MAT 212LectureNote
MAT 212LectureNote
LECTURE NOTES ON
LINEAR ALGEBRA
B. O. Onasanya
University of Ibadan.
Copyright © 2024 by B. O. Onasanya
All rights reserved.
Certification
I certify that .
i
Dedication
ii
Acknowledgements
We...
iii
Abstract
iv
TABLE OF CONTENTS
Title Page 1
Certification i
Dedication ii
Aknowledgement iii
Abstract iv
1 MATRICES AND OTHER CONCEPTS 7
1.1 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.1 What is a matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.2 Some Operations on Matrices . . . . . . . . . . . . . . . . . . . 8
1.1.3 Cofactors of a square matrix . . . . . . . . . . . . . . . . . . . . 9
1.2 Elementary operations on Matrices . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Elementary matrices . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Row reduced echelon matrix . . . . . . . . . . . . . . . . . . . . 10
2 PERMUTATION AND/OR SYMMETRIC GROUP 12
2.1 Permutation Group and Its Operation . . . . . . . . . . . . . . . . . . 12
2.2 Polynomial in Relation to Permutation . . . . . . . . . . . . . . . . . . 15
3 DETERMINANTS 16
3.1 Determinant of a square matrix . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Determinant in Relation to Symmetric Group . . . . . . . . . . . . . . 17
3.3 Some properties of determinants . . . . . . . . . . . . . . . . . . . . . . 18
3.4 Inverse of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4.1 Inverse by elementary operations . . . . . . . . . . . . . . . . . 21
4 SYSTEM OF LINEAR EQUATIONS 22
4.1 System of Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1.1 Describing a System of equations . . . . . . . . . . . . . . . . . 22
4.1.2 Augmented matrix . . . . . . . . . . . . . . . . . . . . . . . . . 23
5
4.2 Solution to System of Equations . . . . . . . . . . . . . . . . . . . . . . 24
4.2.1 Solving homogenous linear systems . . . . . . . . . . . . . . . . 24
4.2.2 Solving non-homogenous linear systems . . . . . . . . . . . . . . 25
5 EIGENVALUES AND EIGENVECTORS 29
5.1 Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2 Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6 CAYLEY-HAMILTON THEOREM AND MINIMAL POLYNOMIAL 34
6.1 Cayley-Hamilton Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 34
6.1.1 Minimal Polynomial . . . . . . . . . . . . . . . . . . . . . . . . 36
7 SIMILAR, DIAGONAL AND TRIANGULAR MATRICES 41
7.1 Similar Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
8 SIMILAR AND TRIANGULAR MATRICES 44
8.1 Similar Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
8.2 Triangular Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
8.2.1 Upper and lower triangular matrices . . . . . . . . . . . . . . . 45
8.3 Reducing a System to Triangular Matrix . . . . . . . . . . . . . . . . . 47
9 VECTOR SPACES AND LINEAR TRANSFORMATIONS 49
9.1 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
9.2 Some Results in Vector Space . . . . . . . . . . . . . . . . . . . . . . . 52
9.3 Linear Mappings/Transformations . . . . . . . . . . . . . . . . . . . . . 52
9.4 Change of Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
10 BILINEAR, QUADRATIC AND CANONICAL FORMS 60
10.1 Bilinear Form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
10.2 Orthogonal Diagonalization of a Real Symmetric Matrix . . . . . . . . 64
10.3 Quadratic Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
10.4 Canonical Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6
Chapter 1
1.1 Matrices
7
Remark 1.1.2 The square matrix
0, if i ̸= j,
B = (bij ) = (1.1.1)
α ∈ F, a field, for i = j
is diagonal matrix.
The n × n square matrix
0, if i ̸= j,
In = (aij ) = (1.1.2)
1, for i = j
is an identity matrix.
Remark 1.1.3 Note that GLn (IR) is the set of all n × n matrices over a field IR. For
any A, B, C ∈ GLn (IR)
(i) AB ∈ GLn (IR)
(ii) A(BC) = (AB)C
(iii) In is the identity element
(iv) A−1 is the inverse of A
(vi) AB ̸= BA
Our conclusion about the set GLn (IR) under matrices multiplication should not be dif-
ficult.
Also, consider Mn (IR) as the set of all n × n matrices over a field IR. For any
A, B, C ∈ Mn (IR)
8
(i) A + B ∈ Mn (IR)
(ii) A + (B + C) = (A + B) + C
(iii) O matrix is the identity element
(iv) −A is the inverse of A
(vi) A + B = B + A.
Our conclusion about the set Mn (IR) under matrices addition should not be difficult.
9
0 0 1
E13 = 0 1 0 and
1 0 0
2 0 0
E1 (2)E2 (3)E3 (4) = 0 3 0
4 0 4
Definition 1.2.1 A matrix is in its canonical or normal form if it is in its row reduced
echelon form and the leading entry in the first row is in its first column
Example 1.2.3 Find the row reduced echelon form of each of A and B and state if it
is in its canonical form.
10
1 0 2
(a) A = 2 −1 3
4 1 8
0 1 3
(b) B = 2 1 −4 .
2 3 2
−7
1 0 2
Solution: For A it is I3 and for B it is B = 0 1 3 .
0 0 0
Remark 1.2.4 The number of non zero rows in the row reduced echelon form of a
matrix is its rank. Rank(A) = 3 and Rank(B) = 2.
11
Chapter 2
PERMUTATION AND/OR
SYMMETRIC GROUP
Example 2.1.2 Let X = {1, 2, 3} in which case n = 3 and n! = 6. We can have the
set of permutations on X denoted by S3 as
1 2 3 1 2 3
ϕ= = = (132),
ϕ(1) ϕ(2) ϕ(3) 3 1 2
1 2 3 1 2 3
β= = = (123),
β(1) β(2) β(3) 2 3 1
1 2 3 1 2 3
ρ= = = (12),
ρ(1) ρ(2) ρ(3) 2 1 3
12
1 2 1 3 2 3
α= = = (13),
α(1) α(2) α(3) 3 2 1
1 2 3 1 2 3
λ= = = (23)
λ(1) λ(2) λ(3) 1 3 2
and
1 2 3
e= = (1).
1 2 3
S3 = {e, ρ, α, λ, β, ϕ} and |S3 | = 6 = 3!.
Remark 2.1.3 The permutation which does not change the elements of X is an identity
permutation.
1 2 3
e= .
1 2 3
13
◦ e ρ α λ β ϕ
e e ρ α λ β ϕ
ρ ρ e ϕ β λ α
α α β e ϕ ρ λ
λ λ ϕ β e α ρ
β β α λ ρ ϕ e
ϕ ϕ λ ρ α e β
Lemma 2.1.6 The set of all permutations on the set X under the composition of per-
mutations form a group.
We will illustrate this with an example.
Example 2.1.7 Let X = {1, 2, 3} in which case S3 permutations and 3! = 6. We can
have the permutations on X as S3 = {e, ρ, α, λ, β, ϕ} for
1 2 3 1 2 3 1 2 3 1 2 3
ϕ= β = ρ = α =
3 1 2 2 3 1 2 1 3 3 2 1
1 2 3
λ=
1 3 2
and
1 2 3
e= .
1 2 3
This can be shown in a Caley’s table.
Example 2.1.8 Let G = {e, ρ, α, λ, β, ϕ} and define an operation on it by the table
below. This group is not commutative.
We can compute the composition of the permutations as:
1 2 3 1 2 3 1 2 3
= .
3 1 2 2 3 1 1 2 3
1 2 3 1 2 3 1 2 3
= .
2 3 1 1 3 2 2 1 3
14
2.2 Polynomial in Relation to Permutation
Let Y
Pn (x1 x2 x3 · · · xn ) = (xi − xj ) (2.2.1)
i>j
and δ ∈ Sn ,
(ii) If δ = (132),
Remark 2.2.2 δ̂(Pn ) is either Pn (in which case δ is said to be even) or −Pn (in which
case δ is said to be odd). When the permutation δ is even, we write sign(δ) = +1 and
when odd, sign(δ) = −1. The following properties are obvious:
(i) sign(δ1 δ2 ) = sign(δ1 )sign(δ2 )
(ii) sign(δ −1 ) = sign(δ)
(i) sign(12) = −1
(ii) sign(123) = 1
Example 2.2.3
15
Chapter 3
DETERMINANTS
Example 3.1.3
−2 −2 3
M = 1 1 −1
−1 −2 2
16
|M | = (−1)1+1 (−2)M11 + (−1)1+2 (−2)M12 + (−1)1+3 (3)M13
= −2M11 + 2M12 + 3M13
= −2(0) + 2(1) + 3(−1) .
= 0+2−3
= −1
17
det(M ) = (3)(1)(−2)
+ (−2)(1)(2)
− (−2)(1)(2)
− (3)(1)(−1)
− (−2)(−1)(−2)
+ (−2)(−1)(−1)
Therefore,
det(M ) = −6 − 4 + 4 + 3 + 4 − 2 = −1.
Example 3.2.2 Show that if A is a n×n upper triangular matrix, det(A) = a11 a22 a33 · · · ann .
Solution:
If δ = (1), δ(i) = i and aiδ(i) = aii . If δ ̸= (1), we have i > δ(i) for some 1 ≤ i ≤ n and,
in such case, aiδ(i) = 0. Hence,
and X
det(B) = (signδ)a1δ(1) a2δ(2) · · · λarδ(r) · · · anδ(n) .
δ∈Sn
Then,
det(B) = λdet(A).
18
Consequently,
i. det(λA) = λn det(A)
ii. If one row in A is zero, det(A) = 0 since
X
det(A) = (signδ)a1δ(1) a2δ(2) · · · 0arδ(r) · · · anδ(n) = 0det(A) = 0.
δ∈Sn
5. det(AB) = det(A)det(B)
6. det(AT ) = det(A)
7. Interchanging rows and columns of a determinant does not alter its value.
8. If two rows of A are identical, det(A) = 0.
19
3 −1
Example 3.4.1 A = , |A| = 8,
2 2
2 −2
C(A) = ,
1 3
2 1
A∗ = ,
−2 3
∗ ∗ 2 1
A A 1
A−1 = = =
|A| 8 8 −2 3
Example 3.4.2
8 −1 −3
B = −5 1 ,
2
10 −1 −4
|B| = −1,
−2 0 −5
C(B) = −1 −2 −2 ,
1 −1 3
−2 −1 1
B ∗ = 0 −2 −1 ,
−5 −2 3
2 1 −1
∗ ∗
B B
B −1 = = = 0 2 1
−1 |B|
5 2 −3
20
3.4.1 Inverse by elementary operations
Theorem 3.4.3 If a square matrix A can be reduced to an identity matrix by a sequence
of elementary row operations, then it is invertible. Hence, its inverse can be obtained
by performing the same sequence of elementary row operations on the identity matrix
of the same dimension.
Perform elementary operations on the two sides until matrix A on the left is reduced to
an identity matrix. Whatever matrix the identity matrix on the right transforms into
is now the inverse of A.
A I
A−1 A A−1 I
I A−1
21
Chapter 4
SYSTEM OF LINEAR
EQUATIONS
22
AX = B.
A is matrix of coefficient. If B = 0, the system is said to be homogeneous. Otherwise,
it is non-homogeneous.
Example 4.1.1
x1 + 6x2 + 2x3 = 1
−4x1 + x3 = 6
2x1 + 5x2 − 3x3 = 0
has
1 6 2 x1 1
−4 0 1 x2 = 6
2 5 −3 x3 0
1 6 2 1
with augmented matrix −4 0 1 6 is non-homogeneous.
2 5 −3 0
Example 4.1.2
5x1 − x2 + 2x3 = 0
x1 + 2x2 + 3x3 = 0
−x1 + 6x2 = 0
23
has
5 −1 2 x 0
1
2 3 x2 = 0
1
−1 6 0 x3 0
5 −1 2 0
with augmented matrix 1 2 3 0 is homogeneous.
−1 6 0 0
x1 + 2x3 = 0
2x1 − x2 + 3x3 = 0
4x1 + x2 + 8x3 = 0
has
1 0 2
A = 2 −1 3
4 1 8
and its row reduced form
1 0 0
AR = 0 1 0 .
0 0 1
24
x 0
1
Hence, x2 = 0
x3 0
Example 4.2.2 The linear system
x1 + 2x2 + 8x3 = 0
x1 − 3x2 − 7x3 = 0
−x1 + x2 + x3 = 0
has
1 2 8
A = 1 −3 −7
−1 1 1
and its row reduced form
1 0 2
AR = 0 1 3 .
0 0 0
x1 + 2x3 = 0 ⇒ x1 = −2x3
x2 + 3x3 = 0 ⇒ x2 = −3x3
Hence,
x −2x3 −2
1
x2 = −3x3 = λ −3 ,
x3 x3 1
when x3 = λ.
25
Example 4.2.3 The linear system
x1 + 3x2 + x3 = 0
2x1 − x2 − x3 = 1
x1 − 4x2 − 2x3 = 2
has
1 3 1 0
[A|B] = 2 −1 −1 1
1 −4 −2 2
and its row reduced form
3
1 0 −2 7
[A|B]R = 0 1 73 − 17 .
1
0 0 0 −7
x1 + 2x2 + 8x3 = 1
x1 − 3x2 − 7x3 = 6
−x1 + x2 + x3 = −4
has
1 2 8 1
[A|B] = 1 −3 −7 6
−1 1 1 −4
and its row reduced form
1 0 2 3
[A|B]R = 0 1 3 −1 .
0 0 0 0
26
Hence, r = k = 2 < n = 3 implies that the system has infinitely many solutions.
x1 + 2x3 = 3 ⇒ x1 = −2x3 + 3
x2 + 3x3 = −1 ⇒ x2 = −3x3 − 1
Hence,
x1 −2x3 + 3 −2x3 3
x2 = −3x3 − 1 = −3x3 + −1 .
x3 x3 x3 0
When x3 = λ.
−2 3
X=λ −3 + −1
1 0
| {z } | {z }
Gen. Sol. of AX=0 P articular sol. of AX=B
x+y−z = 2
x + 2y + z = 3
x + y + (α2 − 2)z = 2α
to have
(i) no solution
(ii) unique solution
(iii) infinitely many solution
27
Example 4.2.7 Check whether or not the system
2x + 4y − 2z = 4
−x + 4y + 3z = 2
x − y + (α4 − 18)z = 2α
to have
(i) no solution
(ii) unique solution
(iii) infinitely many solution
28
Chapter 5
EIGENVALUES AND
EIGENVECTORS
5.1 Eigenvalues
Let A be × n matrix, I an identity matrix of the same dimension and x is a column
an n
x1
x2
matrix
..
. Then,
.
xn
Ax = λx (5.1.1)
which implies
(A − λI)x = 0 (5.1.2)
corresponds to the system of n homogeneous linear equations. It is easy to see that
x = 0 is a trivial solution of Eqn (5.1.2). However, if
n
X
CA (λ) = ai λi = |A − λI| = 0, (5.1.3)
i=0
then, Eqn (5.1.2) has other non-trivial solutions. This implies that if A − λI is singular,
Eqn (5.1.2) has non-trivial solution(s). The polynomial obtained from Eqn (5.1.3) is
29
called Characteristic Polynomial. If Eqn(5.1.3) is of degree n (that is A is of dimension
n × n), then there are n eigenvalues or characteristic roots λ1 , λ2 , λ3 , · · · λn .
3 1
Example 5.1.1 Find the eigenvalue of A = .
2 2
λ 0
λIn = .
0 λ
3−λ 1
|A − λI2 | = = (3 − λ)(2 − λ) − 2 = 0
2 2−λ
Solve (λ − 4)(λ − 1) = 0 ⇒ λ1 = 4, λ2 = 1.
3 −4
Example 5.1.2 Find the eigenvalue of A = .
2 −6
λ 0
λIn = .
0 λ
3−λ −4
|A − λI2 | = = (3 − λ)(−6 − λ) + 8 = 0
2 −6 − λ
Solve (λ − 2)(λ + 5) = 0 ⇒ λ1 = 2, λ2 = −5.
2 1 1
Example 5.1.3 Find the eigenvalue of A = 1 2 1 .
1 1 2
λ 0 0
λIn = 0 λ 0 .
0 0 λ
30
2−λ 1 1
|A − λI3 | = 1 2−λ 1
1 1 2−λ
= (2 − λ)[(2 − λ)2 − 1] − 1[(2 − λ) − 1] + 1[1 − (2 − λ)] = 0
Solve (λ − 1)2 (λ − 4) = 0 ⇒ λ1 = 1, λ2 = 1, λ3 = 4.
5.2 Eigenvectors
To find eigenvectors, we solve for each case of λi the equation
(A − λi In )x = 0. (5.2.1)
The set of all these eigenvectors of A is called the eigenspace or row null space of A.
4 1
Example 5.2.1 Let A = . λ1 = 1, λ2 = 5.
3 2
Case λ1
4−1 1 x 0 3 1 x
1 = = 1
3 2−1 x2 0 3 1 x2
3x1 + x2 = 0 ⇒ x2 = −3x1 ,
x1 x1 1
= = x1 = ν1
x2 −3x1 −3
1
is the eigenvector corresponding to λ1 = 1. Simply . Case λ2 = 5
−3
4−5 1 x 0 −1 1 x
1 = = 1
3 2−5 x2 0 3 −3 x2
−x1 + x2 = 0 ⇒ x2 = x1 ,
x x 1 1
1 = 1 = x 1 = x 2 = ν2
x2 x1 1 1
31
1
is the eigenvector corresponding to λ2 = 5. Simply .
1
Example 5.2.2 Let
2 0 1
A = −1 4 −1 ,
−1 2 0
λ1 = 1, λ2 = 2, λ3 = 3.
Case λ1 = 1
1 0 1 x 0
1
(A − λi In )X = −1 3 −1 x2 = 0
−1 2 −1 x3 0
x1 + x3 = 0 ⇒ x1 = −x3 ,
− x1 + 3x2 − x3 = 0 ⇒ −x1 + 3x2 − (−x1 ) = 0 ⇒ x2 = 0.
x x −x3 −1
1 1
ν1 = x2 = 0 = 0 = x3 0
x3 x3 x3 1
or
x x 1 1
1 1
ν1 = x 2 = 0 = x1 0 = 0 .
x3 −x1 −1 −1
Case λ2 = 2
0 0 1 x1 0
(A − λi In )X = −1 2 −1 x2 = 0
−1 2 −2 x3 0
x3 = 0, −x1 + 2x2 − x3 = 0 ⇒ −x1 + 2x2 = 0 ⇒ x1 = 2x2 .
x1 2x2 2
ν2 = x 2 = x 2 = x 2 1 .
x3 0 0
32
Case λ3 = 3
−1 0 1 x 0
1
(A − λi In )X = −1 1 −1 x2 = 0
−1 2 −3 x3 0
−x1 + x3 = 0 ⇒ x1 = x3 ,
− x1 + x2 − x3 = 0 ⇒ −x1 + x2 − x1 = 0 ⇒ −2x1 + x2 = 0 ⇒ 2x1 = x2 .
x x 1
1 1
ν3 = x2 = 2x1 = x1 2 .
x3 x1 1
33
Chapter 6
CAYLEY-HAMILTON THEOREM
AND MINIMAL POLYNOMIAL
Also, let
n−1
X
J(λ) = Ji λi ,
i=0
where J(λ) is the adjoint of A − λI (i.e. (A − λI)∗ ) and Ji are n × n matrices with
entries independent of λ. By Theorem 6.1.1,
(A − λI)J(λ) = |A − λI|I
34
so that
(A − λI)(Jn−1 λn−1 + · · · + J1 λ + J0 ) = (an λn + an−1 λn−1 + · · · + a1 λ + a0 )I.
35
2 1 1
Example 6.1.5 Let A = 1 2 1 .
1 1 2
2−λ 1 1
|A − λI3 | = 1 2−λ 1
1 1 2−λ
Remark 6.1.6 Note that if A is a diagonal matrix with the diagonal entries d1 , d2 , · · · , dn ,
then
|A − λI| = (d1 − λ)(d2 − λ)(d3 − λ) · · · (dn − λ).
is such that divides the characteristic polynomial CA (x). It can be obtained by finding
for which first i does the equation
n
X
Ai = αi−1 Ai−1 I
i=1
then
A2 = α0 I + α1 A,
then
A3 = α0 I + α1 A + α2 A2 ,
and so on. The first of these to equal zero is the minimal polynomial.
36
3 1
Example 6.1.7 Let A = .
2 2
−(−4+) − (5)x + x2 = 4 − 5x + x2 .
2 1 1
Example 6.1.8 Let A = 1 2 1 .
1 1 2
A2 = α0 I + α1 A,
α0 = −4 and α1 = 5 so that the minimal polynomial is
−(−4+) − (5)x + x2 = 4 − 5x + x2 .
37
ii. Use m(x) = 4−5x+x2 to divide x3 +2x2 +4 and we have x +7
x2 − 5x + 4 x3 + 2x2 −x +4
− x3 + 5x2 − 4x
7x2 − 5x + 4
− 7x2 + 35x − 28
30x − 24
in which case,
4I − 5A + A2 = 0
4A−1 − 5I + A = 0
4A−1 = 5I − A
1
A−1 = (5I − A)
4
Therefore,
−1 −1
3
1
A−1 = −1 3 −1
4
−1 −1 3
38
Consider −8+10x+2x2 −5x3 +x4 and divide by m(x) to have x2 −
x2 − 5x + 4 x4 − 5x3 + 2x2 + 10x −
− x4 + 5x3 − 4x2
− 2x2 + 10x −
2x2 − 10x +
Therefore,
−8 + 10x + 2x2 − 5x3 + x4 = (4 − 5x + x2 )(x2 − 2).
Thus,
−8I + 10A + 2A2 − 5A3 + A4 = (4 − 5A + A2 )(A2 − 2).
Since
4 − 5A + A2 = 0,
−8I + 10A + 2A2 − 5A3 + A4 = 0.
Now multiply by A−4 and we have
−8A−4 + 10A−3 + 2A−2 − 5A−1 + I = 0.
Solution
Since C is a diagonal blocked matrix, the characteristic polynomial of C is the product
of the characteristic polynomials of its blocked diagonals. The diagonal matrices in C
are
2 8
C1 =
0 2
and
4 2
C2 = .
1 3
39
Their respective characteristic polynomials are
χ(C1 ) = (λ − 2)2
and
χ(C2 ) = (λ − 2)(λ − 5).
Hence, the characteristic polynomial of C, the product of χ(C1 ) and χ(C2 ) is
m(C1 ) = (λ − 2)2
and
m(C2 ) = (λ − 2)(λ − 5).
Hence, minimal polynomial of C, the LCM of m(C1 ) and m(C2 ) is
40
Chapter 7
B = P −1 AP.
Remark 7.1.2 Consider the eigenvectors associated with the eigenvalues of a matrix
A to be
a b c
1 1 1
a2 , b2 , c2 .
a3 b3 c3
The required matrix P can be obtained by arranging the eigenvectors of those eigenvalues
as the columns of P so that
a b c
1 1 1
P = a2 b 2 c 2
a3 b 3 c 3
41
Example 7.1.3 Let
8 −8 −2
A = 4 −3 −2 .
3 −4 1
The eigenvalues are λ1 = 1, λ2 = 2, λ3 = 3 and the eigenvectors are
4 3 2
3 , 2 ,
1
2 1 1
the matrix
4 3 2
P = 3 2 1 ,
2 1 1
−1 1 1
P −1 = 1 0 −2
1 −2 1
1 0 0
−1
and D = P AP = 0 2 0
0 0 3
42
1 1
eigenvectors are , the matrix
−2 1
1
1 1 − 13
P = , P −1 = 3
2 1
−2 1 3 3
and
1 0 λ1 0
D = P −1 AP = = .
0 4 0 λ2
4 1
Example 7.1.6 Let A = . The eigenvalues are λ1 = 1, λ2 = 5 and the
3 2
1 1
eigenvectors are , the matrix
−3 1
1
1 1 − 14
P = , P −1 = 4
3 1
−3 1 4 4
and
1 0 λ1 0
D = P −1 AP = = .
0 5 0 λ2
Remark 7.1.7 Note that similar matrices have the same determinant, characteristic
roots, rank and characteristic polynomials.
43
Chapter 8
B = P −1 AP.
a1 b1 c1
Remark 8.1.2 Consider the eigenvectors of a matrix to be a2 , b2 , c 2
a3 b3 c3
a a a
1 2 3
the matrix P = b1 b2 b3
c1 c2 c3
8 −8 −2
Example 8.1.3 Let A = 4 −3 −2 . The eigenvalues are λ1 = 1, λ2 = 2, λ3 = 3
3 −4 1
44
4 3 2 4 3 2
and the eigenvectors are 3 , 2 , 1 the matrix P = 3 2 1 and
2 1 1 2 1 1
1 0 0
D = P −1 AP = 0
2 0
0 0 3
λ 0 0
1
Remark 8.1.4 It can be observed that D = P −1 AP = 0 λ2 0
0 0 λ3
3 1
Example 8.1.5 Let A = . The eigenvalues are λ1 = 1, λ2 = 4 and the
2 2
1
1 1 1 1 3
− 31
eigenvectors are , the matrix P = , P −1 =
2 1
−2 1 −2 1 3 3
1 0 λ 0
and D = P −1 AP = = 1 .
0 4 0 λ2
Remark 8.1.6 Note that similar matrices have the same determinant, characteristic
roots, rank and characteristic polynomials.
45
matrices such that
A = L1 U1
or
A = U2 L2 .
2 −1 3
Example 8.2.1 Let A = 1 −1 . Decompose A into
2
1 1 1
2 0 0 a b c 2 −1 3
A = L1 U1 = 1 0 f d = 1 2 −1
g 0
1 1 1 0 0 e 1 1 1
and use this decomposition to solve the system
2x − y + 3z = 7 (8.2.1)
x + 2y − z = 1 (8.2.2)
x+y+z = 3 (8.2.3)
Solution:
Solving simultaneously the equations 2a = 2, 2b = −1,
2c = 3, b +
f = 1, b + gf =
2 0 0 1 − 21 32
2, c + dg = −1, c + d + e = 1, we obtain L1 = 1 35 0 and U1 = 0 23 − 32
1 1 1 0 0 1
Note that
AX = Y ⇒ L1 U1 X = Y ⇒ U1 X = L−1 1 Y.
1
0 0
2
L−1
3 3
= − 10 5 0 .
1
− 15 − 35 1
Furthermore,
1 − 21 32 x 1
0 0 7 7
2 2
0 23 − 32 y = − 10
3 3
0 1 = − 23 .
5
1 3
0 0 1 z −5 −5 1 3 1
Hence, x = 2, y = 0, z = 1.
46
a b c 2 0 0 2 −1 3
Example 8.2.2 Also, A = U2 L2 = 0 f d 1 g 0 = 1 2 −1 .
0 0 e 1 1 1 1 1 1
5 8
− 3
6 3
Obtain and solve the appropriate simultaneous equations and we get U2 = 0 2 −1
0 0 1
2 0 0
3
and L2 = 1 2 0
1 1 1
47
2 −1 3 x 7
0 25 − 52 y = − 25 .
0 0 1 z 1
Hence, x = 2, y = 0, z = 1.
48
Chapter 9
49
Example 9.1.4 (ii) The set of all n × m matrices with entries from a field F denoted
Mn,m (F) is also a vector space if we define addition and scalar multiplication respectively
as addition of matrices and scalar multiplication of matrices. That is for A = (aij ), B =
(bij ) ∈ Mn,m (F) and λ ∈ F,
A + B = (aij + bij ) and λA = (λaij ).
Example 9.1.5 (iii) Consider a non empty set X and a field F. Let F (X) be the set
of all functions of X with values in F. It is also a vector space if addition and scalar
multiplication are defined as follow: ∀ f (x), g(x) ∈ F (X) and λ ∈ F,
f (x) + g(x) = (f + g)(x)
λf (x) = (λf )(x),
where g0 (x) = 0 ∀x ∈ X is the zero function and for any f (x), there is a −f (x) in
F (X) such that −f (x) = (−f )(x) ∀x ∈ X.
Definition 9.1.6 Let v1 , v2 , · · · , vn be vectors in a vector space V over a field F. A
vector v ∈ V is said to be a linear combination of v1 , v2 , · · · , vn if for λ1 , λ2 , · · · , λn ∈ F,
n
X
v = λ1 v1 + λ2 v2 + · · · λn vn = λi vi .
i=1
50
Example 9.1.9 Let (3, 7, −4) = λ1 (1, 2, 3) + λ2 (2, 3, 7) + λ3 (3, 5, 6). Find λ1 , λ2 , λ3 .
Definition 9.1.10 If every member of V can be expressed as a linear combination of the
set v1 , v2 , · · · , vn ∈ V, then the set is set to span or generate V. That is, v1 , v2 , · · · , vn ∈
V spans or generates V if λ1 , λ2 , · · · , λn ∈ F exist such that
n
X
v= λi vi ∀v ∈ V.
i=1
Example 9.1.14 1. The standard basis for Fn is the set {ei = (0, 0, 0, · · · , 1, 0, · · · , 0)}
where the i-th entry is 1 and 0 elsewhere.
For instance, standard basis for R3 is e1 = (1, 0, 0), e2 = (0, 1, 0) and e3 = (0, 0, 1)
2. The basis for M2×2 (F) is
1 0 0 1 0 0 0 0
, , , .
0 0 0 0 1 0 0 1
Remark 9.1.15 The dimension of a vector space is the number of elements in its basis.
Definition 9.1.16 A subset B of a vector space V over a field F is a subspace of V if
it is also a vector space over F under the same binary operations.
Definition 9.1.17 Let u, v ∈ B and α, β ∈ F. Then B is a subspace of V if and only if
αu + βv ∈ B.
Remark 9.1.18 The set span(v1 , v2 , · · · , vn ) = { ni=1 λi vi |λi ∈ F, vi ∈ V} is a sub-
P
space of V generated by {v1 , v2 , · · · , vn }. It is the intersection of all the subspaces of V
containing {v1 , v2 , · · · , vn }.
51
9.2 Some Results in Vector Space
Theorem 9.2.1 The non zero vectors {vi }ni=1 in a vector space V over a field F are
linearly dependent if one of the vectors is a linear combination of the others.
Lemma 9.2.3 Let {vi }ni=1 be non zero in a vector space V over a field F. They are
linearly dependent if and only if one of them is a linear combination of the preceding
ones.
Proposition 9.2.4 Any two basis for finitely generated vector space have the same
number of elements.
Definition 9.3.2 The set of all linear transformations from Vn (F) to Vm (F) is denoted
L(Vn (F), Vm (F)).
52
Remark 9.3.3 Let T1 , T2 ∈ L, then, T1 + T2 ∈ L. This is because for v ∈ Vn (F) and
λ∈F
(T1 + T2 )(v) = T1 (v) + T2 (v) ∈ L
(λ(T1 + T2 ))(v) = (λT1 + λT2 )(v) = λT1 (v) + λT2 (v) ∈ L
L can be seen as a set of homomorphisms of n-dimensional vector space Vn (F) to
m-dimensional space Vm (F) denoted HomF (Vn (F), Vm (F)). Any element of HomF is
called F-homomorphism.
53
(ii) Let V be a collection of functions of t. The derivative mapping D : V → V defined
by D(v) = dvdt
is a linear transformation.
Consider u, v ∈ V and α, β ∈ F.
d(αu + βv)
D(αu + βv) = (9.3.6)
dt
du dv
= α +β (9.3.7)
dt dt
= αD(u) + βD(v) (9.3.8)
(iii) Let VR be a collection of functions of t. The integral mapping I : V → R defined
1
by I(f (t)) = 0 f (t)dt is a linear transformation.
Consider u(t), v(t) ∈ V and α, β ∈ F.
Z 1
I(αu + βv) = (αu + βv)dt (9.3.9)
0
Z 1 Z 1
= αu dt + βv dt (9.3.10)
0 0
Z 1 Z 1
= α u dt + β v dt (9.3.11)
0 0
= αI(u) + βI(v) (9.3.12)
54
and
P = AT
is the transition or change of basis matrix from {ci } to {fi }. Furthermore, Q = P −1 is
the transition or change of basis matrix from {fi } to {ci }. If on the other hand,
Then,
b11 b12 · · · b1n
..
b21 · · · ···
.
B=
.. .. .. ..
. . . .
bn1 · · · · · · bnn
and
Q = P −1 = B T
is the transition or change of basis matrix from {fi } to {ci }.
Example 9.4.1 Consider that ei = {(1, 0), (0, 1)} and fi = {(2, 3), (1, 5)} are bases in
IR2 . Then, Find the transition matrix that changes the basis of V from
i. {ei } to {fi }
ii. {fi } to {ei }
Solution
i.
f1 = (2, 3) = α1 e1 + α2 e2
.
f2 = (1, 5) = β1 e1 + β2 e2
Hence,
(2, 3) = α1 (1, 0) + α2 (0, 1)
,
(1, 5) = β1 (1, 0) + β2 (0, 1)
55
from where α1 = 2, α2 = 3, β1 = 1, β2 = 5. Thus,
2 3
A=
1 5
and
2 1
P = .
3 5
ii.
5 −3
7 7
Q=
−1 2
7 7
Example 9.4.2 Consider that ei = {(1, 0, 0), (0, 1, 0), (0, 0, 1)} and fi = {(3, 3, 4), (1, 2, 5), (2, 2, 2)}
are bases in IR3 . Then, Find the transition matrix that changes the basis of V from
i. {ei } to {fi }
ii. {fi } to {ei }
Solution
i.
(3, 3, 4) = a11 (1, 0, 0) + a12 (0, 1, 0) + a13 (0, 0, 1)
(1, 2, 5) = a21 (1, 0, 0) + a22 (0, 1, 0) + a23 (0, 0, 1) .
(2, 2, 2) = a31 (1, 0, 0) + a32 (0, 1, 0) + a33 (0, 0, 1)
Hence, a11 = a12 = 3, a13 = 4, a21 = 1, a23 = 5, a22 = a31 = a32 = a33 = 2 Thus,
3 3 4
A= 1
2 5
2 2 2
and
3 1 2
AT = P = 3 2 2
4 5 2
56
ii.
3 −4 1
Q = −1 1
0
−7 11 −3
2 2 2
Example 9.4.3 Consider that ei = {(1, 0), (0, 1)} and fi = {(1, 3), (2, 5)} are bases in
IR2 . Then, Find the transition matrix that changes the basis of V from
i. {ei } to {fi }
ii. {fi } to {ei }
Example 9.4.4 Consider that ci = {(1, 1, 0), (1, 2, 3), (1, 3, 5)} and fi = {(3, 3, 4), (−4, 7, 6), (−8, 10, 7)}
are bases in IR3 . Then, Find the transition matrix that changes the basis of V from
i. {ci } to {fi }
ii. {fi } to {ci }
Theorem 9.4.5 Let P be transition matrix that changes the basis of V from {ei } to
{fi } in a vector space V. Then, any linear operator T on V is such that
[T ]f = P −1 [T ]e P,
where P −1 transition matrix that changes the basis of V from {fi } to {ei } in the vector
space.
This theorem will be illustrated with an example.
Example 9.4.6 Consider that ei = {(1, 0, 0), (0, 1, 0), (0, 0, 1)} and fi = {(3, 3, 4), (1, 2, 5), (2, 2, 2)}
are bases in IR3 and T is the linear transformation
T : IR3 −→ IR3
defined by T (x, y, z) = (2x, z, y). Show that
[T ]f = P −1 [T ]e P,
where P −1 is transition matrix that changes the basis of V from {fi } to {ei } in the
vector space.
Solution
57
From these, the matrix of coefficients is
5 −2 −7
2
−12 3
35 .
2
6 −2 −6
Its transpose is
2 0 0
[T ]e = 0 0 1 .
0 1 0
Hence,
3 −4 1 2 0 0 3 1 2 5 −12 6
P −1 [T ]e P = −1 = −2 −2 = [T ]f .
1 0 0 0 1 3 2 2 3
−7 11 −3 −7 35
2 2 2
0 1 0 4 5 2 2 2
−6
58
Example 9.4.7 Consider that ei = {(1, 0), (0, 1)} and fi = {(1, 3), (2, 5)} are bases in
IR2 and T is the linear transformation
T : IR2 −→ IR2
[T ]f = P −1 [T ]e P,
where P −1 transition matrix that changes the basis of V from {fi } to {ei } in the vector
space.
Example 9.4.8 Consider that ei = {(1, 0, 0), (0, 1, 0), (0, 0, 1)} and fi = {(1, 1, 1), (1, 1, 0), (1, 0, 0)}
are bases in IR3 and T is the linear transformation
T : IR3 −→ IR3
defined by T (x, y, z) = (2y + z, x − 4y, 3x). Show that
[T ]f = P −1 [T ]e P,
where P −1 transition matrix that changes the basis of V from {fi } to {ei } in the vector
space.
59
Chapter 10
60
Furthermore,
f (u, αv1 + βv2 ) = h(u)g(αv1 + βv2 )
= h(u)[g(αv1 ) + g(βv2 )]
= h(u)[αg(v1 ) + βg(v2 )] .
= αh(u)g(v1 ) + βh(u)g(v2 )
= αf (u, v1 ) + βf (u, v2 )
Hence, f is a bilinear form on V.
Proposition 10.1.3 Let f, g : V × V −→ IK be a bilinear form on a vector field V
over a field IK. Then, for any α, β, λ ∈ IK
i. f + g is also bilinear
ii. λf is also bilinear
Proof 10.1.4 i. We show that f + g is also bilinear.
Furthermore,
61
Furthermore,
f (X, Y ) = X T AY,
62
which can be written as a matrix form
a11 a12 · · · a1n y1
..
a21 · · · ···
. y2
x1 x2 · · · xn
.. .. .. ..
..
. . . . .
an1 · · · · · · ann n
63
Example 10.1.11 Define the bilinear form on IR2 by
f [(x1 , x2 ), (y1 , y2 ) = 2x1 y1 − 3x1 y2 + x2 y2 .
ii. In A, a11 = f [v1 , v1 ] = f [(0, 1), (0, 1)] = 1, a12 = f [v1 , v2 ] = f [(0, 1), (−1, 0)] = 0,
a21 = f [v2 , v1 ] =
f [(−1,0), (0, 1)] = 3 and a22 = f [v2 , v2 ] = f [(−1, 0), (−1, 0)] =
1 0
2. Hence, A = .
3 2
64
iii. is orthonormal if it is orthogonal and each of its column is normalized or of length
1
Definition 10.2.2 Consider a square matrix A. It is said to be orthogonally diago-
nalizable if we can find an orthogonal matrix P such that A is similar to a diagonal
matrix. That is,
D = P −1 AP = P T AP
Remark 10.2.3 If we have a matrix P̄ whose columns are the eigenvectors corre-
sponding to the eigenvalues of the symmetric matrix A, we can obtain an orthogonal
(orthonormal) matrix P whose columns are the normalized columns of P̄ , in which case,
P will diagonalize A.
3 −6 0
Example 10.2.4 Let A = −6 0 6 . Obtain an orthogonal matrix P which
0 6 −3
diagonalizes A.
Solution
Obviously, A is symmetric. Also,
1
1 − 2 −2
1
P̄ = 2 −1 2
1 1 1
and
2 −1 −2
1
P = 1 −2 2 .
3
2 2 1
Hence,
0 0 0
D = P T AP = 0 −9 0 .
0 0 9
1 −2
Example 10.2.5 Let A = . Obtain an orthogonal matrix P which diag-
−2 1
onalizes A.
65
Solution
Obviously, A is symmetric. Also,
1 −1
P̄ =
1 1
and
1 1 −1
P =√ .
2 1 1
Hence,
−1 0
D = P T AP = .
0 3
xT Ax = a11 x21 + a22 x22 + a33 x23 + 2a12 x1 x2 + 2a13 x1 x3 + 2a23 x2 x3 . (10.3.1)
To simply put,
n X
X n
xT Ax = aij xi xj . (10.3.2)
i=1 j=1
66
10.4 Canonical Forms
When we let x = P y, where
y1
y2
y=
..
.
yn
and P is the orthogonal matrix that diagonalizes A, then,
i.
xT Ax = 2x21 + 4x22 + 5x23 − 4x1 x3 .
ii.
√2 0 √1
5 5
P = 0
1 0
√1 0 − √25
5
67
and
1 0 0
P T AP = 0 4 0 .
0 0 6
Hence,
y T (P T AP )y = y12 + 4y22 + 6y32 .
68
Bibliography
[1] S.A. Ilori and O. Akinyele, Elementary Abstract and Linear Algebra, Univeristy
Press, Ibadan (1986).
69