Original Linear Algebra-@ Richie

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 111

MC/EL/CE/RN 169

LINEAR ALGEBRA

BSc

Compiled By

HENRY OTOO (PhD)

AUGUST, 2019
LINEAR ALGEBRA
MC / EL / CE /RN 169

INSTRUCTOR: Dr Henry Otoo

Email: otoohenry@gmail.com
OFFICE: Mathematical Sciences Department / Phone: 0541414406

OBJECTIVES: The goal of the course is to develop a fundamental understanding in


mathematical describing and studying complex numbers, matrices, vectors. We will look at
theorems and how to apply such theorems for solving problems.

PREREQUISITES: It is assumed that the student has some background in elective


mathematics and core mathematics.

GRADING CRITERIA and EVALUATION PROCEDURES: The grade for the course will be
based on homework, a class test and a final exam.
1. Attendance: All students should make it a point to attend classes. Random
attendance will be taken to constitute 10% of the final grade.
2. Homework: Homework is worth 5% of the final grade. Homework will be assigned
on Fridays and will be due at before 8 AM the following Monday.
3. Class test: There shall be unannounced short class tests at any time within the
semester which will constitute 15% of the continuous assessment. Also, one main
test based on theoretical techniques worth 10% of the grade will be given during
class. The exam date will be announced one week in advance.
4. Final exams: A final exam is worth 60% of the final grade.

ASSESSMENT OF COURSE

Lecture Note prepared by Dr Henry Otoo Page 1


 Assessment of students

The student’s assessment will be in two forms:

Continuous Assessment [40%] and

End of semester examination [60%]

 Assessment of Lecturer

At the end of the course each student will be required to evaluate the course and the
lecturer’s performance by answering a questionnaire specifically prepared to obtain
the views and opinions of the student about the course and lecturer. Please be sincere
and frank.

STUDENT RESPONSIBILITY: It is assumed that each student attends the lectures and works
all the assigned problems. The student is responsible for ALL material covered in class and
any assigned reading. Students may discuss homework problems but you are responsible for
writing your individual answers. If your name does not appear on the final class roll, then you
will not receive a grade for this course.

EXAM POLICY: No makeup for missed work will normally be given, unless extenuating
circumstances occur. Travel plans are not extenuating circumstances. Acceptable medical
excuses must state explicitly that the student should be excused from class.

Lecture Note prepared by Dr Henry Otoo Page 2


TABLE OF CONTENT

CHAPTER ONE Complex Numbers 6

1.0 Introduction 6
1.1 Complex numbers and the complex plane 7
1.2 Equality of Complex Number 8
1.1.1 Set of Complex Numbers 8
1.3 Addition and Subtraction of complex number 9
1.4 Multiplication of complex numbers 10
1.5 Square root of a complex number 11
1.5.1 Multiplication by I 12
1.6 Conjugate of a complex number 13
1.6.1 Modulus of a complex number 14
1.6.2 Conjugate properties 14
1.7 Division of complex number 15
1.8 The modulus argument and polar form of a complex number 16
1.8.1 Principal value of a argument 18
1.9 De Moivre’s Theorem 21

CHAPTER TWO Matrix Algebra 25

2.0 Introduction 25
2.1 Important concepts in matrices 26
2.2 Special matrices 30
2.2.1 Addition of matrices 33
2.2.2 Subtraction of matrices 33
2.3 Multiplying a scalar to a matrix 34
2.4 Multiplication of matrices 35
2.5 Determinants 37
2.5.1 Determinants of Matrices of Higher Order 37
2.6 Invertible matrices 39
2.7 Linear Dependence and Independence of matrices 42
2.8 Rank of a Matrix 43
2.9 System of Linear Equation 44
2.9.1 Matrix Representation of a Linear System 45
2.9.2 Gaussian Elimination method 46
2.9.3 Eigenvalues and Eigenvectors 50

Lecture Note prepared by Dr Henry Otoo Page 3


CHAPTER THREE Vectors 59

3.0 Introduction 59
3.1 Vector Element or Components in a coordinate frame 59
3.2 Basic Properties 60
3.3 Vector Addition and Subtraction 63
3.4 Multiplication of a vector by a scalar 65
3.5 Scalar, Dot, or Inner product 65
3.5.1 Geometrical interpretation of scalar product 66
3.6 Projection of one vector onto the other 68
3.7 Vector or Cross product 70

3.7.1 Geometrical interpretation of vector product 72

3.8 Scalar triple product 73


3.8.1 Geometrical interpretation of scalar triple product 73

Lecture Note prepared by Dr Henry Otoo Page 4


CHAPTER 1

COMPLEX NUMBERS

Objectives
At the end of this chapter you will be able to:
 Understand that i stand for - 1 and be able to reduce powers of i to ± i or ± 1 .
 Understand that all complex numbers are in the form (real part )+ i (imaginary part ).

 Add, subtract and multiply complex numbers.


 Find the complex conjugate of a complex number.
 Divide complex numbers.
 Draw complex numbers.
 Convert a complex number from Cartesian to polar form and vice versa.
 Write a complex number in its exponential form.

1.0 Introduction

Usually in mathematics the first number system that is encountered is the set of natural
numbers ( ), i.e. 1, 2, 3,

This set is subsequently enlarged to cope with various difficulties:

 The set of integers (Z ) allows subtractions such as 3  5  2  2  N but  2  Z  .

 The set of rationals (Q ) enables divisions such as 4  6  2


3  2 3  Z but 2 3  Q  .
 The set of reals (R ) provides for solutions to certain quadratic equations. For example,

x2  2  0 . The solutions are x = 


2  2  Q but  2  R . 
The set of complex numbers is an extension of the set of real numbers. An expression of the
form z = x + iy , where x and y are real numbers, is called a complex number, and the

symbol i is called the imaginary unit: i 2 = - 1 . The numbers x and y are called, respectively,
the real part and imaginary part of z and denoted by
x = Âe z and y = Ám z.

Lecture Note prepared by Dr Henry Otoo Page 5


However the set R will not allow solutions to all quadratic equations. For example, x2  1  0
. This implies that x 2  1 , which has no real solution since x 2 is always positive for real
numbers. This difficulty can be overcome by enlarging the set R to include a new 'number'
denoted i with the property that i 2  1 .
1.1 Complex numbers and the complex plane

The introduction of this new number, i , overcome the problem of finding the square roots of
a negative number.

Example: Find the square roots of -25

Answer: 25  25  1  25  i 2   5  i 
2

So this gives 25  5i

With the existence of the square roots of a negative number, it is possible to find the solutions
of any quadratic equation of the form ax2  bx  c  0 using the quadratic formula. If the
discriminant  b2  4ac   0 , the equation has no real roots.

Example: Solve the quadratic equation x2 - 6x + 25 = 0

Answer: (here a = 1, b = -6, and c = 25)

b2  b2  4ac
The quadratic formula x  gives
2a

62  36  100 6  64
x  i.e. x  3  4i and x  3  4i
2 2

Self Check: Solve the following quadratics equation using the quadratic formula.

a. x2  2x  3  0
b. x2  2x  6  0
c. x2  x  1  0

Lecture Note prepared by Dr Henry Otoo Page 6


Notation
A typical complex number is usually denoted by z, where z is defined in terms of an ordered

pair (x, y ) of real number and the imaginary number i = - 1 such that z = (x, y ) = x + iy

and component notation z = (x, y ).

1.2 Equality of Complex Numbers


Two complex numbers (x1 , y1 ) and (x2 , y2 ) are said to be equal if and only if x1 = x2 and

y1 = y2 .

Examples
1. Given that z1  4  i76 and z2  4  i76 , then z1  z2 since x1  x2  4 y1  y2  76 z1  z2

2. z1  4  i 6 , z2  4  i76 , then z1  z2 since x1  x2  4 and y1  y2 i.e. y1  6, y 2  76

1.2.1 Set of complex numbers

The set of complex numbers is written as C  a  bi : a, b  R

Note that it is customary to denote a complex number by the letter z .

If z  a  bi then:

 The number a is called the real part of z.


 The number b is called the imaginary part of z.

These are sometimes denoted e  z  for a and m  z  for b. That is, e  a  bi   a and
m  a  bi   b .

The real numbers can be represented on the number line.

Is there a similar representation for the complex numbers?

Lecture Note prepared by Dr Henry Otoo Page 7


The definition of a complex number involves two real numbers. Two real numbers give a
point on a plane. So complex numbers can be plotted in a plane by using the x-axis for the
real part and the y-axis for the imaginary part.

This plane is called The Complex Plane or Argand Diagram.

Example: Show the complex numbers z = -2 + 3i and w = 1 - 2i on an Argand diagram.

Answer: The complex numbers z and w are shown on the following diagram.

a  bi  0 if and only if a  b  0

1.3 Addition and subtraction of complex numbers

The addition of two complex numbers, z1 and z2 , in general gives another complex number.

The real components and the imaginary components are added separately and in a like
manner to the familiar addition of real numbers:

 Rules for adding complex numbers

 Add the real parts.

Lecture Note prepared by Dr Henry Otoo Page 8


 Add the imaginary parts.
i.e.  a  bi    c  di    a  c    b  d  i

Example

1. Add 3 + 4i and 2 - i

Answer: The rule for adding these numbers is straightforward:

 Add the real parts together to give the real part 5 ( i.e. 3 + 2 = 5)
 Add the imaginary parts together to give the imaginary part 3 ( i.e. 4 + (-1) = 3)

So  3  4i    2  i   3  2   4  1 i  5  3i

2. Given z1  6  i9 and z2  12  i7 , evaluate z1  z2


Solution
z1  z2  (6  12)  i(9  7)  18  i16

 Rule for subtracting complex numbers

 Subtract the real parts.


 Subtract the imaginary parts.
i.e.  a  bi    c  di    a  c    b  d  i

Example
1. Given that z1  7  i and z2  3  i evaluate z1  z2

Solution:
z1  z2   7  i    3  i    7  3  1   1  i  4  2i

2. Evaluate z1  z2 given that z1  16  i 4, z2  12  i 4

Solution:

z1  z2  (16  i 4)  (12  i 4)  16  12  i 4  i 4  4

1.4 Multiplication of complex numbers

 Rule for multiplying complex numbers


Use the technique for multiplying out two brackets.  a  bi  c  di    ac  bd    bc  ad  i

Lecture Note prepared by Dr Henry Otoo Page 9


The technique for multiplying two complex numbers is similar to that used when multiplying
out two brackets. Now multiply the two complex numbers (3 + 2i) and (4 + 5i)

To multiply these together take the same approach.

This will give

 3  4    3  5i    2i  4    2i  5i 
 12  15i  8i  10i 2
 12  23i  10  Remember that 10i 2  10  1  10 
 2  23i

Self check
(i) Multiply (4 - 2i) and (2 + 3i)
(ii) Multiply (1 + 2i) and (2 - 3i)

1.5 Square roots of a complex number

Earlier examples showed that the square roots of a negative real number could be found in
terms of i in the set of complex numbers. It is also possible to find the square roots of any
complex number.

Example: Find the square roots of the complex number 5 + 12i

Answer: Let the square root of 5  12i be the complex number  a  bi  so  a  bi   5  12i
2

Note: the trick is to equate the real parts and the imaginary parts to give two equations, which
can be solved simultaneously.

First of all multiply out the brackets.  a  bi  a  bi   a 2  b2  2abi

So  a 2  b2   2abi  5  12i

Lecture Note prepared by Dr Henry Otoo Page 10


Equate the real parts to give a 2  b2  5 (call this equation 1).

Equate the imaginary parts to give 2ab  12 (call this equation 2).

Rearranging equation 2 to make b the subject gives b6


a

Substitute in equation 1 to gives a 4  5a 2  36  0

Hence  a 2  9  a 2  4   0  a 2  4 or 9

Since a 2  4  0 then a 2  4 is impossible. This leaves a 2  9 which gives a  3

Substitute a  3 in equation 2 to give b  2 and then substitute a  3 in equation 2 to give


b  2

Hence the square roots are 3  2i and 3  2i

Self check: Find the square root of 3+4i

1.5.1 Multiplication by i

Multiplication by i has an interesting geometric interpretation. The following examples should


demonstrate what happens.

Example: Take the complex number 1 + 7i and multiply it by i

Answer: (1 + 7i)i = i + 7i2 = -7 + i

Lecture Note prepared by Dr Henry Otoo Page 11


Example: Take the complex number -6 - 2i and multiply it by i

Answer: (-6 - 2i)i = -6i - 2i2 = 2 - 6i

Self check Take the complex number 8 - 4i and multiply it by i

In each of the examples it can be seen that the effect of multiplying a complex number by i is

a rotation of the point on the Argand diagram through 90° or

1.6 Conjugate of a complex number


The conjugate of the complex number z  a  bi is denoted as z and defined by z  a  bi
The conjugate is sometimes denoted as z * .

Geometrically the conjugate of a complex number can be shown in an Argand diagram as


the reflection of the point in the x-axis.

Lecture Note prepared by Dr Henry Otoo Page 12


Example Give the conjugate of the complex number z  6  8i and show z and z on an
Argand diagram.

Answer: The conjugate is = -6 - 8i

1.6.1 Modulus of a complex number

The modulus r of a complex number z  a  bi is written z and defined by z  a 2  b2

1.6.2 Conjugate properties

There are special identities which apply to complex numbers and their conjugates. These are
listed below.

Let the two complex numbers be z and w with conjugates z and w respectively.

1. z  z
2. z  z where z  R
3. z  z  2Re  z 

4. z  z

5. z z  z
2

6.  z  w  z  w
7.  zw  zw
8. zw  z w

Lecture Note prepared by Dr Henry Otoo Page 13


9. z  w  z  w . This is commonly known as the triangle inequality.

Proofs of Special Identities

Proof (1): zz

Let z  a  bi for a and b  R , By the definition of a conjugate z  a  bi  a   b  i .

Again using the conjugate definition z  a   b  i  a   b  i  a  bi  z .

Proof (2): z  z where z  R

If z  R then z  a  0i  a (since 0i = 0) for some a  R

By the definition of a conjugate z  a  0i , but 0i = 0, so z  a  z

Proof (3): z  z  2e  z 

Let z  a  bi then z  a  bi , therefore, z  z  a  bi  a  bi  2a  2e  z 

Proof (4): z  z

Let z  a  bi where z  a 2  b2 and z  a  bi so z  a 2   b   a 2  b2  z


2

Proof (5): z z  z
2

Let z  a  bi and z  a  bi , Then zz   a  bi  a  bi   a 2  b2

But z  a 2  b2 so z  a 2  b2 , Thus z z  a 2  b2  z
2 2

Lecture Note prepared by Dr Henry Otoo Page 14


Self check

1. Prove properties 6 to 9.
n n
2. Prove the Triangle inequality for n complex numbers zj   zj
j 1 j 1

1.7 Division of complex numbers

Consider, the two complex numbers z1 = a + bi and z2 = c + di . Writing the quotient in

component form we obtain


z1 a + bi
=
z2 c + di

 Rule for dividing complex numbers

 Find the conjugate of the denominator.


 Multiply the complex fraction, both top and bottom, by this conjugate to give an
integer on the denominator.
 Express the answer in the form a + bi

a  bi  a  bi  c  di   ac  bd    bc  ad  i
 
c  di  c  di  c  di   c2  d 2 

 ac  bd    bc  ad  i
c 2
 d2 c 2
 d2

Example 1
3  7i  3  7i  2  6i  6  14i  18i  42
 
2  6i  2  6i  2  6i   22  62 
36 32
  i
40 40

Lecture Note prepared by Dr Henry Otoo Page 15


Example2
1000  1000i 1000  1000i 10  10i  20000i
 
10  10i 10  10i 10  10i  200
 100i

Self check

i. Divide 5 – 2i by 4 + 3i
z1
ii. If z1 = 2 – 3i and z2 = 3 – 4i, find
z2

 3  2i   3  2i 
iii. Find Re   and Im  .
 1  4i   1  4i 

1.8 The modulus, argument and polar form of a complex number

Remember the Argand diagram in which the point (a, b) corresponds to the complex number
z  a  bi

When the complex number is written as a + bi where a and b are real numbers, this is known
as the Cartesian form.

This point (a, b) can also be specified by giving the distance, r, of the point from the origin
and the angle, , between the line joining the point to the origin and the positive x – axis.

Lecture Note prepared by Dr Henry Otoo Page 16


By some simple trigonometry it follows that a  r cos and b  r sin

Thus the complex number z can be written as r cos  i r sin  . This is known as the polar form
of a complex number. r is called the modulus of z and is the argument of z.

Argument of a complex number: The argument of a complex number is the angle between
the positive x-axis and the line representing the complex number on an Argand diagram. It is
denoted arg (z).
Polar form of a complex number: The polar form of a complex number is z  r  cos  i sin  
where r is the modulus and is the argument.

There will be times when conversion between these forms is necessary. Given a modulus ( r)
and argument ( ) of a complex number it is easy to find the number in Cartesian form.

Lecture Note prepared by Dr Henry Otoo Page 17


Use the following steps to do this:

 Evaluate ' a '  r cos and ' b '  r sin 


 Write down the number in the form a  bi


Example: If a complex number z has modulus of 2 and argument of , express z in the
6
form a  bi and plot the point which represents the number in an Argand diagram.

Answer:

   3    1
We have a  2cos    2  3 and b  2sin    2  1 , So a  bi  3  i .
 6  2  6  2

Self check

11
Plot the complex number with modulus 2 and argument in an Argand diagram.
6

1.8.1 Principal value of an argument


The principal value of an argument is the value which lies between  and  .

The values of the principal argument for a complex number in each quadrant are shown on
the following diagrams.

Lecture Note prepared by Dr Henry Otoo Page 18


b b
arg z   where tan   , i.e   tan 1  
a a

b b
arg z   where     a and tan a  , i.e a  tan 1  
a a

In the previous two examples the complex number is represented by a point in either the first
or second quadrant and the principle argument is positive (lying between 0 and ). In the
next two examples when the complex number is represented in the third and fourth quadrants
the principal argument is negative (lying between 0 and - ).

b b
arg z   where     a and tan a  , i.e a  tan 1  
a a

Lecture Note prepared by Dr Henry Otoo Page 19


b b
arg z   where  a and tan a  , i.e a  tan 1  
a a

b
NB: Taking tan 1   without the modulus sign on a calculator may give a different value than
a
required.

Example 1

 Rule for multiplying two complex numbers in polar form

 Multiply the moduli.


 Add the arguments.

Take two complex numbers z  r1  cos  i sin   and w  r2  cos  i sin   and multiply them
together. Thus zw  r1r2  cos cos  sin  sin    i r1r2  sin  cos  sin  cos 

Using trig. formulae this simplifies to r1r2 cos      i sin    

So the modulus of the product, r1r2 , is the product of the moduli of z and w, namely r1 and r2

The argument of the product,    , is the sum of the arguments of z and w, namely  and 

Example: If z  3 cos300  i sin 300  and w  4  cos600  i sin 600  , find zw

Answer:  0 0

12 cos  30  60  i sin  30  60   12  cos900  i sin900   12i

Lecture Note prepared by Dr Henry Otoo Page 20


 Rule for dividing two complex numbers in polar form

 Divide the moduli.


 Subtract the arguments.

If z  r1  cos  i sin   and w  r2  cos  i sin   then

z r1  cos  i sin  

w r2  cos   i sin  
r1  cos  i sin   cos   i sin  
 
r2  cos   i sin   cos   i sin  
r1 cos cos   sin  sin   i  sin  cos   cos sin  
 
r2  cos2   sin 2  
cos      i sin    
r1

r2

1.9 De Moivre's Theorem


Repeated use of multiplication shows how to compute powers of a complex number.
If
z = r (cos q + i sin q)

then
z 2 = r 2 (cos 2q + i sin 2q)

and
z 3 = zz 2 = r 3 (cos3q + i sin3q)

In general, we obtain the following result, which is named after the French mathematician
Abraham De Moivre.

Theorem: For all rational values of n, if z = r (cos q + i sin q) , then z n = éêër n (cos nq + i sin nq)ùúû

This says that to take the nth power of a complex number we take the nth power of the
modulus and multiply the argument by n.

Lecture Note prepared by Dr Henry Otoo Page 21


Proof

The proof of De Moivre's theorem is in threefold.

Case I: n as a positive integer


The theorem states that if z  r  cos  i sin   then z n  r n  cos n  i sin n  for all n  N .

This is a proof by induction.

For n  1

r  cos  i sin   r  cos  i sin    r1 cos1  i sin1  .


1

And so it is true for n  1

Now suppose that the result is true for n  k then

r  cos  i sin   r k cos k  i sin k 


k

Consider n  k  1 then
r  cos  i sin    r  cos  i sin  r  cos  i sin  
k 1 k

 r  cos  i sin  r k  cos k  i sin k 


 r k 1  cos cos k  sin  sin k   i  sin  cos k  cos  sin k 
 r k 1 cos   k   i sin   k 


 r k 1 cos   k  1   i sin   k  1  
So the result is true for n  k  1 if it is true for n  k . Since it is also true for n  1 , then it is
true for all n  N .

Consider the product of z with itself  z 2  if z  r cos  i r sin 

The rule of multiplying the moduli and adding the arguments gives

z 2  r  r cos      i sin      r 2  cos 2  i sin 2 

Now consider z 3

Lecture Note prepared by Dr Henry Otoo Page 22


Take z 2  r 2  cos 2  i sin 2  and multiply this by z  r cos  i r sin 

This gives z 3   r 2  cos 2  i sin 2    r  cos  i sin     r 3  cos 3  i sin 3 

A pattern has emerged.

If z  r cos  i r sin  , then z n  r n  cos n  i sin n  for all n  N

Case II: n as a negative integer


Let n = - m where m is a positive integer.
n 1 1
{r (cos q + i sin q)} = - n
= m
{r (cos q + i sin q)} {r (cos q + i sin q)}
1
= rationalising gives
{r (cos mq + i sin mq)}
m

= r - m {cos (- mq)+ i sin (- mq)}


= r n (cos nq + i sin nq)

Case III: n as a fraction


p
Let n = were p and q are integers. By the result in (Case I) and (Case II), We can write
q
p
æq ö æq ö 1 é æq ö æ öù
cos çç ÷ ÷ çç ÷÷ ( )q such that êcos çç ÷
÷ ççq ÷ ÷ú p q
÷
çè q ø
÷
+ i sin ÷
èç q ø
÷
= cos q + i sin q
ê çè q ø÷
÷
+ i sin ÷ú = (cos q + i sin q)
çè q ø
÷
ë û

Again, by the result in (Case I) and (Case II), we have


p
      p   p 
 cos  i sin    cos     i sin    
 q q   q   q 

Thus, for all rational values of n.

Example
5
    
Calculate 2  cos  i sin  
  5 5 

Lecture Note prepared by Dr Henry Otoo Page 23


Answer: Using De Moivre's theorem

5
     5     
2  cos  i sin    2 cos  5    i sin  5   
  5 5    5  5 
 32  cos   i sin  
 32

Example
1
   3
Calculate  cos  i sin 
 4 4

Answer: By De Moivre's theorem for fractional powers

1
   3
 1    1  
 cos  i sin   cos     i sin    
 4 4  3 4   3 4 
   
  cos  i sin 
 12 12 

Self check
Calculate
1
     3
1. 27  cos  i sin  
  2 2 
2
     3
2. 8  cos  i sin  
  4 4 
Activity One (Basic Operation of Complex Number

Write the complex numbers in standard form

2.  3i 2  i
1. 3   9 i 2  1
 3  3i  3(1)  i
 3i

Lecture Note prepared by Dr Henry Otoo Page 24


3.)   75 
2


 5i 3 2

 25i 2 * 3
 75

Perform the addition or subtract and write the result in standard form.

4. 4  i   7  2i  5. 11  2i    3  6i 
 11  i  8  4i

11  2i    3  6i 
 8  4i

  
6. 7   18  3  3i 2 
7. 13i  14  7i 
 7  3i 2   3  3i 2 
 14  20i
4

3 7  5 1 
9.    i     i 
4 5  6 6 
8. 22   5  8i   10i 3 7 5 1
  i 
 17  18i 4 5 6 6
 19 37
  i
12 30

3 7  5 1 
9.    i     i 
4 5  6 6 
3 7 5 1
  i 
4 5 6 6
 19 37
  i
12 30

Perform the multiplication and write the result in standard form.

10.  6 *  2
 12
2 3

Lecture Note prepared by Dr Henry Otoo Page 25


11. 1  i 3  2i 
 3  2i  3i  2(i ) 2
but (i ) 2  1
 5i

12. 6  2i 2  3i 
 66  2i   2i2  3i 
 6  22i

13.  14  i 10  14  i 10 
  14   i 10 
2 2

 14  (10)
 24

Find the product of the number and its conjugate

14. 4  3i 4  3i 
 
15.  3  i 2  3  i 2 
 16  9i 2
 9  i2 * 2
 16  9
 11
 25

Perform the division and write result in standard form

6 i 4 4  5i 
17.
16. *
i i 4  5i  4  5i 
 6i 16  20i
 2 
i 16  25i
 6i 16  20i

41

8  7i  1  2i  1
18.   19.
1  2i  1  2i  4  5i 2
22  9i 1  41  40i 
   
5 41  40i   41  41i 
41  40i
 2
41  40 2
41  40i

3281

Lecture Note prepared by Dr Henry Otoo Page 26


20.
2  3i 5i  21.
2i

5
2  3i 2i 2i

15  10i 2  3i  
2  4i  10  5i
2  3i 2  3i  2  i 2  i 
60  25i 12  9i
 
2 2  32 4 1
60  25i 12  9i
 
13 5

Simplify the complex numbers and write it in the standard form

22. 4i 2  2i 3 23.  5i 5
 4(1)  2i (1)  5(1)(1)i
 4  2i  5i

24.  2 
6

 
 i 2
6

8

Lecture Note prepared by Dr Henry Otoo Page 27


Activity 2 - Complex Numbers on the Complex Plane
Calculate the absolute value of each number and then graph each number on the complex
plane.

5. 2

6. 4  2i

Lecture Note prepared by Dr Henry Otoo Page 28


7. 1  i

1  i   1   1  2
2 2

8. 1  i

1  i   1  1  2
2 2

9. 2  2i

2  2i   2   2  82 2
2 2

10. 2  2i

Lecture Note prepared by Dr Henry Otoo Page 29


11. 3  i 3

 3   3
2 2
3 i 3   6

12.  5  i 5

 5    5 
2 2
 5 i 5   10

Lecture Note prepared by Dr Henry Otoo Page 30


13. 3i
3i  02   3  3
2

14. 2  2i 2

 2  
2
2  2i 2   2 2  10
2

15. 5  5i 5

 5  
2
5  5i 5   5 5  150  5 6
2

Lecture Note prepared by Dr Henry Otoo Page 31


Find each sum graphically
16.  4  i    4  5i   0  6i

17.  3  2i    2  4i   1  6i

18.  6  i    3  2i   9  3i

Lecture Note prepared by Dr Henry Otoo Page 32


Write the complex number in trigonometric form.

1.

2.

Represent the complex number graphically and the trigonometric form of the number

Lecture Note prepared by Dr Henry Otoo Page 33


1. 5  5i

2. 3  i


3. 2 1  i 3 

Lecture Note prepared by Dr Henry Otoo Page 34


4. 8i

5. 7  4i

7. 2  cos120  i sin120 

Lecture Note prepared by Dr Henry Otoo Page 35


8.
3
2
 cos 330  i sin 330 

 3 3 
9. 3.75  cos  i sin 
 4 4 

Lecture Note prepared by Dr Henry Otoo Page 36


 3 3 
10. 4  cos  i sin 
 2 2 

Perform and leave the result in trigonometric form.

         
12.3 cos  i sin  4 cos  i sin 
  3 3    6 6 

3         
13.  cos  i sin  6 cos  i sin 
2  6 6    4 4 

5
  2
  
14. cos 140  i sin 140   cos 60  i sin 60  
3  3 

 
15. cos 5  i sin cos 20  i sin 20 
cos 50  i sin 50 
2 cos 120  i sin 120 
16.
cos 20  i sin 20
17.

4 cos 40  i sin 40 
 7   7 
cos   i sin  
18.  4   4 
cos   i sin 


18 cos 54  i sin 54  
9 cos 20  i sin 20 
19.

3 cos 102  i sin 102  20.

5 cos 75  i sin 75 

Lecture Note prepared by Dr Henry Otoo Page 37


Use De Moivre’s Theorem to find the indicated power of the complex number. Express the
result in standard form.

1  i 
3
1.

 1  i 
10
2.

 
5
3. 2 3 i

 3  2i 
5
4.

5. 5  cos 20  i sin cos 20 


3

5 5 
10

6.  cos  i sin 
 4 4 

7. 3  cos150  i sin150 


4

8. 4  cos 2.8  i sin 2.8


5

Use the complex Roots Theorem to find the indicated roots of the complex number and then
represent each of the roots graphically, express the roots in standard form

 4 4 
9. 16  cos  i sin 
 3 3 

Lecture Note prepared by Dr Henry Otoo Page 38


 5 5 
10 Fifth roots 32  cos  i sin 
 6 6 

11 Fifth roots i

12 Fifth roots 1

Lecture Note prepared by Dr Henry Otoo Page 39


Use the complex Roots Theorem to find all the solutions of the equation and represent the
solutions graphically.

1. x4  i  0

2. x5  243  0

3. x3  64i  0

Lecture Note prepared by Dr Henry Otoo Page 40


Exercise

i4
1. Compute the real and imaginary part of z 
2i  3
2. Compute the absolute value and the conjugate of
z  1  i  w  i17
6

3. Write in the “algebraic” form a  ib  the following complex numbers


z  i5  i  1 w  3  3i 
8

4. Write in the “trigonometric” form  cos   i sin   the following complex numbers
 
7

a).8 b.6i c.  cos  i sin 
 3 3
5. Siplify

1 i 3i
a)  1  2i 2  2i  
1 i 1 i

b. 2ii  1   3  i   1  i1  i
3

6. Compute the square root of z  1  i

7. Compute the cube root of z  8

8. prove that there is no complex numbers such that z  z  i

9. Find z  such that

a) z  iz  1 b. z 2 * z  z c. z  3i  3 z

Question 2

Solve the equation


2 z 2  2iz  5  0 z 
3 1
Answer z   i
2 2
Question
 9  3i
w Find the modulus and the argument of the complex number w.
1  2i
 3
Answer w 3 2 , arg w 
4

Lecture Note prepared by Dr Henry Otoo Page 41


Question 3

Find the value of x and the value of y in the equation, Given further that x  , y  
x  iy 2  i   3  i
Answer x, y   1,1

Question 4

  4i
z ,   .
1  i

Given that z is a real number, find the possible values of .

Answer   2

Question 5

Find the values of x and y in the equation

x1  i   y2  i   3  10 x x  , y  
2 2

Answer

x  7, y  1

Question 6

Find the values of x and y in the equation, , Given further that x  , y  


x  iy 3  4i   3  4i
Answer

x, y     7 ,  24 
 25 25 

The complex number z satisfies the equation

Lecture Note prepared by Dr Henry Otoo Page 42


1  18i
4 z  3z  , where z denotes the conjugate of z .solve the equation, given the
2i
answer in the form x  iy , where x and y are real numbers.
Answer
z  4i

Question 7

z  3  4i and zw  14  2i
By showing clear workings, find…….
a) w in the form a  bi , where a and b are real numbers.
b)The modulus and the argument of w

Question 8
z
z  22  4i and  6  8i
w

By showing clear workings, find…….


c) w in the form a  bi , where a and b are real numbers.
d) The modulus and the argument of w

Answer

w  1 2i w 5 arg w  1.11c

Question 9
7  4i
z  2  i   8
2

2i

Express z in the form x  iy where x and ,are real numbers.

Answer=  3  7i

Question 10

The complex conjugate of z is denoted by z.

Solve the equation

Lecture Note prepared by Dr Henry Otoo Page 43


Giving the answer in the form x  iy , where x and y real numbers.
 27  23i
2 z  3z 
1 i Answer

z  2  5i

Question 11

Solve the following equation

z 2  21  20i, z  C.

Give the answers in the form a  bi, where a   and b  

ANSWER

Z  5  2i 

Question 12

The Cubic equation

2 z 3  5z 2  cz  5  0 c   has a solution 0f z  1 2i

Find in any order

a) The other two solutions of the equation


b) The value of c
Answer
z 2  1  2i
1
z3 
2
c  12

Lecture Note prepared by Dr Henry Otoo Page 44


QUESTION

z  8  i(7  2 z), z  C.

The complex conjugate of z is denoted by z.

Determine the value of z in the above equation, given your answer in the form
x  iy where x and y are real numbers.

Lecture Note prepared by Dr Henry Otoo Page 45


CHAPTER TWO

MATRIX ALGEBRA

Objectives
At the end of this chapter you will be able to:
 State the conditions for the equality of two matrices.
 Types of matrices.
 Addition, Subtraction and Multiplication of matrices.
 Multiplying a scalar to a matrix.
 Finding the determinant of a matrix.
 Solving systems of linear equations.
 Computing the eigenvalues and eigenvectors of a matrix.

2.0 Introduction

This chapter investigates matrices and algebraic operations defined on them. These matrices
may be viewed as rectangular arrays of elements where each entry depends on two subscripts
(as compared with vectors, where each entry depended on only one subscript). Systems of
linear equations and their solutions may be efficiently investigated using the language of
matrices.

DEFINITION
(Matrix) A rectangular array of numbers is called a matrix. We shall mostly be concerned
with matrices having real numbers as entries. The horizontal arrays of a matrix are called its
ROWS and the vertical arrays are called its COLUMNS. A matrix having -rows and -
columns is said to have the order m  n

A matrix of ORDER m  n can be represented in the following form:

Lecture Note prepared by Dr Henry Otoo Page 46


 a11 a12 a1n 
a a22 a2 n 
A  aij   21 ,
 
 
 am1 am 2 amn 

where aij is the entry at the intersection of the i th row and j th column.

In a more concise manner, we also denote the matrix by  aij  by suppressing its order.

1 3 7 
Let A    , then a11  1, a12  3, a13  7, a21  4, a22  5 and a23  6
4 5 6
A matrix having only one column is called a COLUMN VECTOR; and a matrix with only one
row is called a ROW VECTOR.
Example 1
Give the size of the matrices below
Solution
 
4 3 0 6 1 0
 
0 2 4 7 1 3 
 1 
 6 1 15 1 0 
 2 
Size 3*6

7 10 1 
8 0  2 

9 3 0 
Size 3*3
In this matrix the number of rows is equal to the number of columns. Matrices that have the same number
of rows as columns are called Squared Matrices.

Lecture Note prepared by Dr Henry Otoo Page 47


12 
 4 
 
2 
 
 17 
This matrix has a single column and often called Column Matrix.

3  1 12 0  9
This matrix has a single Row and often called Row Matrix.

 2
Often when dealing with 1*1 matrix we drop the surrounding bracket and just write -2
2.1 IMPORTANT CONCEPTS IN MATRICES
 Principal Diagonal
The diagonal entries containing the elements a11 , a22 ,K , ann of a square matrix of order n, is

called the principal diagonal.


E.g.

Example

 3 88 5 
 
A   20 15 69  The principal diagonal of A is given by a11  3, a22  15, a33  23
 6 69 23 
 

 Equality of two Matrices


Two matrices A   aij  and B  bij  having the same order m  n are equal if aij  bij for

each i  1,2, , m and j  1,2, , n .

In other words, two matrices are said to be equal if they have the same order and their
corresponding entries are equal.

Lecture Note prepared by Dr Henry Otoo Page 48


Example:

10 8 5  10 8 5 
   
(i ) A 2 5 6  B 2 5 6 
 6 2 23   6 2 23 
   

A  B since both have the same order 3  3 and their individual entries are equal i.e.
a11  b11  10, a12  b12  8, a13  b13  5, a21  b21  2, a22  b22  5, a23  b23  6
a31  b31  6, a32  b32  2, a33  b33  23

10 8 5  10 8 5 
   
(ii) A   2 15 6  B   2 5 6 
 6 2 23   6 2 23 
   

In the above matrix, A  B even though both matrices have the same order. Here their entries
are not the same i.e. a22  b22 since a22  15 b22  5

 Sub – Matrix
Any matrix obtained by omitting some row and columns from a given matrix A is called a
sub-matrix of A.
 a11 a12 a13 
 
 a21 a22 a23 
   11 12 
A A
A   a31 a32
  A21 A22 

a33

 a41 a42 a43 
 
 a51 a52 a53 

Example:

10 8 5 
  15 6 
Given that A   2 15 6  then C    is a sub-matrix of A
 6 2 23   2 23 
 

Lecture Note prepared by Dr Henry Otoo Page 49


 Zero-Matrix/ Null Matrix

A matrix in which each entry is zero is called a zero-matrix, denoted by 0. For example,

0 0 0 0 0
0 2 2    and 0 23  0 0 0
0 0  

 Nilpotent Matrix

A nilpotent matrix is a square matrix N such that N k  0 , for some positive integer k. The
smallest k is sometimes called the degree of N.

Example:

0 1 
The matrix M    , is nilpotent, since M 2  0 . Thus the degree of matrix M  2
0 0

 Square Matrices

A square matrix is a matrix which has the same number of rows and columns. A n  by  n
matrix is known as a square matrix of order n.

10 8 5 
 
A   2 15 6  is an example of a square matrix of order 3  3
 6 2 23 
 

Any two square matrices of the same order can be added and multiplied. A square matrix A
is called invertible or non-singular if there exists a matrix B such that

AB  I n

Where In is the identity element.

Lecture Note prepared by Dr Henry Otoo Page 50


Example:

2 1
  
 1 1  3 3
Given that A    has an inverse matrix B   1  then
 1 2   1
 
3 3
2 1
 
AB  I n   1 1   3 3 1 0

1 0
where I  
   
 1 2   1 1  0 1 0 1
 
3 3

 Identity Matrix

The identity matrix or unit matrix of size n is the n  by  n square matrix with ones on the

main diagonal and zeros elsewhere. It is denoted by I n , or simply by I if the size is immaterial

or can be trivially determined by the context.

1 0 0
1 0 0  0 1
1 0  0 
I1  1 , I 2    , I 3  0 1 0  , , In  
0 1   
0 0 1   
0 0 1

The important property of matrix multiplication of identity matrix is that for m  by  n matrix
A.

I m A  AI n  A.

Example:

1 4 7
 
A   2 6 1  , Show that I m A  AI n  A.
If
 3 2 4
 
Solution
Since A is a square matrix with order 3X 3 the expression becomes
I3 A  AI3  A

Lecture Note prepared by Dr Henry Otoo Page 51


Therefore
 1 0 0  1 4 7   1 4 7  1 0 0   1 4 7 
       
 0 1 0  2 6 1    2 6 1  0 1 0    2 6 1 
 0 0 1  3 2 4   3 2 4  0 0 1   3 2 4 
       
1 0  0 4  0  0 7  0  0 1 0  0 0  4  0 0  0  7  1 4 7 
     
0  2  0 0  6  0 0 1 0    2  0  0 0  6  0 0  0 1   2 6 1 
 0  0  3 0  0  2 0  0  4 3  0  0 0  0  2 0  0  4 3 2 4
     

1 4 7 1 4 7 1 4 7
     
2 6 1  2 6 1  2 6 1
 3 2 4  3 2 4 3 2 4
     

 Transpose of a Matrix
The transposed matrix AT or A¢ of a matrix A is defined to be the matrix which has rows
identical with the columns of A. Thus if A = (ai j ) then AT = (a j i ) . In short AT is obtained by

interchanging the rows and columns of A. Consider the matrix

æa11 a12 a13 ö


÷ æa11 a21 a31 ö
÷
çç ÷ çç ÷
A = çça21 a22 a23 ÷
÷
÷ then AT
= çça12 a22 a32 ÷
÷
÷
çç ÷
÷ çç ÷
÷
çèa31 a32 ÷
a33 ø èça13 a23 ÷
a33 ø

Example 1:

1 4 7 1 2 3
   
Given that A   2 6 1  then AT   4 6 2 
 3 2 4 7 1 4
   

Example 2:

Lecture Note prepared by Dr Henry Otoo Page 52


 5 2 3 5 4 8
A   4 7 1  A   2 7 5
  T

8 5 9  3 1 9 

Theorem Given matrices A and B, then the following holds

T
(i ) (A + B) = AT + BT
T
(ii ) é(A)T ù = A
êë úû
T
(iii ) (kA) = kAT where k is a scalar.
T
(iv) (AB) = BT AT

Observe in (iv) that the transpose of a product is the product of transposes, but in the reverse

order.

 Conjugate Transpose

The conjugate transpose of an m  by  n matrix A with complex entries is the n  by  m

matrix A* obtained from A by taking the transpose and then taking the complex conjugate of
each entry (i.e. negating their imaginary parts but not their real parts). The conjugate
transpose is formally defined by

A 
*
ij
 Aji

This definition can also be written as

 
T
A*  A  AT

where AT denotes the transpose and A denotes the matrix with complex conjugated entries.

Lecture Note prepared by Dr Henry Otoo Page 53


Example

 3  i 5 3  i 2  2i 
If A    then A*  
 2  2i i   5 i 

 Normal Matrix

A complex square matrix A is a normal matrix if

A* A  AA*

where A* is the conjugate transpose of A. That is, a matrix is normal if it commutes with its
conjugate transpose.

If A is a real matrix, then A  AT ; it is normal if AT A  AAT .

 Symmetric Matrix

A symmetric matrix is a square matrix, A, that is equal to its transpose

A  AT

The entries of a symmetric matrix are symmetric with respect to the main diagonal (top left to
bottom right). So if the entries are written as A   aij  , then aij  a ji for all indices i and j. The

following 3×3 matrix is symmetric:

1 2 3 
 2 4 5
 
 3 5 6 

 Skew-Symmetric Matrix

A skew-symmetric (or antisymmetric or antimetric) matrix is a square matrix A whose


transpose is also its negative; that is, it satisfies the equation:

Lecture Note prepared by Dr Henry Otoo Page 54


AT   A

or in component form, if A   aij  ; then aij  a ji for all i and j. For example, the following

matrix is skew-symmetric:

 0 2 1
 2 0 4 
 
 1 4 0 

2.2 Special Matrices: Triangular, Diagonal

We have seen that a matrix is a block of entries or two dimensional data. The size of the
matrix is given by the number of rows and the number of columns. If the two numbers are
the same, we called such matrix a square matrix.

To square matrices we associate what we call the main diagonal (in short the diagonal).
Indeed, consider the matrix

a b
A 
c d
Its diagonal is given by the numbers a and d. For the matrix

a b c
 
A  d e f
g h k 

its diagonal consists of a, e, and k. In general, if A is a square matrix of order n and if aij is

the number in the i th - row and j th - column, then the diagonal is given by the numbers aii ,

for i  1, ,n .

The diagonal of a square matrix helps define two type of matrices: upper – triangular and
lower – triangular. Indeed, the diagonal subdivides the matrix into two blocks: one above the
diagonal and the other one below it. If the lower-block consists of zeros, we call such a matrix

Lecture Note prepared by Dr Henry Otoo Page 55


upper – triangular. If the upper-block consists of zeros, we call such a matrix lower – triangular.
For example, the matrices

a b e
a b  
  and  0 e f
0 d  0 0 k 

are upper-triangular, while the matrices

a 0 0
a 0  
  and  d e 0
c d g h k 

are lower-triangular. Now consider the two matrices

a 0 0 a d g
   
A  d e 0  and B   0 e h.
g h k  0 0 k 
 
The matrices A and B are triangular.

A diagonal matrix is a symmetric matrix with all of its entries equal to zero except may be the
ones on the diagonal. So a diagonal matrix has at most n different numbers. For example,
the matrices

 a 0 0
 a 0  
  and  0 0 0 
0 b 0 0 b
 
are diagonal matrices. Identity matrices are examples of diagonal matrices.

Example. Consider the diagonal matrix

 a 0
A 
 0 b
Define the power-matrices of A by
A0  I 2 , A1  A, A2  AA, A3  AAA etc.

Find the power matrices of A and then evaluate the matrices

Lecture Note prepared by Dr Henry Otoo Page 56


1 1 1 n
I2  A  A2   A
1! 2! n!
for n  1,2,

Answer. We have

 a 0  a 0   a 2 0 
A 
2
  2
.
 0 b  0 b   0 b 
and
 a 2 0   a 0   a3 0 
A3  A2 A   2   3
.
 0 b   0 b   0 b 
By induction, one may easily show that
 an 0
An   .
0 bn 

for every natural number n.

2.2.1 Addition of Matrices: In order to add two matrices, we add the entries one by one.
Note: Matrices involved in the addition operation must have the same size.

So how do we add matrices? The answer is to add entries one by one. For example, we have
a b c       a  b c  
   .
d e f       d   e  f 
Clearly, if you want to double a matrix, it is enough to add the matrix to itself. So we have
double of which implies
a b c a b c  aa bb c  c   2a 2b 2c 
    .
d e f  d e f  d d ee f  f   2d 2e 2f 
Example

1 4 7 6 9 6
   
Given that A   2 6 1  and B   5 6 4  . Evaluate A  B
 3 2 4  2 2 3
   

Lecture Note prepared by Dr Henry Otoo Page 57


Solution

 1 4 7   6 9 6   7 13 13 
     
A  B   2 6 1    5 6 4    7 12 5 
 3 2 4  2 2 3  5 4 7 
     

This suggests the following rule

Let A, B and C be matrices of order m  n , and let k ,  R . Then

1. A  B  B  A (commutativity).
2.  A  B  C  A   B  C  (associativity).

3. k  A   k A.
4.  k   A  kA  A

5. A    A    A  A  0

6. A  0  A

2.2.2 Subtraction of matrices: If M and N are two matrices, then we will/can write
M  N  M   1 N , we subtract the entries one by one.

Note: Matrices involved in the subtraction operation must have the same size.

So how do we subtract matrices? The answer is to subtract entries one by one. For example,
we have
a b c       a  b c  
   .
d e f       d   e  f 
Clearly, if you want to get a zero matrix, it is enough to subtract the matrix from itself. So we
have
a b c a b c   aa b b c  c   0 0 0
    .
d e f  d e f  d d ee f  f   0 0 0

Lecture Note prepared by Dr Henry Otoo Page 58


Scalar Product. Consider the 3x1 matrices

a  
   
X   b  and Y     .
c  
   
The scalar product of X and Y is defined by

 
 
X Y   a b c      a  b  c .
T

 
 
Example 1
For the following matrices perform the following operation if possible.

2 0 2 
 5 
2 0  3 2  0  4  7 2  C   4 9
A  B
 1 8 10  5 12 3 7 9  6 0  6 

a A  B
b B  C
c A  C
Solution
a. Both A and B are of the same size and so we know the addition can be done in this case. Once we
know addition can be done there isn’t much to do than to just add.

 2  4  10 4 
A B   
11 11 17 4 
b. Again since A and B are of the same size we can do the difference as the previous one, there
isn’t really much to do. All we need to be careful with is the order. Just like with real number
arithmetic B-C is different from A-B. So in this case we will subtract the entries of A from entries

 2  4  4 0
B A  
of B 13  5  3 14 
c. In this case because A and C are different sizes the addition cannot be done likewise, A-C,C-
A,B+C,C-B and B-C cannot be done for the same reason.
2.3 Multiplying a Scalar to a Matrix.
Let A   aij  be an m  n matrix. Then for any element k  R we define kA   kaij  .

Lecture Note prepared by Dr Henry Otoo Page 59


1 4 5  5 20 25
For example, if A    and k  5 then 5 A   .
0 1 2 0 5 10 

In order to multiply a matrix by a number, you multiply every entry by the given number.

Keep in mind that we always write numbers to the left and matrices to the right (in the case
of multiplication).

a b c   2a 2b 2c 
2  
d e f   2d 2e 2f 
Example 1 Given that

0 9 8 1 2 3
A   2 3 B   7 0  C   2 5 
 
 1 1   4 1 10 6 

1
Compute 3 A  2 B 
C2
SOLUTION
So we are really being asked to compute a linear combination here. We will do that by first computing the
scalar multiplies and the perform the addition and subtraction. Note as well that in the case of the third
1
scalar multiple we are going to consider the scalar to be a positive and leave the miunus sign out in
2
front of the mtrix. Here is the work for the problem.

 3   55 
 0 27   16 2  1 2  15 2
1         
3 A  2 B    6 9    14 0   1 5  7 23
2  2   2 
 3 3   8 3  5 3   0 4 
   

2.4 Multiplication of Matrices


DEFINITION: (Matrix Multiplication / Product). Let A   aij  be an m  n matrix and B  bij 

be an n  r matrix. The product AB is a matrix C  cij  of order m  r .

Lecture Note prepared by Dr Henry Otoo Page 60


Observe that the product AB is defined if and only if number of columns of A is equal to the
number of rows of B.

1 2 1 
1 2 3 
For example, if A    and B  0 0 3  then
2 4 1 1 0 4 

 1 0  3 200 1  6  12   4 2 19 
AB   
 2  0 1 400 2  12  4  3 4 18

In other words, we have


 a b  e f   ae  bg af  bh 
   
 c d  g h   ce  dg cf  dh 
In fact, we do not need to have two matrices of the same size to multiply them. Above, we
did multiply a  2 by 2  matrix with a  2 by 1 matrix (which gave a  2 by 1 matrix). In

fact, the general rule says that in order to perform the multiplication AB, where A is a  m  n 

matrix and B is a  k  1 matrix, then we must have n  k . The result will be a  m  1 matrix.

For example, we have

 
a b c     a  b  c 
    
d e f     d  e  f  
 

Remember that though we were able to perform the above multiplication, it is not possible to
perform the multiplication
 
  a b c
   d .
  e f
 

Lecture Note prepared by Dr Henry Otoo Page 61


Two square matrices A and B are said to commute if AB  BA .

1. Note that if A is a square matrix of order n then AI n  I n A . Also, a scalar matrix of

order n commutes with any square matrix of order n .


2. In general, the matrix product is not commutative. For example, consider the following
1 1  1 0 
two matrices A    and B    . Then check that the matrix product
0 0 1 0 

 2 0 1 1
AB     1 1  BA .
 0 0   

Properties involving Addition and Multiplication.

1. Let A, B and C be three matrices. If you can perform the appropriate products, then
we have
 A  B  C  AC  BC
and
A  B  C   AB  AC

2. If  and  are numbers, A and B are matrices, then we have

  A  B   A   B

and
    A   A   A

Example 4. Consider the matrices

 0 1 2
A  , B    , and C   0 1 5
 1 0   1

Lecture Note prepared by Dr Henry Otoo Page 62


Evaluate  AB  C and A  BC  . Check that you get the same matrix.

EXAMPLE 5
Compute AC and CA for the following two matrices. If possible
8 5 3
 3 10 2 
1 3 0 4 
A  C 
 2 5 8 9   2 0 4 
 
 1 7 5 
SOLUTION
Okay lets first do AC . Here are the sizes for A and C
AC= AC

2*4 4*3 2*3


So the inner numbers (4 and 4) are the same and so the multiplication can be done and we
can see that the new size of the matrix is 2*3. Now lets actually do the multiplication. We
will go through the first couple of entries in the product in detail and the remaining entries a
little quicker. To get the number in the first row and first column of AC we will multiply the
first row of A and the first column of C (1)(8)+(-3)(-3)+(0)(2)+(4)(-1)=13

If we want the entry in the first row and second column AC we will multiply the first row of A by the
second column of B as follows,

(1)(5)+(-3)(10)+(0)(0)+(4)(-7)=-53

Okay, at this point lets stop and insert these into the product so we can make sure that we’ve got our
bearings. Here is the product so far

8 5 3
 
1 3 0 4   3 10 2  13 53 
 2 5 8 9   2 0 4    
   
 
 1 7 5 

As we can see we’ve got four entries left to compute. For these we will give row and column
multiplications but leave it to you to make sure we used the correct row/column and put the result in
the correct place her is the remaining work.

Lecture Note prepared by Dr Henry Otoo Page 63


1 3   3 2    0  4    4  5   17
 2 8   5 3   8 2    9  1  56
 2  5   510    8 ()0   9  7   23
 2  3   5 2    8 4    9  5   81
Here is a complete product

8 5 3
 
1 3 0 4   3 10 2  13 53 17 
 2 5 8 9   2 0 4    56 23 81
   
 
 1 7 5 

Now lets do CA. here are the sizes of the product

C A  CA

4*3 2*3  N / A

Okay in this case the two inner numbers (3 and 2) are not the same and so this product can’t be done.

Example 6 Compute BD and DB for the given matrices, if possible

 3 1 7   1 4 9 
B  10 1 8 D   6 2 1
 
 5 2 4   7 4 7 

SOLUTION

First notice that both of these matrices are 3*3 matrices and so both BD and DB are defined.

Again its worth pointing out that this example differs from the previous example in that both the
products are defined in this example rather than only being defined as in the previous example. Also
know that in both case the product will be a new 3*3 matrix. In this example we are going to leave the
work of verifying the products to you it is good practice so you should try and verify at least one of the
following products.

 3 1 7   1 4 9   40 38 77 
BD = 10 1 8  6 2 1 =  60 10 33 
 5 2 4   7 4 7   45 0 19 

Lecture Note prepared by Dr Henry Otoo Page 64


 1 4 9   3 1 7   8 23 3
DB   6 2 1 10 1 8 =  43 6 22 
 7 4 7   5 2 4   26 11 45 

2.5 Determinants

For any square matrix of order 2, we have found a necessary and sufficient condition for
invertibility. Indeed, consider the matrix

a b
A 
c d
a b a b a b
determinant of    det    ad  bc
c d  c d c d
The matrix A is invertible if and only if ad  bc  0 . We called this number the determinant
of A. It is clear from this that we would like to have a similar result for bigger matrices (meaning
higher orders). So is there a similar notion of determinant for any square matrix, which
determines whether a square matrix is invertible or not?

A matrix A is said to be a singular matrix if det  A  0 It is called non-singular if det  A  0

2.5.1 Determinants of Matrices of Higher Order

These occur in systems of three linear equations with three unknowns x1 , x2 and x3 . The

determinant is defined by the equation


a11 a12 a13
a22 a23 a12 a13 a12 a13
D = a21 a22 a23 = a11 - a21 + a31 .
a32 a33 a32 a33 a22 a23
a31 a32 a33

Lecture Note prepared by Dr Henry Otoo Page 65


That is, D = a11M11 - a21M 21 + a31M 31 . The determinants M i j , obtained by deleting one row

and one column of D is called the minor of the elements that belong to the deleted row and
column.

Cofactors: The cofactors of the elements in D in the i-th row and j-th column are defined as

(- 1) M i j = Ci j , i.e. (- 1) times the minor of that element. The signs form a


i+ j i+ j

æ+ - +÷ö
çç ÷
checkerboard pattern çç- + ÷
- ÷ . We may write D = a11C11 + a21C21 + a31C31, where C11 is
çç ÷
÷
+÷÷
èç+ - ø

the cofactor of the element a11 in D.

Example. Evaluate

3 2 1
2 1 3 .
4 0 1

We will use the general formula along the third row. We have
3 2 1
2 1 3 1 3 2
2 1 3  4 0 1  4  6  1  1 3  4   29 .
1 3 2 3 2 1
4 0 1

Which technique to evaluate a determinant is easier ? The answer depends on the person
who is evaluating the determinant. Some like the elementary row operations and some like
the general formula. All that matters is to get the correct answer.

Properties of the Determinant

1. Any matrix A and its transpose have the same determinant, meaning
det A  det AT

Lecture Note prepared by Dr Henry Otoo Page 66


2. The determinant of a triangular matrix is the product of the entries on the diagonal,
that is
a b a 0
  ad .
0 d b d

3. If we interchange two rows, the determinant of the new matrix is the opposite of the
old one, that is
a b c d
 .
c d a b
4. If we multiply one row with a constant, the determinant of the new matrix is the
determinant of the old one multiplied by the constant, that is
 a b a b a b
  .
c d c d c  d
In particular, if all the entries in one row are zero, then the determinant is zero.
5. If we add one row to another one multiplied by a constant, the determinant of the new
matrix is the same as the old one, that is
a  c b   d a b a b
  .
c d c d c   a d  b

Note that whenever you want to replace a row by something (through elementary
operations), do not multiply the row itself by a constant. Otherwise, you will easily
make errors (due to Property 4).
6. We have
det  AB   det  A det  B  .

In particular, if A is invertible (which happens if and only if det  A  0 ), then

det  A1  
1
det  A

If A and B are similar, then det  A  det  B  .

Lecture Note prepared by Dr Henry Otoo Page 67


2.6 Invertible Matrices

Invertible matrices are very important in many areas of science. For example, decrypting a
coded message uses invertible matrices

Definition. An n  n matrix A is called nonsingular or invertible iff there exists an n  n matrix


B such that

AB  BA  I n

where In is the identity matrix. The matrix B is called the inverse matrix of A. The inverse of
an n by n matrix is another n by n matrix. If the first matrix is A, its inverse is written A1 (and
pronounced "A inverse").

Example. Let

 2 3  1 3 2 
A  and B   
 2 2  1 1 
Solution
 2 3  1 3 2   1 0 
AB  BA       I 2 . Hence A is invertible and B is its inverse.
 2 2  1 1   0 1 

Notation. A common notation for the inverse of a matrix A is A1 . So AA1  A1 A  I n .

Method 1

Example. Find the inverse of

 1 1
A .
 1 2 
Solution:

Lecture Note prepared by Dr Henry Otoo Page 68


a b
Write A1   
c d 
Since
 ac bd 
AA1     I2 .
 a  2c b  2d 
we get
 a  c  1
a  2c  0


 b  d  0
 b  2d  1
Therefore,
2 1 1 1
a , b , c , d  .
3 3 3 3

2 1
 3 3
Hence A1   
1 1 
 
3 3 

The inverse matrix is unique when it exists. So if A is invertible, then A1 is also invertible and

A 
1 1
 A.

The following basic property is very important:

If A and B are invertible matrices, then AB is also invertible and  AB   B 1 A1


1

Method 2

If A  0 , the inverse of a 2  2 matrix A may be obtained from A as follows:

(i) Interchange the two elements on the diagonal.


(ii) Take the negatives of the other two elements.

Lecture Note prepared by Dr Henry Otoo Page 69


1
(iii) Multiply the resulting matrix by or, equivalently, divide each element by A .
A

In case A  0 , the matrix A is not invertible.

Thus,

1
A1  adj  A
A

Example: Given

a b
A 
c d

Solution:

 d c   d b 
T

adj  A      and det A  ad  bc .


 b a   c a 

which gives
1  d b 
A1   .
ad  bc  c a 

2.7 Linear Dependence and Independence of Matrices

Lecture Note prepared by Dr Henry Otoo Page 70


Definition. A set of n, row or column vectors v1 , v2 ,K , vn are said to be linearly dependent if

there exist n scalars a 1 , a 2 ,K , a n , not all equal to zero, such that a 1v1 + a 2v2 + K + a nvn = 0 .

When no such scalars exist such that the above relationship is only true when
a 1 = a 2 = K = a n = 0 , then N matrix is said to be linearly independent.

Test for Linear Dependence


Rows and columns of a square matrix A are said to be linear independent if A = 0 . Similarly,

if A ¹ 0 , then the rows and columns are of matrix A are said to be linearly independent.

é1 4 3ù
ê ú
Example. Consider the matrix A = ê- 2 18 7ú where A = 0 . This implies linear
ê ú
ê 4 - 6 1ú
ë û
dependence exists between either rows or columns of A. Again the rows or columns are
linearly dependent because c2 = 2(c3 - c1 ) where ci denotes column i. Consider another

é1 1 0ù
ê ú
example, B = ê3 2 1úwhere B = - 3 . So the rows or columns of B are linearly
ê ú
ê1 1 3ú
ë û
independent.

2.8 Rank of a Matrix


There are two equivalent definitions of a rank of a matrix. Firstly, the rank of a matrix may
be defined in terms of linearly independent vectors, row or vector. Suppose that the columns
of an m´ n matrix are interpreted as the components in a given basis of n (n-component)
vectors v1 , v2 ,K , vn , as follows:

é­ ­ L ­ù
ê ú
A = êv1 v2 L vn ú
ê ú
ê¯ ¯ L ¯ú
ë û

Lecture Note prepared by Dr Henry Otoo Page 71


The rank of A, denoted by rank A or by R (A), is defined as the number of linearly

independent vectors in the set v1 , v2 ,K , vn and equals the dimension of the vector space

spanned by those vectors.

Secondly, A second (equivalent) definition may be given of the rank of a matrix and uses the
concept of sub-matrices. A sub-matrix of A is any matrix that can be formed from the elements
of A by ignoring one, or more than one, row or column. It may be shown that the rank of a
general m´ n matrix is equal to the size of the largest square sub-matrix of A whose
determinant is non-zero. Therefore, if a matrix A has an r ´ r sub-matrix S with S ¹ 0 , but

no (r + 1)´ (r + 1) sub-matrix with non-zero determinant then the rank of the matrix is r. From

either definition it is clear that the rank of A is less than or equal to the smaller of m and n.

Example. Find the rank of each of the following matrices:

é1 1 0 - 1ù é1 3 - 1ù
ê ú ê ú
(a) A = êê2 0 2 2 úú and (b) B = ê8 9 4 ú
ê ú
ê4 1 3 1 ú ê2 1 2 ú
ë û ë û

Solution

1 1 0 1 1 - 2 1 0 - 2 1 0 - 2
2 0 2= 2 0 2 = 2 2 2 = 0 2 2 = 0.
4 1 3 4 1 1 4 3 1 1 3 1

This implies the rank of A is not 3. The next largest square sub-matrices of A are of dimension
2´ 2 . Consider, for example, the 2´ 2 sub-matrix formed by ignoring the third row and the
1 1
third and fourth columns of A; this has determinant = (1´ 0)­ (2 ´ 1) = ­ 2. Thus, A is
2 0

of rank 2 and we need not consider any other 2´ 2 sub-matrices.

2.9 Systems of Linear Equations: Gaussian Elimination

Lecture Note prepared by Dr Henry Otoo Page 72


Definition: The equation ax  by  cz  dw  h where a, b, c, d and h are known numbers,
while x, y, z and w are unknown numbers, is called a linear equation. If h  0 , the linear
equation is said to be homogeneous. A linear system is a set of linear equations and a
homogeneous linear system is a set of homogeneous linear equations.

For example,

x yz  1

 x  3 y  3z   2

are linear systems, while

2 x  3 y 2  1

x  y  z  2

is a nonlinear system (because of y2). The system is a non homogeneous linear system.

Remark. In more the general case in which m = n but det A = 0 , the inverse does not exist

and so any procedure using A- 1 would not work. In such circumstances, we consider more
carefully what solution means. Generally, when a solution vector x exists whose elements
simultaneously satisfy all the equations in the system – the equation will be said to be
consistent. Otherwise it is inconsistent

2.9.1 Matrix Representation of a Linear System

Matrices are helpful in rewriting a linear system in a very simple form. The algebraic properties
of matrices may then be used to solve systems. First, consider the linear system

 ax  by  cz  dw  e
 fx  gy  hz  iw  j

 .
kx  ly  mz  nw  p
 qx  ry  sz  tw  u

Lecture Note prepared by Dr Henry Otoo Page 73


Set the matrices
a b c d e  x
     
f g h i j y
A , C  , and X   
k l m n  p z
     
q r s t u  w
Using matrix multiplications, we can rewrite the linear system above as the matrix equation

A X  C

The matrix A is called the matrix coefficient of the linear system. The matrix C is called the
nonhomogeneous term. When C  0 , the linear system is homogeneous. The matrix X is the
unknown matrix. Its entries are the unknowns of the linear system. The augmented matrix
associated with the system is the matrix  A | C  , where

a b c d e
 
f g h i j
 A | C   
k l m n p
 
q r s t u 

In general if the linear system has n equations with m unknowns, then the matrix coefficient
will be a n  m matrix and the augmented matrix an n   m  1 matrix. Now we turn our

attention to the solutions of a system.

Definition: Two linear systems with n unknowns are said to be equivalent if and only if they
have the same set of solutions.

2.9.2 Gaussian Elimination. Consider a linear system.

1. Construct the augmented matrix for the system;


2. Use elementary row operations to transform the augmented matrix into a triangular
one;

Lecture Note prepared by Dr Henry Otoo Page 74


3. Write down the new linear system for which the triangular matrix is the associated
augmented matrix;
4. Solve the new system. You may need to assign some parametric values to some
unknowns, and then apply the method of back substitution to solve the new system.

Elementary Row Operations

1. Interchange two rows.


2. Multiply a row with a nonzero number.
3. Add a row to another one multiplied by a number.

ie. DEFINITION (Forward/Gauss Elimination Method) Gaussian elimination is a method of


solving a linear system Ax  b (consisting of m equations in n unknowns) by bringing the
augmented matrix

 a11 a12 a1n b1 


 
a a a2 n b2 
 A | b   21 22 
 
 am1 am 2 amn bm 

to an upper triangular form

c11 c12 c1n d1 


 
 0 c22 c2 n d2 
 
 
 0 0 cmn d m 

This elimination process is also called the forward elimination method.

For example, consider the matrix equation

Lecture Note prepared by Dr Henry Otoo Page 75


 9 3 4  x1   7 
    
 4 3 4  x2    8 
 1 1 1  x   3 
  3   

In augmented form, this becomes

9 3 4 7   x1 
  
 4 3 4 8   x2 
1 1 1 3   x3 

Switching the first and third rows (without switching the elements in the right-hand column
vector) gives

1 1 1 3   x1 
  
 4 3 4 8   x2 
9 3 4 7   x3 

Subtracting 9 times the first row from the third row gives

1 1 1 3   x1 
 
4 3 4 8   x2 
0 6 5 20   x3 

Subtracting 4 times the first row from the second row gives

1 1 1 3   x1 
  
0 1 0 4   x2 
0 6 5 20   x3 

Finally, adding -6 times the second row to the third row gives

Lecture Note prepared by Dr Henry Otoo Page 76


1 1 1 3   x1 
  
0 1 0 4   x2 
0 0 5 4   x3 

Restoring the transformed matrix equation gives

1 1 1   x1   3 
0 1 0   x    4
  2  
0 0 5   x3   4 

4
which can be solved immediately to give x3  , back-substituting to obtain x2  4 (which
5
1
actually follows trivially in this example), and then again back-substituting to find x1   .
5

EXAMPLE: Solve the linear system by Gaussian elimination method.

y  z 2
2x  3z  5
x  y  z 3

Solution:

In this case, the augmented matrix is

0 1 1 2
 
2 0 3 5
1 1 1 3 

The method proceeds along the following steps.

1. Interchange 1st and 3rd equation.

Lecture Note prepared by Dr Henry Otoo Page 77


x  y  z 3 1 1 1 3
 
2x z 5 2 0 3 5
y  z 3 0 1 1 2 

2. Add -2 times the 1st equation to the 2nd equation.

x yz  3 1 1 1 3
 
 2 y  z  1 0 2 1 1
yz  2 0 1 1 2 

3. Add 2 times the 3rd equation to the 2nd equation.

x yz  3 1 1 1 3
 
 2 y  z  1 0 2 1 1
3z  3 0 0 3 3 

4. Divide 2nd equation through by -2 and 3rd equation through by 3.

x yz  3 1 1 1 3
 
y  12 z  1
2 0 1  2
1 1
2

z  1 0 0 1 1 

The last equation gives z  1 the second equation now gives y  1 . Finally the first equation

gives x  1 . Hence the set of solutions is  x, y, z   1,1,1 , A UNIQUE SOLUTION.


t t

Definition. A linear system is called inconsistent or overdetermined if it does not have a


solution. In other words, the set of solutions is empty. Otherwise the linear system is called
consistent.

2.9.3 Eigenvalues and Eigenvectors

Lecture Note prepared by Dr Henry Otoo Page 78


The eigenvalue or characteristic value and its corresponding eigenvector or characteristic
vector of an N  N matrix A are defined as a scalar λ and a nonzero vector v satisfying
Av  v   A   I n  v  0  v  0  
where   , v  is called an eigenpair and there are N eigenpairs for the N  N matrix A.

How do we get them? Noting that


 in order for the above equation to hold for any nonzero vector v, the matrix  A   I n 

should be singular – that is, its determinant should be zero  A  I n  0  and

 the determinant of the matrix is a polynomial of degree N in terms of λ,


we first must find the eigenvalue i ' s by solving the so-called characteristic equation

A   I n   N  aN 1 N 1   a1  a0  0  

and then substitute the i ' s , one by one, into Eq.   to solve it for the eigenvector vi ' s .

 Computation of Eigenvalues

For a square matrix A of order n, the number  is an eigenvalue if and only if there exists a
non-zero vector x such that

Ax   x
Using the matrix multiplication properties, we obtain

 A   In  x  0
This is a linear system for which the matrix coefficient is A   I n . We also know that this

system has one solution if and only if the matrix coefficient is invertible, i.e. det  A   I n   0

. Since the zero-vector is a solution and x is not the zero vector, then we must have
det  A   I n   0 .

Lecture Note prepared by Dr Henry Otoo Page 79


In general, for a square matrix A of order n, the equation det  A   I n   0 will give the

eigenvalues of A. This equation is called the characteristic equation or characteristic


polynomial of A. It is a polynomial function in  of degree n. So we know that this equation
will not have more than n roots or solutions. So a square matrix A of order n will not have
more than n eigenvalues.

Example. Consider the matrix

 1 2 
A .
 2 0 

The equation det  A   I n   0 translates into

1  2
 1    0     4  0
2 0

which is equivalent to the quadratic equation

2    4  0

Solving this equation leads to

1  17 1  17
 , and  
2 2
In other words, the matrix A has only two eigenvalues.

Example. Find the eigenvalues/eigenvectors of the matrix

0 1 
A 
0 1
Lecture Note prepared by Dr Henry Otoo Page 80
First, we find its eigenvalues as

  1 
A  I     2    0
 0 1   
    1  0, 1  0, 2  1
and then, get the corresponding eigenvectors as

0 1   v11   v21  0 
A  1I v1        
0 1 v21   v21  0
    1  0, 1  0, 2  1
where we have chosen v11, v12, and v22 so that the norms of the eigenvectors
become one.

Example. Consider the diagonal matrix

a 0 0 0
 
0 b 0 0
D
0 0 c 0
 
0 0 0 d

Its characteristic polynomial is


a 0 0 0
0 b 0 0
det  D   I n     a    b    c    d     0
0 0 c 0
0 0 0 d 

So the eigenvalues of D are a , b , c and d , i.e. the entries on the diagonal. .

We have some properties of the eigenvalues of a matrix.

Theorem. Let A be a square matrix of order n. If is an eigenvalue of A, then:

Lecture Note prepared by Dr Henry Otoo Page 81


1.  m is an eigenvalue of Am , for m  1, 2,
1
2. If A is invertible, then is an eigenvalue of A1 .

3. A is not invertible if and only if   0 is an eigenvalue of A.
4. If  is any number, then    is an eigenvalue of A   I n .

5. If A and B are similar, then they have the same characteristic polynomial (which
implies they also have the same eigenvalues).
 Computation of Eigenvectors

Let A be a square matrix of order n and  one of its eigenvalues. Let x be an eigenvector of
A associated to  . We must have

Ax   x or  A  In  x  0 .
This is a linear system for which the matrix coefficient is A   I n . Since the zero-vector is a

solution, the system is consistent.

Remark. It is quite easy to notice that if x is a vector which satisfies Ax   x , then the vector
y  cx (for any arbitrary number c) satisfies the same equation, i.e. Ay   y . In other words,
if we know that x is an eigenvector, then cx is also an eigenvector associated to the same
eigenvalue.

Let us start with an example.

Example: Find the eigenvectors and eigenvalues of the linear transformation u  Ax which
has the component form

u1  3x1  4 x2 ,
u2  5 x1  2 x2 .

Solution: Solving the characteristic equation

Lecture Note prepared by Dr Henry Otoo Page 82


3 4
0
5 2

or

 2  5  14  0

we find the eigenvalues

1  2 2  7.

For 1  2 the eigenvector can be determined from the system

5 x1  4 x2  0,
5 x1  4 x2  0,

which implies

x1 4
 .
x2 5

Similarly, for 2  7 we have

4 x1  4 x2  0,
5 x1  5 x2  0,

which implies

x1
 1.
x2

Thus the eigenvectors of the transformation A are

a1   4 , 5 , a2   1 , 1  ,

Lecture Note prepared by Dr Henry Otoo Page 83


and all vectors collinear with a1 and a2 .

Remark. In general, the eigenvalues of a matrix are not all distinct from each other (they can
be equal root/repeated root or complex root). In the next two examples, we discuss this
problem.

 1 4 
Example. Consider the matrix  .
 4 7 

The characteristic equation is given by


1  4
    3  0
2

4 7  
Hence the matrix A has one eigenvalue, i.e. -3. Let us find the associated eigenvectors. These
are given by the linear system
AX   3 X or  A  3I 2  X  0

which may be rewritten by


 4x  4 y  0

 4x  4 y  0
This system is equivalent to the one equation-system x  y  0 .

So if we set x = c, then any eigenvector X of A associated to the eigenvalue -3 is given by


 x   1
X     c   .-
 y   1

Let us summarize what we did in the above examples.

Summary: Let A be a square matrix. Assume  is an eigenvalue of A. In order to find the


associated eigenvectors, we do the following steps:

1. Write down the associated linear system

Lecture Note prepared by Dr Henry Otoo Page 84


AX   X or  A  In  X  0
2. Solve the system.
3. Rewrite the unknown vector X as a linear combination of known vectors.

The above examples assume that the eigenvalue  is real number. So one may wonder
whether any eigenvalue is always real. In general, this is not the case except for symmetric
matrices. The proof of this is very complicated. For square matrices of order 2, the proof is
quite easy. Let us give it here for the sake of being little complete.

Consider the symmetric square matrix

a b
A 
b c

Its characteristic equation is given by

a b
det  A   I 2     2   a  c    ac  b 2  0
b c

This is a quadratic equation. The nature of its roots (which are the eigenvalues of A) depends
on the sign of the discriminant

   a  c   4  ac  b2  .
2

Using algebraic manipulations, we get

   a  c   4b2
2

Therefore, is a positive number which implies that the eigenvalues of A are real numbers.

Remark. Note that the matrix A will have one eigenvalue, i.e. one double root, if and only if
  0 . But this is possible only if a  c and b  0 . In other words, we have A  aI 2 .

Lecture Note prepared by Dr Henry Otoo Page 85


 The Case of Complex Eigenvalues

First let us convince ourselves that there exist matrices with complex eigenvalues.

Example. Consider the matrix

 3 2 
A 
 4 1 
The characteristic equation is given by

3 2
  2  2  5  0 .
4 1  

This quadratic equation has complex roots given by

2  i 16
  1  2i
2

Therefore the matrix A has only complex eigenvalues.

The trick is to treat the complex eigenvalue as a real one. Meaning we deal with it as a number
and do the normal calculations for the eigenvectors. Let us see how it works on the above
example.

We will do the calculations for   1  2i . The associated eigenvectors are given by the linear
system

AX  1  2i  X

which may be rewritten as



  2  2i  x  2 y  0


 4 x   2  2i  y  0

Lecture Note prepared by Dr Henry Otoo Page 86


In fact the two equations are identical since  2  2i  2  2i   8 . So the system reduces to one

equation
1  i  x  y  0
Set x  c , then y  1  i  c . Therefore, we have

 x  c   1 
X     c 
 y   c 1  i    1  i  
where c is an arbitrary number.

Remark. It is clear that one should expect to have complex entries in the eigenvectors.

We have seen that 1  2i  is also an eigenvalue of the above matrix. Since the entries of the

matrix A are real, then one may easily show that if  is a complex eigenvalue, then its

conjugate  is also an eigenvalue. Moreover, if X is an eigenvector of A associated to  ,

then the vector X , obtained from X by taking the complex-conjugate of the entries of X, is

an eigenvector associated to  . So the eigenvectors of the above matrix A associated to the


eigenvalue 1  2i  are given by

 1 
X  c 
 1  i  
where c is an arbitrary number.

Lecture Note prepared by Dr Henry Otoo Page 87


CHAPTER THREE
VECTORS
Objectives
At the end of this chapter you will be able to:
 Vector elements or Components in a coordinate frame.
 Vector magnitude and unit vectors.
 Addition, Subtraction and Multiplication of vectors.
 Geometrical interpretation of scalar product.
 Projection of one vector onto the other.
 Geometrical interpretation of vector product.
 Geometrical interpretation of scalar triple product.

Lecture Note prepared by Dr Henry Otoo Page 88


3.0 Introduction

A vector is a geometric entity characterized by a magnitude and a direction. A vector possesses


a definite initial point and terminal point. Such a vector is called a bound vector. In other
situations, when only the magnitude and direction of the vector matter, then the particular
initial point is of no importance, and the vector is called a free vector. Thus two arrows AB
and AB in space represent the same free vector if they have the same magnitude and
direction: equivalently, they are equivalent if the quadrilateral ABB′A′ is a parallelogram. If
the Euclidean space is equipped with a choice of origin, then a free vector is equivalent to the
bound vector of the same magnitude and direction whose initial point is the origin.

3.1 Vector elements or components in a coordinate frame

A method of representing a vector is to list the values of its elements or components in a


sufficient number of different (preferably mutually perpendicular) directions, depending on
the dimension of the vector. These specified directions define a coordinate frame. In this
course we will mostly restrict our attention to the 3-dimensional Cartesian coordinate frame
O  x, y, z  . When we come to examine vector fields later in the course you will use curvilinear

coordinate frames, especially 3D spherical and cylindrical polars, and 2D plane polar,
coordinate systems.

Figure 1: Vector components

Lecture Note prepared by Dr Henry Otoo Page 89


In a Cartesian coordinate frame we write
a   a1, a2 , a3    x2  x1, y2  y1, z2  z1  or a  ax , a y , az 

as sketched in Figure 1. Defining iˆ, ˆj, kˆ as unit vectors in the x, y, z directions.

iˆ  1,0,0 ˆj  0,1,0 kˆ  0,0,1

we could also write


a  a1iˆ  a2 ˆj  a3kˆ .

3.2 Basic properties

The following section uses the Cartesian coordinate system with basis vectors

e1  1,0,0  , e2   0,1,0  , e3   0,0,1

and assume that all vectors have the origin as a common base point. A vector a will be written
as

a  a1e1  a2e2  a3e3

 Negative vector

Since PQ  QP  0 , you can write QP   PQ. That is, QP is the Negative of vector PQ . So

the vector PQ has the same magnitude as the vector PQ but its direction is exactly opposite

to that PQ .

 Vector equality

Two vectors are said to be equal if they have the same magnitude and direction. Equivalently
they will be equal if their coordinates are equal. So two vectors a  a1e1  a2e2  a3e3 and

b  b1e1  b2e2  b3e3 are equal if a1  b1, a2  b2 , a3  b3 .

Lecture Note prepared by Dr Henry Otoo Page 90


 Parallel Vectors

From the above you should be able to see that if two vectors a and b are parallel then one is
scalar multiple of the other, that is: a  b

If  is positive then a is in the same direction as b. if  is negative then a is in the opposite


direction to b.

 Vector magnitude and unit vectors


Provided we use an orthogonal coordinate system, the magnitude of a 3-vector is

a  a  a12  a22  a32

a
To find the unit vector in the direction of a, simply divide by its magnitude aˆ  .
a

For example
If A  3xˆ  4 yˆ

then A  32  42  9  16  25  5

 Scalar multiplication

Scalar multiplication of a vector by a factor of 3 stretches the vector out.

Lecture Note prepared by Dr Henry Otoo Page 91


The scalar multiplications 2a and −a of a vector a

A vector may also be multiplied, or re-scaled, by a real number r. In the context of


conventional vector algebra, these real numbers are often called scalars (from scale) to
distinguish them from vectors. The operation of multiplying a vector by a scalar is called scalar
multiplication. The resulting vector is

ra   ra1  e1   ra2  e2   ra3  e3

Intuitively, multiplying by a scalar r stretches a vector out by a factor of r. Geometrically, this


can be visualized (at least in the case when r is an integer) as placing r copies of the vector in
a line where the endpoint of one vector is the initial point of the next vector.

If r is negative, then the vector changes direction: it flips around by an angle of 180°. Two
examples  r  1 and r  2  are given below:

Scalar multiplication is distributive over vector addition in the following sense:


r  a  b   ra  rb for all vectors a and b and all scalars r. One can also show that

a  b  a   1 b .

3.3 Vector Addition and Subtraction

Lecture Note prepared by Dr Henry Otoo Page 92


Vectors are added/subtracted by adding/subtracting corresponding components, exactly as
for matrices. Thus
a  b   a1  b1 , a2  b2 , a3  b3 

Addition follows the parallelogram construction of Figure 1.3(a). Subtraction  a  b is

defined as the addition  a   b   . It is useful to remember that the vector a  b goes from b

to a.

The following results follow immediately from the above definition of vector addition:
a) a  b  b  a  commutativity  Figure 1.3(a) 
b)  a  b   c  a  b  c   a  b  c  associativity  Figure 1.3(b) 
c) a  0  0  a  a, where the zero vector is 0  0,0,0 .

d) a   a   0

 Unit vector

A unit vector in a normed vector space is a vector (often a spatial vector) whose length is 1
(the unit length). A unit vector is often denoted by a lowercase letter with a superscribed caret
or “hat”, like this: (pronounced "i-hat").

Lecture Note prepared by Dr Henry Otoo Page 93


In Euclidean space, the dot product of two unit vectors is simply the cosine of the angle
between them. This follows from the formula for the dot product, since the lengths are both
1.

The normalized vector or versor u of a non-zero vector u is the unit vector co-directional with
u , i.e.,

u
u
u

where u is the norm (or length) of u . The term normalized vector is sometimes used as a

synonym for unit vector.

The normalization of a vector a into a unit vector a .

To normalize a vector a   a1 , a2 , a3  , scale the vector by the reciprocal of its length a . That

is:

a a a a
a  1 e1  2 e2  3 e3
a a a a

 Null vector

Lecture Note prepared by Dr Henry Otoo Page 94


The null vector (or zero vector) is the vector with length zero. Written out in coordinates, the

vector is  0,0,0  , and it is commonly denoted 0 , or 0, or simply 0. Unlike any other vector,

it does not have a direction, and cannot be normalized (that is, there is no unit vector which
is a multiple of the null vector). The sum of the null vector with any vector a is a (that is,
0  a  a ).

3.4 Multiplication of a vector by a scalar: (Not the scalar product!)


Just as for matrices, multiplication of a vector a by a scalar c is defined as multiplication of
each component by c , so that
ca  ca1, ca2 , ca3  .

It follows that:

ca   ca1    ca2    ca3   c a .


2 2 2

The direction of the vector will reverse if is negative, but otherwise is unaffected. (By the way,
a vector where the sign is uncertain is called a director)

Position Vectors: If you have a fixed origin O and a point A, then the vector OA is defined to

be the position vector of the point A. The line segment representing OA starts a O and ends

at A, so the vector OA uniquely defines the position of A.

Suppose you have two points A and B. the position vector of A is OA  a . The position vector

of B is OB  b . From the vector triangle, you can see that the vector AB is b  a . Likewise,

the vector BA is a  b .

3.5 Scalar, dot, or inner product


This is a product of two vectors results in a scalar quantity and is defined as follows for 3 –
component vectors:
a  b  a1b1  a2b2  a3b3 .

Note that

a  a  a12  a22  a32  a  a2 .


2

Lecture Note prepared by Dr Henry Otoo Page 95


The following laws of multiplication follow immediately from the definition:
a) a  b  b  a (commutativity).
b) a   b  c   a  b  a  c (distributivity with respect to vector addition).

c)  a   b    a  b   a   b  (scalar multiple of a scalar product of two vectors).

3.5.1 Geometrical interpretation of scalar product

Projection of

a  b  a  b  a  b
2

 a  a  b  b  2a  b
 a 2  b2  2  a  b 

But, by the cosine rule for the triangle OAB (figure 1.4a), the length AB 2 is given by

a b  a 2  b2  2ab cos
2

Where is the angle between the two vectors. It fellows that


a  b  ab cos ,

Which is independent of the co-ordinate system used, and that a  b  ab . Conversely, the

cosine of the angle between the two vectors a and b is given by cos  a  b ab .

Lecture Note prepared by Dr Henry Otoo Page 96


 a b 
  arc cos   .
a b 
The dot product of two vectors a   a1 , a2 , , an  and b  b1 , b2 , , bn  is defined as:

n
a  b   aibi  a1b1  a2b2   anbn
i 1

where  denotes summation notation and n is the dimension of the vectors.

For example, the dot product of two three-dimensional vectors 1 3 5 and  4 2 1

is

1 3 5   4 2 1  1 4   3 2    5 1  3

Example

Find the angle between the vectors a  2i  3 j  k and b  i  5 j  4k , giving your answer
to the nearest tenth of a degree.

a  b  a b cos where  , is the required angle

a   4  9  1  14

b  1  25  16  42

Also a  b   2i  3 j  k    i  5 j  4k 

 2  15  4
 13

So: 14 42 cos   13

13
cos    0.5361
588

Lecture Note prepared by Dr Henry Otoo Page 97


and:   122.4

3.6 Projection of one vector onto the other

Another way of describing the scalar product is as the product of the magnitude of one vector
and the component of the other in the direction of the first, since b cos is the component of
b in the direction of a and vice versa (Figure 1.4b)
Projection is particularly useful when the second vector is a unit vector – a  iˆ is the component

of a in the direction of iˆ .
Notice that if we wanted the vector component of b in the direction of a, we would write
b  a  a .
 b  aˆ  aˆ 
a2
In the particular case a  b  0 , the angle between the two vectors is a right angle and the
vectors are said to be mutually orthogonal or perpendicular – neither vector has any
component in the direction of the other.
An orthonormal coordinate system is characterized by iˆ  iˆ  ˆj  ˆj  kˆ  kˆ  1 ; and

iˆ  ˆj  ˆj  kˆ  kˆ  iˆ  0 .

 A scalar product is an “inner product”


So far we have been writing our vectors as row vectors a   a1 , a2 , a3  . This is convenient

because it takes up less room than writing column vectors

 a1 
a   a2 
 a3 

In matrix algebra vectors are more usually defined as column vectors, as in


 M 11 M 12 M 13   a1   v1 
M M 22 M 23   a   v 
 21  2  2
 M 31 M 32 M 33   a3   v3 

Lecture Note prepared by Dr Henry Otoo Page 98


and a row vector is written as aT . Now for most of our work we can be quite relaxed about
this minor difference, but here let us be fussy.
Why? Simply to point out at that scalar product is also the inner product more commonly
used in linear algebra. Defined as aT b when vectors are column vector as

 b1 
a  b  a b   a1 , a2 , a3  b2   a1b1  a2b2  a3b3
T

b3 

Here we treat a n – dimensional column vector as n  1 matrix.


(Remember that if you multiply two matrices M mn Nn p then M must have the same columns

as N has rows (here denoted by n) and the result has size (rows x columns) of m  p . So for

n – dimensional column vectors a and b, aT is a 1  n matrix and b is n  1 matrix, so the


product aT b is a 1  1 matrix, which is (at last!) a scalar.)

 Direction cosines use projection


Direction cosines are commonly used in the field of crystallography. The quantities

a  iˆ a  ˆj a  kˆ
 ,  , 
a a a
represent the cosines of the angle which the vector a makes with the co-ordinate vectors

iˆ, ˆj, kˆ and are known as the direction cosines of the vector a. Since a  iˆ  a1 etc, it follows

  1
immediately that a  a iˆ   ˆj   kˆ and  2   2   2  2  a12  a22  a32   1
a

Lecture Note prepared by Dr Henry Otoo Page 99


3.7 Vector or cross product
The cross product is a binary operation on two vectors in a three-dimensional Euclidean space
that results in another vector which is perpendicular to the plane containing the two input
vectors.
The vector product of two vectors a and b is denoted by a  b and is defined as follows:
a  b   a2b3  a3b2  iˆ   a3b1  a1b3  ˆj   a1b2  a2b1  kˆ .

It is much more easily remembered in terms of the pseudo – determinant

iˆ ˆj kˆ
a  b  a1 a2 a3
b1 b2 b3

where the top row consists of the vectors iˆ, ˆj, kˆ rather than scalars.
Since a determinant with two equal rows has value zero, it follows that a  a  0 . It is also
easily verified that  a  b   a   a  b   b  0 , so that a  b is orthogonal (perpendicular) to both

a and b, as shown in Figure 1.6.


Note that iˆ  ˆj  kˆ, ˆj  kˆ  iˆ , and kˆ  iˆ  ˆj .
The magnitude of the vector product can be obtained by showing that

a  b   a  b   a 2b2
2 2

from which it follows that

Lecture Note prepared by Dr Henry Otoo Page 100


a  b  ab sin  ,

which is again independent of the co-ordinate system used. This is left as an exercise. Unlike
the scalar product, the vector product does not satisfy commutativity but is in fact anti-
commutative, in that a  b  b  a . Moreover the vector product does not satisfy the
associative law of multiplication either since, as we shall see later a   b  c    a  b   c .

Since the vector product is known to be orthogonal to both the vectors which form the
product, it merely remains to specify its sense with respect to these vectors. Assuming that the
co-ordinate vectors form a right-handed set in the order i , j , k it can be seen that the sense
of the vector product is also right handed, i.e. the vector product has the same sense as the
co-ordinate system used.

iˆ ˆj kˆ
iˆ  ˆj  1 0 0  kˆ
0 1 0

In practice, figure out the direction from a right-handed screw twisted from the first to second
vector as shown in Figure below.

Lecture Note prepared by Dr Henry Otoo Page 101


Example

Simplify  i  2 j  5k    2i  j  3k  (a) directly (b) by evaluating an appropriate

determinant.

(a)  i  2 j  5k    2i  j  3k 

 2  i  i    i  j   3  i  k   4  j  i   2  j  j   6  j  k   10  k  i   5  k  j   15  k  k 
 0  k  3 j  4k  0  6i  10 j  5i  0
 i  7 j  3k

iˆ ˆj kˆ
1 2 5
2 1 3

 i  6  5  j  3  10   k 1  4 
 i  7 j  3k

3.7.1 Geometrical interpretation of vector product


The magnitude of the vector product  a  b  is equal to the area of the parallelogram whose

sides are parallel to, and have lengths equal to the magnitudes of, the vectors a and b (Figure
1.6b). Its direction is perpendicular to the parallelogram.
Example

g is vector from A1, 2, 3 to B 3, 4, 5 . ˆ is the unit vector in the direction from O to A. Find

m̂ , a UNIT vector along g  ˆ . Verify that m̂ is perpendicular to ˆ . Find n̂ , the third member

of a right – handed coordinate set ˆ , m


ˆ , nˆ .

Lecture Note prepared by Dr Henry Otoo Page 102


Answer
g  3, 4, 5  1, 2, 3   2, 2, 2
1
 1, 2, 3
14
iˆ ˆj kˆ
1 1
g  2 2 2   2,  4, 2
14 14
1 2 3

Hence
1 1
ˆ
m   2,  4, 2
14 24
and
1 1
ˆ
m   2,  4, 2
14 24

3.8 Scalar triple product


This is the scalar product of a vector product and a third vector, i.e a   b  c  . Using the

pseudo – determinant expression for the vector product, we see that the scalar triple product
can be represented as then true determinant
a1 a2 a3
a   b  c   b1 b2 b3
c1 c2 c3

This says that


(i) a   b  c   b   c  a   c   a  b   Called cylic permutation.

(ii) a   b  c   b   a  c  and so on.  Called anti  cylic permutation.

(iii) The fact that a   b  c    a  b   c allows the scalar triple product to be written as
 a, b, c. This notation is not very helpful, and we will try to avoid it below.

Lecture Note prepared by Dr Henry Otoo Page 103


3.8.1 Geometrical interpretation of scalar triple product
The scalar triple product gives the volume of the parallelepiped whose sides are represented
by the vectors a, b, and c.
We saw earlier that the vector product  a  b  has magnitude equal to the area of the base,

and direction perpendicular to the base. The component of c in this direction is equal to the
height of the parallelepiped shown in Figure 2.1(a)

Trial Questions
Let u   1,2,3 , v  2,1,2, and w  0,3,1 be vectors in R 3 . Find u * v * w, w * u * v  and
v * w * u  . What do you notice.

i j k
V *W  2 1  1
0 3 1
1 2 2 2 2 1
i j k
3 1 0 1 0 3
 i 1  6   j 2  0   k  6  0 
 i  5  j  2   k  6 
  5,2,6
U * V * W    2,2,3 *  5,2,6 
 5  4  18
 17

Lecture Note prepared by Dr Henry Otoo Page 104


i j k
u * v  1 2 3
2 1 2
 i  4  3  j 2  6   k  1  4 
 7i  4 j  5 k
  7,4,5
 w * u * v   0,3,1 *  7,4,5
 0  12  5
 17
i j k
w*u  0  3 1
1 2 3
 i  9  2   j 0  1  k 0  3
 11i  1 j  3k
  11,1,3
V * w * u   2,1,2  * 11,1,3
 22  1  6
 17

^ ^ ^
Example 1 Let a  i  2 j and b  2 i  j .Is a  b ? Are the vectors a and b equal?

Solution

a  12  2 2  5

b  2 2  12  5

ab

But , the two vectors are not equal since their corresponding components are distinct.
Example 2
^ ^ ^
Find unit vector in the direction of vector a  2 i  3 j  k

Lecture Note prepared by Dr Henry Otoo Page 105


^ 1
Solution the unit vector in relation of a vector a is given by a  a
a

a  2 2  3 2  12  14

^ 1  ^ ^ ^
 2 ^ 3 ^ 1 ^ ^ ^
Therefore a   2 i 3 j k   i j k a  i  2 j that has magnitude 7 units
14   14 14 14
^ ^
Example 3. Find a vector in the direction of vector a  i  2 j

Solution The unit vector in the direction of the given vector a is


^ 1 1 ^ ^
 1 ^ 2 ^
a a  i 2 j   i j
a 5  5 5

Therefore, the vector having magnitude equal to 7 and in the direction of a is


^  1 ^ 2 ^
7 a  7 i j 
 5 5 
 7 ^ 14 ^ 
  i j 
 5 5 
Example 4
Find the unit vector in the direction of the sum of the vectors,
^ ^ ^ ^ ^ ^
a  2 i  2 j  5 k and b  2 i j 3 k
Solution the sum of the given vectors is

 
a  b  c, say  4i  3 j  2k

And c  4 2  3 2   2  29
2

Thus, the required unit is


^ 1 1  ^ ^ ^
 4 ^ 3 ^ 2
c c  4 i 3 j 2 k   i j k
c 29   29 29 29

Example 5

Write the direction ratios of the vector a  i  j  2k and hence calculate its direction cosines.

Lecture Note prepared by Dr Henry Otoo Page 106


^ ^ ^
Solution Note that the direction ratios a, b, c of a vector r  x i  y j  z k are just the respective
components x, y and z of the vector. So the given vector, we have a  1, b  1 and c  2.
Further, if l, m and n are the direction cosines of the given vector, then
a 1 1 c 2
l  , m , n  as 6
r 6 6 r 6

 1 1  2
Thus, the direction cosines are  , , 
 6 6 6
^ ^ ^ ^ ^ ^
Example Find the angle  between the vectors a  i  j  k and b  i  j  k

Solution the angle  between two vectors a and b is given by

a *b
cos  
ab

Now a * b   i  j  k  *  i  j  k   1  1  1  1
^ ^ ^ ^ ^ ^

   
1
Therefore, we have cos  
3
1
Hence the required angel is   cos 1
3

^ ^ ^ ^ ^ ^
Example if a  5 i  j  3 k and b  i  3 j  5 k , then show that the vectors

a  b and a  b are perpendicular.


Solution We know that nonzero vectors are perpendicular if their scalar product is zero.
 ^ ^ ^
 ^ ^ ^
 ^ ^ ^
Here a  b   5 i  j  3 k    i  3 j  5 k   6i  2 j  8 k
   

 ^ ^ ^
 ^ ^ ^
 ^ ^ ^
And a  b   5 i  j  3 k    i  3 j  5 k   4i  4 j  2 k
   

  ^ ^

^
  ^
 
^ ^

So ( (a  b) * a  b   6 i  2 j  8 k  *  4 i  4 j  2 k   24  8  16  0

Lecture Note prepared by Dr Henry Otoo Page 107


Hence a  b and a  b are perpendicular vectors

^ ^ ^ ^ ^
Example Find the projection of the vector a  2 i ^  3 j  2 k on the vector b  i  2 j  k .

Solution The projection of the vector a is given by


1
a * b  2 *1  3 * 2  2 *1  106  53 6
b (1) 2  (2) 2  (1) 2

Example Find a  b , if two vectors a and b are such that a  2 , b  3 and a * b  4

2
ab  ab * ab  
 a *a  a *b  b*a  b*b

We have
2
 a  2 a *b  b  2

 2  24   3


2 2

ab  5

^ ^ ^ ^ ^
Example Find a  b , if a  2 i  j  3 k and b  3 i  5 j  2 k

We have
^ ^ ^
i j k
a *b  2 1 3
3 5 2

 i  2  15   4  9 j  (10  3)k


^

^ ^ ^
 17 i  13 j  7 k

 17 2  132  7 2
^ ^
a b 
Hence
 507
Example Find the area of a triangle having the points A1,1,1, B1,2,3 and C 2,3,1 and its
vertices.

Lecture Note prepared by Dr Henry Otoo Page 108


^ ^ ^ ^ 1
Solution we have AB  j  2 k and AC  i  2 j The area of the given triangle is AB * AC .
2
^
^ ^
i j k
NOW AB * AC  0 1 2
1 2 0

 AB * AC  16  4  1  21

1
The required area is 21
2

Example find the area of a parallelogram whose adjacent sides are given by the vectors
^ ^ ^ ^ ^ ^
a  3 i  j  4 k and b  i  j  k

Solution the area of a parallelogram with a and b as its adjacent sides given by a * b .

^
^ ^
i j k
^ ^ ^
a *b  3 1 4  5 i j 4 k
1 1 1

Now  a * b  25  1  16

 42

Hence the required area is 42

Lecture Note prepared by Dr Henry Otoo Page 109


REFERENCES:
1. J. Bird, (2006), Higher Engineering Mathematics, 5th Edition, Elsevier Ltd.
2. F. Ayres, Matrices. Schaum's Outline Series, McGraw-Hill, 1968. This is a book of
problems and solutions.
3. H. Anton, Elementary Linear Algebra, 4th edition. John Wiley, 1984.
4. D. T. Finkbeiner, Elements of Linear Algebra, 3rd edition. Freeman, 1978.
5. B. Kolman, Elementary Linear Algebra, 4th edition. Collier Macmillan, 1986.
6. I. Reiner, Introduction to Linear Algebra and Matrix Theory. Holt, Rinehart & Winston,
1971.
7. P. J. Kelly & E. G. Straus, Elements of Analytical Geometry. Scott Foresman, 1970.
8. J. H. Kindle, Plane and Solid Analytic Geometry. Schaum's Outline Series, McGraw-
Hill, 1950.
These two books, as their titles suggest, are about geometry rather than algebra, but
they may be useful as background and/or further reading for the more geometrical
aspects of this book.

Lecture Note prepared by Dr Henry Otoo Page 110

You might also like