for any : .
Thus in order to establish that some statement 1(:) is true for all : it is sucient
to check that 1(1) is true and to establish 1(/+1) assuming 1(/) is true. The statement
1(1) is called the base of induction. When proving 1(/+1), assuming 1(/) is true, we are
doing the step of induction. The statement 1(/) is then called the inductive hypothesis.
Proposition 2.7. For all positive integers :,
a

=1
:
2
=
1
6
:(: + 1)(2: + 1).
Proof. Let 1(:) denote the statement that
a

=1
:
2
=
1
6
:(: + 1)(2: + 1).
21
When : = 1,
LHS =
1

=1
:
2
= 1
2
= 1.
RHS =
1
6
1 2 3 = 1 = LHS.
So 1(1) is true.
Suppose now that 1(:) is true when : = /, i.e. suppose
I

=1
:
2
=
1
6
/(/ + 1)(2/ + 1).
Then
I+1

=1
:
2
=
I

=1
:
2
+ (/ + 1)
2
=
1
6
/(/ + 1)(2/ + 1) + (/ + 1)
2
by the inductive hypothesis.
=
1
6
(/ + 1) [/(2/ + 1) + 6(/ + 1)] =
1
6
(/ + 1)(2/
2
+ 7/ + 6) =
=
1
6
(/ + 1)(/ + 2)(2/ + 3) =
1
6
((/ + 1)(/ + 1) + 1) (2(/ + 1) + 1) .
so that 1(:) is true when : = / + 1.
Thus 1(1) is true and 1(/ + 1) is true whenever 1(/) is true. Therefore by the
principle of mathematical induction 1(:) is true for all positive integers :.
Example 2.6. A sequence of numbers c
1
. c
2
. c
3
. . . . is dened by
c
1
= 2 and c
a+1
= 2c
a
: for all : 1.
Find c
2
. c
3
. c
4
. Guess the general formula for c
a
and prove it by induction.
Solution. We have
c
1
= 2.
c
2
= 2c
1
1 = 3.
c
3
= 2c
2
2 = 4.
c
4
= 2c
3
3 = 5.
This suggests that c
a
= :+1 for all :. Let 1(:) denote the statement c
a
= :+1. When
: = 1 LHS= 2 (as we are given c
1
= 2) and RHS= 2 =LHS. So 1(:) is true when : = 1.
Suppose now that 1(:) is true when : = / , i.e. suppose
c
I
= / + 1.
Then
c
I+1
= 2c
I
/ by the denition of the sequence.
hence c
I+1
= 2(/ + 1) / by the inductive hypothesis.
Thus c
I+1
= / + 2 = (/ + 1) + 1.
22
We get 1(:) is true when : = / + 1.
So 1(1) is true and 1(/ + 1) is true whenever 1(/) is true. Therefore 1(:) is true
for all positive integers : by the principle of mathematical induction.
Example 2.7. A sequence of numbers c
1
. c
2
. c
3
. . . . is dened by
c
1
= 1 and c
a
=
1
2
c
a1
+ 1 for all : 2.
Find c
2
and c
3
. Prove that, for all positive integers n,
c
a
= 2
1
2
a1
. (2.1)
Solution. We have c
2
=
1
2
c
1
+1 =
3
2
, c
3
=
1
2
c
2
+1 =
3
4
+1 =
7
4
(which is in agreement
with (2.1)).
Let 1(:) denote the statement (2.1). When : = 1, LHS= c
1
= 2 and RHS= 2
1
2
0
=
2 1 1 =LHS. So 1(:) is true when : = 1.
Suppose now that 1(:) is true when : = /, i.e. suppose
c
I
= 2
1
2
I1
.
Then
c
I+1
=
1
2
c
I
+ 1 by the denition of the sequence,
hence c
I+1
=
1
2
_
2
1
2
I1
_
+ 1 by the inductive hypothesis.
Thus c
I+1
= 1
1
2
I
+ 1 = 2
1
2
I
= 2
1
2
(I+1)1
so that 1(:) is true when : = / + 1.
So 1(1) is true and 1(/ + 1) is true whenever 1(/) is true. Therefore 1(:) is true
for all positive integers :.
Example 2.8. Recall that :! = : (: 1) (: 2) . . . 3 2 1 so that 2! = 2 1,
3! = 3 2 1, 4! = 4 3 2 1, etc. Calculate
a

=1
: (:!)
when : = 1. 2. 3, guess the general formula and prove it by induction.
23
Solution.
1

=1
: (:!) = 1 (1!) = 1 1 = 1.
2

=1
: (:!) = 1 (1!) + 2 (2!) = 1 1 + 2 2 = 5.
3

=1
: (:!) = 1 (1!) + 2 (2!) + 3 (3!) = 1 1 + 2 2 + 3 6 = 23.
This suggests that
a

=1
: (:!) = (: + 1)! 1. (2.2)
Let 1(:) denote the statement (2.2). When : = 1,
LHS =
1

=1
: (:!) = 1 (1!) = 1 1 = 1 and
RHS = 2! 1 = 2 1 = 1 = LHS.
So 1(:) is true when : = 1.
Suppose now that 1(:) is true when : = / , i.e. suppose
I

=1
: (:!) = (/ + 1)! 1.
Then
I+1

=1
: (:!) =
I

=1
: (:!) + (/ + 1) (/ + 1)! by the denition of the sum,
hence
I+1

=1
: (:!) = (/ + 1)! 1 + (/ + 1) (/ + 1)! by the inductive hypothesis.
Thus
I+1

=1
: (:!) = (/ + 2) (/ + 1)!
. .
(I+2)(I+1)I(I1)...321
1 = (/ + 2)! 1 = ((/ + 1) + 1)! 1
so that 1(:) is true in the case : = / + 1.
So, 1(1) is true and 1(/ + 1) is true whenever 1(/) is true. Therefore 1(:) is true
for all positive integers :.
Example 2.9. Prove that, for all positive integers :,
7
2a
+ 3 is divisible by 4.
24
Solution. Let 1(:) denote the statement that 7
2a
+3 is divisible by 4. When : = 1,
7
2a
+ 3 = 7
2
+ 3 = 52, which is divisible by 4. So 1(:) is true when : = 1.
Suppose now that 1(:) is true when : = /, i.e. suppose
7
2I
+ 3 is divisible by 4,
so that 7
2I
+ 3 = 4: for some : . Then
7
2(I+1)
+ 3 = 7
2I+2
+ 3 = 7
2
7
2I
+ 3 = 49(4:3) + 3
= 4 49:147 + 3 = 4 49:144.
Thus 7
2(I+1)
+ 3 = 4(49:36).
which is divisible by 4 since 49:36 , so that 1(:) is true when : = / + 1.
So 1(1) is true and 1(/ + 1) is true whenever 1(/) is true. Therefore 1(:) is true
for all positive integers :.
2.6 Variations of the Principle of Induction
We consider two other versions of the principle of mathematical induction.
A. Dierent Starting Point
Here we suppose that some statement 1(:), : , is known to be true for : = /
0
where
/
0
. We also suppose that 1(/ + 1) is true whenever 1(/) is true where / ,
/ /
0
. Then the modied principle of mathematical induction states that 1(:) is true
for all : such that : /
0
.
This is a generalization of the principle of mathematical induction from Section 2.5
where we had the same principle with /
0
= 1.
Example 2.10. Prove that 2
a
:
2
for all integers : 5.
Note. The result is true when : = 1 but not true when : = 2. 3. 4.
Solution. Let 1(:) denote the statement 2
a
:
2
. When : = 5, 2
a
= 2
5
= 32,
:
2
= 5
2
= 25 and 32 25 so that 1(:) is true when : = 5.
Suppose now that 1(/) is true (and / 5), i.e. suppose
2
I
/
2
.
Then
2
I+1
= 2 2
I
2 /
2
.
25
So to prove that 2
I+1
(/ + 1)
2
it will be sucient to prove that 2/
2
(/ + 1)
2
.
2/
2
(/ + 1)
2
2/
2
(/ + 1)
2
0 /
2
2/ 1 0
(/ 1)
2
1 1 0 (/ 1)
2
2 / 1

2
/ 1 +

2.
Since / 5 this is true and hence 1(/ + 1) is true.
So 1(5) is true and 1(/ +1) is true whenever 1(/) is true for / 5. Therefore 1(:)
is true for all integers : 5.
B. Inductive Step Involving More Than One Smaller Value
This version of the principle can be formulated as follows. Suppose that a statement
1(:) is true for : = 1. Suppose also that 1(/ +1) is true whenever 1(:) is true for all :
such that 1 : /. Then the modied principle of mathematical induction states that
1(:) is true for any : .
As you see the dierence with the principle of mathematical induction from Section
2.5 is that we allow ourselves to assume at the inductive step that 1(:) is true not just
for : = / but for all (or some) smaller values of : as well.
Example 2.11. A sequence of numbers c
1
. c
2
. c
3
. . . . is dened by
c
1
= 1, c
2
= 1 and c
a
= c
a1
+ 6c
a2
for all : 3.
Prove that, for all positive integers :,
c
a
=
1
5
(3
a
(2)
a
). (2.3)
Solution. Let 1(:) denote the statement (2.3).
When : = 1,
LHS = c
1
= 1 and RHS =
1
5
(3 (2)) =
1
5
5 = 1 = LHS.
When : = 2,
LHS = c
2
= 1 and RHS =
1
5
(3
2
(2)
2
) =
1
5
(9 4) = 1 = LHS.
So 1(:) is true when : = 1 and when : = 2. Consider now 1(/), where / 3 and
suppose that 1(:) is true for all values of : less than /. We have for the LHS of 1(/)
by denition
c
I
= c
I1
+ 6c
I2
=
1
5
_
3
I1
(2)
I1
_
+ 6
1
5
_
3
I2
(2)
I2
_
26
by the inductive hypothesis. Thus
c
I
=
1
5
__
3
I1
+ 6 3
I2
_

_
(2)
I1
+ 6 (2)
I2
_
=
=
1
5
__
3
I1
+ 2 3
I1
_

_
(2)
I1
3 (2)
I1
_
=
=
1
5
_
3 3
I1

_
(2) (2)
I1
_
=
1
5
(3
I
(2)
I
) = RHS of 1(/)
so that 1(/) is true.
So 1(1) and 1(2) are true and, for / 3, 1(/) is true whenever 1(:) is true for all
positive integer values of : < /. Therefore 1(:) is true for all positive integers :.
27
Chapter 3
Sequences and Series
3.1 Standard Series
Let us repeat some denitions from the beginning of Section 2.5. A sequence is a list of
number c
1
. c
2
. c
3
. . . . (sometimes denoted {c
a
}
a
). The n-th term of the sequence is c
a
.
A series is a sum of terms from a sequence, e.g.

25
=2
c

= c
2
+c
3
+c
4
+. . . +c
24
+c
25
.
The series,
a

=1
: = 1 + 2 + 3 + . . . + :.
a

=1
:
2
= 1
2
+ 2
2
+ 3
2
+ . . . + :
2
.
a

=1
:
3
= 1
3
+ 2
3
+ 3
3
+ . . . + :
3
.
are sometimes called the standard series. The following formulae for them can be proven
by induction
a

=1
: =
1
2
:(: + 1).
a

=1
:
2
=
1
6
:(: + 1)(2: + 1).
a

=1
:
3
=
1
4
:
2
(: + 1)
2
(we did

a
=1
:
2
in Proposition 2.7).
28
Remark 3.1. Notice that the formula for

:
3
is the square of the formula for

:.
Does anything similar happen for

a
=1
:
5
?
Using the sums of the standard series more series can be summed up.
Example 3.1. Find the sum of the rst : terms of the sequence 2 3. 3 4. 4 5. . . ..
Solution. The :-th term is (: + 1)(: + 2). So the required sum is
a

=1
(: + 1)(: + 2) =
a

=1
(:
2
+ 3: + 2) =
=
a

=1
:
2
+
a

=1
3: +
a

=1
2 =
a

=1
:
2
+ 3
a

=1
: +
a

=1
2 =
=
1
6
:(: + 1)(2: + 1) +
3
2
:(: + 1) + 2: =
1
6
:[(: + 1)(2: + 1) + 9(: + 1) + 12] =
=
1
6
:(2:
2
+ 12: + 22) =
1
3
:(:
2
+ 6: + 11)
Example 3.2. The triangular numbers are the numbers of dots in the following
diagrams:
etc.
t
1
= 1 t
2
= 3 t
3
= 6 t
4
= 10
Find a formula for the :-th triangular number t
a
and also for the sum of the rst :
triangular numbers.
Solution.
t
a
= 1 + 2 + 3 + . . . + : =
a

=1
: =
1
2
:(: + 1).
Sum of the rst : triangular numbers is
t
1
+ t
2
+ . . . t
a
=
a

=1
t

=
a

=1
1
2
:(: + 1) =
1
2
a

=1
(:
2
+ :) =
=
1
2
_
1
6
:(: + 1)(2: + 1) +
1
2
:(: + 1)
_
=
1
12
:(: + 1) [(2: + 1) + 3] =
=
1
12
:(: + 1)(2: + 4) =
1
6
:(: + 1)(: + 2).
29
3.2 Arithmetic Series
An arithmetic progression is a sequence of the form
c. c + d. c + 2d. c + 3d. . . . (3.1)
where c. d in general. The rst element c is called the initial term.The dierence
between any two successive elements is d and it is called the common dierence. The
:-th term of an arithmetic progression (3.1) is then given by c
a
= c + (: 1)d.
An arithmetic series is a sum of terms from an arithmetic progression. Let o
a
be the
sum of the rst : terms of the arithmetic series (3.1). Then
o
a
= c + (c + d) + (c + 2d) + . . . + (c + (: 1)d) =
a1

)=0
(c + ,d).
Proposition 3.1. The sum of the rst : terms of the arithmetic progression (3.1) is
o
a
=
:
2
(2c + (: 1)d).
Equivalently, o
a
equals : times the average of the rst and the :-th terms of the pro-
gression.
Proof. Let us consider
2o
a
=
a1

)=0
(c + ,d) +
a1

)=0
(c + ,d).
The second sum runs over , = 0. . . . . : 1. Let us replace index , with i = : 1 ,,
then the range for i is i = 0. 1. . . . . : 1. Thus
2o
a
=
a1

)=0
(c + ,d) +
a1

i=0
(c + (: 1 i)d).
Let us now rename i to , this is just a summation index so it does not matter how we
call it. We get,
2o
a
=
a1

)=0
(c + ,d) +
a1

)=0
(c + (: 1 ,)d) =
a1

)=0
(c + ,d + c + (: 1 ,)d) =
=
a1

)=0
(2c + (: 1)d) = (2c + (: 1)d):
since the terms in the last sum do not depend on , and there are : terms in the sum.
Hence o
a
=
1
2
(2c + (: 1)d) as required.
30
Notice that what happened in the proof was that we just represented o
a
as
o
a
= c +(c + d) +. . . +c + (: 1)d. (3.2)
and o
a
= c + (: 1)d +c + (: 2)d +. . . +c. (3.3)
Then the pairwise summation (that is the rst element in (3.2) with the rst element
in (3.3), then the second element in (3.2) with the second element in (3.3), etc.) gives
2o
a
= (2c + (: 1)d) + (2c + (: 1)d) + . . . + (2c + (: 1)d) = :(2c + (: 1)d).
as required. It is however important to be able to use freely summation notations as we
do in the proof above.
Exercise 3. Prove Proposition 3.1 by proving rstly that

a
=1
: =
1
2
:(: + 1).
Example 3.3. Find the sum 2 + 5 + 8 + 11 + . . . + 59.
Solution. This is an arithmetic series with (in the usual notation) c = 2 and d = 3.
To nd the number of terms we note that 59 = c + (: 1)d = 2 + 3(: 1) so that
: 1 =
57
3
= 19 and hence : = 20. Thus the required sum is
o
a
= o
20
=
:
2
[2c + (: 1)d] = 10[4 + 19 3] = 610.
Example 3.4. The sum of the rst : terms of an arithmetic progression with initial
term 0 and common dierence 32 is 1760. Find :.
Solution. In the usual notation
1760 =
:
2
[2c + (: 1)d] =
:
2
[32(: 1)] = 16:(: 1)
hence :(: 1) = 110. Therefore : = 11.
Example 3.5. (Higher Maths 1916) Find the sum of the rst 6 terms of an arithmetic
progression of which n is the 3rd term and is the 5th term.
Solution. In the usual notation for an arithmetic progression the :-th term is
c
a
= c + (: 1)d, so the 3rd term equals c + 2d and the 5th term equals c + 4d. Thus
c + 2d = n (3.4)
c + 4d = . (3.5)
(3.5)(3.4) gives
2d = n. i.e. d =
n
2
31
and 2(3.4)(3.5) gives
c = 2n .
(3.6)
Hence the sum of the rst 6 terms equals
o
6
= 3(2c + 5d) = 3
_
2(2n ) + 5
_
n
2
__
= 3
_
8n 4 + 5 5n
2
_
=
3
2
(3n + ).
3.3 Geometric Series
A geometric progression is a sequence of the form
c. c:. c:
2
. c:
3
. . . . . (3.7)
where c. : in general. Then c is called the initial term and : is called the common
ratio it is the ratio of any two successive terms. The :-th term is then given by
c
a
= c:
a1
.
We assume : = 1 as the case : = 1 is trivial. A geometric series is a sum of terms
from a geometric progression. Let o
a
be the sum of the rst : terms of the geometric
progression (3.7). Then
o
a
= c + c: + c:
2
+ c:
3
+ . . . + c:
a1
.
Proposition 3.2. The sum of the rst : terms of the geometric progression (3.7) equals
o
a
=
c(1 :
a
)
1 :
.
Proof. We have
o
a
=
a1

i=0
c:
i
.
Also consider
:o
a
=
a1

i=0
:c:
i
=
a1

i=0
c:
i+1
=
a

)=1
c:
)
as we introduced the summation index , = i + 1; , changes in the range 1. . . . . :. Hence
o
a
:o
a
=
a1

i=0
c:
i

i=1
c:
i
=
_
c:
0
+
a1

i=1
c:
i
_

_
a1

i=1
c:
i
+ c:
a
_
=
= c:
0
c:
a
= c(1 :
a
).
Thus o
a
=
c(1 :
a
)
1 :
as required.
32
Example 3.6. Find the sum of the rst : terms of the series 1 +
1
2
+
1
2
2
+
1
2
3
+ . . ..
Solution. This is a geometric series. In the usual notation c = 1, : =
1
2
and the
required sum is
o
a
=
c(1 :
a
)
1 :
=
1
_
1
1
2

_
1
1
2
= 2
_
1
1
2
a
_
.
Idea of Convergence
Notice that as : increases to innity the sum o
a
approaches 2 since the term
1
2
a
ap-
proaches 0. We write o
a
2 as : . We say that the series has sum to innity 2
and we also write o

= 2.
In general, let c
1
+ c
2
+ c
3
+ . . . be any series and let o
a
be the sum of the rst :
terms, i.e. o
a
=

a
=1
c

. If o
a
tends to a xed number C as : tends to (o
a
as
: ) we say that a series converges and has sum to innity C, and we write o

= C
and

=1
c

= C. On the other hand, if o


a
does not tend to any xed number as :
we say that the series diverges, does not have a sum to innity and that

=1
c

does
not exist.
Proposition 3.3. For : < 1, the geometric series
c + c: + c:
2
+ c:
3
+ . . .
converges and has sum to innity
o
1
.
Sketch of Proof. If : < 1 it can be shown that :
a
0 as : . So
o
a
=
c(1 :
a
)
1 :

c
1 :
as : .
A representation of rational numbers
Rational numbers are often given in the form of periodic decimal expressions. For
instance, the number 3.c/cdccdccdc . . ., where c. /. c. d. c are some integers such that
0 c. /. c. d. c 9, can also be denoted in the form 3.c/ cd c where the dots above c
and c mean that the part between c and c, that is the part cdc, is a periodic (innitely
repeating) part.
Sums of geometric series can be used to transfer the rational numbers from periodic
decimal forms into the form of a ratio of two integers.
Example 3.7. Express the repeating decimal 0.1

3(= 0.12323232323 . . .) as a rational


number in its lowest terms.
33
Solution. We have
0.1

3 = 0.123232323 . . . =
1
10
+
23
10
3
+
23
10
5
+
23
10
7
+ . . . =
1
10
+
23
10
3
_
1 +
1
10
2
+
1
10
4
+ . . .
_
=
=
1
10
+
23
1000

i=0
(10
2
)
i
=
1
10
+
23
1000
_
1
1
1
100
_
=
1
10
+
23
990
=
122
990
=
61
495
.
A Special Case of the geometric series
Consider the series
1 + r + r
2
+ r
3
+ . . . (3.8)
i.e. a geometric series with c = 1 and : = r. Then
o
a
=
a1

)=0
r
)
= 1 + r + . . . + r
a1
=
r
a
1
r 1
by above. This can be rewritten as:
r
a
1 = (r 1)(r
a1
+ . . . + r + 1).
So, for example,
r
3
1 = (r 1)(r
2
+ r + 1).
r
4
1 = (r 1)(r
3
+ r
2
+ r + 1).
r
5
1 = (r 1)(r
4
+ r
3
+ r
2
+ r + 1).
Also, provided r < 1, the geometric series (3.8) has sum to innity
1
1 r
, i.e.
1
1 r
= 1 + r + r
2
+ r
3
+ . . . .
Replacing r by r we obtain
1
1 + r
= 1 r + r
2
r
3
+ . . . (provided r < 1).
If we were allowed to integrate innite series term by term we would have got a series for
logarithm:
ln(1 + r) = r
1
2
r
2
+
1
3
r
3

1
4
r
4
+ . . . (provided r < 1).
This can actually be justied see later courses.
34
3.4 Two Special Series
1. Consider the series

=1
1
:
2
=
1
1
2
+
1
2
2
+
1
3
2
+
1
4
2
+ . . .
Let o
a
be the sum of the rst : terms. Then
o
a
=
1
1
2
+
1
2
2
+
1
3
2
+
1
4
2
+ . . . +
1
:
2
and
o
a
=
a

=1
1
:
2
< 1 +
a

=2
1
:(: 1)
as
1
:
2
<
1
:(: 1)
for : 2.
Since
1
:(: 1)
=
1
: 1

1
:
, we have
o
a
< 1 +
a

=2
_
1
: 1

1
:
_
= 1 +
a

=2
1
: 1

a

=2
1
:
= 1 + 1 +
a1

=2
1
:

_
a1

=2
1
:
+
1
:
_
= 2
1
:
.
So the sum never gets as big as 2. It can be shown (wait for a later course) that
the series converges and has sum to innity

2
6
, i.e.,

=1
1
:
2
=

2
6
.
2. Consider the series

=1
1
:
= 1 +
1
2
+
1
3
+
1
4
+ . . .
This series is called the harmonic series. Let o
a
be the sum of the rst : terms.
Then
o
a
=
a

=1
1
:
= 1 +
1
2
+
1
3
+
1
4
+ . . . +
1
:
.
Consider
o
2
= 1 +
a1

i=0
2
+1

=2

+1
1
:
.
The double sum corresponds to the grouping of terms as
o
2
= 1 +
1
2
+
_
1
3
+
1
4
_
. .
2 terms
+
_
1
5
+ . . . +
1
8
_
. .
4 terms
+. . . +
_
1
2
a1
+ 1
+ . . . +
1
2
a
_
. .
2
1
terms
.
35
Then we estimate each bracket by replacing the terms with the smallest term in
the same bracket:
o
2

a1

i=0
2
+1

=2

+1
1
:

a1

i=0
2
+1

=2

+1
1
2
i+1
=
a1

i=0
(2
i+1
2
i
)
1
2
i+1
=
a1

i=0
1
2
=
:
2
as the sum from 2
i
+ 1 to 2
i+1
contains 2
i+1
2
i
= 2
i
terms. Thus o
2

:
2
. So as
: gets arbitrarily large so too does the sum, i.e. o
a
as : . Thus, the
series diverges, i.e.

=1
1

does not exist.


36
Chapter 4
The Binomial Theorem and
Applications
4.1 Permutations and Combinations
Suppose we have : dierent objects (here : ). If we choose : of them (0 : :),
the : objects are called a combination of : objects from :. If instead we choose : of the :
objects and place them in a denite order (so that we have a rst object, a second object,
etc.) the : objects are called a permutation (or an ordered combination) of : objects from
:.
The number of permuations of : objects from : is denoted
a
1

. The number of
combinations of : objects from : is denoted
a
C

or more usually
_
:
:
_
(spoken : choose
:).
Proposition 4.1. The number of permutations is
a
1

= :(: 1)(: 2) . . . (: : +1).


Proof. To form a permutation of : objects from :, we need to choose a rst object. This
can be any of the : objects. So we have : choices. For each of these choices we can then
choose our second objects from the : 1 remaining objects, i.e. we have : 1 choices.
At the third step we have :2 choices, etc. Thus the total number of choices for : steps
is
a
1

= :(: 1)(: 2) . . . (: : + 1).


For example,
7
1
3
= 7 6 5 = 210.
Recall now that for : one denes :! = 1 2 . . . : and by denition 0! = 1. The
number :! is pronounced as : factorial.
Corollary 4.1. The number of permutations of : objects from : is

1

= :!.
Corollary 4.2. One can rearrange the number of permutations as
a
1

=
:!
(: :)!
.
37
Proposition 4.2. The number of combinations of : objects from : (: , : ,
0 : :) is given by
_
:
:
_
=
:!
:!(: :)!
=
:(: 1)(: 2) (: : + 1)
:(: 1)(: 2) 1
.
(That is we take the product of : successive numbers down from : and divide it by :!).
Proof of Proposition 4.2. To form a permutation of : objects from : we have to choose
the objects and put them into order. We separate these two stages. First choose the :
objects, i.e. form a combination of : objects from :, then put them into order, i.e. form
a permutation of : objects from :. This tells us that
a
1

=
_
:
:
_

.
Thus
_
:
:
_
=
a
1

=
:(: 1) . . . (: : + 1)
:(: 1) . . . 1
by Proposition 4.1 and Corollary 4.2, which gives us one of the required forms. This can
also be rearranged as
_
:
:
_
=
:(: 1) . . . (: : + 1) (: :)!
:(: 1) . . . 1 (: :)!
=
:!
:!(: :)!
as needed.
For example,
The number of combinations of 3 out of 7 is
_
7
3
_
=
7 6 5
3 2 1
= 35.
The number of combinations of 6 balls from 49 is
_
49
6
_
=
49 48 47 46 45 44
6 5 4 3 2 1
= 13. 983. 816.
So, by the way, the chance of winning the lottery is about 1 in 14 million.
Now we discuss some properties of the number of combinations
_
:
:
_
.
Proposition 4.3. For any :. : , 0 : :,
_
:
:
_
=
_
:
: :
_
.
Proof.
_
:
:
_
=
:!
:!(: :)!
.
_
:
: :
_
=
:!
(: :)!(: (: :))!
=
:!
(: :)!:!
.
Hence result.
38
An alternative proof which does not require any calculations is as follows. When we
remove : objects from a collection of : we leave : : behind (and viceversa). So the
number of ways of choosing : is the same as the number of ways of choosing ::. Thus
Proposition 4.3 holds.
Proposition 4.3 can simplify calculations. For example,
_
21
19
_
=
21 20 19 . . . 3
19 18 17 . . .
can also be represented as
_
21
19
_
=
_
21
2
_
=
21 20
2 1
= 210
which is easier to compute. Note also that
_
:
0
_
=
_
:
:
_
= 1 (4.1)
for any : . That is there is only one way of choosing : objects from : (or, equivalently,
of choosing none objects).
Proposition 4.4. Let : . Then
_
: + 1
:
_
=
_
:
:
_
+
_
:
: 1
_
for any : such
that 1 : :.
Proof.
RHS =
:!
:!(: :)!
+
:!
(: 1)!(: : + 1)!
=
(: : + 1) :!
:! (: : + 1) (: :)!
+
: :!
: (: 1)!(: : + 1)!
=
(: : + 1) :!
:!(: : + 1)!
+
: :!
:!(: : + 1)!
=
(: : + 1 + :) :!
:!(: : + 1)!
=
(: + 1) :!
:!(: : + 1)!
=
(: + 1)!
:!(: + 1 :)!
= LHS.
Exercise 4. Give a combinatorial argument that proves Proposition 4.4 without
calculations.
Let us also formally dene
_
:
:
_
= 0 if : < 0 or : :+1. Then it is easy to check that
Proposition (4.4) is true for any : , : . Propositions 4.4 and formula (4.1) imply
that the number of combinations can be read o the Pascal triangle. In the following
diagramme we present Pascals triangle up to and including row 4.
39
1 Row 0
1
!!
C
C
C
C
C
C
C
C
C
add
1
}}{
{
{
{
{
{
{
{
{
Row 1
1

@
@
@
@
@
@
@
@
add
2
}}{
{
{
{
{
{
{
{
{
!!
C
C
C
C
C
C
C
C
C
add
1
~~~
~
~
~
~
~
~
~
Row 2
1 3 3 1 Row 3
1 4 6 4 1 Row 4
In general, the :-th row consists of the numbers
_
:
0
_
.
_
:
1
_
.
_
:
2
_
. . . . .
_
:
:
_
.
4.2 The Binomial Theorem and Applications
Theorem 4.1 (The Binomial Theorem). For any c. / , : one has the expansion
(c + /)
a
=
a

=0
_
:
:
_
c
a
/

. (4.2)
Equivalently
(c + /)
a
=
_
:
0
_
c
a
+
_
:
1
_
c
a1
/ +
_
:
2
_
c
a2
/
2
+ . . . +
_
:
: 1
_
c/
a1
+
_
:
:
_
/
a
. (4.3)
Combinatorial Proof. We need to expand the brackets in (c+/)
a
. The result is the sum of
monomials

a
=0
`

c
a
/

for some coecients `


0
. . . . . `
a
. Indeed the monomial c
a
/

appears in the expansion when we choose / in : brackets and c in the remaining


(: :) brackets when expanding the power; and no other monomials can arise. This
monomial c
a
/

appears as many times as we can choose : objects (in this case the
expressions c + /) out of :. Thus `

=
_
a

_
as required.
Algebraic Proof. We prove the theorem by induction. Let 1(:) be the statement given
by Equation (4.2). For : = 1 we have
c + / =
1

=0
_
1
:
_
c
1
/

= c + /
so 1(1) is true. Suppose now that 1(/) is true. That is
(c + /)
I
=
_
/
0
_
c
I
+
_
/
1
_
c
I1
/ + . . . +
_
/
/
_
/
I
. for some / .
40
Then
(c + /)
I+1
= (c + /)(c + /)
I
= (c + /)
__
/
0
_
c
I
+
_
/
1
_
c
I1
/ + . . . +
_
/
/
_
/
I
_
=
=
_
/
0
_
c
I+1
+
__
/
1
_
+
_
/
0
__
c
I
/ +
__
/
2
_
+
_
/
1
__
c
I1
/
2
+ . . .
. . . +
__
/
/
_
+
_
/
/ 1
__
c/
I
+
_
/
/
_
/
I+1
=
=
_
/ + 1
0
_
c
I+1
+
_
/ + 1
1
_
c
I
/ +
_
/ + 1
2
_
c
I1
/
2
+ . . . +
_
/ + 1
/
_
c/
I
+
_
/ + 1
/ + 1
_
/
I+1
.
where we used Proposition 4.4 and the formulas
_
/
0
_
= 1 =
_
/ + 1
0
_
.
_
/
/
_
= 1 =
_
/ + 1
/ + 1
_
.
Thus the statement 1(/ + 1) is true. Since 1(1) is true and 1(/ + 1) is true whenever
1(/) is true we conclude that 1(:) is true for any : by the principle of mathematical
induction.
Remark 4.1. The equivalent formulation is (c + /)
a
=

a
=0
_
:
:
_
c

/
a
. Indeed, let
us replace the summation index : with : : in equation (4.2). Then : : varies from 0
to :, so : varies from 0 to : and we get
(c + /)
a
=
a

=0
_
:
: :
_
c

/
a
=
a

=0
_
:
:
_
c

/
a
by Proposition 4.3.
Remark 4.2. Due to the Binomial theorem the numbers of combinations
_
:
:
_
are also
called the binomial coecients.
Note that the coecients in the RHS of equation (4.2) are the numbers in row : of
Pascals Triangle.
Example 4.1. Expand (r + )
4
.
Solution. Row 4 of Pascals Triangle is 1 4 6 4 1. So
(r + )
4
= r
4
+ 4r
3
+ 6r
2

2
+ 4r
3
+
4
.
Example 4.2. Expand
_
r

2
_
4
.
41
Solution. By modifying the expression from Example 4.1 we get
_
r

2
_
2
= r
4
+ 4r
3
_

2
_
+ 6r
2
_

2
_
2
+ 4r
_

2
_
3
+
_

2
_
4
=
= r
4
2r
3
+
3
2
r
2

1
2
r
3
+
1
16

4
.
Example 4.3. Find the term independent of r in the expansion of
_
3r +
2
r
2
_
33
.
Solution. By the binomial theorem
_
3r +
2
r
2
_
33
=
33

=0
_
33
:
_
(3r)
33
_
2
r
2
_

=
33

=0
_
33
:
_
3
33
2

r
333
.
The term independent of r is the term with 33 3: = 0, i.e. : = 11, and therefore this
term equals
_
33
11
_
3
22
2
11
.
Binomial theorem can also be used to simplify various expressions. We start with the
following proposition.
Proposition 4.5. For any : , r one has the expansion
(1 + r)
a
=
a

=0
_
:
:
_
r

. (4.4)
Proof. This follows immediately from the Binomial Theorem at c = 1, / = r.
Example 4.4. Find the sum
1 +
1
3
_
7
1
_
+
1
9
_
7
2
_
+ . . . +
1
3
7
_
7
7
_
.
Solution. Notice that the required sum equals
_
1 +
1
3
_
7
(see Proposition 4.5 with
r = 1,3, : = 7). Thus, the sum equals
_
4
3
_
7
=
16384
2187
.
42
Let us put r = 1 and r = 1 in Proposition 4.5, formula (4.4). We obtain the
following relations
_
:
0
_
+
_
:
1
_
+ . . . +
_
:
:
_
= 2
a
.
_
:
0
_

_
:
1
_
+
_
:
2
_

_
:
3
_
+ . . . + (1)
a
_
:
:
_
= 0.
Thus the sum of the entries in the :-th row of Pascals triangle is 2
a
and the alternating
sum of these entries is zero. For example, for : = 4 we get
1 + 4 + 6 + 4 + 1 = 16 = 2
4
.
and 1 4 + 6 4 + 1 = 0.
Let us further explore Proposition 4.5, to derive more identities with binomial coe-
cients. By dierentiating (4.4) we get
:(1 + r)
a1
=
a

=0
_
:
:
_
:r
1
. (4.5)
Let us put r = 1 in (4.5). Note that the LHS of (4.5) equals 0 at r = 1 under the
extra assumption : 2. We get the following result.
Proposition 4.6. Let : . Then
a

I=1
/
_
:
/
_
= : 2
a1
.
and
a

I=1
(1)
I
/
_
:
/
_
= 0 if : 2.
Remark 4.3. Proposition 4.5 is also useful for some approximations. Indeed, in the
expansion
(1 + r)
a
=
_
:
0
_
+
_
:
1
_
r +
_
:
2
_
r
2
+ . . . +
_
:
:
_
r
a
the terms on the RHS get smaller and smaller if r 1. So the rst few terms may give
a reasonable approximation to (1 + r)
a
in this case.
Example 4.5. Write down the binomial theorem for (1+r)
a
, where r . : . Dif-
ferentiate the result to show that
a

I=1
/
2
_
:
/
_
= :(:+1)2
a2
. Show that
a

I=1
(1)
I
/
2
_
:
/
_
=
0 if : 3.
43
Solution. We have
(1 + r)
a
=
a

I=1
r
I
_
:
/
_
(4.6)
by the Binomial Theorem. By dierentiating (4.6) we get
:(1 + r)
a1
=
a

I=1
/
_
:
/
_
r
I1
. (4.7)
and by dierentiating (4.7) we get
:(: 1)(1 + r)
a2
=
a

I=1
/(/ 1)
_
:
/
_
r
I2
. (4.8)
Now we put r = 1 and sum (4.8) and (4.7):
:2
a1
+ :(: 1)2
a2
=
a

I=1
/
2
_
:
/
_
.
which is the required relation as
:2
a1
+ :(: 1)2
a2
= :2
a2
(2 + : 1) = :(: + 1)2
a2
.
Now let us put r = 1 and consider the dierence of (4.7) and (4.8). Assuming : 3
we get
0 =
a

I=1
/
_
:
/
_
(1)
I1

I=1
/(/ 1)
_
:
/
_
(1)
I2
=
=
a

I=1
(1)
I1
/
_
:
/
_
(1 (/ 1)(1)) =
a

I=1
(1)
I1
/
2
_
:
/
_
. (4.9)
By multiplying (4.9) by (1) we get the required relation.
Example 4.6. Simplify
a

I=1
/
3
_
:
/
_
where : .
Solution. We have (1 + r)
a
=
a

I=0
_
:
/
_
r
I
by the binomial theorem. Dierentiation
44
gives
:(1 + r)
a1
=
a

I=0
/
_
:
/
_
r
I1
. (4.10)
:(: 1)(1 + r)
a2
=
a

I=0
/(/ 1)
_
:
/
_
r
I2
. (4.11)
:(: 1)(: 2)(1 + r)
a3
=
a

I=0
/(/ 1)(/ 2)
_
:
/
_
r
I3
=
=
a

I=0
(/
3
3/
2
+ 2/)
_
:
/
_
r
I3
. (4.12)
Let us put r = 1 and let us add the equations (4.11) multiplied by 3 to the equation
(4.12) (so that /
2
terms cancel). We get
:(: 1)(: 2)2
a3
+ 3:(: 1)2
a2
=
a

I=0
(/
3
/)
_
:
/
_
. (4.13)
Let us now add (4.10) and (4.13) (so that the / terms cancel). We have
:(: 1)2
a3
(: 2 + 3 2) + : 2
a1
=
a

I=0
/
3
_
:
/
_
.
Thus
a

I=0
/
3
_
:
/
_
= :(: 1)2
a3
(: + 4) + : 2
a1
= :2
a3
(:
2
+ 3: 4 + 4) = :
2
(: + 3)2
a3
.
Example 4.7

. Prove that
a

I=1
(1)
I
1(/)
_
:
/
_
= 0 if 1(r) is a polynomial of degree at
most : 1. (For the denition of the degree of a polynomial see Section 5.1 below).
Sketch of Solution. Write down the Binomial Theorem for (1+r)
a
and its rst deg 1
derivatives. Substitute r = 1 to get 0 in the lefthand sides of the identities. Show
that there is a combination of the obtained identities which gives
a

I=0
(1)
I
1(/)
_
:
/
_
in
the righthand side for any given polynomial 1(/).
45
4.3 Application to Trigonometric Calculations
We are going to obtain some trigonometric identities with the help of complex numbers.
Recall from Section 1.2.3 that
cos :o + i sin :o = (cos o + i sin o)
a
by De Moivres Theorem. This theorem can be used to express cosine and sine of a
multiple angle through cosine and sine of the angle.
Example 4.8. Express cos 4o as a polynomial in cos o.
Solution. By De Moivres Theorem
cos 4o + i sin 4o = (co:o + i sin o)
4
.
By the binomial theorem
(cos o + i sin o)
4
= cos
4
o + 4 cos
3
o(i sin o) + 6 cos
2
o(i sin o)
2
+ 4 cos o(i sin o)
3
+ (i sin o)
4
=
= cos
4
o + 4i cos
3
o sin o 6 cos
2
o sin
2
o 4i cos o sin
3
o + sin
4
o =
= (cos
4
o 6 cos
2
o sin
2
o + sin
4
o) + i(4 cos
3
o sin o 4 cos o sin
3
o).
By equating the real parts we obtain
cos 4o = cos
4
o 6 cos
2
o sin
2
o + sin
4
o = cos
4
o 6 cos
2
o(1 cos
2
o) + (1 cos
2
o)
2
=
= cos
4
o 6 cos
2
o + 6 cos
4
o + 1 2 cos
2
o + cos
4
o = 8 cos
4
o 8 cos
2
o + 1
Note that by taking imaginary parts we can also express sin
4
o as a polynomial of cos o
and sin o. Further we can reverse the process in order to express powers of cosine and
sine of an angle as a combination of the trigonometric functions of the multiple angles.
Recall from Section 1.2.3 that if . = cos o + i sin o, then
.
a
= cos :o + i sin :o.
1
.
a
= cos :o i sin :o
(see de Moivres theorem and relation (1.12)) for any : . This implies that
.
a
+
1
.
a
= 2 cos :o. (4.14)
.
a

1
.
a
= 2i sin :o. (4.15)
Example 4.9. Express cos
4
o in the form c cos 4o + / cos 2o + c.
46
Solution. Let . = cos o + i sin o. Then using (4.14) and the binomial theorem we get
(2 cos o)
4
=
_
. +
1
.
_
4
= .
4
+ 4.
3
1
.
+ 6.
2
1
.
2
+ 4.
1
.
3
+
1
.
4
=
= .
4
+ 4.
2
+ 6 +
4
.
2
+
1
.
4
=
_
.
4
+
1
.
4
_
+ 4
_
.
2
+
1
.
2
_
+ 6.
Hence
16 cos
4
o = 2 cos 4o + 8 cos 2o + 6.
and
cos
4
o =
1
8
cos 4o +
1
2
cos 2o +
3
8
.
Example 4.10. Express sin
5
o in the form c sin 5o + / sin 3o + c sin o.
Solution. Let . = cos o + i sin o. Then using (4.15) and the binomial theorem we get
(2i sin o)
5
=
_
.
1
.
_
5
= .
5
+ 5.
4
_

1
.
_
+ 10.
3
_

1
.
_
2
+ 10.
2
_

1
.
_
3
+ 5.
_

1
.
_
4
+
_

1
.
_
5
=
= .
5
5.
3
+ 10.
10
.
+
5
.
3

1
.
5
=
_
.
5

1
.
5
_
5
_
.
3

1
.
3
_
+ 10
_
.
1
.
_
.
Hence
32i sin
5
o = 2i sin 5o 10i sin 3o + 20i sin o.
and
sin
5
o =
1
16
sin 5o
5
16
sin 3o +
5
8
sin o.
Exercise 5. Simplify the sum C = 1 + cos 2o + cos 4o + cos 6o + cos 8o + cos 10o.
Hint. Find a combination of C and o where
o = 1 + sin 2o + sin 4o + sin 6o + sin 8o + sin 10o.
47
Chapter 5
Polynomials and roots in the polar
form
5.1 Polynomials
Denition 5.1. A polynomial j(r) is a function of the form
j(r) = c
a
r
a
+ c
a1
r
a1
+ . . . + c
1
r + c
0
where : and c
i
for i = 0. 1. . . . . :. We say that a polynomial j(r) is real if
c
i
for all i = 0. 1. . . . . :. The polynomial j(r) has degree : if c
a
= 0, this is denoted
as deg j(r) = :. In this case c
a
is called the leading term of j(r).
Remark 5.1. In the case j(r) 0, i.e. all of the coecients c
i
= 0, we say j(r) is a
zero polynomial. We do not dene a degree of j in this case.
Remark 5.2. Polynomials of degree 1 are also called linear polynomials.
We considered division of integers with remainder in Section 1.1. There is a version
of a division with remainder for polynomials.
Euclidean Algorithm for Polynomials
Let )(r) and p(r) be polynomials where p(r) is not a zero polynomial. Then there exist
polynomials (r) and :(r) such that )(r) = (.) p(r) + :(r), with deg :(r) < deg p(r),
or :(r) is a zero polynomial.
To obtain these we use long division, e.g. let us take )(r) = 2r
4
+ r
3
+ r
2
3r 2,
and p(r) = r
2
+ 2r + 3. We have the following process:
48
We stop as deg(4r 5) = 1 < 2 = deg p(r). So (r) = 2r
2
3r + 1 and :(r) = 4r 5
(and indeed )(r) = (r) p(r) + :(r)).
Denition 5.2. A number c is called a root of a polynomial j(r) if j(c) = 0.
We have the following Theorem as a corollary from the Euclidean Algorithm.
Theorem 5.1. A number c is a root of the nonzero polynomial )(r) if and only if
)(r) is divisible by r c, i.e. )(r) = (r c)(r) for some polynomial (r).
Proof. By the Euclidean algorithm )(r) = (r c)(r) + :(r) where deg :(r) < 1 or
:(r) is a zero polynomial, i.e. in both cases :(r) = constant = :. Then )(c) = :, so
)(c) = 0 : = 0.
Recall now that if . = r+i, where r. then . = ri is the complex conjugate
of .. Note also that
. + . = 2r = 2Re.. .. = r
2
+
2
= .
2
.
Also for any .. u we have
. + u = . + u. .u = . u. (5.1)
By iterating the last equality we also have
.
a
= .
a
(5.2)
for any : . Furthermore, the following lemma takes place.
Lemma 5.2. Let )(r) be a real polynomial. Then )(c) = )(c) for any c .
Proof. Let )(r) have the form )(r) =
a

I=0
c
I
r
I
where c
I
. By applying the properties
(5.1) and (5.2) we get
)(c) = c
a
c
a
+ c
a1
c
a1
+ . . . + c
1
c + c
0
= c
a
c
a
+ c
a1
c
a1
+ . . . + c
1
c + c
0
=
= c
a
c
a
+ c
a1
c
a1
+ . . . + c
1
c + c
0
= )(c)
as required.
49
As a corollary from this lemma we get the following.
Proposition 5.1. Let )(r) be a real polynomial. If c is a root of )(r) then c is
also a root of )(r).
Proof. Let c be a root of )(r), so )(c) = 0. By Lemma 5.2 we have )(c) = 0 as
required.
This leads us to the following proposition.
Proposition 5.2. Let c be a nonreal root of a real polynomial )(r). Then
)(r) = (r
2
2Re(c)r +c
2
)p(r) (5.3)
where p(r) is some real polynomial.
Proof. By Theorem 5.1 we can represent )(r) = (rc)(r) where (r) is a polynomial.
We have )(c) = (c c)(c) = 0 by Proposition 5.1. As c we have c c = 0, so
(c) = 0. By Theorem 5.1 (r) = (r c)p(r) for some polynomial p(r). Thus
)(r) = (r c)(r c)p(r) = (r
2
(c + c)r + cc))p(r)
as required.
Exercise 6. Show that the polynomial p(r) from (5.3) is a real polynomial.
Example 5.1. Given that 2 + 3i is a root of the equation
r
4
+ 7r
2
12r + 130 = 0.
nd all the roots of the equation.
Solution. Since 2 + 3i is a root, so is 2 3i and
r
2
2Re(2 + 3i)r + 2 + 3i
2
= r
2
+ 4r + 13
is a factor of the polynomial. Hence
r
4
+ 7r
2
12r + 130 = (r
2
+ 4r + 13)(r
2
4r + 10).
So the remaining roots are given by r
2
4r + 10 = 0, i.e.
r =
4

16 40
2
=
4

24
2
=
4

1
2
=
4 2

6i
2
= 2

6i.
Therefore the required roots are 2 3i, 2

6i.
We conclude by stating the following important result.
Theorem 5.3 (Fundamental Theorem of Algebra). Any polynomial j(r) has a root in
.
50
We do not prove Theorem 5.3 in this course. The theorem has the following immediate
consequences.
1. The number system is large enough to deal with the solutions of polynomial
equations.
2. Every complex polynomial of degree : can be factorised into : linear factors. (Ex-
ercise).
3. Every real polynomial can be factorised into a product of real linear factors (cor-
responding to real roots) and real quadratic factors (corresponding to pairs c. c of
complex roots as in Proposition 5.2).
5.2 Roots in the Polar Form
Let o . One can dene exponents of imaginary numbers and it appears that the
following formula takes place:
c
i0
= cos o + i sin o. (5.4)
While we do note prove the relation (5.4) we show that it is consistent with a number of
familiar properties of exponential and trigonometric functions.
1. Let o = 0. Then c
i0
= c
0
= 1 and cos o + i sin o = 1, so (5.4) indeed holds in this
case.
2. The known relation cos o i sin o =
1
cos o + i sin o
can be rearranged using (5.4)
into c
i0
=
1
c
i0
which should hold for the exponential function.
3. De Moivres theorem in the form (cos o + i sin o)
a
= cos :o + i sin :o is rearranged
using (5.4) into (c
i0
)
a
= c
ia0
which is one of the index laws.
4. De Moivres theorem in the form
(cos o + i sin o)(cos + i sin ) = cos(o + ) + i sin(o + )
takes the form c
i0
c
i
= c
i(0+)
which is a familiar property of exponential functions.
One way to establish relation (5.4) would be to use the power series for the functions
c
a
. cos r and sin r. It will be discussed in the later courses that the following expansion
take place:
c
a
=

a=0
r
a
:!
.
so for r = io,
c
i0
=

a=0
(io)
a
:!
= 1 + io
o
2
2!
i
o
3
3!
+
o
4
4!
+ i
o
5
5!
+ . . . (5.5)
51
And also
cos o = 1
o
2
2!
+
o
4
4!

o
6
6!
+ . . . (5.6)
and
sin o = o
o
3
3!
+
o
5
5!

o
7
7!
+ . . . (5.7)
Then formula (5.4) can be seen by comparing (5.5) with the sum of the series (5.6) and
the series (5.7) multiplied by i.
Example 5.2. c
2i
= cos 2 + i sin 2 = 1. Similarly, c
2ia
= 1 for any : . Also
clearly the opposite holds: if c
i
= 1 for some then = 2: for some : .
Formula (5.4) is useful for nding the roots of complex numbers. Let . = :(cos o +
i sin o) = :c
i0
.
Proposition 5.3. For all positive integers :, the :-th roots of . = :c
i0
are c
I
=
:
1

c
i
(
+2

)
, where / = 0. 1. 2. . . . . : 1.
Remark 5.3.
1. In Proposition 5.3 :
1

means the unique real nonnegative :-th root of the real


nonnegative number :.
2. If : = 0, that is . = 0, the only :-th root is 0.
3. Proposition 5.3 states in particular that any complex number . has : distinct :-th
roots in unless . = 0.
Proof of Proposition 5.3. Let c be the :-th root of ., that is c
a
= .. Let c have modulus
: and argument o, so c = :c
ic
. By de Moivres theorem c
a
has modulus :
a
and argument
:o. So:
c
a
= . :
a
= : and :o = o + 2/ for some /
: = :
1

and o =
o + 2/
:
c = :
1

c
i
(
+2

)
.
Also we note that for : = 0 we have
c
I
= c
|
c
i
(
+2

)
= c
i
(
+2

)
c
2i
()

= 1
/ |
:
.
Hence all the roots c
0
. . . . . c
a1
are pairwise dierent, and any :-th root, c
|
, satises
c
|
= c
I
for some /, 0 / : 1. (We just divide | by :, that is | = : : + / where
:. / , 0 / : 1.)
Let us now sketch the :-th roots on a diagram. Let c
I
= :
1

c
i
(
+2

)
, : = 0.
52
Figure 5.1
A sketch of general complex roots on the complex plane;

=
1

(
+2

)
. The gure shows a circle of
radius
1

with its centre located at the origin.


Figure 5.1 shows that c
0
. c
1
. . . . . c
a1
are distinct roots and that c
I
for other values of
/ gives repetitions of these roots.
Example 5.3. Solve the equation .
2
= 4i.
Solution. Note that 4i = 4c
i

2
in the polar form. By Proposition 5.3 the square roots
of 4i are
c
I
= :
1
2
c
i
(

2
+2
2
)
. / = 0. 1.
That is
c
0
= 2c
i

4
and c
1
= 2c
i
5
4
.
Thus
c
0
= 2
_
cos

4
+ i sin

4
_
= 2
_

2
2
+ i

2
2
_
=

2(1 + i).
Similarly, c
1
= 2
_
cos
5
4
+ i sin
5
4
_
=

2(1+i). Thus the roots are

2(1+i).
Example 5.4. Find the cube roots of unity, i.e. solve the equation r
3
= 1.
Solution. We have 1 = 1c
i0
in the polar form. So the cube roots of unity are
c
I
= 1
1
3
c
i
(
0+2
3
)
. / = 0. 1. 2.
53
So we have
c
0
= 1. c
1
= c
2
3
= cos
2
3
+ i sin
2
3
=
1
2
+ i

3
2
.
and
c
2
= c
i
4
3
= cos
4
3
+ i sin
4
3
=
1
2
i

3
2
.
Note. The two nonreal cubic roots of 1 are conjugates of each other. This can also be
seen algebraically because r
3
= 1, i.e. r
3
1 = 0, is a real polynomial equation so that
Theorem 5.1 applies. Geometrically the cubic roots can be seen sketched on Figure 5.2.
Figure 5.2
A diagram of the cube roots of unity on the complex plane.
Proposition 5.4. The sum of all : :-th roots of unity is zero.
Proof. The :-th roots of unity are 1. u. u
2
. . . . . u
a1
where u = c
i
2

. Notice that
1 + u + u
2
+ . . . + u
a1
=
u
a
1
u 1
(5.8)
because the LHS is the sum of the rst : terms of a geometric series. Since u
a
= 1, the
RHS of (5.8) equals 0 as required.
54
Chapter 6
Matrices
6.1 Scalars, Matrices and Arithmetic
Scalar is an alternative name for a number. It is used when we want to restrict all the
numbers in our considerations to be of a specic type. For example our scalars will be
real numbers means that all the numbers in the calculations are assumed to be real.
A matrix is a rectangular array of scalars. An : : matrix (spoken : by :) is a
matrix with : rows and : columns. Such a matrix contains :: scalars. For example,
_
3 0 1
2 5 4
_
is a 2 3 matrix; it has 6 scalars.
For an :: matrix , :: is called the type (or the size) of the matrix, the rows
are numbered from the top and the columns are numbered from the left. The scalar in
the i-th row and the ,-th column is called the (i. ,)-th entry, it is denoted c
i)
or ()
i)
.
For example, if =
_
3 0 1
2 5 4
_
then c
12
= 0, c
22
= 5, c
21
= 2, etc. Two matrices
and 1 are said to be equal ( = 1), if they have same type :: and all the corre-
sponding entries are equal: c
i)
= /
i)
for all i = 1. . . . . : and , = 1. . . . . :.
There are some important operations which we can perform with matrices.
Addition
Matrices of the same type can be added entry by entry.
Denition 6.1. Let and 1 be : : matrices. Then + 1 is dened to be the
:: matrix with the entries
( + 1)
i)
= c
i)
+ /
i)
for all i and ,. 1 i :. 1 , :.
For example,
_
3 0 1
2 5 4
_
+
_
1 2 4
4 1 0
_
=
_
4 2 5
6 6 4
_
.
Remark 6.1. If and 1 are not of the same type then + 1 is not dened.
55
Scalar Multiplication
To multiply a matrix by a scalar we multiply all the entries by that scalar.
Denition 6.2. If is an :: matrix and ` is a scalar then ` is the :: matrix
with the entries
(`)
i)
= `c
i)
for all i and ,. 1 i :. 1 , :.
For example,
3
_
3 0 1
2 5 4
_
=
_
9 0 3
6 15 12
_
.
Subtraction
Having matrix addition and scalar multiplication we also get matrix subtraction. Indeed
we dene 1 = +(1)1, provided and 1 have the same type ::. The matrix
elements can then be found as
( 1)
i)
= c
i)
/
i)
for all i, , . 1 i :. 1 , :.
For example,
_
3 0 1
2 5 4
_

_
1 2 4
4 1 0
_
=
_
2 2 3
2 4 4
_
.
Rules of Arithmetic
The operations of addition and multiplication by scalars satisfy the following rules:
(a) Commutativity of addition: + 1 = 1 + ,
(b) Associativity of addition: ( + 1) + C = + (1 + C),
(c) `( + 1) = ` + `1,
(d) `(j) = (`j),
where . 1. C are matrices of the same type and `. j are scalars.
We prove property (a), proofs of the other properties are similar.
Proof of (a). Note that + 1 and 1 + are both dened. We have
( + 1)
i)
= c
i)
+ /
i)
= /
i)
+ c
i)
= (1 + )
i)
.
so all the corresponding matrix entries in + 1 and 1 + are equal, hence + 1 =
1 + .
56
By using the rules (a) (d) further properties can be obtained. For example, we have
(1 + C) = ( 1) C (6.1)
if . 1 and C are matrices of the same type. It is easy to verify (6.1) directly, however
let us derive it from the previous properties. By the denition of subtraction, we have
(1 + C) = + (1)(1 + C) = +
_
(1)1 + (1)C
_
by (c). Using associativity (b), we get
+
_
(1)1 + (1)C
_
= ( + (1)1) + (1)C = ( 1) C
by the denition of subtraction, thus (6.1) is established.
Remark 6.2. Due to associativity (b), we can omit brackets when considering expres-
sions with a few matrix additions.
Example 6.1. Simplify 2(31) +3(+21), where the matrices and 1 have the
same type.
Solution. We have
2(3 1) + 3( + 21) = 2(3) + 2((1)1) + 3 + 3(21) =
= 6 + (2)1 + 3 + 61 = 9 + 41.
Zero Matrix
The : : zero matrix (denoted C
n,a
or just C) is the : : matrix with all entries
zero. For example, C
2,3
=
_
0 0 0
0 0 0
_
. The zero matrix has the following important
properties:
+ C
n,a
= = C
n,a
+ ,
0 = C
n,a
(here in the LHS 0 is a zero scalar),
`C
n,a
= C
n,a
,
for all :: matrices and all scalars `.
Just as a zero matrix has all entries 0, a zero row of a matrix means a row all of whose
entries are zero, and a nonzero row means a row not all of whose entries are zero (like-
wise for columns). So, for example, in the matrix

1 2 1 3
0 0 0 0
1 2 0 0
0 0 0 1

, row 2 is a zero row,


rows 1,3 and 4 are nonzero rows and all columns are non-zero.
57
Matrix Multiplication
The next very important operation on matrices is multiplication. In order to dene
it, it is convenient to have the notion of a scalar product. Let r
1
. r
2
. r
3
. . . . . r
n
, and

1
.
2
.
3
. . . . .
n
be two sequences of scalars of length : . Then their scalar product
is dened to be
n

i=1
r
i

i
. For example, the scalar product of 1. 3. 2 and 4. 1. 2 is
(1 4) + 3 (1) + 2 (2) = 3.
Let be an | : matrix and 1 an :: matrix (so that the number of columns of
equals the number of rows of 1). Then the matrix product, 1, is dened to be the
| : matrix with the matrix entries
(1)
i)
= scalar product of the entries in the i-th row of and the ,-th column of 1
for all i and ,, 1 i |, 1 , :.
Notice that since the entries of the i-th row of are c
i1
. c
i2
. . . . c
in
and the entries of
the ,-th column of 1 are /
1)
. /
2)
. . . . . /
n)
, we have
(1)
i)
= c
i1
/
1)
+ c
i2
/
2)
+ . . . + c
in
/
n)
.
i.e.
(1)
i)
=
n

I=1
c
iI
/
I)
.
Remark 6.3. The product 1 is not dened if the number of columns in is not
equal to the number of rows in 1.
Example 6.2. Let =
_
1 2
4 1
_
and 1 =
_
3 0 1
1 2 5
_
. Find 1 and 1 if they
exist.
Solution. is a 2 2 matrix, 1 is a 2 3 matrix. So 1 exists and is a 2 3 matrix.
We have
1 =
_
1 2
4 1
_ _
3 0 1
1 2 5
_
=
_
5 4 11
11 2 1
_
.
However, as the number of columns in 1 (which is 3) is not equal to the number of rows
in (which is 2), 1 does not exist.
Even in situations where 1 and 1 both exist it is possible that 1 = 1. So we
cant talk about multiplied by 1 but rather we must say is postmultiplied by 1
(for 1) and is premultiplied by 1 (for 1). If it so happens that 1 = 1 we say
that and 1 commute.
Example 6.3. Let =
_
1 1
0 1
_
. Find all the matrices that commute with .
58
Solution. Let 1 be any matrix and suppose 1 commutes with . Then 1 = 1. In
particular 1 and 1 both exist. Since is 2 2 this means 1 must have 2 rows and
2 columns, i.e. 1 is 2 2. Say 1 =
_
c /
c d
_
for some scalars c. /. c. d.
Then 1 = 1 if and only if
_
1 1
0 1
_ _
c /
c d
_
=
_
c /
c d
_ _
1 1
0 1
_

_
c + c / + d
c d
_
=
_
c c + /
c c + d
_

c + c = c
/ + d = c + /
c = c
d = c + d.
(6.2)
(6.3)
(6.4)
(6.5)
The equation (6.2) is equivalent to c = 0, (6.3) gives d = c, the equation (6.4) is
automatic and (6.5) gives c = 0. Hence the matrices that commute with are the
matrices of the form
_
c /
0 c
_
, where c. / are arbitrary.
Other surprises of matrix multiplication include the situation that we can have 1 = 0
with = 0. 1 = 0. For example, this is the case for =
_
1 1
1 1
_
, 1 =
_
1 1
1 1
_
.
However, we have the following friendly rules:
1. Associative law: (1C) = (1)C.
2. Distributive laws: (1 + C) = 1 + C, ( + 1)C = C + 1C.
3. For any scalar `, (`)1 = (`1) = `(1).
In each case the law holds provided the matrices . 1. and C are of types that permit
the necessary addition and multiplication to take place. For example, for the associative
law we need to be | :, 1 to be :: and C to be : j for some |. :. :. j .
Denition 6.3. A matrix is called a square matrix if it has the same number of rows
and columns.
For square matrices we can form powers:
1
= ,
2
= ,
3
= () = (),
etc. We also have index laws:

q
=
j+q
. (
j
)
q
=
jq
for all positive integers j. .
In practice, the laws mean that we can manipulate brackets in the usual way, provided
we keep multiplicative factors in the order they come in (need to do that because in general
1 = 1). For example,
(1 C + 41) = 1 C + 41. ( + 1)(C + 1) = C + 1 + 1C + 11.
59
Note in particular the following calculation which holds provided and 1 are square
matrices of the same type:
( + 1)
2
= ( + 1)( + 1) =
2
+ 1 + 1 + 1
2
.
Notice that 1 + 1 = 21 unless and 1 happen to commute. Similarly, (1)
a
=

a
1
a
in general. For example, for : = 2, (1)
2
= 11, whereas
2
1
2
= 11.
Example 6.4. Let =
_
2 0
1 2
_
. Prove that, for all positive integers :,

a
=
_
2
a
0
:2
a1
2
a
_
.
Solution. Let 1(:) denote the statement that
a
=
_
2
a
0
:2
a1
2
a
_
. When : = 1,
LHS=
1
= =
_
2 0
1 2
_
and RHS=
_
2 0
1 2
_
. So 1(:) is true when : = 1.
Suppose now that 1(:) is true when : = /, i.e. suppose that
I
=
_
2
I
0
/2
I1
2
I
_
.
Then

I+1
=
I
=
_
2
I
0
/2
I1
2
I
_ _
2 0
1 2
_
=
_
2
I+1
0
/2
I
+ 2
I
2
I+1
_
=
=
_
2
I+1
0
(/ + 1)2
I
2
I+1
_
=
_
2
I+1
0
(/ + 1)2
(I+1)1
2
I+1
_
so that 1(:) is true when : = / + 1.
So 1(1) is true, and 1(/ + 1) is true whenever 1(/) is true. Therefore 1(:) is true
for all positive integers : by the principle of mathematical induction.
Identity Matrices
The identity matrix, 1
a
(sometimes just denoted by 1), is the : : matrix with 1s on
its main diagonal and 0s elsewhere, i.e.
1
a
=

1 0 0
0 1
.
.
.
.
.
.
.
.
.
0
0 0 1

.
For example, 1
3
=

1 0 0
0 1 0
0 0 1

.
Formally, 1
a
is the : : matrix whose (i. ,)-th entry is o
i)
for all i. ,, 1 i. , :,
where o
i)
is the Kroneckers delta symbol, i.e.
o
i)
=
_
1 if i = ,
0 if i = ,
.
60
The key property of 1
a
is the following.
Proposition 6.1. For any : : matrix ,
1
a
= = 1
a
.
Proof. is ::, 1
a
is :: so 1
a
exists and is ::. Also, for all i and ,, 1 i. , :,
we have
(1
a
)
i)
=
a

I=1
()
iI
(1
a
)
I)
=
a

I=1
c
iI
o
I)
= c
i)
since o
))
= 1 and o
I)
= 0 when / = ,. Therefore 1
a
= . Similarly 1
a
= .
As a consequence we note that factorisation of matrix polynomials becomes possible.
For example
( 21) =
2
(21) =
2
2(1) =
2
2. (6.6)
Also for example

2
5 + 61 = ( 21)( 31). (6.7)
Indeed, using distributivity and other laws,
( 21)( 31) = ( 31) (21)( 31) =
2
(31) 2(1 31) =
=
2
3 2( 31) =
2
5 + 61.
which is the LHS of (6.7).
Remark 6.4. Notice we write 21 and not 2 which is meaningless.
6.2 Systems of Linear Equations
Matrices are particularly useful for considering systems of linear equations. Any system
of : linear equations in : unknowns has the form
c
11
r
1
+ c
12
r
2
+ . . . + c
1a
r
a
=
1
.
c
21
r
1
+ c
22
r
2
+ . . . + c
2a
r
a
=
2
.
.
.
.
.
.
.
.
.
.
.
.
.
c
n1
r
1
+ c
n2
r
2
+ . . . + c
na
r
a
=
n
.
It can be written in matrix form as A = H where
=

c
11
. . . c
1a
.
.
.
.
.
.
c
n1
. . . c
na

is the coecient matrix.


61
A =

r
1
r
2
.
.
.
r
a

is the column of unknowns, and H =

2
.
.
.

is the column of constants.


Here all c
i)
and
i
are some scalars.
We also dene the augmented matrix. It is denoted by [H] and has the form
[H] =

c
11
. . . c
1a

1
.
.
.
.
.
.
.
.
.
c
n1
. . . c
na

n

.
This matrix contains all the numerical data from the system with each row corresponding
to an equation.
We are going to solve a system of equations by manipulating equations. Equivalently
we manipulate the rows of the matrix [H]. We allow ourselves the following operations
called elementary row operations (eros). Below, 1
i
denotes the i-th row of the augmented
matrix.
Type 1 ero. Interchange 1
i
and 1
)
, for i = , (denoted 1
i
1
)
).
Type 2 ero. Multiply 1
i
by a nonzero scalar ` (denoted 1
i
`1
i
).
Type 3 ero. Add to 1
i
the row 1
)
multiplied by a scalar `, for i = , (denoted 1
i

1
i
+ `1
)
).
Eros correspond to invertible operations on the system of equations. Indeed the transfor-
mation 1
i
1
)
undoes the eect of 1
i
1
)
, the transformation 1
i

1
`
1
i
(which is de-
ned because ` = 0) undoes the eect of 1
i
`1
i
and the transformation 1
i
1
i
`1
)
undoes the eect of 1
i
1
i
+ `1
)
. It is clear that solutions of the initial system of
equations are also solutions of the transformed system. Invertibility of eros implies that
solutions of the transformed system are also solutions of the initial system, thus the set
of solutions of the initial system coincides with the set of solutions of the transformed
system.
So our method of solving equations will be to apply eros to change the augmented
matrix into a convenient form where one can see the solutions immediately.
We write
c
1 if the matrix 1 is obtained from by applying particular ero c.
Example 6.5. Solve the system of equations
r
1
+ r
2
r
3
= 1.
r
1
+ r
2
= 0.
r
1
r
2
+ 3r
3
= 1.
62
Solution. In the usual notation,
[H] =

1 1 1 1
1 1 0 0
1 1 3 1


1
2
1
2
1
1

1 1 1 1
0 0 1 1
1 1 3 1

1
3
1
3
1
1

1 1 1 1
0 0 1 1
0 2 4 2


1
2
1
3

1 1 1 1
0 2 4 2
0 0 1 1

1
2

1
2
1
2

1 1 1 1
0 1 2 1
0 0 1 1

.
This corresponds to the equations
r
1
+ r
2
r
3
= 1.
r
2
2r
3
= 1.
r
3
= 1.
so r
3
= 1, r
2
= 1 +2r
3
= 1, r
1
= 1 r
2
+r
3
= 1. Therefore the solution of the given
system of equations is
r
1
= 1. r
2
= 1. r
3
= 1.
Denition 6.4. An echelon matrix is a matrix in which all zero rows are at the bottom
and for the nonzero rows, the number of zeros at the beginning of each row increases as
we go down the matrix.
For example, the matrix

1 1 0 1 0 1
0 1 1 2 1 1
0 0 0 1 1 1
0 0 0 0 1 1
0 0 0 0 0 0

is echelon. In Example 6.5, transforming the augmented matrix to the echelon matrix

1 1 1 1
0 1 2 1
0 0 1 1

enabled us to nd the solution, but we still had some work to do. The method can be
improved by transforming the matrix further to its reduced echelon form.
63
Denition 6.5. A reduced echelon matrix is an echelon matrix in which all the leading
entries, i.e. the rst nonzero entry in each row, are 1, and all entries above (and below)
each leading entry are 0.
For example the matrix,

1 0 2 0 0 1
0 1 1 0 0 1
0 0 0 1 0 1
0 0 0 0 1 2
0 0 0 0 0 0

is reduced echelon. From such a matrix solutions of equations can simply be read o. For
example, in Example 6.5 we arrived at the matrix

1 1 1 1
0 1 2 1
0 0 1 1

. We can continue
applying eros as follows:

1 1 1 1
0 1 2 1
0 0 1 1


1
1
1
1
1
2

1 0 1 0
0 1 2 1
0 0 1 1

1
1
1
1
1
3
then 1
2
1
2
+21
3

1 0 0 1
0 1 0 1
0 0 1 1

which is reduced echelon and corresponds to the equations,


r
1
= 1.
r
2
= 1.
r
3
= 1.
i.e. the solution is seen immediately indeed.
Algorithm for Arriving at Reduced Echelon Form
Any matrix can be transformed to reduced echelon form by performing the following
steps.
1. Take the rst nonzero column.
2. If the upmost element is zero change the rows (i.e. apply eros of Type 1) so that
it becomes nonzero. This entry will be the leading entry.
3. Apply eros of Type 3 so that the entries below the leading entry become 0.
4. Consider the matrix without the upmost row and repeat steps 1 - 4 while it is
possible.
5. Make all the leading entries 1 by eros of Type 2.
64
6. Make all the entries above the leading entries 0 by eros of Type 3 (start with the
leading entry at row 2, then consider the leading entry at row 3 etc.).
Note that after step 4 one arrives at the echelon form. Note also that various modications
of the above algorithm are possible. It may be useful to apply extra eros in order to
simplify the matrix, e.g. to cancel the common denominator or to get rid of factors.
Example 6.6. Transform the matrix

1 2 1 7
2 1 1 0
2 3 4 19

to reduced echelon form and


hence solve the equations
r
1
+ 2r
2
+ r
3
= 7.
2r
1
r
2
r
3
= 0.
2r
1
+ 3r
2
+ 4r
3
= 19.
Solution. We have

1 2 1 7
2 1 1 0
2 3 4 19


1
2
1
2
21
1
then 1
3
1
3
21
1

1 2 1 7
0 5 3 14
0 1 2 5

1
2
1
3

1 2 1 7
0 1 2 5
0 5 3 14


1
3
1
3
51
2

1 2 1 7
0 1 2 5
0 0 13 39

1
2
1
2
then 1
3

1
13
1
3

1 2 1 7
0 1 2 5
0 0 1 3


1
1
1
1
21
2

1 0 5 17
0 1 2 5
0 0 1 3

1
1
1
1
51
3
then 1
2
1
2
+21
3

1 0 0 2
0 1 0 1
0 0 1 3

which is the reduced echelon form. This matrix is the augmented matrix of the system
of equations,
r
1
= 2. r
2
= 1. r
3
= 3. (6.8)
So (6.8) is the solution of the original system.
In the previous examples the systems of linear equations have a unique solution. It
can also be the case that the system of equations has innitely many solutions. In this
case by general solution we mean a formula which denes all solutions, it should normally
contain paramter(s) which can take arbitrary values.
65
Example 6.7. Find the general solution to the system of equations
r
1
+ r
2
+ r
3
= 1.
2r
1
+ 2r
2
+ r
3
= 3.
3r
1
+ 3r
2
+ 2r
3
= 4.
Solution. We have, in the usual notation,
[H] =

1 1 1 1
2 2 1 3
3 3 2 4


1
2
1
2
21
1
then 1
3
1
3
31
1

1 1 1 1
0 0 1 1
0 0 1 1

1
3
1
3
1
2

1 1 1 1
0 0 1 1
0 0 0 0


1
2
1
2

1 1 1 1
0 0 1 1
0 0 0 0

1
1
1
1
1
2

1 1 0 2
0 0 1 1
0 0 0 0

which is reduced echelon form corresponding to the system


r
1
+ r
2
= 2.
r
3
= 1.
Note that this tells us in eect that r
1
can be obtained from r
2
, and there is no restriction
on r
2
. Therefore the general solution is
r
1
= 2 c. r
2
= c. r
3
= 1.
where c is arbitrary.
Notice that in Example 6.7 the variable r
2
which corresponds to the column of the
reduced echelon form not containing the leading entries, takes arbitrary values. This can
be done in general: in the case of innitely many solutions the variables corresponding
to the columns which do not contain leading entries, can take arbitrary values, whereas
values of other variables are xed then.
Consider now the opposite situation when a system of equations appears to have no
solutions.
Example 6.8. Solve the equations
2r
1
+ r
2
+ 8r
3
= 0.
r
2
+ 2r
3
= 3.
r
1
r
2
+ r
3
= 1.
66
Solution. In the usual notation,
[H] =

2 1 8 0
0 1 2 3
1 1 1 1


1
1
1
3

1 1 1 1
0 1 2 3
2 1 8 0

1
3
1
3
21
1

1 1 1 1
0 1 2 3
0 3 6 2


1
3
1
3
31
2

1 1 1 1
0 1 2 3
0 0 0 11

which is the echelon form. The bottom row corresponds to the equation
0r
1
+ 0r
2
+ 0r
3
= 11. i.e. 0 = 11.
This is impossible. So the equations have no solution.
Denition 6.6. A system of equations is said to be consistent if it has at least one
solution. A system of equations is said to be inconsistent if it has no solutions.
Denition 6.7. For any matrix `, the rank of `, denoted rk(`) is the number of
nonzero rows in the echelon matrix obtained from `.
Remark 6.5. The notion of rank is correctly dened. That is it appears that the
number of nonzero rows in the echelon form of a matrix ` does not depend on the
particular echelon form.
Exercise 7. Show that the system of equations A = H is inconsistent if rk() =
rk([H]).
Example 6.9. In Example 6.8 above, we saw that the system was inconsistent. We
also have rk() = 2 and rk([H]) = 3.
Actually the following statement takes place.
Theorem 6.1. A system of linear equations A = H is consistent if and only if rk() =
rk[H].
Example 6.10. Find the value of / for which the following system of equations is
consistent, and solve it for that value of /.
r
1
+ r
2
+ r
3
+ r
4
= 2.
r
1
+ 2r
2
+ 3r
3
+ 2r
4
= 6.
2r
1
+ 7r
2
+ 12r
3
+ 7r
4
= /.
67
Solution. The augmented matrix is,
[H] =

1 1 1 1 2
1 2 3 2 6
2 7 12 7 /


1
2
1
2
1
1
then 1
3
1
3
21
1

1 1 1 1 2
0 1 2 1 4
0 5 10 5 / 4

1
3
1
3
51
2

1 1 1 1 2
0 1 2 1 4
0 0 0 0 / 24


1
1
1
1
1
2

1 0 1 0 2
0 1 2 1 4
0 0 0 0 / 24

.
The bottom row corresponds to the equation 0 = / 24. So, for consistency, we must
have / = 24. In that case the last matrix is reduced echelon and it corresponds to the
system of equations
r
1
r
3
= 2.
r
2
+ 2r
3
+ r
4
= 4.
In eect these equations say r
1
and r
2
can be obtained from r
3
and r
4
, which could take
any values, say r
4
= c and r
3
= . So the general solution is
r
1
= 2 + . r
2
= 4 c 2. r
3
= . r
4
= c
where c. .
6.3 Inverse
Denition 6.8. Let be an : : matrix. is said to be invertible if there is an
: : matrix 1 such that 1 = 1 = 1
a
. The matrix 1 is called the inverse of , it is
denoted
1
.
Notice that the inverse matrix is uniquely determined. Indeed, suppose two matrices
1
1
and 1
2
satisfy 1
1
= 1
1
= 1, 1
2
= 1
2
= 1. Consider 1
2
(1
1
) = 1
2
. On the
other hand consider 1
2
(1
1
) = (1
2
)1
1
= 1
1
, so it follows that 1
1
= 1
2
.
There are noninvertible matrices. For example, any matrix with a zero row is not
invertible. Indeed, if has i-th row zero, then for all 1, 1 has ith row zero so 1 = 1.
We need methods for deciding invertibility, noninvertibility and nding inverses.
Remark 6.6. If and 1 are square matrices such that 1 = 1 then it actually already
follows that 1 =
1
(i.e. it follows that 1 = 1 even though apriori and 1 dont
commute). Similarly, if 1 = 1 then it already follows that 1 =
1
.
68
Row Reduction Method of nding inverse
Let be an : : matrix. We present a method to nd
1
, when it exists. Form the
: (2:) matrix, [1
a
] so that its i-th row consists of the i-th row of followed by the
i-th row of 1
a
, (i = 1. . . . :). Apply eros to the double matrix [1
a
] so that transforms
to its reduced echelon form. If the reduced echelon form of is 1
a
then is invertible.
Moreover when the left half of the : (2:) matrix has been transformed to 1
a
then the
right half will have been transformed to
1
. Otherwise rk() = :, the reduced echelon
form of has a zero row and is not invertible.
Example 6.11. Let =

3 4 5
1 1 2
2 1 3

. Show that is invertible and nd


1
.
Solution.
[1
3
] =

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


1
1
1
2

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

1
2
1
2
31
1
then 1
3
1
3
21
1

1 1 2 0 1 0
0 7 1 1 3 0
0 3 1 0 2 1


1
2
1
2
21
3

1 1 2 0 1 0
0 1 1 1 1 2
0 3 1 0 2 1

1
3
1
3
31
2

1 1 2 0 1 0
0 1 1 1 1 2
0 0 4 3 5 7


1
3

1
4
1
3

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

7
4

1
1
1
1
+1
2

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

7
4


1
1
1
1
31
3
then 1
2
1
2
1
3

1 0 0
5
4

7
4
13
4
0 1 0
1
4

1
4

1
4
0 0 1
3
4
5
4

7
4

.
Therefore is invertible and
1
=

5
4

7
4
13
4
1
4

1
4

1
4
3
4
5
4

7
4

.
Example 6.12. Let =

1 1 5
2 1 8
3 2 15

. Show that is invertible and nd


1
.
69
Solution.
[1
3
] =

1 1 5 1 0 0
2 1 8 0 1 0
3 2 15 0 0 1


1
2
1
2
21
1
then 1
3
1
3
31
1

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

1
2
1
2
then 1
3
1
3

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


1
3
1
3
1
2

1 1 5 1 0 0
0 1 2 2 1 0
0 0 2 1 1 1

1
3

1
2
1
3

1 1 5 1 0 0
0 1 2 2 1 0
0 0 1
1
2

1
2
1
2


1
1
1
1
1
2

1 0 3 1 1 0
0 1 2 2 1 0
0 0 1
1
2

1
2
1
2

1
2
1
2
21
3

1 0 3 1 1 0
0 1 0 3 0 1
0 0 1
1
2

1
2
1
2


1
1
1
1
31
3

1 0 0
1
2
5
2

3
2
0 1 0 3 0 1
0 0 1
1
2

1
2
1
2

.
Thus is invertible and
1
=

1
2
5
2

3
2
3 0 1

1
2

1
2
1
2

.
Example 6.13. Let =

1 2 3
1 1 0
0 2 t + 1

. For which t is invertible? Find


1
when it exists.
Solution.
[1
3
] =

1 2 3 1 0 0
1 1 0 0 1 0
0 2 t + 1 0 0 1


1
2
1
2
+1
1

1 2 3 1 0 0
0 3 3 1 1 0
0 2 t + 1 0 0 1

1
3
1
3

2
3
1
2

1 2 3 1 0 0
0 3 3 1 1 0
0 0 t 1
2
3

2
3
1

= 1 (6.9)
The reduced echelon form of is 1 t = 1 is invertible. Suppose now that
t = 1. We continue (6.9) as follows,
1
1
3

1
1
1
3
then 1
2

1
3
1
2

1 2 3 1 0 0
0 1 1
1
3
1
3
0
0 0 1
2
3(|1)
2
3(|1)
1
|1


1
1
1
1
21
2

1 0 1
1
3

2
3
0
0 1 1
1
3
1
3
0
0 0 1
2
3(|1)
2
3(|1)
1
|1

1
1
1
1
1
3
then 1
2
1
2
1
3

1 0 0
1
3
_
1 +
2
|1
_

2
3
_
1
1
|1
_

1
|1
0 1 0
1
3
_
1 +
2
|1
_
1
3
_
1 +
2
|1
_

1
|1
0 0 1
2
3(|1)

2
3(|1)
1
|1

.
70
So that is invertible when t = 1 and
1
=

|+1
3(|1)
2|
3(|1)
1
1|
|+1
3(|1)
|+1
3(|1)
1
1|
2
3(1|)
2
3(1|)
1
|1

Explanation of Why the Method Works


Let t be any ero. Consider the action of t on an : : matrix . Let t() denote the
resulting matrix. Then there exists an : : matrix 1
t
such that the matrix product
1
t
= t().
Exercise 8.
1. Let t be the ero 1
i
`1
i
. Then
1
t
=

1 0
.
.
.
A
.
.
.
0 1

.
That is 1
t
is the identity matrix where 1 in the (i. i)-th entry is replaced with `.
2. Let t be the ero 1
i
1
)
. Then (1
t
)
I|
=

1. / = | = i and / = | = ,.
1. / = i. | = , or / = ,. | = i.
0. otherwise.
3. Let t be the ero 1
i
1
i
+ `1
)
. Then (1
t
)
I|
=

1. / = |.
`. / = i. | = ,.
0. otherwise.
Let now t
1
. . . . . t
I
be a sequence of eros so that is transformed to the identity matrix.
Denote 1 = 1
t

1
t
1
. . . 1
t
1
. We have that 1 = 1. During the process of applying eros
the : (2:) matrix [1] is transformed to [111] = [1. 11]. Since 1 = 1 we have
1 =
1
. Therefore the resulting matrix is [11] = [1
1
] as required.
Polynomial Method of nding inverse
This method of nding inverse matrix can be applied in the special situation when a
polynomial equation for a matrix is known. We demonstrate it by example.
Example 6.14. A square matrix satises the equation

3
+ 3
2
2 1 = 0.
Show that is invertible and express
1
as a polynomial in .
Solution. We have
3
+3
2
2 = 1, hence (
2
+321) = (
2
+321) = 1.
Therefore is invertible and
1
=
2
+ 3 21.
71
Remark 6.7. In fact, any :: matrix satises an equation 1() = 0 for a suitable
polynomial 1 of degree :.
Exercise 9. Let =
_
c /
c d
_
. Find the coecients c. in the polynomial 1(r) =
r
2
+ cr + such that 1() =
2
+ c + 1 = C
22
.
Further Statements on Inverse Matrices
Proposition 6.2. If is invertible then so is
1
and (
1
)
1
= .
Proof. We have
1
=
1
= 1. But this exactly tells us (in terms of
1
) that
1
is inveritble and (
1
)
1
= .
Proposition 6.3. If and 1 are invertible :: matrices then so is 1 and (1)
1
=
1
1

1
.
Proof. Since and 1 are invertible,
1
and 1
1
exist. Then
(1)(1
1

1
) = (11
1
)
1
= 1
1
=
1
= 1
and
(1
1

1
)(1) = 1
1
(
1
)1 = 1
1
11 = 1
1
1 = 1.
Therefore 1 is invertible and (1)
1
= 1
1

1
.
Theorem 6.2. Let A = H be a system of equations in which the coecient matrix
is square and invertible. Then the system has a unique solution, namely A =
1
H.
Proof. A = H =
1
A =
1
H = 1A =
1
H = A =
1
H, and
conversely, A =
1
H = A =
1
H = A = 1H = A = H. Hence
result.
6.4 Transpose
Denition 6.9. Let be an :: matrix. Then
T
, the transpose of is dened to
be the : : matrix such that
(
T
)
i)
= c
)i
. 1 i :. 1 , :.
For example, if =
_
1 2 3
4 5 6
_
then
T
=

1 4
2 5
3 6

.
72
Note that the i-th row of becomes the i-th column of
T
and the ,-th column of
becomes the ,-th row of
T
.
The transpose matrices satisfy the following properties:
(a) (
T
)
T
= ,
(b) ( + 1)
T
=
T
+ 1
T
,
(c) (`)
T
= `
T
,
(d) (1)
T
= 1
T

T
,
where ` is a scalar, and the matrices are assumed to have sizes such that the operations
in the lefthand sides of the identities are dened.
Exercise 10. Prove properties (a) - (d).
Note that properties (a) - (d) have further immediate corollaries for matrix operations
and transposition, e.g. ( 1)
T
=
T
1
T
. Indeed, ( 1)
T
= ( + (1)1)
T
=

T
+ (1)1
T
=
T
1
T
.
Exercise 11. Let be an invertible matrix. Show that
T
is invertible and that
(
T
)
1
= (
1
)
T
.
Denition 6.10. A square matrix is called symmetric if
T
= . A square matrix
is called skewsymmetric if
T
= .
For example,

1 2 3
2 4 5
3 5 6

is symmetric and

0 1 2
1 0 3
2 3 0

is skewsymmetric.
Note that in a skewsymmetric matrix , c
i)
= c
)i
for all admissible i and ,. In
particular c
ii
= c
ii
, i.e. c
ii
= 0, so that the diagonal entries are all zeros.
The properties (a) - (d) can often be used to establish symmetry or skew symmetry.
Example 6.15. Show that, for every square matrix , +
T
is symmetric.
Solution. ( +
T
)
T
=
T
+ (
T
)
T
=
T
+ = +
T
.
Denition 6.11. A square matrix is called orthogonal if
T
= 1.
For example, the rotation matrix, 1
0
=
_
cos o sin o
sin o cos o
_
(discussed also below), is
orthogonal for any o . Indeed, 1
T
0
=
_
cos o sin o
sin o cos o
_
and 1
0
1
T
0
= 1.
Remark 6.8. As it is sucient to check the invertibility of by one side relation
1 = 1, an orthogonal matrix has to be invertible, with
1
=
T
.
Remark 6.9. is orthogonal
T
= 1. Indeed, if is orthogonal then by
Remark 6.8 we have 1 =
1
=
T
. Also if
T
= 1 then is invertible with

1
=
T
, and hence
T
= 1, so is orthogonal.
73
6.5 Transformations of the Plane
Orthogonal matrices are related to rotations. Let be a 22 matrix. If we think of
_
r

_
as being the coordinates of a point in the plane then the equation
_
r

_
=
_
r

_
represents
a transformation of the plane: a point with coordinates (r. ) is mapped to a point with
coordinates (r

) given as above. For some matrices , the resulting transformation


has simple geometric meaning. For example, let
1
0
=
_
cos o sin o
sin o cos o
_
.
Then the equation
_
r

_
= 1
0
_
r

_
represents anticlockwise rotation about the origin
through the angle o, so 1
0
is called a rotation matrix.
Proof. Let 1 be the point with coordinates (r. ). Let 1

be the point with coordinates


(r

). Suppose 1

is obtained from 1 by rotating about the origin through the angle o


anticlockwise.
Figure 6.1
Rotation of the plane.
Using the complex numbers,
r + i = :(cos c + i sin c)
and
r

+ i

= :(cos(c + o) + i sin(c + o)) = :(cos c + i sin c)(cos o + i sin o) =


= (r + i)(cos o + i sin o) = (r cos o sin o) + i(r sin o + cos o).
Equating real and imaginary parts we obtain
r

= r cos o sin o.

= r sin o + cos o.
74
i.e.
_
r

_
=
_
cos o sin o
sin o cos o
_ _
r

_
.
as required.
Proposition 6.4. (The geometric meaning of matrix multiplication.) The composition
of transformations corresponds to the product of matrices.
For example, for rotations on the plane Proposition 6.4 states that 1

1
0
is the matrix
of rotation by o + . Let us verify this. We have
1

1
0
=
_
cos sin
sin cos
_ _
cos o sin o
sin o cos o
_
=
_
cos cos o sin sin o cos sin o sin cos o
sin cos o + cos sin o sin sin o + cos cos o
_
=
=
_
cos( + o) sin( + o)
sin( + o) cos( + o)
_
= 1
+0
as required. This can be generalised from rotations of the plane to more general trans-
formations as follows. Suppose we have an expression involving variables r
1
. r
2
. . . . . r
a
and we decide to change to variables
1
.
2
. . .
n
where

1
= c
11
r
1
+ c
12
r
2
+ . . . + c
1a
r
a
.

2
= c
21
r
1
+ c
22
r
2
+ . . . + c
2a
r
a
.
.
.
.

n
= c
n1
r
1
+ c
n2
r
2
+ . . . + c
na
r
a
.
for some constants c
i)
. In terms of matrices we have 1 = A where
1 =

1
.
.
.

. A =

r
1
.
.
.
r
a

. and

c
11
. . . c
1n
.
.
.
.
.
.
c
n1
. . . c
na

.
Let us make another change of variables from the s to .
1
. .
2
. . . .
|
where
.
1
= /
11

1
+ . . . + /
1n

n
.
.
.
.
.
|
= /
|1

1
+ . . . + /
|n

n
.
In terms of matrices, if we let 2 =

.
1
.
.
.
.
|

then 2 = 11 .
We can also get directly from A to 2 by observing that 2 = 11 = 1(A) = (1)A,
where the matrix product 1 is used. This establishes Proposition 6.4.
75
6.6 Determinants
Denition 6.12. Let be a 2 2 matrix, =
_
c /
c d
_
. Then the determinant of ,
denoted det , , det
_
c /
c d
_
or

c /
c d

, is dened to be cd /c.
For example, if =
_
3 5
1 2
_
then det = 3 2 5 (1) = 11.
Denition 6.13. Let be a 3 3 matrix, =

c / c
d c )
p i

. Then the determinant


of , denoted det , , det

c / c
d c )
p i

or

c / c
d c )
p i

is dened to be
= c det
_
c )
i
_
/ det
_
d )
p i
_
+ c det
_
d c
p
_
. (6.10)
For example, if =

3 1 2
1 2 4
1 1 3

, then
det = 3 det
_
2 4
1 3
_
1 det
_
1 4
1 3
_
+ 2 det
_
1 2
1 1
_
=
= 3 (6 4) (3 4) + 2(1 2) = 6 + 7 6 = 7.
Remark 6.10. The determinant of any : : matrix can be dened. One way to
do it is to give an inductive denition using determinants of (: 1) (: 1) matrices
generalizing the denition (6.10).
Determinants are particularly useful for checking the invertibility of matrices.
Proposition 6.5. Let be a square matrix. Then
1
exists det = 0.
Remark 6.11. A square matrix with det = 0 is also called nonsingular, and if
det = 0 then is called singular. So Proposition 6.5 can be rephrased as is invertible
if and only if is nonsingular.
The inverse matrix can itself also be written in terms of determinants. Let =
_
c /
c d
_
. If det = 0 then

1
=
1
det
_
d /
c c
_
=
1
cd /c
_
d /
c c
_
. (6.11)
Exercise 12. Check the formula (6.11).
76
Exercise 13

. Let =

c / c
d c )
p /

such that det = 1. Find


1
.
77

You might also like